You are on page 1of 15

Engineering Applications of Computational Fluid Mechanics Vol. 8, No. 3, pp.

407421 (2014)

MULTI-REGIME SHAPE OPTIMIZATION OF FAN VANES FOR


ENERGY CONVERSION EFFICIENCY USING CFD, 3D OPTICAL
SCANNING AND PARAMETERIZATION
Zoran Milas, Damir Vuina* and Ivo Marini-Kragi
FESB-Department of Electrical Engineering, Mechanical Engineering and Naval Architecture, University
of Split, R. Boskovica 32, Split, Croatia
* E-Mail: vucina@fesb.hr (Corresponding Author)

ABSTRACT: An enhanced reverse engineering procedure was developed for roof fan re-design. An original
numerical workflow for robust shape optimization based on maximum energy conversion efficiency was developed.
It operates using a sample of multiple operating regimes coupled with CFD simulations. The initial shape solution
was originally obtained in point cloud form by optical 3D scanning and subsequent B-spline based parameterization
of shape. The CFD simulation of the scanned shape using 3D RANS based software was shown to agree very well
with the measured features, experimentally obtained in our lab with the actual initial-shape fan. By manipulating the
control points of parametric curves, the developed evolutionary optimization workflow was subsequently able to
create shape-optimized vanes. This original procedure was applied to cases of constant-thickness and profiled single
curvature vanes, both for single-regime and robust multi-point operating conditions. The corresponding increase in
efficiency gained by our computational procedure was correlated with respective velocity and pressure distributions
and suppression of flow separation. The novel numerical procedure developed here therefore provides a numerical
framework for generic object geometry to re-shape itself autonomously. The change in shape ensures maximum
energy conversion efficiency for a given composition of operating regimes. The gain in efficiency with optimized
vane shapes proves to be significant in the wide range of flow rates around the best efficiency point.
Keywords:

1.

CFD, centrifugal fan, shape optimization, performance, efficiency

downstream during the course of calculation,


which requires a proper exchange of flow
variables at the interfaces.
The flow in turbomachines is unsteady due to the
interaction between the rotating and stationary
parts, e.g. between the impeller and the diffuser or
the scrolled housing of centrifugal turbomachines.
Transient modeling at the interface between the
impeller zone and stationary zones (upstreamdownstream) accounts for physical sliding
between these zones. This results in unsteady
flow models in which the velocity and pressure
oscillations are resolved during one impeller
revolution in very short time steps.
Circumferential periodicity can usually not be
assumed and hence the entire impeller must be
modeled. Due to excessive computational time,
this is inappropriate in the design phase (Brost et
al., 2002). The frozen rotor approach is a
compromise whereby the impeller is fixed in a
certain angular position with respect to the
stationary part. The resulting steady flow model is
much simpler for calculation. The prediction of
turbomachine performance using the frozen rotor
simulation is typically fairly good when compared
with transient simulations (Gugau, 2002).

INTRODUCTION

Fans make a significant class of machines


consuming energy if their usage of electric energy
is considered on the large scale. Despite the fact
of usually being small in size or unit power, the
respective energy consumption of these machines
is of growing interest because of their multitude.
The efficiency of these lightweight turbomachines
is getting into the focus since there is a large
margin for respective improvements and
consequently potential for energy consumption
reductions.
In the last three decades, CFD modeling of
turbomachinery flows has evolved from quasi 3D
potential flow models to fully 3D viscous models
(Keck and Sick, 2008; Menter et al., 2004).
The development of computer resources and
efficient numerical codes allowed for the
integration
of
stationary
and
rotating
turbomachine components into a common
computational domain. This has eliminated the
need for imposing boundary conditions at the
interfaces between turbomachine components.
The respective influence of the flow in each of the
components can freely propagate upstream and
Received: 19 Jan. 2014; Revised: 2 Apr. 2014; Accepted: 2 May. 2014
407

Engineering Applications of Computational Fluid Mechanics Vol. 8, No. 3 (2014)

However, the quality of the prediction


deteriorates at flow rates far away from the design
or best efficiency flow rates. The fixed position of
the impeller allows that any non-uniformity at the
impeller outlet is carried away too deep inside the
stationary zone in comparison with experimental
evidence.
Preliminary design or sizing of the turbomachines
is traditionally based on 1D flow analysis and
statistical correlations between dimensionless
parameters that are derived from previous
experimental evidence on high efficient designs
(Bois, 2005; Braembussche, 2005). CFD
calculations have largely replaced elaborate
experiments because they can model details of
very complex geometries.
Numerical optimization using non-gradient,
gradient and especially evolutionary algorithms
has opened entirely new horizons for the
engineering designers, (Arora, 1989; Rao, 1996;
Deb, 2000; Goldberg, 1989; Papadrakakis et al.,
2004; Derakhshan et al., 2013). Adapting
methods
and
algorithms
known
from
computational geometry and employing them to
parametrically represent geometric shapes of
objects has provided the missing link between
CFD and numerical optimization, (Hardee et al.,
1999; Dai et al., 2005; Rogers and Adams, 1976;
Bohm et al., 1984; Farin, 1993). These three
basic ingredients, numerical optimization
algorithms, parametric geometry representation
and CFD simulation have been coupled in this
paper to provide full-scale shape optimization of
roof fan vanes.
In complex flow simulations, soft computing
approaches including surrogate models for the
evaluation of excellence and constraints based on
simulation models are applied frequently along
with heuristic optimizers (Taormina et al., 2012;
Zhang et al., 2009; Cheng et al., 2005). These
methods can sometimes replace deterministic
evaluation
based
on
simulations
with
corresponding response provided by advanced
numerical approximation methods. Once trained
using corresponding data, methods such as neural
networks can deliver adequate approximation
results very fast. On the other hand, soft
optimizers such as genetic algorithms or particle
swarm methods operate consistently without
gradient information and without getting trapped
in local optima, however demanding more
computational resources. There are numerous
examples of such approaches in different
engineering disciplines.
Traditionally, vanes have been designed for a
predefined nominal operating point. Nevertheless,

