Sei sulla pagina 1di 9

Jing-Sheng Liu

Department of Engineering,
University of Hull,
Hull, HU6 7RX, UK
e-mail: J.S.Liu@hull.ac.uk

Geoffrey T. Parks
e-mail: gtp@eng.cam.ac.uk

P. John Clarkson
e-mail: pjc10@eng.cam.ac.uk
Department of Engineering,
University of Cambridge,
Cambridge, CB2 1PZ, UK

Optimization of Turbine Disk


Profiles by Metamorphic
Development
A novel topology/shape optimization method, Metamorphic Development, is applied to an
axisymmetric thermo-elasticity design problem. Based on solid modeling and finite element analysis, optimal profiles of minimum mass turbine disks are sought by growing and
degenerating simple initial structures subject to both response and geometric constraints.
Radial stress, axial stress, hoop stress and von Mises stress are analyzed throughout the
optimization and a constraint is imposed on von Mises stress everywhere in the disk. The
optimal structures are developed metamorphically in specified infinite design domains
using both quadrilateral and triangular axisymmetric finite elements. Comparisons are
made of the results obtained for different optimization scenarios: (a) with and without
thermal loading; (b) with and without centrifugal body forces; (c) with and without a fit
pressure on the inner surface of the hub; and (d) operating at different rotational
speeds. DOI: 10.1115/1.1467079

Introduction

High-speed rotating disks are commonly used in rotating machinery such as flywheels, gears and rotors in turbines and compressors. A major constraining factor on the development of these
components is the magnitude and distribution of stresses under
operating conditions. Common sense dictates that using a uniform
cross-section for rotating disks is very uneconomic.
Design problems of this sort fall within the category of the
structural shape optimization. However, it is generally difficult
and expensive to calculate accurate gradients of the functions for
sensitivity analysis, since rotating disks are generally axisymmetric continuum structures and their responses to changes of
topology/shape are usually non-linear and implicit functions of the
variables. Given the importance of minimizing their mass and the
potentially catastrophic effects of disk failures, the optimal design
of high-speed rotating disks to achieve low cost and high performance has long been a significant topic in industry, especially the
gas turbine industry. In an early approach, Donath 1 approximated the actual disk using a series of rings with uniform thickness. Through a comprehensive analysis of the problem, Stodola
2 suggested a hyperbolic curve for the profile of the disk. Rowe
3 investigated the effects of peripheral loading and gave charts
for the design of such disks. With the advent of the computer age,
Mathematical Programming MP methods were used to solve the
design problem, for example:
The Gradient Search GS method, see Seireg and Surana 4,
in which the disk was defined by a system of finite rings with
different thicknesses and the stresses in each ring were calculated from approximate expressions.
the Sequential Linear Programming SLP method, see Bhavikatti and Ramakrishnan 5, in which thicknesses at selected radial levels were used as design variables and the
shape of the disk cross-section was assumed to be an algebraic function of theses design variables;
the Complex method, see Luchi et al. 6, in which the disk
profile was defined by spline interpolation;
the Feasible Direction FD method, see Cheu 7, in which
the coordinates of the selected points on the disk contours
were used as design variables.
Contributed by the Design Automation Committee for publication in the JOURNAL OF MECHANICAL DESIGN. Manuscript received January 2001. Associate Editor: G. M. Fadel.

