Paper Submitted

Download as pdf or txt
Download as pdf or txt
You are on page 1of 21

Turbomachinery design by a swarm-based optimization method

coupled with a CFD solver


1∗
Enrico Ampellio1c, Francesco Bertini2b, Andrea Ferrero , Francesco Larocca1a and
Luca Vassio3c
1
Department of Mechanical and Aerospace Engineering, Politecnico di Torino, Corso Duca degli Abruzzi 24,
Torino, Italy
2
GE Avio S.r.l., Rivalta di Torino, Italy
3
Department of Electronics and Telecommunications, Politecnico di Torino, Corso Duca degli Abruzzi 24,
Torino, Italy

(Received keep as blank , Revised keep as blank , Accepted keep as blank )

Abstract. Multi-Disciplinary Optimization (MDO) is widely used to suitably handle the advanced
design in several engineering applications. Such applications are commonly simulation-based, in order
to properly capture the physics of the phenomena under study. This framework demands fast
optimization algorithms as well as trustworthy numerical analyses, and a synergic integration between
the two is required to obtain an efficient design process. In order to meet these needs, an adaptive
Computational Fluid Dynamics (CFD) solver and a fast optimization algorithm have been developed
and combined by the authors. The CFD solver is based on a high-order discontinuous Galerkin
discretization while the optimization algorithm is a high-performance version of the Artificial Bee
Colony method. In this work, they are used to address a typical aero-mechanical problem encountered
in turbomachinery design. Interesting achievements in the considered test case are illustrated,
highlighting the potential applicability of the proposed approach to other engineering problems.

Keywords: MDO, swarm intelligence, discontinuous Galerkin, turbomachinery, CFD

1. Introduction

In advanced engineering, computational tools assume an essential role when dealing with
complex design problems (Rao 2009). This field requires efficient improvement strategies,
extensive automation of the design process and includes appropriate investigations of the involved
physical phenomena. In general, a compromise between accuracy and time/cost is mandatory for
the whole design system, especially for the computational simulations adopted to describe the
physics. In this regard, a proper tuning and validation of numerical models on reference problems
must be accomplished. Furthermore, the adopted optimization strategy should efficiently satisfy
the pursued goals while respecting all the constraints. In conclusion, the optimization algorithm
and the physical description should be combined by an automatic process which leads to an
efficient design platform.

∗Corresponding author, PhD, E-mail: [email protected]


a
Professor
b
Ph.D.
c
Ph.D. Student
In this framework, several modern applications in engineering require detailed, multipurpose and
fast optimizations. The Multi-Disciplinary Optimization (MDO) approach is regularly adopted to
manage such problems (Vanderplaats 2007), returning a set of non-dominated candidate solutions,
i.e., the Pareto Front. In the present work, the MDO methodology is used to optimize a
turbomachinery test case, the well-known T106c low pressure turbine cascade. Its geometry and
experimental results are openly available (Michalek et al. 2010; Hillewaert et al. 2013). A
representative simulation-based MDO problem is then set up, estimating both performance and
mechanical properties. The objective is to minimize the aero losses while respecting some
structural constraints.
In order to address such a problem, a fully automatic MDO platform is developed. It integrates the
optimization algorithm directly with the Computational Fluid Dynamics (CFD) simulations,
without using surrogates. Although surrogates are exploited in a large class of applications (Jones
2001; Koziel and Leifsson 2013; McCullagh and Nelder 1989) their training phase is a delicate
matter (Forrester et al. 2008) and hence could not be not cost-effective in terms of Function
Evaluations (FEs) when the target functions have high dimensionality and are very complex, noisy
and multimodal. Therefore, this work exploits an efficient direct search technique explicitly
designed to limit the total number of FEs.
There are several architectures and algorithms suitable for MDO, also when each function
evaluation is costly: gradient-based methods, trust region approaches, nature-based algorithms,
adjoint techniques and many others (Martins and Lambe 2013; Vestraete and Periaux 2012; Koziel
and Yang 2011; Iollo et al. 2001). Among them, the meta-heuristic plays a predominant role when
dealing with many variables and noisy target functions, whose properties are not known a priori
(Talbi 2009). Besides the well-known evolution-based strategies, there are many recent bio-
inspired and efficient methods (Yang 2010), such as Particle Swarm Optimization (PSO) and other
swarm-based techniques (Bonabeau et al. 1999; Kennedy 2010). One of the most promising is the
Artificial Bee Colony (ABC) algorithm (Karaboga 2007). Starting from it, the authors developed a
high-performance single-objective strategy called AsBeC, which also offers the possibility to
provide a partial Pareto set for MDO problems by using the weighted sum approach. The AsBeC
algorithm was already successfully employed for real-world turbomachinery multi-objective
optimizations (Bertini et al. 2013), giving advantages in comparison to other forefront methods.
The flow field around the T106c blade is studied by means of an adaptive discontinuous Galerkin
CFD method, developed by the authors for solving the compressible Reynolds Averaged Navier-
Stokes (RANS) equations in two-dimensional domains. This solver can use high-order
discretizations in both space and time and it can manage both structured and unstructured meshes
and different physical models (Ferrero and Larocca 2013; Ferrero and Larocca 2015).
Details about the adopted CFD solver and the AsBeC algorithm are provided in Section 2 and 3
respectively, while Section 4 covers the engineering test case with the optimization settings and
results.