the approach developed here has provided for


multi-regime shape optimization as the optimizer
engages the CFD simulator over a range of
regimes. These were generated as a random
sample corresponding to the predicted distribution
of operating regimes of the roof fan. This
approach results in the optimum design for the
overall operation of the fan rather than peak
optimum performance at the nominal point, which
is usually accompanied by a sharp decrease of
performance beyond the nominal point.
The objective was to provide a complete generic
re-engineering procedure. CAD models typically
consist of many interconnected partitioned
geometric entities with corresponding parameters,
relationships and constraints, which are not
appropriate for shape optimization. We have
therefore based our approach on the integral
parametric geometric model to provide a compact
and scalable overall geometric set for the shape
optimization model.
There are several novel aspects of the developed
procedure. The first novel aspect is the shape of
the vane which is modeled using B-spline
surfaces allowing the vane to assume virtually
any shape in 2D or 3D space by manipulating the
corresponding parametric form. The second novel
aspect is the fact that by virtue of the developed
workflow, the vane has the capacity to re-shape
itself autonomously for maximum energy
conversion efficiency. Re-shaping is achieved by
means of an evolutionary optimizer, numerically
coupled with a CFD simulator. The latter is set-up
as a numerical service provided to our
computational workflow by an external node
delivered by independent software. The automatic
re-shaping of generic vanes for maximum energy
conversion efficiency is based on multi-point
operating regimes sampled to represent the
overall life-cycle operation of the fan. Another
novel aspect is the fact that the vane is modeled as
a completely generic shape and its initial
geometry is provided for by optical 3D scanning
of an existing vane for increased numerical
efficiency due to an adequate starting solution.
The response of the CFD simulator and the CFD
model used in the workflow were validated by
carrying out extensive laboratory tests.
Numerical shape optimization of vanes is a rather
complex undertaking as it involves many
challenges. Computational modeling of the
respective geometry is one of the difficulties as it
needs to be capable of modeling global and local
variations of shape by using only a modest-size
data-set of shape parameters, since otherwise the
dimensionality of the subsequent optimization
408

Engineering Applications of Computational Fluid Mechanics Vol. 8, No. 3 (2014)

space is very high. CFD models are another major


concern, as appropriate turbulence models and
domain discretization exhibit a major influence on
the simulation results. Numerical integration in
the form of a workflow is another important asset
as it needs to encapsulate both process flows with
synchronization of processes and the necessary
data-mining. The data transfers within the
workflow include results for all candidate
designs, in particular pressure distributions and
velocities in selected regions of the domain. On
the input side of the CFD, the current shapes of
the candidate designs need to be communicated to
the simulators (data-burying). The definition of
the excellence criteria is also crucial. As the fan
operates across a range of operating regimes,
single-point shape optimization is of limited
value.
More
realistic
optimization
is
accomplished by robust optimization for a given
distribution of the operating regimes. Moreover,
the design variables may include additional
parameters beyond those which control the shape
of the vanes. For example, the rotational velocity
of the fan may be a controllable parameter for all
regimes under the conditions of prescribed flow
rates and output pressure.
The roadmap of the paper includes the following
elements:
- basics of the background physics and CFD
models
- basics of numerical modeling of generic shape
- key elements of the optimization procedures
developed in the paper
- presentation of optimization/simulation cases
and discussion of results
2. ROOF
FAN
PERFORMANCE

DESIGN

motor and impeller, such that it doesn't obstruct


the outflow from the impeller. This allows the fan
outflow to be dispersed horizontally above the
roof. Vertical outflow from the roof fan is another
viable solution. A more complex housing is
required in order to divert the radial outflow from
the impeller into the upward swirl-like flow. In
both cases the far field flow conditions are
axisymmetrical.

Fig. 1 Roof fan.

The most indicative fan parameters are the


increase of the total pressure in fan pt and the
fan efficiency . Roof fans belong to the class of
exhaust fans without pressure duct. The outlet
dynamic pressure pd o v o2 / 2 for this class of
fans is ignored according to the rules of AMCA,
Eurovent. Correspondingly the expression for
pt reduces to:
pt

po (v o dAo ) / V ( pi 2 v i )(v i dAi ) / V

Ao

(1)

Ai

where p o and p i are the static pressures, v o and

v i the absolute velocities at the fan outlet Ao

AND

and fan inlet section Ai respectively,

the air

density, and V the volume flow rate.


The fan overall efficiency is Vpt / P , where

Roof fans make a significant part of the fan


market, approximately 30% of fans for nonresidential ventilation are roof fans (Radgen et al.,
2008). By unit power they belong to the group of
low power fans, 2 kW of installed power is an
upper limit for most roof fan types. Their overall
efficiency (accounting for the electric motor
losses) rapidly increases with fan size-power. For
most below- kilowatt- range power fans, the peak
overall efficiency ranges between 0.3-0.5.
A specific feature of roof fan design is that the
scroll housing is no longer required. Roof fans
can expel the waste air into the ambient directly
from the impeller, as shown in Fig 1. No volutelike housing is necessary which is reduced to
being merely a weather shield. It can be a simple
axisymmetrical cap (cowl) above the electric

P is the fan power equal to the electric motor net


output. For the purpose of CFD-based prediction
of efficiency , the fan power P is calculated
as:
P = M i M fd Pv Pm
(2)
M i is the torque of the aerodynamic (hydraulic)
forces acting on the impeller interior surfaces.
M fd is the torque of the shear stress (friction)
forces on the impeller hub and shroud outer
surface. The last two terms, Pv and Pm , account
for the volumetric and mechanical losses of
power respectively. Due to modifications to the
roof fan that was investigated in the experimental
409

Engineering Applications of Computational Fluid Mechanics Vol. 8, No. 3 (2014)

part of this work, both terms only exhibit minor


impact on the overall efficiency.
The torque of the aerodynamic forces is
calculated by integrating pressure and shear stress
along the walls of the fan impeller according to:



