You are on page 1of 10

Author's personal copy

Composite Structures 92 (2010) 244253

Contents lists available at ScienceDirect

Composite Structures
journal homepage: www.elsevier.com/locate/compstruct

Reliability-based design and optimization of adaptive marine structures


Y.L. Young a,*, J.W. Baker c, M.R. Motley b
a
Department of Naval Architecture and Marine Engineering, University of Michigan, Ann Arbor, MI 48109, USA
b
Department of Civil and Environmental Engineering, Princeton University, Princeton, NJ 08544, USA
c
Department of Civil and Environmental Engineering, Stanford University, Stanford, CA 94305, USA

a r t i c l e i n f o a b s t r a c t

Article history: The objectives of this work are to quantify the inuence of material and operational uncertainties on the
Available online 28 July 2009 performance of self-adaptive marine rotors, and to develop a reliability-based design and optimization
methodology for adaptive marine structures. Using a previously validated 3D uidstructure interaction
Keywords: model, performance functions are obtained and used to generate characteristic response surfaces. A rst-
Reliability-based design and optimization order reliability method is used to evaluate the inuence of uncertainties in material and load parameters
Adaptive marine structure and thus optimize the design parameters. The results demonstrate the viability of the proposed reliabil-
Fluidstructure interaction
ity-based design and optimization methodology, and demonstrate that a probabilistic approach is more
Composite blade
Propeller
appropriate than a deterministic approach for the design and optimization of adaptive marine structures
Turbine that rely on uidstructure interaction for performance improvement.
2009 Elsevier Ltd. All rights reserved.

1. Introduction rise. In Gowing et al. [11], experimental studies demonstrated that


load-induced deformations of composite elliptic hydrofoils helped
Self-adaptive structures are those which change in shape or to delay cavitation inception, while maintaining the overall lift and
property via active and/or passive control mechanisms to automat- drag. Numerical [1217] and experimental [18] studies have also
ically adjust to the changing environment. Active, passive, and hy- shown that passive pitch adjustments through the use of aniso-
brid control mechanisms have been exploited to achieve adaptive/ tropic laminated composites helped to increase the fuel-efciency
intelligent/smart designs. Among the passive control mecha- of marine propellers over a range of operating conditions. Similar
nisms, one important class involves the use of uidstructure passive pitch adjustment strategies have also been explored to in-
interactions. crease the energy capture of marine/current turbines in Nicholls-
In aerospace engineering, utilization of uidstructure interac- Lee and Turnock [19].
tions for performance enhancement has been well documented. The focus of the current work is on passive, self-adaptive mar-
In Khan [1], aeroelastic behaviors of composite helicopter blades ine structures that utilize uidstructure interactions. Because
were investigated to improve propeller performance, including the performance of these structures depends on uidstructure
thrust, power, and efciency. Studies have also shown that aero- interactions, they may be more sensitive to random variations in
elastic tailoring and optimization of composite helicopter blades material and load uncertainties. Hence, the objective of this work
can improve stability, reduce vibration, and reduce hub and dy- is to develop a reliability-based design and optimization methodol-
namic blade loads [26]. In Yang [7], uidstructure interaction ogy to improve the performance and reliability of adaptive marine
analysis was performed for a composite canard structure, where structures. To demonstrate the methodology, results are shown for
the possibility of achieving higher efciency and better maneuver- a self-adaptive composite marine propeller, but the methodology is
ability was numerically demonstrated. Passive control technolo- generally applicable to other adaptive structures that undergo
gies have also been considered for wind turbines, where material uidstructure interaction.
loaddeformation coupling was used to reduce fatigue damage
through load-mitigation at high winds (see Refs. Lee and Flay [8], 1.1. Reliability-based design and optimization of self-adaptive
Lobitz et al. [9]), and to improve energy capture (see Ref. Lobitz structures
and Veers [10]).
In marine and ocean engineering, investigations into methodol- Reliability-based design and optimization is a common practice
ogies that utilize uidstructure interactions have also been on the for many rigid and/or non-adaptive structural engineering sys-
tems. The objective is to ensure the level of required reliability is
* Corresponding author. Tel.: +1 734 936 7636; fax: +1 734 936 8820. achieved with respect to uncertainties in structural parameters
E-mail addresses: ylyoung@umich.edu, yyoung@princeton.edu (Y.L. Young). and operating conditions.