2. Simulation environment

2.1 Discontinuous Galerkin CFD


The blade performance is evaluated by numerically solving the equations of fluid dynamics
through a time marching approach. Navier-Stokes equations or Reynolds-Averaged Navier-Stokes
equations are integrated on general shaped two-dimensional domains. As far as the RANS
approach is concerned, the low Reynolds k-omega model proposed by Wilcox (2006) is used. This
model can give good results for turbomachinery applications in which separation phenomena have
to be taken in account (Pacciani et al. 2011). The model is implemented in a non-standard manner
following the approach of Bassi et al. (2005). The physical description is completed by the use of a
transition model based on the laminar kinetic energy concept, following the work of Pacciani et al.
(2011).
The numerical technique is based on a discontinuous Galerkin spatial discretization. The most
attractive feature of this choice is the possibility to implement robust high order schemes also in
the presence of complex geometries with unstructured meshes. Furthermore, the local nature of the
discontinuous Galerkin reconstruction makes it possible to deal with very stretched and distorted
elements, such those which are generally used for high Reynolds number flows.
Briefly, the solution inside an element is represented by means of a high order reconstruction
projected on a hierarchical modal basis. In this way, several degrees of freedom are used inside
each element. In this work a basis obtained from a tensor product of Legendre polynomials is
chosen. The basis functions are orthonormalized with the modified Gram-Schmidt procedure,
following the approach of Bassi et al.(2012). This last aspect is particularly important for
preserving the accuracy order when distorted elements are employed (Bassi et al. 2012).
Curvilinear elements are used at solid walls in order to match the accuracy of the geometrical
representation with the accuracy of the solution discretization. In particular, Serendipity mappings
are implemented for quadrilateral elements up to the fourth order and mappings up to the third
order are used for triangular elements (Onate 2009). The convective fluxes are evaluated according
to Osher and Solomon (1982) and Pandolfi (1984).
Diffusive fluxes are computed by a recovery-based scheme (Ferrero et al. 2015). For what
concerns time integration, both explicit and implicit schemes have been implemented. Steady state
simulations, like those considered in this work, are accelerated by means of a fully implicit
backward Euler scheme. In this case, the Jacobian of the system is evaluated analytically. The
resulting linear system is solved at each time step through the restarted Generalized Minimal
Residual algorithm (Saad 2003) with an Incomplete Lower Upper preconditioner.
The time step in the performed simulations is chosen according to the ramping strategy
described by Colombo (2011). An additional control is introduced in the present implementation,
in order to switch from implicit to explicit (first order) time integration when the time step size
drops below a certain threshold. This feature is particularly important during the initial transient in
which the cost of the implicit integration would not be compensated by a large time step. The
simulations were stopped when the L2 norm of the residuals drops down a fixed tolerance (10e-5).

2.2 Grid generator


The simulation environment is completed by the use of the GNU GPL mesh generation
software Gmsh (Geuzaine and Remacle 2009). The meshes employed in this work are composed
by a clustered O-type structured grid near the blade, surrounded by an unstructured grid (Fig.1).
All the elements are quadrilateral. The distance between the inlet and the leading edge is equal to
0.7 chords, while the outlet boundary is positioned one chord downstream the trailing edge.
Since the governing equations are discretized by a high-order representation, it is fundamental
to employ a sufficiently high-order description for the wall boundaries. In particular, curvilinear
elements with cubic edges are used. The high-aspect ratio elements inside the boundary layer
region can become significantly distorted when the wall edge is curved. In order to prevent the
jacobian of the mapping to become negative, the high-order optimization tool available in Gmsh is
used. In this way, the final mesh contains curvilinear edges not only at wall (Fig.2a) but also inside
the first layers near it (Fig. 2b).

Y
Z X

Fig. 1 Example of mesh

Y Y
Z X Z X

(a) Without wall optimization (b) With wall optimization


Fig. 2 Curvilinear mesh generation close to the wall (Detail of the trailing edge)

The hierarchical nature of the chosen basis was exploited to develop a p-adaptive algorithm.
This is a useful feature in the framework of automatic optimization problems; indeed, the
algorithm refines the solution without the need to change the grid. In particular, an entropy-based
sensor is used to detect the boundary layer and the wake and to increase the reconstruction order in
the involved elements. In this way it is possible to use a relatively coarse and uniform mesh in the
wake region. The order of each element can range from two to four according to the sensor
response (see Fig. 7).
3. Optimization environment

3.1 The AsBeC algorithm


The applicative field of engineering optimization is usually characterized by simulation-based
problems (Martins et al. 2005; Larocca 2008) in which computational analyses are heavily time
consuming. Some of the most widespread techniques in the single-objective optimization context
lay in the class of meta-heuristic (Glover and Kochenberger 2003). Among them, one of the
newest and most promising is the nature-inspired Artificial Bee Colony algorithm (ABC)
(Karaboga 2007), which combines the principles of Particle Swarm Optimization theory (Kennedy
and Eberhart 1995) and Differential Evolution (Price et al. 2005). If compared with other
competitive methods, ABC demonstrates high quality, robustness and flexibility for a great variety
of optimization problems (Karaboga and Akay 2009). Besides, a lot of technical dissertations, test
case applications and improvement works have been done in the recent years (Bolaji et al. 2013).
Karaboga firstly developed the Artificial Bee Colony (ABC) algorithm in 2005. The algorithm
reproduces the behaviour of a honeybee colony searching for the best nectar source into a target
area. Some bees, i.e., the employees, are each assigned to a food source and search the space near
it. Then they come back to the hive and communicate the position of the best found food sources
to other bees, i.e., the onlookers, that help the employees in the most promising regions. Nectar
sources that reveal themselves non-productive are abandoned in place of eventual new fruitful
positions, which are investigated by a bee moving in the whole target area, i.e., the scout. In the
optimization context, nectar sources represent the input configurations and their nectar amount is
the objective value to optimize; non-productive sources are those not improving for some time.
The ABC algorithm is recognized to be simple to implement, easy to be effectively parallelized
and hybridized, driven by few control parameters, flexible and robust over a wide range of
problems, according to many authors (e.g., Karaboga et al. 2014). On the other hand, its local
search and refinement skills are less efficient with respect to the global search attitude. Moreover,
it does not exploit the history of the best points found.