M i M fd (r n ) y pdA (r w ) y dA (3)

where n is the normal vector of the impeller wall

surface A , r is the corresponding radius vector

and w is the wall shear stress. Subscript y


indicates the components in the direction of y
coordinate axis coinciding with the axis of
rotation.
CFD prediction of the fan performance has to be
validated by comparing CFD results with
experimental ones. This requires that the fan
geometry and boundary conditions in CFD
analysis are the same as in the experiment. The
elements of geometry required for the simulation
of the centrifugal roof fan impeller with a
horizontal outlet are axisymmetrical hub, shroud
and electric motor with external rotor casing.
Together with the vane, the geometry of impeller
is fully defined as shown in Fig. 2. The fan
geometry data are given in Table 1, and they refer
to the impeller design of the standard roof fan
with backward inclined flat vanes of constant
thickness
For the purpose of more comprehensive
laboratory testing, the standard roof fan was
slightly modified. The supporting structure of the
fan was removed in order to provide an unobstructed outflow from the impeller and easy
access for the velocity probe to the impeller outlet
(Fig. 3).
The electric motor of the impeller with an
external rotor was attached to the suspension plate
as in the original design. The suspension plate
affects the flow around the impeller.
Consequently, the flow in the clearance between
the impeller hub and suspension plate had to be
modeled in CFD simulation.
The sealing of the gap between the impeller and
the intake pipe was significantly improved by
using the comb type labyrinth seal. As a result,
the sealing gap flow only had minor impact on the
fan performance, and hence the gap flow was not
modeled in CFD analysis. The intake pipe has the
same diameter as the impeller eye.
The measurements of the velocity close to the
impeller inlet allowed for the control of proper
inlet boundary conditions to be applied in
impeller inlet allowed for the control of proper
inlet boundary conditions to be applied in CFD

Fig. 2 Roof fan impeller.


Table 1 Impeller geometry data.
Impeller diameter (outlet)

D (mm)

325

Impeller eye

Do (mm)

235

External rotor
(el. motor)

der (mm)

138

Impeller width

b (mm)

75

Nr. of vanes

14

Vane outlet angle


Vane thickness

(O)

45

t (mm)

1,5

Fig. 3 Roof fan laboratory test stand, 1-impeller, 2suspension plate, 3-labyrinth seal, 4-intake
pipe, 5-bellmouth with inlet screen.

analysis. A rather uniform inlet velocity


distribution was measured during experiments.
Fig. 3 illustrates the fan characteristics
determined by the measurements in the University
laboratory of fluid mechanics. Angular velocity of

410

Engineering Applications of Computational Fluid Mechanics Vol. 8, No. 3 (2014)

the fan impeller was constant throughout the


measurements =103 s-1. The impeller Reynolds
number ( D 2 / ) was about 7E+5, where is
the viscosity of air.
The pressure at the fan inlet and from the
(pressure) velocity probes was measured using the
Digitron 2080P differential pressure transducers
with the measuring range of 2500 Pa. The
pressure differences were mostly below 100 Pa.
Accounting for the factory calibration of the
transducers, the maximum uncertainty of the
measured pressure was about 1 Pa with 95%
confidence.
The fan flow rate was determined and calculated
by means of the velocity profiles that were
measured in the intake pipe cross section. The air
velocity was measured using the two point Pitot
tube and cross flow probe from Airflow
Instruments and Turboinstitute respectively. The
uncertainty of the calculated flow rates was less
than 25 m3/h. The fan efficiency was determined
based on the net output of the fan electric motor.
The net input electric power was measured by two
meter method using OEL 0120 Watt-meters from
Iskra Inc. It was corrected for motor losses using
the motor characteristic as specified by the motor
supplier Marvent-Systemair. The uncertainty of
the calculated fan efficiency was 1%.
The measured pressure and flow rate at the best
efficiency point (BEP) of the experimental fan
were 80 Pa and 1000 m3/h respectively. These
values are assumed to be the design ones for the
optimization procedure in Section 4.

restrains the selection to the class of two equation


models of turbulent viscosity: k and k
models or their hybrid SST model. In numerical
terms, robust 2-eq. turbulence models moderately
contribute to the expansion of the system of
equations governing the flow.
The continuity and momentum equations, as well
as those for the transport of turbulence kinetic
energy k and its dissipation , can be cast into
the general transport equation. For the stationary
flow of incompressible fluid the general transport
equation is:

( v j
) S
x j
x j

(4)

where is the Reynolds averaged transport


variable, the diffusion coefficient, S the

v j the Cartesian velocity


components ( j =1, 2, 3). For the k turbulence
source-sink term and

model , , S are specified as follows:


1, v j ,k ,

0, E , t / k , t /

v j
p

S 0, E
( E
) f i , Pk , (C 1 Pk C 2 )

x
k
i
j
i

in

which

pE

is

the

effective

pressure

( pE = p 2 / 3 k ) , Pk the production of k , E
the effective viscosity consisting of molecular
and turbulent viscosity t .

k , ,

C 2 are the model constants.

the

f i stands for

Coriolis
force:
i j x j j j xi 2eijk j v k in the rotating

3. CFD MODELING OF FAN FLOW AND


VALIDATION

centrifugal

C 1 and

and

reference frame. v k is the relative velocity,

Numerical modeling of the impeller flow


generally requires a 3D non-stationary viscous
model in order to correctly capture the physics of
the flow for a wide range of flow rates.
The development of boundary layer along the
impeller vanes can be accompanied by the
separation of the leading edge and followed by
reattachment. The rotation and curved vane
passages contribute to the secondary flows
associated with secondary losses.
At off-design conditions the recirculation zones
are large and they affect the impeller flow.
The fan flow is highly turbulent, and the
modeling of turbulence significantly affects the
numerical prediction of separation and swirl flow
as well. In principle, this requires more advanced
turbulence models, e.g. the (direct) Reynolds
stress model. Its numerical complexity usually

the angular velocity components and eijk the


permutation symbol.
The second order accurate scheme is used for the
spatial discretisation of convective fluxes in the
finite volume formulation of Eq. (4).
The computational domain required for the CFD
analysis consists of the intake channel, impeller,
and outflow zone as shown in Fig. 4. The
suspension plate which is part of the outflow zone
geometry also has to be modeled since it
influences fluid flow.
The frozen rotor approach is applied to the fan
impeller analysis. Flow inside any of the vane
channels is the same regardless of the impeller
angular position due to the axisymmetric far-field
conditions. The impeller flow can be analyzed in
a single vane channel or impeller segment
411

Engineering Applications of Computational Fluid Mechanics Vol. 8, No. 3 (2014)

defined by means of the velocity scale u * based


on turbulence kinetic energy u * C1/ 4 k 1/ 2 .
The grid characteristics are presented in Table 2
and Fig. 5. The optimization described in section
4 is accompanied by many variations in the fan
geometry. In order to keep the meshing
computational effort relatively low, the impeller
is meshed using tetrahedral cells.
Close to the vanes, the impeller grid is refined by
applying prism layers. The structure of the grid
along the periodic boundaries was identical.
Therefore no interpolation was required for
matching the values between the periodic
boundaries.