192 Vol. 124, JUNE 2002

In most of these methods, the Finite Element FE method has


been a popular analysis tool. As an alternative to FE, the boundary
element BE method based shape design sensitivity analysis
SDSA method has also successfully applied to the design problem Lee 8.
The last decade has seen a proliferation of structural topology
and shape optimization methods based on heuristic interpretations
of natural processes. The original Evolutionary Structural Optimization ESO method Xie and Steven 9,10 mimics the intuitively simple concept of evolutionary processes in nature, in
which unnecessary material is gradually removed from a structure
and the residual structure evolves towards an optimum. In its initial implementations the criterion for material removal was based
on the maximum Xie and Steven 9 or mean Hinton and Sienz
11 von Mises stress. These ideas have been investigated by a
number of researchers, who have all introduced variants on the
original ESO concept Rosko 12; Papadrakakis et al. 13; Van
Keulen and Hinton 14; Reynolds et al. 15. Other naturally
inspired heuristic methods have been inspired by botanical growth
Mattheck and Burkhardt 16 and bone growth Tanaka et al.
17.
A potential disadvantage of the ESO approach is the computational effort required in solving problems. To overcome this the
Bi-directional ESO BESO method was developed Querin et al.
18. The BESO entails adding material where the structure is
over-stressed and removing material where the structure is understressed. However, since the element removal ratio and the inclusion ratio of BESO are small, the optimization process is still
slow. Metamorphic Development MD Liu et al. 19; Liu et al.
20, which was developed independently, adopts an approach
conceptually similar to that of BESO, but uses a dynamic growth
factor, which is an adaptive function of the structural response
constraints, to control the growth and degeneration each iteration,
thus accelerating convergence.
The MD optimization procedure starts from a minimal number
of nodes and elements connecting the applied loads and support
points. The structure is then developed using finite elements that
can be of any specified sizes and a design domain may or may not
be specified. The optimum is sought through simultaneous growth
and degeneration, i.e., by adding to and removing from the structure both nodes and elements. Growth is guided to occur in areas
of high strain energy or high stress. One important feature of the
method is that it allows the introduction and re-introduction if
they have been removed in previous iterations of nodes and ele-

Copyright 2002 by ASME

Transactions of the ASME

Downloaded 05 Feb 2008 to 129.5.128.34. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

Fig. 1 Schematic diagrams of the turbine disk geometries and loadings, boundary conditions and geometric
restrictions: a disk geometry, mechanical loading, and displacement boundary conditions; b thermal
boundary conditions and geometric restriction for optimization

ments, enabling the topology design space to be robustly explored. Another original feature of MD is its use of both quadrilateral and triangular elements that are ideally suited for modeling
continua involving curved boundaries. The MD method can be
used to solve different kinds of topology/shape optimization problems, such as

The geometry parameters can best be represented in the symmetric section on the (r,z) plane using a cylindrical coordinate system, and the following notation is used to define the disk geometry:
R h inner radius of hub
Llength between inner radius of hub and outer radius of rim
L h thickness of hub
L r thickness of rim
Z h half-width of hub
Z r half-width of rim

minimizing mass subject to structural response constraints,


and
minimizing mean compliance or mean stress/strain energy
subject to mass/structural response constraints.
The effectiveness and efficiency of the MD method in solving
structural topology/shape optimization problems has encouraged
the application of the method to the rotating disk design problem.
This paper presents the MD method as applied to axisymmetric
elastic and thermo-elastic design problems. This capability is
demonstrated on the shape optimization of a turbine disk rotating
at a high speed in a specific temperature environment. The task is
to find the minimum mass profile of variable thickness of a turbine disk that connects the blades to the rotating shaft and eliminates the stress peaks in the disk caused by the thermo-mechanical
loads.
The optimization takes into account the centrifugal-type body
forces, external forces, fit pressures and the thermal loading. As
the loading and boundary conditions are shape-dependent, they
are updated in structural reanalysis to accommodate the changes.
Since the displacement field and thermal field are significantly
influenced by a shape change of the disk, a coupled steady-state
thermal-stress analysis is repeatedly performed. Axisymmetric
quadrilateral and triangular finite elements are employed. The optimization results and the metamorphic development histories are
presented for the mass minimization problem of a turbine disk
with high temperature on the tip and cooling on the lateral faces in
addition to the mechanical loading. Radial stress, axial stress,
hoop stress and von Mises stress are analyzed and the optimization is subject to a von Mises stress constraint. Optimal shapes are
obtained and compared for different loading conditions: with and
without thermal loading, with and without centrifugal body forces,
with and without fit pressure from the hub, and different rotational
speeds.
The MD method has been developed to perform both topology
and shape optimization. In this paper, only shape optimization is
considered. Shape optimization is a specific case addressed by the
presented method in that the metamorphic development is restricted to take place only on existing design surfaces and no new
design surfaces are created during the optimization.