1: Produce random food sources in the search area


2: Assigm each food source to a different employee
3: Repeat
4: Employee Group phase: produce new configurations near their food sources updating them
whenever any improvement is found
5: Assign each onlooker to one of the food sources, depending on their quality
6: Onlooker Group phase: produce new configurations near their food sources updating them
whenever any improvement is found
7: If a food source is not improving for some time then replace it with a new random selected
configuration (scout)
8: Until requirements are met
Fig. 3 ABC/AsBeC pseudocode

Since the original version of the Artificial Bee Colony many researches on the topic were
developed. Despite the great amount of available literature on ABC variants and modifications
(Bolaji et al. 2013), to the best of our knowledge no paper underlines a performance gain even
with very few function evaluations. This aspect is of major interest in a CFD-based context like the
one herein presented, in which only some hundreds of FEs are available to limit the machine
times.
A new algorithm named Artificial super-Bee enhanced Colony (AsBeC) is then proposed to
deal with costly optimization. The AsBeC is a modification of the original ABC aimed at
improving its speed and solution accuracy for problems where function evaluations have to be
limited below 103. To accomplish these tasks, enhancements of the basic structure and
hybridizations with local interpolation are used. These techniques speed up the convergence of the
best solutions in their neighbourhood without clustering the swarm at the same time. As a result,
the local search skills of the original ABC (exploitation) is improved without worsening its global
attitude (exploration), especially during the first search phases. The improving techniques are
classified in the two groups outlined below.
• Hybridizations: super-bee concept. These modifications alter the original pseudo-
random movement of the bees, accelerating the optimization process and its accuracy
through simple local interpolation hybrids.
a. Whenever one bee does not improve its nectar source with the original pseudo-
random mutation, then it tries the opposite movement. This simple linear local
estimator is inspired to the concept of opposition based learning (Tizhoosh
2005);
b. Each bee can estimates the local and directional curvature of the objective
function, acting as a second order optimization method with partial Hessian
computation. The multi-dimensional parabola passing through three previous
positions of the bee (starting food source point, first random movement and
opposite position) is calculated and its minimum is evaluated. This local
parabolic estimator follows the basic principles of convex optimization (Boyd
2004);
c. The data knowledge about the history of the best solutions evolving in time is
used to make a prediction. A bee is then guided towards the foreseen next best
search direction, computed through a weighted average of the last directions of
improvement.
• Enhancements. These techniques do not alter the architecture of the original ABC, but
they make it work differently in order to boost the exploitation from the early
optimization phases.
a. Each group of bees have more time to evolve their nectar sources, having also
more chances to use its super-bee skills;
b. The exploitation of the best food sources is strongly privileged while the worst
ones are always penalized. This is particularly beneficial with small colonies;
c. The scout is relocated in a dynamic range that depends on the current position
of the food sources, allowing also exiting the starting boundaries and following
the colony movements.
A short pseudo-code of the ABC/AsBeC is outlined in Fig. 3. In the presented test case, the
AsBeC algorithm will include the baseline configuration as starting point into the random
generated initial food sources, in order to fasten the optimization process.

3.2 Partial pareto


The AsBeC algorithm was originally designed to deal with Single-Objective problems (SO),
nevertheless it can be successfully used to extract a partial Pareto Front, which is an approximation
of the global one, when solving Multi-Objective (MO) problems. First of all, the weighted sum
approach is adopted to transform the MO problem into a SO one (Deb 2001). Then, since the
algorithm stores all the configurations tested and their results, the information about the
contributions to the weighted sum is kept and evolved during the entire optimization process.
Therefore, an estimated Pareto front is built a posteriori by identifying the non-dominated points
from the collective hive memory. This partial Pareto is more reliable and close to convergence in
the regions presenting higher weighted sum.
Demonstrative tests on multi-objective analytical functions show that the AsBeC can extract a
consistent partial Pareto. In particular, the authors followed the same approach of Deb et al.
(2002), by using benchmark problems widely adopted in literature to test Multi Objectives Genetic
Algorithms (MOGAs). All problems in this testbed have two objective functions to be minimized,
they are unconstrained and their dimensionality, n, varies from 1 to 30. Six heterogeneous
problems over the nine described by Deb at al. (2002), namely SCH, FON, KUR, ZDT3, ZDT4
and ZDT6, have been selected for the purpose of validating the AsBeC. The total number of FEs is
chosen to be 25000, as in many other papers on the topic (e.g Toffolo and Benini 2003). The
weighted sum approach has been applied to the two functions and using unitary weights and
the corresponding results are presented in Fig. 4. Each optimization process is repeated 20 times,
due to the random nature of the algorithm, then averaged results are shown.
It is evident that the AsBeC partial Pareto is able to approximately describe the Pareto-optimal
Front in the zones presenting minimal weighted sum, which are reported in the Table 1 below.

Table 1 Minimal weighted sum regions for the adopted benchmark in the two objectives Pareto space
SCH FON KUR ZDT3 ZDT4 ZDT6
[0.002, 0.978] [-14.522, - [1.0, 0.0]
[1.0, 1.0] [0.850, -0.773] [0.250, 0.500]
[0.978, 0.002] 11.583] [0.3884, 0.8492]

The resulting AsBeC performance seems valuable in comparison with well-known genetic and
evolutionary multi objective algorithms, like NSGA, NSGA-II, SPEA, PAES and GeDEA ( Deb et
al. 2002; Toffolo and Benini 2003).
Another important test is the evaluation of the Pareto convergence as function of FEs. This
helps to understand how many function evaluations are needed to approach the optimal Pareto, at
least with few points. The convergence check for the AsBeC is performed in a 5D search space
(same dimensionality of Section 4) by using two of the selected six test functions, FON and ZDT3,
which present different types of Pareto Front: one is non-convex and continuous while the other is
convex but disconnected. Partial Pareto results are reported in Fig. 5a-c for FON Fig. 5d-f for
ZDT3, testing 5000, 600 and 250 FEs. In both cases it is clear that Pareto convergence quickly
degrades reducing function evaluations. Nevertheless, even 250 FEs seems to be sufficient to
obtain at least some configurations close to the analytical Pareto.
The results of this paragraph support the use of the AsBeC partial Pareto to interpret the
optimization results in Section 4.
(a) SCH, n=1 (b) FON, n=3