Fig. 4 Computational domain: a) 3D view, b)


impeller segment.

embedding a single vane, Fig.4.


The stationary flow model can thus be applied to
the segment of the computational domain.
The treatment of boundary conditions for the fan
flow, i.e. along the outlet of the computational
domain is of particular interest. Boundary
conditions must provide for free development of
flow downstream of the impeller outlet (Zhang et
al., 1996; Moradnia et al., 2014) and upstream of
the impeller as well. The computational domain is
therefore extended downstream of the impeller
exit by an additional outflow zone. The constant
pressure boundary conditions are most
appropriate for the flow at the exit from the
outflow zone. The periodical boundary conditions
are imposed in the circumferential direction. An
additional inflow zone upstream of the fan
impeller represents a part of the laboratory intake
pipe. The fan flow rate is prescribed at the inlet
cross section located eight pipe diameters away
from the impeller. The turbulence kinetic energy
and dissipation at the inlet are based on the
turbulence intensity (estimated 5%) and eddy
length scale (equal to the intake pipe diameter)
respectively.
Simulations of the flow were executed using
commercial CFD software (ANSYS). Initially,
SST and k - model were used for modeling

Fig. 5 Grid characteristics.


Table 2 Grid characteristics.
Sub-domain
Intake pipe
Impeller
Outflow zone

turbulence. The results with standard k - model


agreed better with the experimental ones. In all
simulations further on the standard k - model
with a scalable wall function for smooth walls
was used. Scalable wall functions overcome the

Grid type
structured,
hexahedral cells
unstructured,
tetrahedral cells
unstructured,
tetrahedral cells

Cell
number

Reference
frame

1944

stationary

51 908

rotating

23449

stationary

The dependence of the solution with regard to the


grid resolution is tested. Fig. 6 shows that gridindependent solution for the fan efficiency and
pressure is achieved using a grid with N 75000
cells.
The convergence criteria for the solution of the
discretized equations were defined as maximum
residual of 10-4.
Numerical results for the fan pressure capacity
performance follow the slope of the experimental
characteristic as shown in Fig. 7. At the design

problem associated with low y values of the


near-wall grid. This occurs either with
excessively fine grids or with moderately refined
grids in the area close to the separation point.
The y value of the grid points closest to the wall
can no longer be in the log law region as
presumed by wall functions. Scalable wall

flow rate of V = 1000 m3/h fan pressure as


predicted by CFD
agrees well with the experimental value. There is
a minor under-prediction of the fan pressure (less

functions limit the y values above the lower


edge of the log layer ( y 11 ). The y is
412

Engineering Applications of Computational Fluid Mechanics Vol. 8, No. 3 (2014)

than 5%) for partial flows. The standard deviation


of CFD results relative to the experimental
pressure curve is 10%. This is mainly due to
overshooting of the CFD prediction at overflow
rates.
The CFD prediction of fan efficiency deviates
from the experiment almost in the same way as
for the fan pressure. The standard deviation of the
CFD predicted values with respect to
experimental curve is about 10%.

Adams, 1976; Bohm et al., 1984; Farin, 1993).


The authors of this paper have also studied a
number of parameterizations (Vucina et al.,
2012). The approach applied in this paper is to
use 2D parametric entities to represent shape, in
particular B-spline curves.
A B-spline curve of degree d using (n+1) 2D
control points Q is defined as
(5)
x(t ) N (t ) Q , t 0,1
n

i 0

i ,d

1 , ti t ti 1
N i ,0 (t )
,0i nd
0 , otherwise

(6)

t
t
t ti
N i , j 1 (t ) i j 1
N i 1, j 1 (t ) , 1 j d , 0 i n d j
ti j ti
ti j 1 ti 1

N i , j (t )

where N are the basis functions defined


recursively using a non-decreasing sequence of
scalars- knots ti:
0 ,0i d

i d

ti
, d 1 i n
n 1 d

1 , n 1 i n d 1

The B-spline curve possesses the property of local


control since Ni,j(t) have non-zero values only in
the interval [ti, ti+j+1], hence the curve is formed
locally by a few adjacent control points in Q only.
The numerical procedure in this paper starts by
generating B-spline curves that represent the
given existing shape of the blades accordingly,
which provides the initial solution for the
optimizer. The implementation of this step was
carried out numerically by fitting B-spline curves
according to the following procedure. A straightforward method of fitting a B-spline curve to a
point cloud which was originally obtained by
optical scanning was used in this paper. The
(m+1) data points P are assumed to be ordered
with increasing sample times sequence sk
according to the parameter values according to
tk (sk s0 ) / (sm s0 ) .
The curve fitting procedure results in determining
the control points Q such that the least square
error

Fig. 6 Impact of grid resolution on CFD prediction of


fan pressure pt and efficiency at
m3/h.

(7)

V =1000

Fig. 7 Comparison of experimental fan performance


and CFD prediction for initial design with flat
vanes.