Optimization Problem

A typical rotating disk with hub and rim is shown in Fig. 1. The
disk body material is assumed to be isotropic and homogeneous.
Journal of Mechanical Design

In this study, a typical turbine disk rotating at high speed in a


thermo-elastic environment is considered for shape optimization
using the MD method. For high-speed rotating turbine and compressor disks, mass minimization is one of the most important
design goals, and the magnitude of the stresses must be maintained below allowable limits, depending on the disk material.
Hence the task is to find a minimum mass structure with an optimal disk section profile under multiple thermo-mechanical loadings and with both temperature and displacement/rotation boundary conditions imposed. In addition to the external blades
loading, the body forces and thermal loading should be taken into
account in the shape optimization. In some cases, the inner diameter of the hub needs to be designed slightly smaller than the
diameter of the shaft, so that the disk after being heated, slotted
onto the shaft and then cooled fits tightly. This would give rise to
a fit pressure between the disk and shaft, which should be considered as another type of loading for the disk design. Therefore, the
loadings to be considered are the blade load due to high-speed
rotation, the disk centrifugal body force, the temperature distribution, and possibly the fit pressure between hub and shaft. It is
obvious that both loadings and boundary conditions are shape
dependent. However, the design task can be considered as an axisymmetric problem on account of the symmetry of the loadings
and the geometry. Generally, von Mises stress can be used as a
stress constraint criterion in shape optimization problems. A
coupled steady-state thermal-stress analysis should be performed
to reveal the distributions of the stresses and temperature and their
mutual effects.

Optimization Methodology

In this study, the optimal shapes of the structures under consideration are determined using the Metamorphic Development optimization procedure. The design problem can be stated as:
Minimize compliance f 1 T i and/or mass f 2 T i

(1a)

subject to geometric and response constraints

(1b)

The optimization method is powered with mechanisms for both


growing and degenerating a structure in order to modify its perJUNE 2002, Vol. 124 193

Downloaded 05 Feb 2008 to 129.5.128.34. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

Fig. 2 Growth cones

formance. Growth is guided to occur only in certain growth


cones, a growth cone being a local section of structural surface where high strain energy, high compliance and/or high
stress occur. The growth grammars implemented through
growth cones, as shown in Fig. 2, are based on the following
reasoning:
Adding new elements can create new load paths;
Adding structural material to these areas can disperse high
strain energy and reduce high compliance;
High stress can be decentralized by adding more elements.
Conversely, elements that carry only a small load are considered to be used inefficiently, and thus can be removed. The local
section of structural surfaces occupied by these less-stressed
elements may be called degeneration cones, and these are essentially inverse growth cones. The grammars of the growth and
degeneration cones are implemented in such a way that geometry
irregularities can not be introduced by their use.
A dynamic growth factor DGF is used to control growth and
degeneration. The general form of the DGF is shown in Fig. 3.
The values of max(i), min(i), min(i) and max(i) vary
throughout the optimization process. These specify the maximum
and minimum number of elements that may be added or removed
each iteration. They depend on factors which can vary in each
iteration such as the scale of the structure numbers of nodes and
elements, the total size of structural surface, and the availability
of structural symmetry in one or more directions. The way in
which these factors are combined is specific to each design problem and may need prior experience of using the MD method.
The DGF is used to regulate dynamically the rates of growth
and degeneration i.e., to control the sizes of the growth cones and
the parts to be removed. The DGF is related to the current structural performance, the closeness to the imposed response constraints and the calculated stress and/or strain energy. In Fig. 3,
G(T i ) represents a hybrid constraint function and is determined
by comparing structural responses such as stress or deflection
with specified limits:
m

G Ti

w g T w R T R *
j1

j j

j1

(2)

where R j (T i ) is j-th structural response, m is the number of the


response constraints, R *j is a user-specified target for j-th response, and w j is a user-selected weighting function dependent on
the importance of the particular constraint. The values of G(T i ) at
points A and B are specified as:
m

w R *
j1

194 Vol. 124, JUNE 2002

(3a)

Fig. 3 Structural dynamic growth factor versus hybrid


constraint function

BA/2

(3b)

Thus the DGF, the value of which depends through simple averaging on the values of the limits max(i), min(i) etc., is a
piecewise linear function of G(T i ). The values and factors that
determine the growth factor may vary in each iteration. Therefore,
the DGF is an adaptive function changing from one structure to
another. The adaptive nature of the DGF improves the algorithms
speed of convergence, allowing more elements to be added or
removed each iteration when the structure is far from optimal,
thus making the MD method very computationally efficient in
comparison to other heuristic methods Liu et al. 20.