0263-8223/$ - see front matter 2009 Elsevier Ltd. All rights reserved.
doi:10.1016/j.compstruct.2009.07.024
Author's personal copy

Y.L. Young et al. / Composite Structures 92 (2010) 244253 245

Although much progress in this eld has been made for rigid (CFRP). In addition to the well-known higher specic stiffness
and/or non-adaptive structures (see [20] for a recent review in this and higher specic strength of CFRP, the intrinsic deformation cou-
area), relatively little work focuses on exible structures that inter- pling behavior of anisotropic composites can be utilized to improve
act with the environment. Reviews of the state-of-the art methods the propeller performance by passive tailoring of the load-induced
in reliability-based design and optimization of aeroelastic struc- deformations according to the changing inow, as demonstrated in
tures can be found in [21]. As noted in [21], only limited work recent numerical (see Lee and Lin [12], Lin and Lee [13], Young
has been done on reliability analysis of structures undergoing et al. [14], Young and Liu [15], Young [16], Motley et al. [17]) and
uidstructure interactions, and most existing methods employed experimental (see Chen et al. [18]) studies. Nevertheless, all of
simplistic linear uid and uidstructure interaction models to the work thus far on self-adaptive composite marine propellers
determine the mechanical response, which introduce epistemic has been limited to deterministic analysis. Since the performance
modeling uncertainty. Hence, [21] introduced a reliability analysis of these structures is more sensitive to material or load uncertain-
method that integrates a coupled Euler ow solver with a struc- ties due to their dependence on uidstructure interaction, a reli-
tural nite element model (FEM) for the deterministic aeroelastic ability-based design and optimization method that can consider
analysis of a 3D wing structure, and employed a rst-order reliabil- natural or man-made variations is needed.
ity method (FORM) to evaluate the performance sensitivities to de-
sign parameters, operating conditions, and modeling uncertainties. 3. Problem denition
A probabilistic design assessment of smart composite structures is
presented in [22] in which sensitivity factors were developed for a To perform a reliability-based evaluation of the structure, we
series of design parameters for a composite wing based on their ef- will need to evaluate two performance measures. First, we must
fects on the angle of attack and impact response of the structure. nd the probability of unsatisfactory performance. This is done by
By improving the reliability of design parameters with the highest dening a limit state function, gX, where X consists of a vector
sensitivity factors, the failure probability was reduced for the of design variables, XD , either deterministic or random, and a vector
structure. The stochastic nature of composite properties has also of random variables, XR , representing uncertain structural proper-
been shown to lead to overestimation of structural reliability. By ties and loading conditions, and g is a function that relates the de-
using a probabilistic design methodology, improvements can be sign variables, random variables and the performance of the
made over traditional deterministic design methods [23]. In [24], structure. The function gX can either be implicit (e.g., the outcome
the reliability of a thin-walled circular composite cylinder was of a numerical BEMFEM code), or explicit (e.g., an approximate
shown to have a strong sensitivity to the applied load and to the equation obtained using the response surface method). The func-
amount of parametric scatter via multiple response surface tion gX is chosen such that gX 0 denes a boundary between
techniques. satisfactory and unsatisfactory performance (with gX < 0 indicat-
It should be noted that all of the above mentioned reliability- ing that the structure has unacceptable performance, and gX > 0
based design and optimization methods focus on adaptive/smart indicating acceptable performance). The performance state associ-
aerospace structures. Similar work is also needed for adaptive mar- ated with the boundary gX 0 is denoted as a limit state. Given
ine structures, where the uid loading tends to be much higher this formulation, the optimization problem herein can be written as
(due to the higher uid density and viscosity), the ow may be
highly unsteady (due to transient structural motion, as well as spa- maxpg obj X > 0 1
XD
tial and temporal variations in the ow eld), and may be suscep-
or minpg obj X 6 0
tible to cavitation damage. XD

where g obj X is the objective function, based on the efciency g of


1.2. Objectives the adaptive composite propeller, which is required to be greater
than a minimum target efciency for all loading conditions, g :
The objectives of this work are to (1) quantify the inuence of
g obj X gX  g 2
material and load uncertainties on the performance of self-adap-
tive marine structures and (2) optimize the design to achieve the subject to two probabilistic limit state functions g prob
1 and g prob
2 .
desired level of reliability in structural performance and, in doing
so, develop a reliability-based design and optimization method g prob
j pfj  pg fj X < 0 P 0; j 1; 2 3
for adaptive marine structures.
where the constraint functions g fj are dened as
P ST X
2. Self-adaptive composite marine propellers g f 1 X 1  4
P rigid J
Dmax DX
Marine propellers are traditionally made of nickelaluminum g f 2 X  5
bronze (NAB) due to its excellent stiffness, yield strength, and D D
anti-biofouling characteristics. They are designed to be rigid, and subject to an acceptable probability of failure, pf pf 1 pf 2 T . We de-
the blade geometry is optimized to yield the maximum efciency note gX as the efciency of the self-twisting propeller. PST X and
at the design ow condition. However, when the advance speed Prigid J represent the power demand of the self-twisting and rigid
or the shaft rotational frequency moves away from the design val- propellers, respectively. Note here that the performance of the rigid
ues, the blade geometry becomes sub-optimal relative to the chan- propeller is a function of the loading condition represented by the
ged inow, and hence leads to decreases in energy efciency. The advance coefcient J V=nD (ratio of mean relative inow velocity
effect is more severe when a rigid propeller is operating behind to rotor tip velocity) only because the objective is to optimize the
an asymmetric wake (caused by interactions with the upstream design variables for the self-twisting propeller such that it yields
hull, inclined shaft, ship maneuvering, etc.) because the resultant equal or better performance compared to the already optimized ri-
inow angle will vary periodically with blade position. Conse- gid propeller. The rigid propeller is only used as a reference to eval-
quently, the efciency of a rigid propeller tends to decrease when uate the performance of the adaptive propeller. It should be noted
operating behind spatially varying wake. This problem can be min- here that V is the propeller advance speed, D is the propeller diam-
imized by using blades made of carbon ber reinforced plastics eter, and n is the propeller rotational frequency.
Author's personal copy