(c) KUR, n=3 (d) ZDT3, n=30

(e) ZDT4, n=10 (f) ZDT6, n=10


Fig. 4 AsBeC performace on the analytical benchmark, 25000 FEs
(a) FON, n=5, 5000 FEs (d) ZDT3, n=5, 5000 FEs

(b) FON, n=5, 600 FEs (e) ZDT3, n=5, 600 FEs

(c) FON, n=5, 250 FEs (f) ZDT3, n=5, 250 FEs
Fig. 5 AsBeC convergence check as function of FEs by using FON and ZDT3
4. Engineering application

This core section is devoted to the engineering application, from pre-processing details to
results presentation.
Airfoil design is a complex topic which still raises a lot of interest in the engineering
community. Coupling the airfoil shape optimization with high-fidelity aero analyses such as
detailed CFD is an established practice, especially for turbomachinery (e.g. Koiro 1999). In
particular, the multi-disciplinary aero-mechanical optimization of turbine blades is standard
problem addressed in industrial contexts (Schabowski et al. 2010; Bertini et al. 2013).
In this work, a renowned turbine test case, i.e. the cascade blade T106c, is chosen as the
baseline to exemplify a real-world oriented MDO, based on contemporary numerical tools. For
that cascade, experimental data on the 2D blade-to-blade plane for several working conditions are
available from the von Karman Institute for Fluid Dynamics (Michalek et al. 2010; Hillewaert et
al. 2013). These data are used in Paragraph 4.3 for validating the simulation environment. In
certain conditions this blade is characterized by an evident recirculation bubble and a strong
diffusion on the suction side, thus offering large margins of improvement. In particular, the MDO
process in this Section considers the inlet swirl angle set to 40 deg while the design value for the
T106c is 32.7 deg (see Table 3). The optimization process will then drive the airfoil shape towards
new profiles suitable for this severe off-design.
The Paragraph 4.1 briefly describes the parametric system adopted to capture the blade shape;
Paragraph 4.2 illustrates the validation of the applied tools, such as the grid convergence and
independence, the estimation of the CFD errors and the parameterization efficacy. The
optimization platform is described in Paragraph 4.3 while Paragraph 4.4 illustrates the problem
definition and settings. Finally, results are reported in Paragraph 4.5.

4.1 Airfoil parametrization


Airfoils have to be finely characterized in order to properly interpret the physical phenomena
investigated by computational fluid dynamics. Indeed, the grid generation is one of the most
delicate aspects when numerical simulations are involved. The mesh generation requires a large
number of points on the profile: they have to be enough to capture all the geometrical peculiarities
and closely spaced near to the areas with the larger curvature.
A parameterization system for the airfoil shape is then introduced. This approach enables to
draw the profile through its coordinates by means of few curves, whose definition is obtained by
defining some fundamental geometrical parameters. The best parameterization is the one with the
lower dimensionality among those capable to properly define any geometric detail. In this way the
airfoil management would be as simple as possible without losing accuracy. A good parametric
system involves only a small number of key parameters, even if a little approximation in the
geometry could be present and accepted. Consequently, the adopted parameterization must be a
balanced compromise between complexity and quality.
A suitable reference parameterization for turbine airfoils could be the one suggested by Anders
and Haarmeyer (2010). It has been adopted In the present work with few simplifications and
modifications. The most straightforward idea for a 2D section in Cartesian coordinates would be to
implement circumferences for the leading edge and the trailing edge connecting them through
polynomial lines, a third degree curve for pressure side and a fourth degree one for suction side.
Seven reference points univocally define these four drawing laws: three for the leading edge, three
for the trailing edge and one on the throat point. Four reference tangents, which define the wedge
angles for upper and lower sides near leading and trailing edges are then used as conditions to
mathematically solve the system. The airfoil curves calculated in this way are continuous and
differentiable and they have no discontinuities up to first order, hence they are in class.
Extra degrees of freedom are added by capturing the same polynomial curves using rational
Bezier Splines (BS) of the same order. Then the weights of the BS control points can later
introduce further flexibility in defining the shape of the airfoil.
The independent parameters needed to draw a parametric airfoil in the system above described
are seventeen (Table 2), assuming that the first and last weights of each BS have to be equal to 1.

Table 2 The seventeen parameters needed to univocally describe a turbine airfoil


∆x, translation of the leading edge along
Cax, axial chord 2nd pressure side BS weight
turbine x axis
∆y translation of the leading edge along
Ct, tangential chord 3rd pressure side BS weight
turbine y axis
xth, x coordinate of throat point βin, inlet blade angle 2nd suction side BS weight
yth, y coordinate of throat point Rle, leading edge radius 3rd suction side BS weight
ωin, inlet wedge angle 4th suction side BS weight
βout, outlet blade angle
Rte, trailing edge radius
ωout, outlet wedge angle

This definition is in line with other typical choices for turbomachinery (Wilson 1991; Anders
and Haarmeyer 2010).
The drawing laws are used to obtain area and minimum principal inertia moment of the airfoil
(Gauss’s formulae). Fig. 6 illustrates the T106c baseline airfoil captured in the parametric system
(thick curve). BS control points (circular see-through markers), BS polygons (thin lines) and the
centroid (black filled marker) are highlighted in the figure.