The Optimization Process

Structural shape optimization is inherently an iterative procedure. The MD optimization process consists of the following
steps:
Step 1: Form an optimization model
Define all the kinematic boundary constraints, loads and material
properties that are expected under service conditions. Specify the
optimization criteria to be used and the design constraints to be
directly and explicitly imposed to optimize the structure.
Step 2: Define a finite or infinite design domain
It is not necessary to specify a finite design domain. This is especially helpful if the best domain for the design is not known
a priori. Nevertheless it is possible to impose restrictions on structural growth both in terms of location and direction, if appropriate.
Step 3: Choose appropriate finite elements
Finite elements appropriate to the design problem under consideration must be chosen. They can be linear or non-linear elements
with various shape functions. In the application detailed in this
paper, both triangular and quadrilateral axisymmetric elements are
used. The use of triangular elements produces relatively smooth
boundaries, thus enabling a relatively coarse mesh to be used
while still capturing detailed shape information Liu et al. 20.
Step 4: Generate an initial structure
The initial FE structure can be the simplest possible geometry
connecting the loads to the supports, providing that it has a nonsingular FE solution. In fact, the optimization process can start
from any degree of structural development from the simplest possible structure to a heavy ground mesh. Like other heuristic methods, MD cannot guarantee a global optimum solution, but it has
been shown to be able to successfully find the optima on standard
Transactions of the ASME

Downloaded 05 Feb 2008 to 129.5.128.34. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

test problems Liu et al. 20, and, based on our experience, in


most cases the choice of initial structure will not make much
difference to the final design obtained. Generally, only a rough
far from optimal initial design is needed. It does not matter too
much if an extremely poor initial solution is used, although the
optimization process will take longer. However, if the user has an
existing design to use as a starting point or can estimate an initial
shape that is close to the optimum, this will certainly reduce the
number of iterations needed to reach the optimum.
Step 5: Perform a structural finite element analysis
FEA is performed to find the stress/displacement distribution over
the whole structure for a strength/stiffness related optimization
problem. Based on the analysis, growth cones, the parts of the
structural surface that are heavily over-stressed, are located.
Meanwhile, heavily under-stressed regions are also located. The
information obtained in this step is used in MD to alter the existing configuration as it evolves towards an optimal design.
Step 6: Growdegenerate the structure
In this step, the structure is allowed to grow and degenerate volumetrically. Growth and degeneration take place simultaneously in
every iteration. Material is added in overloaded locations growth
cones and removed from underloaded locations. The growth rate
is controlled by the DGF, which can be both positive and negative
during the course of an optimization.

Fig. 4 Flow chart of the MD method

Step 7: Update the structure


The FE model is now modified to trace and accommodate the
changes to the structure made in step 6 and to avoid possible
singularities caused by them. This entails updating the structural
elements, nodes etc.

Fig. 5 The metamorphic development convergence history for the turbine disk design: a iteration 0 initial; b
iteration 5; c iteration 10; d iteration 15; e iteration 20; f iteration 25 optimized

Journal of Mechanical Design

JUNE 2002, Vol. 124 195

Downloaded 05 Feb 2008 to 129.5.128.34. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

Fig. 6 Contours of various stresses, temperature and heat flux distributions in the optimized turbine disk; a Mises stress;
b radial stress; c axial stress; d hoop stress; e temperature; f heat flux

Steps 5 to 7 are repeated until a suitably optimized structure has


evolved. This process will result in a homogeneous distribution of
elastic energies or stresses in the structure, indicating that the
peak stresses and/or strain energies are adequately reduced.
However, this distribution does not identify the optimal topology/

shape for the specified loads and design constraints. The optimization process continues until a minimum mass structure satisfying all the structural response constraints has been achieved or
until design limitations prevent a further increase or decrease in
the size of the structure. Convergence is deemed to have occurred