246 Y.L. Young et al. / Composite Structures 92 (2010) 244253

In the application considered here, the vector of random vari- g C CH fug


M MH fu _ Kfug
ables is dened as XR J; E1 ; E2 ; G12 ; m12 ; m21 T . For the sake of sim- fFce g fFco g fFr g 7
plicity, the blades are assumed to be made of a single layer of
orthotropic lamina with material properties E1 ; E2 ; G12 ; m12 ; and where fu g; fug,
_ and fug are the structural nodal acceleration,
m21 and the bers are oriented at angle h counterclockwise relative velocity, and displacement vectors, respectively; M; C, and K
to the local spanwise direction. Again for simplicity, the only de- are the structural mass, damping, and stiffness matrices, respec-
sign variable considered is the ber orientation angle, XD h. tively; MH  and CH  are called the hydrodynamic added mass and
The objective of the optimization problem is to nd the best ber hydrodynamic damping matrices, respectively, because MH  is
orientation angle, h, that maximizes the overall efciency of the associated with fu g and CH  is associated with fug. _ Notice that
self-twisting propeller, as represented by Eq. (2), subject to design fFv g MH fu g  CH fug
_ represents the dynamic hydroelastic
constraints represented by Eqs. (4) and (5). force caused by uidstructure interaction, and can be derived by
Eq. (4) is used to ensure that the expected average power de- application of the pressure and velocity compatibility conditions
mand of the self-twisting propeller is less than that of the rigid at the blade surface [25,16]. fFce g; fFco g, and fFr g are the centrifugal
propeller, which will guarantee that the self-twisting propeller force, the Coriolis force, and the hydrodynamic force (due to rigid
provides higher averaged energy efciency. Further, we dene blades rotation) vectors, respectively. Detailed formulation of these
DX
D
as the blade tip deection D normalized by the propeller matrix identities can be found in [25,16]. Eq. (7) can be solved using
diameter (D), a parameter which is limited by the maximum allow- standard nite element methods (FEM) in the time-domain such as
 