Fig. 6 Baseline airfoil parameterization


4.2 CFD analyses validation
There are several experimental data available in the literature for the T106c blade (Michalek et
al. 2010; Hillewaert et al. 2013), however the CFD solver was tested by studying the reference
configuration defined in Table 3. The purpose of these preliminary simulations is to estimate the
error produced by the CFD solver on the performance evaluation. This is fundamental to identify a
tolerance to stop the optimization process and to understand if the improvements showed by the
algorithm are meaningful or not.
First, a grid convergence and independence analysis is performed to establish the minimum
resolution required to reproduce the performances of the baseline blade with sufficient accuracy.
The refinement procedure is stopped by monitoring the convergence of the average exit losses.
This study is carry out on the baseline configuration only, then the chosen resolution level is also
adopted during the entire optimization process. The final grid contains approximately 7000
quadrilateral elements. If second order (4 degrees of freedom per equations for element)
Discontinuous Galerkin elements were used in the entire domain this discretization would be
equivalent to a finite volume grid with 28000 cells. Since the p-adaptivity algorithm was employed
some elements inside the domain used third order (9 degrees of freedom per equation for each
element) and fourth order (16 degrees of freedom per equation for each element) reconstructions,
so the total number of degrees of freedom is higher. A typical value obtained in a simulation is
approximately equal to 35000 degrees of freedom per equation. Fig. 7 shows an example of the
reconstruction order distribution inside the domain.
In Fig. 8, the wall isentropic Mach number distribution calculated by the present discretization
is compared with the experimental data. The plot reports the numerical results obtained on both the
original geometry and its parametric representation. Notice that the overall behaviour is well
described and the main differences introduced by the parametric representation are localized on the
leading edge and on the first part of the suction side. This is mostly due to the difficult fitting of the
leading edge by a circumference.
In Table 4 the total pressure loss coefficient (Eq. 1) is reported in order to compare experimental
results, CFD on the original geometry and CFD on the parametric geometry:

= (1)

where stays for stagnation pressure and subscripts 1 and 2 indicate respectively inlet and outlet
stations.
Data in Table 4 refer to a control station positioned 0.465 axial chords downstream the trailing
edge, in accord with the experimental results. Loss coefficients are computed through the spatial
average of the losses distribution reported by Hillewaert et al. (2013).
This test was used to calibrate the transition model and to estimate the CFD error.

Table 3 Reference condition for CFD validation


Isentropic exit Re number =1.85e+5 Inlet Turb. Intensity=0.9%
Isentropic exit Mach number =0.65 Inlet angle=32.7 deg
Inlet total temperature =298 K Specific heat ratio γ=1.4
Inlet total pressure =7198 Pa Gas constant R=287.1 J/kgK
Table 4 Comparison on average total pressure loss coefficient
Loss coefficient
Experiment reported by Hillewart et al. (2013) 7.8e-3
CFD original geometry 7.2e-3
CFD parametric geometry 7.0e-3

order
4
3.9
3
2.1
2

Fig. 7 Reconstruction order distribution

Exp.
CFD T106c
1 CFD Param.

0.8
M is

0.6

0.4

0.2

0 0.2 0.4 0.6 0.8 1


x/c ax

Fig. 8 Comparison on wall isentropic Mach number


4.3 Simulation based optimization platform
An automatic platform is developed for the MDO, in which the optimizer and the CFD solver
are integrated. In the specific case, a single airfoil is studied in a 2D fluid environment; however,
the same approach can be extended to multi-row 3D simulations or different applications.
Once the original geometry is captured in the parametric system, the optimization can begin.
The AsBeC MDO algorithm operates the calls to the airfoil drawing, the grid generation and the
CFD solver. These three routines are called through an evaluation function that acts like a “black-
box”: given a certain set of input variables driven by AsBeC, it returns the output objectives to the
main program.
The optimization algorithm settings are: the maximum number of function evaluations, the
number of free-to-change variables, the boundaries, the formulation of the targets and the total
number of bees. When the process finishes, if the achieved solutions are still not good enough for
the designer, the entire procedure can be repeated.
The platform exploits the parallel version of the AsBeC algorithm (see Section 3) hence it is
able to evaluate several configuration in parallel. This feature is of great advantage on multi-core
machines or clusters, since each CFD simulation can be directly associated to a different core or
node. A flow chart of the entire process is reported in Fig. 9.
The AsBeC algorithm, the airfoil drawer/creator, the evaluation function and all the related
scripts are written in GNU Octave language. The GNU GPL software Gmsh is used for the grid
generation and the CFD code is written in Fortran. The optimization is performed on a 6-cores
Intel® i7-3930k machine with GNU/Linux operating system. The authors emphasize that such a
platform can work with a personal computer and free software.