Fig. 7 Stress distributions on the non-design surface of the optimized disk structure

196 Vol. 124, JUNE 2002

Transactions of the ASME

Downloaded 05 Feb 2008 to 129.5.128.34. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

Fig. 8 Stress distributions on the design surface of the optimized disk structure

if the performance of the structure cannot be improved for a specified number of consecutive iterations. In this manner, the initial
design evolves into an optimized design, which is a minimum
mass structure satisfying all the constraints.
The optimization algorithm employs a hierarchical structure.
First, objective f 1 (T i ) the compliance is minimized and objective f 2 (T i ) the mass is disregarded until the structural response
constraints are satisfied. Then f 2 (T i ) is minimized subject to constraints on the structural response. In the first stage, a positive
growth factor is used and more elements are added than removed
each iteration. Conversely, in the next stage, a negative growth
factor is used and more elements are removed than added each
iteration. Depending on the current structural performance, these
two optimization schemes may be adopted alternately. If a well
developed structure is used as a starting design that satisfies the
constraints, the first optimization stage may not be necessary. A
flow chart for the optimization procedure is given in Fig. 4. The
procedure used does not require gradient-based sensitivity analysis, as the surface strain-energy/stress distribution is the sole parameter considered. Although there is no mathematical proof that
this hierarchical approach and other similar approaches e.g.,
Querin et al. 18 will result in convergence to the global optimum, practical experience indicates that they do in fact usually
succeed in doing so Liu et al. 20.
Since this method avoids difficult formulation and programming of complicated relationships between various elements and
shape variables, commercial FE packages can be used as analyzers. The optimization algorithm is used as a design tool in conjunction with commercial FE codes and ABAQUS software 21
is currently used to perform the required FEA.

Examples

5.1 Specifications. The design problem has been outlined in


section 2. Because of symmetry, only one half of the disk (z
0) and the symmetric portion (r0) about the symmetry axis
(r0) of the cross-section of the disk needs to be modeled. Axisymmetric boundary conditions are imposed. The geometric parameters for the disk, as shown in Fig. 1(a), are R h 20 mm, L
100 mm, L h 15 mm, L r 10 mm, Z h 40 mm and Z r
40 mm.
To make the design model as realistic as possible, four different
types of loading are considered:
Centrifugal loading: The disk is assumed to be rotating at a
constant speed 20000 rpm, 18000 rpm, or
16000 rpm. The distributed throughout the disk body
centrifugal load varies in proportion to the distance from the
Journal of Mechanical Design

rotation axis r0, as


pr2r,
where is the material density.

(4)

Blade loading: There is a uniformly distributed load, p


200 MPa, acting radially outwards on the circumferential
surface of the tip. This load simulates the forces induced by
the blades.
Thermal loading: The thermal boundary conditions are
shown in Fig. 1(b). The operating temperature of the disk is
varies from 20C at the hub to 600C at the tip. The constant
high temperature T r 600C at r the outer radius of the
rim is caused by the hot gas deriving from the blades. A
temperature of T h 20C is maintained at h the inner radius of the hub. The condition T/ z0 is imposed on the
symmetric boundary 0 (z0). The disk is cooled by air
flowing past the rotor. A convective boundary condition,
T/ nh(TT ), is imposed on the disk surface s where
cooling is assumed to occur, with the convective heat transfer
coefficient h100 W/m2 K and the ambient temperature T
20C.
Fit pressure loading: A uniformly distributed load resulting
from the disk and shaft fit pressure, q40 MPa, acts radially
outwards from the axis of rotation inwards on the circumferential surface of the hub.
The Poissons ratio, Youngs modulus, density, thermal conductivity and thermal expansion coefficient of the disk material used
are taken to be 0.3, 180.36 GPa, 7706 kg/m3 , 47 W/mK and
1.17105 /K respectively. The von Mises stresses are used as the
stress design criterion. The upper limit set on the maximum allowed von Mises stress is 1.10 GPa over the whole region. A
coupled steady-state thermal-stress analysis is performed in which
two different fields are active. The elements used have both temperature and displacement degrees of freedom.
5.2 Optimization Results. First, a general case is considered for optimization, in which centrifugal ( 20000 rpm),
blade and thermal loadings are included. This general case establishes a basic example for the purpose of comparison with other
loading cases. The disk cross-section used as the starting design
for the optimization is shown in Fig. 5(a). This initial design is
simply a flat plate disk connecting the hub and rim, and the von
Mises stress in it is as high as 3.18 GPa, which violates the stress
constraint by over 270%. The disk can have varying thickness
between hub and rim, and only the disk surface s needs to be
reshaped during the optimization. Engineering design considerations specify geometric parameters Z h and Z r , which represent
JUNE 2002, Vol. 124 197