4. PARAMETERIZATION OF FAN VANE


AND OPTIMIZATION

1 m n
E (Q) N j ,d (tk ) Q j Pk
2 k 0 j 0

(8)

is minimized. The minimum of this quadratic


error function is obtained from the corresponding
necessary conditions which results in a linear
system of equations,

The objective of this paper is enhanced reverse


engineering, whereby the term enhanced refers to
shape optimization of the vanes. Optimum design
refers to changing the shape of some object such
that certain excellence criteria are maximized
within given constraints (Arora, 1989; Rao,
1996). The idea is to use the vanes of the existing
fan as the initial geometric solution for the
process of evolutionary shape optimization, where
different shape representation can be applied
(Hardee et al., 1999; Dai et al., 2005; Rogers and

a
k 0 j 0

a Q j aki Pk 0 , i 0, n

ki kj

k 0

rc

N c ,d ( t r )

(9)

Two different approaches were applied to


represent the geometry of the blade. The first
approach is using a B-spline curve to define the
corresponding shape of a constant-thickness
blade. The second approach is engaging yet
413

Engineering Applications of Computational Fluid Mechanics Vol. 8, No. 3 (2014)

another B-spline function to represent the


thickness distribution (profile) of a more general
blade with arbitrary shape.
This paper considers 2D shape optimization of the
vanes. Generally, the vane surfaces are double
curvature surfaces (two radii of curvature).
However, due to technological limitations, such
types of vanes are more difficult to fabricate than
single curvature vanes. This paper therefore
initially considers single curvature vanes. The
vane geometry is represented by the single Bspline curve for the case of constant vane
thickness, although it is later also parameterized
to allow variable vane thickness profiles.
The paper considers both flat (fixed-thickness)
and profiled vanes as these are interesting from
the industrial application point of view. As of the
optimization point of view in this paper, the
model and the algorithms are the same, with the
only difference being that profiled vanes
introduce more shape variables and require nonpenetration constraints for the vane faces
implemented via move limits of the respective
variables.
Genetic algorithms (Deb, 2000; Goldberg, 1989;
Papadrakakis et al., 2004) were applied as multiobjective numerical optimizers (MOGA) with
SOBOL initialization. Using a multi-objective
optimization algorithm (we have used MOGA-II
algorithm) changes the way in which the
respective fitness value is assigned to an
individual within the population. Instead of
directly using the objective and constraints values
for calculating the fitness, MOGA applies the
formulation based on Pareto optimality.
The optimization variables are the positions of
respective B-spline control points, and the
corresponding outputs are the pressure difference,
efficiency and residuals. The optimization
objective is maximum efficiency. The ranges of
optimization variables were constrained such that
for any current shape of vanes, they physically
remain inside the modeled segment of impeller,
i.e. do not penetrate adjacent periodic segments.
The total fan pressure pt has also been
constrained such that it must amount to be greater
than 70 Pa and lower than 90 Pa at the flow rate
of 1000 m3/h. Since the total fan pressure for
some geometry changes with the flow rate, this
constraint must be applied only at the nominal
rate and ignored for other flow rates. In addition,
it has been noticed that for some highly curved
and wavy vanes, CFD computation didnt achieve
convergence, so those results had to be discarded.
With that in mind, another constraint is the

residual which was restricted to stay below


0.0001.
The shape optimization variables are denoted as
Li as shown in Fig. 8. The variables were hence
the distances of the control points from the x-z
plane. The point L4 (central point of spline closest
to the axis) was fixed such that it is not present in
workflow. This was necessary in order to improve
the numerical efficiency of shape optimization.
Otherwise, equal (in shape) but translated (as
rigid bodies) vanes would be considered different
candidate solutions by the optimizer, thus
deteriorating the convergence properties of shape
optimization. Only the deformation modes
(degrees of freedom) should be allowed for the
vanes and the rigid-body motion modes must be
prevented from taking place during shape
optimization.

Fig. 8 Parameterization of vane.

Beside the already mentioned input variables,


another one designated as C was added. This
input variable enabled the cutting of the leading
edge of the vane as shown in Fig. 8. Changing
that variable changed the length of vane.
Shape optimization is applied only to the vanes of
fan. The rest of geometry stays unchanged, i.e. the
hub and the shroud remain the same as in the
initial design. While the shape of the vane was
changed, the shape of computational domain also
remained the same. This naturally restricted the
range of change of candidate vane shapes such
that stayed within the computational domain.
Changing the vane shape without changing the
computational domain can be a severe restriction,

414

Engineering Applications of Computational Fluid Mechanics Vol. 8, No. 3 (2014)

since it limits the vane between two periodic


boundaries. Numerical tests on the same problem
with the modification as stated next were
conducted to investigate this possible restriction.
The respective modification was that the
computational domain was forced to change in
such a way that the vane is always located at the
center between two periodic boundaries. In order
to do that, the periodic boundaries where
geometrically modeled as B-spline surfaces that
change together with the vane. These test
optimizations yielded the same results as the one
with fixed periodic boundaries. This result
empirically confirms that changing the vane shape
without changing the computational domain
shape is not a severe restriction in the case
presented in this paper.
In order to terminate the optimization procedure,
several convergence criteria in shape optimization
were applied. The optimization process converged
when there was no improvement in the efficiency
during a few generations. Typically for case 1, the
procedure converged after 1000 iterations, as
shown on Fig. 9. If the procedure was continued
only negligible improvements of the energy
efficiency where obtained.

Fig. 9 Fan efficiency () convergence


optimization, Nd- generation.

2. Parameterization of the 3D scanned point


cloud into computational geometry entities, in
particular B-splines, the control points of which
represent the variables for the optimizer, the
vector x. Optionally, deriving 2D shape
representation for 2D cases.
3. Definition and generation of the numerical
sample of operating regimes for the fan,
representative of the future actual distribution.
4. Definition of the excellence criteria and
objective functions for the optimizer such that the
fitness functions for the candidate designs can be
evaluated, in particular the efficiency of energy
conversion, base on (1)-(2). This leads towards
the ultimate objective of this paper, re-engineered
fan vanes for maximized energy conversion
efficiency under the circumstances of multiregime operation.
5. Definition of the optimization constraints, for
example bounds for control points mobility.
6. Launching the genetic algorithms- based
optimizer with corresponding operators and
parameter values, including selection operators,
cross-over operators and probabilities, mutation,
elitism, fitness scaling, etc.
7. Linking an array of CFD simulators to provide
values of excellence and constraints for all
candidate designs represented by corresponding
B-splines curves, across the numerical sample of
flow regimes.
8. Iteratively shape-optimizing the vanes within
the numerical cycle embedding the optimizer,
shape modeler and CFD simulators.

during

Several authors have approached the problem of


optimum design of pumps and fans (Derakhshan
et al., 2013) using different sets of design
variables and procedures. Nevertheless, the
method proposed here is highly generic as it
encompasses very flexible modeling of shape,
initial shapes obtained by scanning existing fans,
and robust multi-regime optimization.
The following is the overall procedure developed
in this paper:
1. 3D optical scanning of the existing rotor to
provide the point cloud representing the initial
shape solution.