Fig. 9 Flow chart of the parallelized simulation-based optimization platform


4.4 MDO set up
Assuming that the optimization algorithm and the physical simulator have been properly tuned,
design parameters and objectives choices are the other key points for the MDO success. The target
functions must combine all the multidisciplinary traits needed to be investigated. The authors
chose two representative objectives to minimize, associated to the real-world design. They are
formulated considering the section under study as a part of the multi-row turbine frame.
Losses
Kinetic energy losses between upstream and downstream averaged quantities are defined as:
!"
!
1−
= ∙ −; =1− !" (2)
!
1−# $
Where stays for static pressure, stays for total pressure, & is the air specific heat ratio
and is the kinetic energy loss coefficient calculated at the exit station. The scaling constant
is described later. The exit total pressure value used in Eq (2) is obtained by means of mass flow
average.
These losses are directly related to aerodynamic efficiency and then they represent a good
performance index.
Dynamic constraints deviations
The 3D shaping of a turbine blade has to be mechanically compliant, so that in each section the
area (A) and minimum inertia moment (Imin) must satisfy the requirements for blade integrity and
safe operability. These two values are set by the structural analyses and are continuously calibrated
while aero-mechanical cycles are performed. For example, increasing area and inertia moment
improve the mechanical properties of the airfoil, but it could lead to weight or resonance issues.
These are related to the 3D blade and are appreciable only by dedicated and comprehensive
analyses. The 2D approach if a fast environment for airfoils optimization, but it cannot involve
directly such verifications. However, you can assume they have been performed by the 3D blade
designer, thus area and moments are already satisfying the margins for structural integrity.
Consequently, the optimization should try not to vary A and Imin from the values obtained by
capturing the baseline airfoil in the parametric system. Then, the second objective is formulated to
penalize their deviations.
Another important constraint that should be kept in mind is the mass flow rate elaborated by the
blade channel ('(). In order not to alter the working point of the turbine stage to which the blade
section belongs, '( must be kept constant. In fact, the Overall Engine Manufacturer decides the
mass flow rate throughout the turbine once the thermodynamic cycle is frozen. The mass flow rate
('() ) related to the baseline airfoil is then chosen as reference value.
In conclusion, the two previous aspects are combined in the second dimensionless Constrained
Objective (CO) that appears as a symmetrical penalization function, including three scaling
constants, , * and + :
. − .) 1234 – 1234,) '( − ')(
, = ∙- -+ *∙0 0+ +∙- - − (3)
.) 1234,) ')(
where |9| operator is the absolute value of 9 and .) , 1234,) and '() are baseline quantities.
Baseline mechanical properties are herein intended as design targets, but the objective in (3) can
be easily reformulated taking care of any specific value for A and Imin, for example returned by
detailed structural analyses.
The coefficients from to + are needed in the context of the weighted sum approach. In
fact, objectives should be written in a form such that they contribute in equal measure to the global
weighted sum, : ∙ + : ∙ , , when the weights : associated to it are unitary. For this
reason, the coefficients should be tuned in order to adjust LOSS and CO values in such a way
that they assume the same order of magnitude during the entire optimization process. It is evident
that a sensitivity on each contribute to the two target functions is needed. This sensitivity is guided
by the particular optimization problem and by its boundaries, and in the specific case leads to:
= 100; = 10; * = 10; + = 10
which makes each target function varies around 1e+0. In this way, the weights : take the role
of importance indexes associated to each objective. The weighted sum can put in evidence one or
another aspect simply by adjusting these importance indexes.
The variables selection and their boundaries are the last key point to be addressed. The decision
to investigate some variables in place of others derives from the common turbine design strategy.
Only tools that solve row interactions, like 1D meanline, 2D crosswise or multi-row CFD tools,
can change some parameters. These are the absolute positioning of the airfoil, the chord, the throat
and the exit blade angle. In addition, the throat regulates the mass flow rate, even if there are
blockage effects due to boundary layer thickness to be also considered. The throat point (xth,yth) is
then maintained constant also to strengthen CO objective. Manufacturing and material limits
impose some restrictions, especially for what regards minimum thickness and bending properties
near the trailing edge. For this reasons, trailing edge radius and exit wedge angle should be
changed cautiously.
As a result, only eight out of seventeen variables are considered for optimization. However,
since the significant problems with this airfoil are localized on the suction side, modifying the
pressure side can hardly have a positive influence. In order to reduce as much as possible the
dimensionality of the problem, the authors have finally selected the following five variables:
• βin, inlet blade angle;
• ωin, inlet wedge angle;
• 2nd suction side BS weight;
• 3rd suction side BS weight;
• 4th suction side BS weight.
During the optimization process these variables are box bounded into an hypercube centred on the
baseline values. The upper and lower boundaries of the interval are chosen according to the
designer experience and specific problem sensitivity. They must be large enough to accept the
changes necessary to achieve the goals, but sufficiently narrow to avoid geometric degeneration.
Moreover, it is important to keep in mind that row interactions and 3D phenomena are not resolved
at this level. As a consequence, the 2D optimization must fall under the assumption of small
perturbations to assure applicability.
For the specific problem boundaries assume the following values:
• Upper\Lower Bound = ±10 deg
• Upper\Lower Bound = ±5 deg
• Upper\Lower Bound* = ±0.3 [-]
• Upper\Lower Bound+ = ±1 [-]
• Upper\Lower BoundL = ±1 [-]
Considering that the CFD simulations are heavily time consuming, the optimization process
has to limit the total number of FEs. Hundreds of function evaluations with the AsBeC algorithm
are enough to approach the global solution when dealing with problems like the one under exam
(see Paragraph 3.2 and Bertini et al. 2013). The optimization is set up by using 12 bees overall,
equally divided in 6 employees and 6 onlookers, and 21 cycles for 264 CFD analyses in total. The
choice of this number of FEs is supported by the tests described in Section 3.
The weights are biased with the aim to drive the optimum search towards the zones with lower
losses, assuming:
: = 5; : 1

4.5 Results
The final optimal solution in terms of weighted sum is reported in Table 5; it is characterized by
a relative losses reduction of about 25.5% with respect to the baseline configuration. On the other
hand, its area, minimum inertia moment and mass flow differ by 1.56%, 28% and 0.83%
respectively from the reference values. Area and mass flow are close to the baseline ones, while
Imin is much higher. Increasing the minimum inertia moment could improve the mechanical
qualities of the airfoil, as it reduces local stresses and enhances blade stiffness. However, only
specific analyses as dynamic response, herein not comprised, can point out if the stiffness increase
produces side effects.

(a) Isentropic Mach number and geometry for


baseline and optimized (best weighted sum) (b) Partial Pareto front
airfoil
Fig. 10 Results
Fig. 10a shows the comparison in terms of geometries and wall isentropic Mach numbers
between the baseline airfoil and the optimized one. Fig. 11a and 11b illustrate the comparison
between the entropy fields around the two airfoils: it is clear that the entropy generation is reduced
in the second configuration. With the considered augmented incidence, the baseline airfoil is
characterized by a large recirculation on the suction side near the leading edge. In contrast, the best
weighted sum configuration avoids this separation and regulates the diffusion after the throat
station. Fig. 12a and 12b show the streamlines in the leading edge region. As a negative aspect, the
best weighted sum airfoil is excessively front loaded with respect to the baseline. To move the
configuration towards more after-loaded profiles, interaction phenomena among adjacent turbine
rows should be included in the simulation.
Besides the best weighted sum airfoil, it is possible to identify other two solutions which
correspond to the best individual goals (Table 5). The one with the best aero performance obtains a
26.5% reduction on losses but that configuration is among the worst ones for CO, while the best
for constraints is obviously the baseline.
The obtained partial Pareto Front in Fig. 10b allows identifying all the best compromise
configurations. The Pareto points are concentrated near the best weighted sum regions found by
the algorithm during the optimization process. The figure also shows the Pareto convergence,
extension and population from 132 FEs (light markers) to the final condition of 264 FEs (dark
markers). Best weighted sum regions are near (LOSS=2.7, CO=3.0) and (LOSS=3.4, CO=0.2) in
the final Pareto, while there was only a best area near (LOSS=3.2, CO=1.6) for the 132 FEs
Pareto. This last area and the one around the baseline could be close to convergence, since they are
densely populated and do not move forward from 132 to 264 Fes.