Downloaded 05 Feb 2008 to 129.5.128.34. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

Fig. 9 Optimal shapes for different loading cases: a bladescentrifugal 20000 rpmthermal loading; b blades
centrifugal 20000 rpm loading; c blades loading only; d bladescentrifugal 18000 rpmthermal loading; e
bladescentrifugal 16000 rpmthermal loading; f bladescentrifugal 18000 rpmthermalfit pressure loading

the half-widths of the hub and rim respectively. However, L h and


L r can vary during the optimization, but they are not allowed to be
less than 5 mm i.e., L h L h min5 mm and L r L r min5 mm.
Another geometric constraint imposed is that material cannot be
removed if it is located within a cone surface z1153r and z
59r see Fig. 1(b).
The optimization takes just 25 iterations to converge. This takes
only 15 minutes to run on a Sun Ultra1 workstation running Solaris 2.6. The history of the optimization is shown in Fig. 5. The
mass of the final structure is 4.036 kg. The maximum Mises
stress, which is 1.10 GPa, occurs at the center of the inner surface
of the hub the bottom left corner of the symmetric model which
is not part of the design surface. From Fig. 6 it can be seen that
the stress constraint is satisfied throughout the structure. Therefore, an optimal turbine disk with a minimum mass and satisfying
the imposed constraints has been achieved for a practical operating environment with a variety of loadings.
Four types of stress, namely radial, axial, hoop and von Mises
stress, are analyzed throughout the optimization. To illustrate the
detailed stress distributions on all the surfaces of the optimized
disk, some specific points are identified in Fig. 10(a) and the
variations of these stresses along these surfaces are shown in Figs.
7 and 8.
Figure 7 shows the changes in the four stresses along most of
198 Vol. 124, JUNE 2002

the non-design surfaces from points a to d, via b and c. From


points a to b, the inner surface of the hub, all the stresses rise from
a value near zero at point a until their turning values at point b,
with the exception of the radial stress that shows almost no
change. The stresses turning at point b is caused by the geometry
change there. The Mises and hoop stresses reach their maximum
values at this point.
Along the symmetric boundary of the disk points b to c, all
the stresses show a gentle, smooth fluctuation. Approaching point
c, however, the hoop and axial stresses exhibit a rapid decrease
and a sudden change at point c. A kink in the Mises stress curve
is also observed at point c. The sudden, steep changes in the
stresses around point c are mainly due to the corner geometry and
the action of the blades loading.
From point c to point d, which is the outer surface of the rim
where the blades loading acts, all the stresses are basically held at
a lower level with little variation.
The stress variations along the disk surface from point d to
point a, via e, f and g are shown in Fig. 8. From points d to e, all
the stresses rise and fall within a limited range of amplitudes.
From points e to f, however, the Mises and radial stresses exhibit
greater fluctuations and a highest radial stress of 1.1 GPa is observed. These cyclic stress variations are mainly due to the slight
geometric discontinuities in the outer surface at transitions beTransactions of the ASME

Downloaded 05 Feb 2008 to 129.5.128.34. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