Fig. 10 Shape optimization variables, vanes with


constant thickness, (MODEFRONTIER).

415

Engineering Applications of Computational Fluid Mechanics Vol. 8, No. 3 (2014)

Since the paper considers 2D shape optimization,


there is a vector of shape variables (L_i and C) as
shown in Fig. 10 and the following is the list of
entities in the figure:
Q flow rate
O_residual - output variable from ANSYS,
residual
O_DP - output variable from ANSYS, fan
pressure pt
O_ef - output variable from ANSYS, fan
efficiency
C_residual - constraint of residual
C_DP - constraint of pressure
B_ef - optimization goal, maximum efficiency
DOE- generator of initial population designs
MOGA-II- multi-objective genetic- algorithm
based optimizer
ANSYSWB74- CFD simulator
The paper goes beyond applying optimization and
CFD codes, as particular codes are not of any
significance here. The actual contributions of the
paper are in fully generic modeling of vane shape
by using parametric surfaces, in particular Bsplines to model vanes rather than some physical
parameters of vanes as it is usually done.
In the basic vane shape optimization approaches
that have been used in other papers only a few
physical features of the vane were varied as shape
optimization variables. Many authors use for
example inlet and outlet vane angles, curvatures
and locations of vane edges as shape variables.
Such variables, while physically sound and
intuitively reasonable, provide very limited shape
modeling capacity, hence the generic geometric
modeling developed here is superior in terms of
modeling freedom.
Use of fully generic modeling of vane shape
enables the optimizer to generate almost arbitrary
geometric shapes and evaluate those using
instances of CFD simulation software operating
on submitted candidate shapes. The developed
generic geometry modeling is adjustable to
particular local requirements. It can range from
basic global shape to high fidelity in local shape
phenomena with corresponding sizes of
parametric sets and resulting dimensionalities of
optimization space.
In order to accomplish this procedure, middleware software was needed for data mining and
coordination of the applications in the numerical
workflow. In a generic numerical workflow, any
optimizers and CFD codes can operate jointly.
Beyond commercial off-the-shelf software,
proprietary and in-house codes have also been
used as well. In-house simulators implementing
the described turbulent model and physical model

equations can be plugged in as the simulation


node in the workflow in Fig. 10.
With more advanced models, the multi-objective
model generally also includes further terms as
objective functions such as the pressure difference
(which may simultaneously be both a constraint
and an excellence criterion), generated noise (both
in absolute terms and respective distribution),
physical dimensions, etc. It is obvious that these
individual excellence criteria are not concurrent.
They can generally be modeled (i) based on the
weighted sum approach with the a-priori
compromise formulation, (ii) using the Pareto
approach and a-posteriori decision-making.
Both the cases of flat (fixed-thickness) and
profiled vanes were considered. In terms of
operating conditions, single-point (nominal)
operation and multi-point (robust) optimization
scenarios were implemented.
In the real world, fans are used not only for one
flow rate, but within some range close to its bestefficiency flow rate. Multi-point optimization is
developed here to provide the best vane shapes
that will be more energy efficient during the
course of real life application of the fan. In order
to specify such operational flow rate range, one
can use statistical distributions defined by mean
flow rate and corresponding standard deviations.
The particular distribution used in the paper is the
normal distribution with mean flow value of 1000
m3/h and with standard deviation of 150 m3/h.
During the optimization process, each candidate
design was evaluated in 10 simulations using
different flow rates. Each design was evaluated
using flow rate values which were elements of a
numerical sample generated randomly based on
the defined normal distribution. In some of the
optimization cases considered the vanes thickness
distribution was another design degree of
freedom. The workflow in that case needs to
accommodate additional shape variables.
Table 3 Optimization cases and specifics.

416

Index

Vane shape

Optimized

Case 1

flat, const.
thickness

Case 2

curved, const.
thickness

Case 3

curved, const
thickness

Case 4

curved, variable
thickness

Constraints
design
flow V =1000 m3/h
optimized for

V =1000 m3/h
optimized for range
of flows
optimized for

V =1000 m3/h

Engineering Applications of Computational Fluid Mechanics Vol. 8, No. 3 (2014)

The vane of the optimized fan (case 2) shown in


Fig. 12 is almost flat close to the leading edge and
gradually changes into being curved towards the
impeller outlet. The vane outlet angle is smaller
(380) compared with the flat vane design (450).
The effect of vane shape optimization on the flow
pattern is of interest. Particularly those aspects of
the flow related to the losses in vane passages.
Fig. 13 presents the wall shear stress distribution
on both sides of the impeller vane for the initial
case 1 and optimized shape, case 2. The analysis
reveals the areas of low shear stresses that may
potentially be accompanied by the separation of
flow.

5. RESULTS OF OPTIMIZATION
In order to distinguish among the results of
optimization the following indexing is applied in
Table 3 and related to the individual shape
optimization scenarios for maximum energy
conversion efficiency:
The case 1 is not interesting in itself, but it
provides the nominal reference values for
benchmarking with optimized vanes. It also
serves the purpose of validation of the CFD
model based on actual experimental results from
our lab.
Pressure and efficiency characteristics of the fan
with vanes optimized for single flow rate (case 2)
are compared with the initial design (case 1) in
Fig. 11. The fan pressure performance curve is
much steeper and the respective efficiency
decreases rapidly at overflow. The flow rate at
best efficiency point (BEP) is slightly lower than
the design flow rate (1000 m3/h).
Fan efficiency close to BEP flow rate is higher.
An increase in peak efficiency (almost 10%) is
much larger than case 1 CFD overshooting. This
implies that the optimization has actually
provided a better design of the impeller vanes.