(a) Baseline airfoil (b) Airfoil with minimum weighted sum


Fig. 11 Entropy field close to the trailing edge

The post processing of the optimization results makes it possible to recognize which are the
variables that strongly influence the losses. In particular, the main benefits are related to the
increase of the inlet blade angle. In fact, the optimization process changed the inlet blade angle in
order to match the inlet flow angle for the chosen working condition. On the other hand, the
reduction of the losses leads to an increment of the minimum inertia moment due to the higher
flow turning and blade cambering. The only way to contain CO objective divergence is improving
efficiency without acting too much on β in, thus regulating BS weights and ωin. However, this leads
to minor loss reduction, since β in has a significant impact on aero performance.
Particular care has to be taken in the interpretation of the obtained results. Indeed, both CFD
simulations and experimental measurements are affected by uncertainties (see Section 4). For this
reason, the main advantage of the optimization is the understanding of the effects and the
sensitivity of the different variables on performance. This knowledge makes it possible to establish
design rules based on a solid physical background.

(a) Baseline airfoil (b) Airfoil with minimum weighted sum


Fig. 12 Streamlines around the leading edge

Table 5 Best configurations for individual goals and weighted sum


LOSS[-] CO[-] Weighted Sum[-]
LOSSBEST 2.68 3.97 2.90
COBEST 3.65 0.00 3.04
Weighted SumBEST 2.72 3.04 2.77

5. Concluding remarks

This paper describes a multidisciplinary approach for turbomachinery design problems. A CFD
solver and an optimization platform have been specifically developed for addressing the
aeromechanical design of a turbine cascade. In order to assess the reliability of the adopted tools,
the meta-heuristic optimization algorithm was tested on analytical functions and the CFD solver
was validated on experimental data.
The results obtained for the chosen test case suggest that this approach is able to efficiently
adapt the airfoil shape according to the working conditions. Above all, some guidelines for
multidisciplinary improvements are derived. Indeed, a Pareto Front approximation is provided and
analysed, obtaining the population of best candidate solutions.
The achievements presented in this paper pave the way for an effective application of such kind
of design platform to other engineering problems, even different from airfoil design. Indeed, the
key features constituting the backbone of the described approach might have a good all-purpose
applicability. Future works on the topic will cover different kinds of engineering applications in
order to assess what are the advantages and limitations when generalizing this methodology.
References