Case Dblades loading plus disk centrifugal loading (


18000 rpm) plus thermal loading;
Case Eblades loading plus disk centrifugal loading (
16000 rpm) plus thermal loading;
Case Fblades loading plus disk centrifugal loading (
18000 rpm) plus thermal loading plus fit pressure loading.
To make the results comparable, the same mesh is used in all
these models. The optimal shapes for loading cases B and C, D
and E, and F and D are compared respectively with that for general case A in Fig. 10. Starting from the same initial structure, as
shown in Fig. 5(a), all these optimizations are subject to the same
response and geometric constraints, but, of course, with different
loading conditions. By plotting these optimal surface profiles together, the effects of the body force, rotation speed, fit pressure
and thermal loading on the final results can be clearly identified.
These comparisons suggest that it could be essential to include all
these loadings when optimizing turbine disks operating in a
thermo-mechanical environment because of the significant effects
each has. It is shown, for instance in Fig. 10(a), that the optimal
disk profile under thermo-mechanical loading requires a thicker
disk than under the mechanical loading only. Table 1 summarizes
all the loading cases in terms of the number of iterations needed
for convergence and the final masses of the optimized structures.

Fig. 10 Comparisons of the optimal shapes for different loading cases: a comparison of optimal shapes for different loadings; b comparison of optimal shapes for different rotational
speeds; c comparison of optimal shapes with and without a fit
pressure

tween triangular and quadrilateral elements. Although quadrilateral and triangular elements can approximate a smooth surface,
they cannot model it perfectly. In this region, hoop and axial
stresses also oscillate but at a relatively lower level. From points f
to g, all four stresses decrease and vanish at point g. Although the
stresses are low in this region, further removal of material is prevented due to the geometric design constraints a cone surface, see
Fig. 1(b). Along the line from points g to a, the disk surface is
basically stress-free.
If we refer to the loading case above as case A, Fig. 9 shows the
optimized structures obtained under five alternative loading cases
B, C, D, E and F respectively:
Case Bblades loading plus disk centrifugal loading (
20000 rpm);
Case Cblades loading only;

Table 1 A summary of the results for all the loading cases


considered
Loading cases

Iterations for convergence


25
25
25
19
18
26
Final mass kg
4.036 3.493 3.120 3.381 3.378 3.465

Journal of Mechanical Design

Conclusions

In this paper the problem of optimizing the design of turbine


disks is tackled by using the MD method, which is a systematic
expansion and contraction approach for adding and removing materials in a structure. A FE-based fully automatic axisymmetric
shape optimization capability is presented.
The design problem is to minimize the volume of the disks
under thermo-mechanical loading with a constraint imposed on
the von Mises stresses over the whole structure. The optimization
seeks to eliminate the stress peaks caused by various loadings.
The disks are subject to thermal loading caused by temperature
distribution, blade loading and centrifugal forces caused by highspeed rotation, and a fit pressure caused by the assembly process.
The loading cases and boundary conditions are shape dependent
and therefore are updated in structural reanalysis. The final structure is developed metamorphically in a specified infinite design
domain. In general, the growth and degeneration are based on the
current objective and constraint functions of the optimization
problem under consideration. The results reveal that the minimum
mass profiles for the thermo-mechanical loading cases exhibit
thicker flanges than those for the mechanical loading case alone,
and similarly for cases with and without body forces. These indicate the significant influence of these factors and the importance
of including them within the loading case when optimizing turbine disks in anger.
The results also show that, by using the MD method, shape
optimization of axisymmetric continuum structures with complicated contour shapes and loading conditions can be achieved with
relatively little computational effort. A marked reduction in any
stress concentrations present is sought and minimum mass structures have been obtained. Both quadrilateral and triangular elements are used to build the structures. The employment of nonrectangular elements makes the surfaces of the resulting structures
relatively smooth. Since in the optimization considered here the
growth and degeneration occur only on the design surfaces, there
are no interior geometric discontinuities in any form. More generally, when MD is used for topology optimization the formulation used for the addition and removal of material through the
growth and degeneration cones automatically avoids the introduction of interior geometry discontinuities, such as checkerboarding.
The linking with commercially available analysis packages enables the method to utilize their versatile capabilities effectively
when dealing with complicated structures. The results demonstrate the success of the method in finding realistic solutions to the
practical design problem under consideration.
JUNE 2002, Vol. 124 199

Downloaded 05 Feb 2008 to 129.5.128.34. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