Fig. 13 Wall shear stress contours along vane surface:


a) case 1- flat vane, b) case 2-, press.- vane
pressure side, suct.- vane suction side, LEvane leading edge, TE vane trailing edge.

Therefore a detailed analysis of the velocity field


in various cross sections of the vane passage was
carried out at the design flow rate. The cross
sections were arranged as planes normal to the
axis of rotation and located at distance y from
the impeller hub. Fig. 14a shows that separation
occurs at the vane suction side close to the shroud
outlet (y =0,75 b2) for the initial flat vane (case
1). There is a large recirculation zone near the
vane passage outlet. For the optimized vane shape
(case 2), the separation is almost completely
suppressed in this area as shown in Fig. 14b.

Fig. 11 Comparison of fan performance for initial flat


vane design (case 1) and optimized vanes (case
2, case 3), design/mean flow rate
m3/h.

=1000

Fig. 14 Velocity vector plot around flat vane (case 1)


and optimized vane (case 2) in x-z plane
normal to axis of rotation, located at distance
y =0,75 b2 from impeller hub.

Fig. 12 Flat vane (case 1) and optimized vane (case 2).


417

Engineering Applications of Computational Fluid Mechanics Vol. 8, No. 3 (2014)

The suppression of separation contributes to the


higher efficiency obtained by the optimized vane
shape (case 2). No separation close to the vane
leading edge can be identified for both vane
shapes.
The pressure distribution along the vane suction
side in Fig. 15 has changed locally, but the
streamwise pressure gradients in the area of
suppressed separation is the same. There is no
area of the minimum pressure stretching along the
vane leading edge as in case 1 and this indicates
that better inflow to the optimized vane is
achieved.
Fig. 17 Velocity vector plot around optimized vane
(case 3),y =0,75 b2.

By allowing the variable thickness distribution of


vane, the highest efficiency was obtained for the
vane shape (case 4) in Fig. 18. The thickness of
the vane has been increased in the mid-section but
has remained almost the same at the leadingtrailing edge. The inner part of the vane profile is
wedge-like with respect to the inflow, while the
outer part is tapered. The vane outlet angle is 330
being the smallest one among those for the
optimized vane shapes.

Fig. 15 Pressure distribution on vane suction side: a)


case 1, b) case 2.

The optimization of the vane shape for the range


of flow rates centered at the mean value of 1000
m3/h (case 3) results in negligibly lower peak
efficiency compared to case 2 in Fig. 11. The top
of the efficiencycapacity curve is more rounded
and the BEP flow rate is the same as for the flat
vane impeller.
The curvature of the vane (case 3) changes
gradually from the leading to the trailing edge as
shown in Fig. 16. The vane outlet angle is
somewhat smaller than for the vane optimized for
single value of flow rate (case 2). The velocity
pattern in the cross section close to the shroud
(Fig. 17) is almost the same as in Fig. 14b and
there is no indication of separation.

Fig. 18 Constant thickness vane (case 2) and profiled


vane (case 4, variable thickness vane)
optimized for design flow rate of V =1000
m3/h.

The profiled vane (case 4) provides the highest


fan efficiency not only at the BEP (=55%) but
almost in the entire range of flow rates as shown
in Fig. 19. The BEP flow rate is the same as for
case 2 and the corresponding fan pressure of 95
Pa is higher than any of BEP pressure values of
the optimized designs.
The overall structure of the novel numerical
procedure developed in this paper is presented
integrally in Fig. 20. The blade is modeled as a
general geometric entity by applying B-spline

Fig. 16 Flat vane (case 1) and vane optimized for multi


values of flow rates (case 3).
418

Engineering Applications of Computational Fluid Mechanics Vol. 8, No. 3 (2014)

surfaces and can assume any shape. The genesis


of new physical designs of blade shapes is
accomplished by the evolutionary optimizer,
taking into account the given geometric feasibility
conditions (e.g. no self-intersecting curves or
surfaces). The corresponding genotypes are
decoded into phenotypes representing 3D shapes
via parametric surfaces and fed into the overall
physical model of the blade and rotor geometry.
Subsequently, the respective geometry is
submitted for CFD simulation for a given sample
of flow regimes. Such a workflow provides for
robust shape optimization. Numerical efficiency
is achieved by starting from an existing shape
design acquired by optical 3D scanning,
partitioning
and
parameterization
into
mathematical surfaces.

composition of regimes, leading to the


respectively optimal shape for the overall
lifecycle-based energy conversion efficiency.
6. CONCLUSIONS
The proposed approach develops a numerical
system which successfully implements shape
optimization of fan vanes combined with CFD
simulation and parameterization of 3D scanned
vanes as the initial shape solution. The procedure
developed in the paper proves that it is indeed
possible to have a numerical procedure that
autonomously re-engineers the shape of the
vanes.
It does this for maximum energy
conversion efficiency, and for the assumed multiregime operation distribution.
The procedure develops a path towards
customization and individualization of energy
conversion devices based on optimization for the
particular operating environment.
The numerical optimization code and CFD code
have been coupled for the purpose of modifying
the roof fan impeller shape according to the given
excellence criteria. The CFD code used for the fan
flow analysis has been validated by experiment.
Fan characteristics predicted by the CFD code
were compared with those obtained by laboratory
experiments in the test facility set up for that
purpose. Very good agreement between the CFD
prediction and experiment is demonstrated.
The SST turbulence model with automatic wall
functions has not provided better agreement with
experimental results compared to the k model
with scalable wall functions. Both turbulence
models over-predict the fan pressure and fan
efficiency at overflow.
The optimized fans have a higher peak efficiency
( = 5-10%) compared to the initial design with
flat vanes. The increase in efficiency is achieved
with optimized vanes of curved 2D shape. This
increase is consistent with the suppression of
separation of flow in vane passages that was
noticed in the velocity field analysis.
The outlet angle of the optimized vanes is smaller
in comparison with the initial flat vane design.
However, the vane inlet angle remained almost
unchanged. Trimming of the vane leading edge
was allowed but only minor displacement of the
leading edge was noticed as the result of
optimization. The meridional section of the
impeller was kept unchanged during optimization
as well as the number of impeller vanes.
The most promising optimized vane shape is the
one obtained for the given range of flow rates. It
displays almost constant curvature and this

Fig. 19 Effect of profiled vane (case 4) on fan


performance (case 2: fan optimized with
constant vane thickness).