1. Anders, J.M., Haarmeyer, J. (2010). A parametric blade design system. VKI lecture series.
2. Bassi, F., Botti, L., Colombo, A., Di Pietro, D.A., Tesini, P. (2012). On the flexibility of agglomeration
based physical space discontinuous Galerkin discretizations. Journal of Computational Physics, 231(1),
45-65.
3. Bassi, F., Crivellini, A., Rebay, S., Savini, M. (2005). Discontinuous Galerkin solution of the Reynolds-
averaged Navier-Stokes and k-omega turbulence model equations. Computers and Fluids, 34, 507-540.
4. Bertini, F., Dal Mas, L., Vassio, L., Ampellio, E. (2013). Multidisciplinary optimization for gas turbines
design. AIDAA XXII conference 2013, 9-12 September, Naples.
5. Bonabeau, E., Dorigo, M., Theraulaz, G. (1999). Swarm Intelligence: From Natural to Artificial Systems.
Oxford University Press US.
6. Boyd, S. and Vandenberghe, L. (2004). Convex Optimization. Cambridge University Press.
7. Binitha, S., Siva Sathya, S. (2012). A Survey of Bio inspired Optimization Algorithms. International
Journal of Soft Computing and Engineering, 2(2), 137-151.
8. Bolaji, A., Khader, A., Al-Betar, M., Awadallah, M. (2013). Artificial Bee Colony Algorithm, its variants
and applications: a survey. Journal of Theoretical and Applied Information Technologhilley, 47(2), 434-
459.
9. Colombo, A. (2011). An agglomeration-based Discontinuous Galerkin method for compressible flows.
PhD Thesis, University of Bergamo, Italy.
10. Deb, K. (2001). Multi-Objective Optimization Using Evolutionary Algorithms. Wiley, ISBN: 978-0-471-
87339-6
11. Deb, K., Pratap, A., Agarwal, S., Meyarivan, T. (2002). A fast and elitist multi-objective genetic
algorithm: NSGA-II. Evolutionary Computation, IEEE Transactions on, 6(2),182-197
12. Ferrero, A., Larocca, F. (2013). Test cases C1.1, C1.2 and C1.6, Second International Workshop on High-
Order CFD Methods, Cologne, Germany, May 2013, http://www.dlr.de/as/hiocfd (Accessed June 2015).
13. Ferrero. A., Larocca, F. (2015). Test cases C1.2 and C1.3, Third International Workshop on High-Order
CFD Methods, Orlando, Florida, USA, January 2015, https://www.grc.nasa.gov/hiocfd/ (Accessed June
2015).
14. Ferrero, A., Larocca, F., Puppo, G. (2015) A robust and adaptive recovery-based discontinuous Galerkin
method for the numerical solution of convection-diffusion equations. Journal for Numerical Methods in
Fluids, 77(2), 63-91.
15. Forrester, A., Sobester, A., Keane, A. (2008). Engineering Design via Surrogate Modelling: A Practical
Guide. Springer, ISBN: 978-0-470-06068-1.
16. Geuzaine, C., Remacle, J.F. (2009). Gmsh: a three-dimensional finite element mesh generator with built-
in pre- and post-processing facilities. International Journal for Numerical Methods in Engineering,
79(11), 1309-1331.
17. Glover, F., Kochenberger, G.A. (2003). Handbook of Metaheuristics. Springer, International Series in
Operations Research & Management Science. ISBN 978-1-4020-7263-5.
18. Hillewaert, K., Carton de Wiart, C., Arts, T. (2013). Test cases C3.7, Second International Workshop on
High-Order CFD Methods, Cologne, May 2013, http://www.dlr.de/as/hiocfd (Accessed June 2015).
19. Iollo, A., Ferlauto, M., Zannetti, L. (2001). An Aerodynamic Optimization Method based on the Inverse
Problem Adjoint Equations, Journal of Computational Physics, 173(1), 87-115.
20. Jones, D.R. (2001), A taxonomy of global optimization methods based on response surfaces, Journal of
Global Optimization, 21(4),345-383
21. Karaboga, D. (2005). An Idea Based on Honey Bee Swarm for Numerical Optimization. Technical Report
TR06, Erciyes University, Turkey.
22. Karaboga, D. (2007). A powerful and efficient algorithm for numerical function optimization: artificial
bee colony (ABC) algorithm. Journal of Global Optimization, 39(3), 459-471.
23. Karaboga, D., Akay, B. (2009). A comparative study of Artificial Bee Colony Algorithm. Applied
Mathematics and Computations, 214(1),108-132.
24. Karaboga D., Gorkemli, B., Ozturk, C., Karaboga, N. (2014). A comprehensive survey: artificial bee
colony (ABC) algorithm and applications. Artificial Intelligence Review, 42(1), 21-57.
25. Kennedy, J., Eberhart, R. (1995). Particle Swarm Optimization. Proceedings of IEEE International
Conference on Neural Networks IV, 1942-1948.
26. Kennedy, J. (2010). Particle Swarm Optimization. Springer, Encyclopedia of Machine Learning, pp 760-
766
27. Koiro, M.J., Myers, R.A., Delaney, R.A. (1999). TADS-A CFD-Based Turbomachinery Analysis and
Design System With GUI. NASA technical report.
28. Koziel, S., Yang, X.S. (2011). Computational Optimization, Methods and Algorithms. Springer.
29. Koziel, S. and Leifsson, L. (2013). Surrogate-Based Modeling and Optimization, Applications in
Engineering. Springer, New York.
30. Larocca, F. (2008). Multiple objective optimization and inverse design of axial turbomachinery blades.
Journal of Propulsion and Power, 24(5),1093-1099.
31. McCullagh, P., Nelder, J. (1989). Generalized Linear Models, Second Edition. Chapman and Hall/CRC.
ISBN0-412-31760-5.
32. Martins, J.R.R.A., Alonso, J.J., Reuthes, J.J. (2005). A Coupled-Adjoint Sensitivity Analysis Method for
High-Fidelity Aero-Structural Design. Optimization and Engineering, 6(1), 33-6.
33. Martins, J.R.R.A., Lambe, A.B. (2013) Multidisciplinary design optimization: A Survey of architectures.
AIAA Journal, 51(9), 2049-2075.
34. Michalek, J., Monaldi, M., Arts, T. (2010). Aerodynamic performance of a very high lift low pressure
turbine airfoil (T106C) at low Reynolds and high Mach number with effect of free stream turbulence
intensity. ASME paper GT2010-22884, Glasgow, UK, 14–18 June 2010.
35. Onate, E. (2009). Structural Analysis with the Finite Element Method: Linear Statics, Volume 1: Basis
and Solids. Springer, Berlin.
36. Pacciani, R., Marconcini, M., Arnone, A., Bertini, F. (2011). An assessment of the laminar kinetic energy
concept for the prediction of high-lift, low-Reynolds number cascade flows, Proceedings of the
Institution of Mechanical Engineers Part A Journal of Power and Energy, 225, 995-1003.
37. Pandolfi, M. (1984). A contribution to the numerical prediction of unsteady flows. AIAA Journal, 22(5),
602-610.
38. Panigrahi, B.K., Shi, Y., Lim, M.H. (2011). Handbook of Swarm Intelligence. Springer.
39. Price, K., Storn, R., Lampinen, J. (2005). Differential Evolution - A Practical Approach to Global
Optimization. Springer, Berlin.
40. Rao, S.S. (2009). Engineering Optimization: Theory and Practice, John Wiley & Sons.
41. Schabowski, Z., Hodson, H., Giacche, D., Power, B., Stokes, M.R. (2010). Aeromechanical Optimisation
of a Winglet-Squealer Tip for an Axial Turbine. ASME Turbo Expo 2010, 7, 1593-1607
42. Saad, Y. (2003). Iterative Methods for Sparse Linear Systems, 2nd Edition, SIAM.
43. Talbi, E.G. (2009). Metaheuristics: from design to implementation. Wiley. ISBN 0-470-27858-7.
44. Toffolo, A., Benini, E. (2003). Genetic Diversity as an Objective in Multi-Objective Evolutionary
Algorithms. Evolutionary Computation 11(2), 151-167.
45. Tizhoosh, H. (2005). Opposition-based learning: A new scheme for machine intelligence. Proceedings of
international Conference on Computational Intelligence for Modelling, Control and Automation,
CIMCA.
46. Vanderplaats, G.N. (2007). Multidiscipline Design Optimization. Vanderplaatz R&D, Inc.
47. Vestraete, T., Periaux, J. (2012). Introduction to optimization and multidisciplinary design in Aeronautics
and Turbomachinery. VKI Lecture series.
48. Wilcox, D.C. (2006). Turbulence Modeling for CFD. 3rd edition, DCW Industries.
49. Wilson, D.G. (1991). The design of high-efficiency turbomachinery and gas turbines. MIT press,
Cambridge, Massachusetts, 5th printing.
50. Yang, X.S. (2010). Nature-Inspired Metaheuristic Algorithms – Second Edition. Luniver press.

You might also like