References
1 Donath, M., 1912, Die Berechnung rotierender Scheiben und Ringe, Berlin.
2 Stodola, A., translated by Loewenstein, L.C., 1927, Steam and Gas Turbines,
6th edition, McGraw-Hill Book Company, New York.
3 Rowe, J.H., 1957, How to Find the Thickness of a Constant Stress Disk,
Prod. Eng. N.Y., 28, pp. 211215.
4 Seireg, A., and Surana, K.S., 1970, Optimum Design of Rotating Disks,
ASME J. Eng. Ind., , pp. 110.
5 Bhavikatti, S.S., and Ramakrishnan, C.V., 1980, Optimum Shape Design of
Rotating Disks, Comp. & Struct., 11, pp. 397 401.
6 Luchi, M.L., Poggialini, A., and Persiani, F., 1980, An Interactive Optimization Procedure Applied to the Design of Gas Turbine Discs, Comp. & Struct.,
11, pp. 629 637.
7 Cheu, T.-C., 1990, Procedures for Shape Optimization of Gas Turbine
Disks, Comp. & Struct., 34, pp. 1 4.
8 Lee, B.Y., 1996, Consideration of Body Forces in Axisymmetric Design Sensitivity Analysis Using the BEM, Comp. & Struct., 61, pp. 587596.
9 Xie, Y.M., and Steven, G.P., 1993, A Simple Evolutionary Procedure for
Structural Optimization, Comp. Struct., 49, pp. 885 896.
10 Xie, Y.M., and Steven, G.P., 1997, Evolutionary Structural Optimization,
Springer-Verlag, Berlin.
11 Hinton, E., and Sienz, J., 1995, Fully Stressed Topological Design of Structures Using an Evolutionary Procedure, Eng. Comput., 12, pp. 229244.
12 Rosko, P., 1995, Three-dimensional Topology Design of Structures Using
Crystal Models, Comp. & Struct., 55, pp. 10771083.
13 Papadrakakis, M., Tsompanakis, Y., Hinton, E., and Sienz, J., 1996, Advanced Solution Methods in Topology Optimization and Shape Sensitivity
Analysis, Eng. Comput., 13, pp. 5790.

200 Vol. 124, JUNE 2002

14 Van Keulen, F., and Hinton, E., 1996, Topology Design of Plate and Shell
Structures Using the Hard Kill Method, Advances in Structural Engineering
Optimization. Edinburgh: Civil-Comp Press, pp. 167176.
15 Reynolds, D., McConnachie, J., Bettess, P., Christie, W.C., and Bull, J.W.,
1999, Reverse AdaptivityA New Evolutionary Tool for Structural Optimization, Int. J. Numer. Methods Eng., 45, pp. 529552.
16 Mattheck, C., and Burkhardt, S., 1990, A New Method of Structural Shape
Optimization Based on Biological Growth, Int. J. Fatigue, 12, pp. 185190.
17 Tanaka, M., Adachi, T., and Tomita, Y., 1995, Optimum Design of Lattice
Continuum Material Suggested by Mechanical Adaptation Model of Cancellous Bone, Proc. 1st World Congress of Structural & Multidisciplinary Optimization held in Goslar, Germany, pp. 185192.
18 Querin, O.M., Steven, G.P., and Xie, Y.M., 1998, Evolutionary Structural
Optimization Using a Bidirectional Algorithm, Eng. Comput., 15, pp. 1031
1048.
19 Liu, J.-S., Parks, G.T., and Clarkson, P.J., 1999, Can a Structure Grow Towards an Optimum Topology Layout?Metamorphic Development: A New
Topology Optimization Method, Proceedings of the 3rd World Congress of
Structural and Multidisciplinary Optimization WCSMO-3 held in Buffalo,
USA, on CD-ROM.
20 Liu, J.-S., Parks, G.T., and Clarkson, P.J., 2000, Metamorphic Development:
A New Topology Optimization Method for Continuum Structures. Struct.
Multidisc. Optim., 20, pp. 288 300.
21 Hibbitt, Karlsson & Sorenson, Inc., 1998, ABAQUS Manual, Version 5.7-5.
Pawtucket RI: Hibbitt, Karlsson & Sorenson, Inc.

Transactions of the ASME

Downloaded 05 Feb 2008 to 129.5.128.34. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

Potrebbero piacerti anche