Fig. 20 Shape re-engineering workflow for optimized


energy conversion.

The excellence criterion thereby is the maximum


efficiency of energy conversion at the blade for
multiple randomly sampled flow regimes. Hence
the workflow in Fig. 20 will let the blade re-shape
itself towards optimal geometry for the given
419

Engineering Applications of Computational Fluid Mechanics Vol. 8, No. 3 (2014)

modification of the flat vane design can be easily


implemented in fan fabrication.
The peak efficiency of the optimized fans still
leaves room for further improvements. Based on
the promising results of the current study, the
planed future work will include full 3D stationary
regime simulations. It will include additional
degrees of freedom including a variable number
of vanes and controllable fan rotational speed.
This will be implemented using the concept of
robust optimization for a selected distribution of
operating regimes.

12. Goldberg D E (1989). Genetic Algorithms


in Search, Optimization and Machine
Learning. Addison Wesley.
13. Gugau M (2002). Transient impeller volute
interaction in a centrifugal pump. Tech. Uni.
Darmstadt, FG Turbomaschinen und
Fluidsantriebtechnik: 1-11.
14. Hardee E, Chang K, Tu J, Choi KK,
Grindeanu I, Yu X (1999). A CAD-based
design
parameterization
for
shape
optimization of elastic solids. Advances in
Engineering Software 30: 18599.
15. Keck H, Sick M (2008). Thirty years of
numerical flow simulation in hydraulic
turbomachines. Acta Mech., Springer 201:
221-29.
16. Menter FR, Langtry R, Ruprecht H (2004).
CFD simulation of turbomachinery flowsverification, validation, modeling. Proceedings
of ECCOMAS 24-28 July, Jyvaskyla
(Finland) 1-14.
17. MODEFRONTIER, www.esteco.com
18. Moradnia P, Golubev M, Chernoray V,
Nilsson H (2014). Flow of cooling air in an
electric generator model-An experimental and
numerical study. J. of Applied Energy
114:644-653.
19. Papadrakakis M, Lagaros ND (2004). Soft
computing methodologies for structural
optimization. J. Applied Soft. Computing 3:
283-300.
20. Radgen P, Oberschmidt J, Cory WT (2008).
EuP Lot11: Fans for ventilation in nonresidential
buildings.
Final
Report,
Fraunhofer Institut ISI. Karlsruhe, Germany.

REFERENCES
1. ANSYS, www.ansys.com.
2. Arora J (1989). Introduction to Optimum
Design. McGraw-Hill.
3. Bohm W, Farin G, Kahmann J (1984) . A
survey of curve and surface methods in
CAGD. Computer Aided Geometric Design
1: 1-60.
4. Bois G (2005). Introduction to design and
analysis of high speed pumps. RTO-EN-AVT
143: 1.1-1.20.
5. Braembussche RA (2005). Optimization of a
radial impeller. RTO-EN-AVT 143-13: 1.11.28.
6. Brost V, Ruprecht A, Maihoefer M (2002).
Rotor stator interactions in a axial turbine, A
comparison of transient and steady state
frozen simulation. Inst. Fluid Mechanics and
Hydraulic Machinery, Uni. Stuttgart: 1-9
7. Cheng CT, Chau KW, Sun YG, Lin JY
(2005). Long-term prediction of discharges in
Manwan Reservoir using artificial neural
network models. Lecture Notes in Computer
Science 3498: 1040-1045.
8. Dai L, Gu Y, Zhao G, Guo Y (2005).
Structural shape optimization based on
parametric dimension-driving and CAD
software integration. 6th World Congress on
Structural and Multidisciplinary Optimization
30 May - 3 June, Rio de Janeiro, Brazil 1-8.
9. Deb K, Goel T (2000). Multi-Objective
Evolutionary Algorithms for Engineering
Shape Design. KanGAL Report, Indian
Institute of Technology Kanpur.
10. Derakhshan
S,
Pourmahdavi
M,
Abdolahnejad E, Reihani A, Ojaghi A (2013).
Numerical shape optimization of a centrifugal
pump impeller using artificial bee colony
algorithm. Computers and Fluids 81:145-51.
11. Farin G (1993). Curves and Surfaces for
Computer
Aided
Geometric
Design.
Academic Press.

21. Rao
SS
(1996).
Engineering
Optimization. Wiley Interscience.
22. Rogers DF, Adams JA (1976). Mathematical
Elements for Computer Graphics. McGrawHill.
23. Taormina R, Chau KW, Sethi, R (2012).
Artificial Neural Network simulation of
hourly groundwater levels in a coastal aquifer
system of the Venice lagoon. Engineering
Applications of Artificial Intelligence 25 (8):
1670-1676.
24. Vucina D, Lozina Z, Pehnec I (2012).
Computational procedure for optimum shape
design based on chained Bezier surfaces
parameterization. Engineering Applications of
Artificial Intelligence 25: 648-67.
25. Vuina D, Milas Z, Pehnec I (2012). Reverse
shape synthesis
of the hydropump
volute using stereo - photogrammetry,
parameterization and geometric modeling.
420

Engineering Applications of Computational Fluid Mechanics Vol. 8, No. 3 (2014)

ASME Journal of Computing and Information


Science in Engineering 12: 021001/1-6.
26. Zhang J, Chau KW (2009). Multilayer
ensemble pruning via novel multi-sub-swarm
particle swarm optimization. Journal of
Universal Computer Science 15 (4): 840-858.
27. Zhang M, Pomfret MJ, Wong CM (1996).
Performance prediction of a backswept
centrifugal impeller at off-design point
conditions. I.J. Num. Methods in Fluids 23:
883-95.

421

You might also like