able normalized blade tip deection Dmax D
. The blade tip deection ABAQUS/Standard [28]. User-developed subroutines are utilized to
needs to be restrained to limit the possibility of blade strength and superimpose the hydrodynamic added mass matrix MH  with the
stiffness failures. It should be noted that composite blades made of structural mass matrix M, and the hydrodynamic damping matrix
CFRP can have many possible material failure modes, as well as CH  with the structural damping matrix C, and to perform itera-
hydroelastic instability failure modes, most of which can be corre- tions between the BEM and FEM solvers to consider nonlinear FSI
lated to the tip deections. As such, the tip deection, Eq. (5), de- effects induced by large blade deformations.
nes the safety limit to ensure structural stability and integrity.
Eq. (4) is used to represent the serviceability limit because its
objective is to minimize power demand. Eqs. (4) and (5) limit the 5. Problem setup
optimal design range and the objective function (Eq. (2)) is used
to nd the ber orientation angle that maximizes the probability The propeller herein is modeled using a single layer for simplic-
of exceeding the minimum target energy efciency of the self- ity, but the actual model will have many layers and will be stacked
twisting propeller. in a sequence such that the loaddeformation characteristics will
be the same as the effective single layer model [29]. However, it
4. Fluidstructure interaction analysis method should be cautioned that such simplication is only appropriate
to determine the loaddeformation characteristics for linear-elas-
The numerical modeling involves a deformable composite pro- tic structural systems. Detailed stress analysis using the actual
peller subjected to a spatially varying inow wake VE . The model multi-layer model should be performed after the hydroelastic opti-
has been validated against analytical, numerical, and experimental mization analysis to verify structural integrity.
results [14,25,15,16]. The material selected is Hexcel IM7-8552 carbon epoxy com-
The governing equation for the uid is the incompressible Euler posite. The mean-load geometry is based on that of propeller
equation in a blade-xed rotating coordinate system: 5474 (Fig. 1), one of the composite propellers manufactured by
AIR Fertigung-Technologie GmbH and designed and tested in coop-
DVt =Dt rp=q g  X  X  x  2X  Vt eration with the Naval Surface Warfare Center, Carderock Division
r  Vt 0 6 (NSWCCD). The propeller has a diameter of D 0:6096 m. The de-
sign rotational frequency is n 780 rpm. The design advance coef-
where Vt is the total velocity, t is the physical time, p is the hydro-
cient is J V=nD 0:66. More details of propeller 5474 can be
dynamic pressure, q is the water density, g is the gravitational
found in [18,15].
acceleration, and X is the propeller rotational speed vector. The to-
tal velocity Vt can be expressed as the sum of the inow velocity
Vin and a perturbation potential velocity rU where the inow
velocity can be decomposed into the effective wake velocity Ve
and the blade rotational velocity: Vin Ve  X  x. The effective
wake velocity Ve is obtained either from experimental measure-
ments [26] or from a coupled RANS/Euler and potential ow solver
[27]. It includes the interaction between the nominal wake vorticity
(in the absence of the propeller) and the vorticity induced by the
propeller [27]. The perturbation ow eld can be treated as incom-
pressible, inviscid, and irrotational. Hence, it satises the Laplace
equation: 52 U 0. Further, the perturbation potential U can be
decomposed into two parts, namely, /, which is due to rigid blade
rotation, and u, which is due to elastic blade deformation. Both /
and u can be formulated as a mixed boundary value problem in
the time-domain and solved using a 3D boundary element method
(BEM) [25,16].
By virtue of the previous decomposition, the nite element dis-
deformed
cretization for structural analysis in the rotating blade-xed coor-
dinate system can be formulated as follows: undeformed

Fig. 1. Deformed and undeformed geometry of the self-twisting propeller.


Author's personal copy

Y.L. Young et al. / Composite Structures 92 (2010) 244253 247

16 0.8

12 0.6 Efficiency,

KT : 10 KQ :
Self-Twisting
tip (o)
Rigid
8 0.4

Torque, 10 KQ
4 Rigid 0.2
Self-Twisting - Undeformed
Self-Twisting - Deformed Thrust, KT
Theoretical Optimal
0 0
0.5 0.6 0.7 0.8 0.9 0.5 0.6 0.7 0.8 0.9
J J

Fig. 2. Comparison of the tip pitch angle (left) and performance curves (right) for the rigid and self-twisting propellers.

Fig. 2 compares the performance of the adaptive self-twisting 0.06


propeller with its rigid counterpart based on deterministic uid

max/D
0.04
structure interaction analysis using the method outlined in Section
Standard Material
4. As shown in the left plot, the adaptable propeller geometry (rep- 0.02 E 1 + 3E1
resented by the tip pitch angle, /tip ) approaches the theoretical E 1 - 3E1
0
optimal propeller geometry, which changes with the advance coef- 0 15 30 45 60 75 90
( )
o
cient (J is inversely proportional to the angle of attack). For the
possible range of J values for forward operations, the self-twisting 0.06
propeller is designed to be overpitched in its unloaded (unde-
max/D

formed) conguration. The self-twisting propeller de-pitches due 0.04

to twisting motion induced by bending deformation caused by 0.02 Standard Material


E 2 + 3E2
the uid loading, which changes with J (operating condition). The E 2 - 3E2
0
design requirements are that: 0 15 30 45 60 75 90
( )
o

1. At J J design 0:66, the deformed geometry of the self-twisting 0.06


propeller matches the optimized rigid propeller geometry to
max/D

0.04
achieve equivalent performance between the two propellers.
Standard Material
2. At all J J design , the self-twisting propeller should yield higher 0.02 G 12 + 3G12
energy efciency than its rigid counterpart. 0
G 12 - 3G12

0 15 30 45 60 75 90
( )
o
The efciency is dened as g TV=2pnQ JK T =2pK Q which
corresponds to the ratio of the thrust power to the available shaft
Fig. 3. Effect of variations in material properties on the blade tip deection for the
power, with thrust coefcient K T T=qn2 D4 and torque coefcient self-twisting propeller.
K Q Q =qn2 D5 . T and Q are the dimensional thrust and torque,
respectively. As shown in Fig. 2, the rigid and the self-twisting pro-
pellers exhibit similar performance at the design condition 45
Standard Material
J 0:66. The efciency of the self-twisting propeller is higher than 40
P (kW)

E 1 + 3E1
its rigid counterpart for all J 0:66. The efciency improvement 35
E 1 - 3E1

increases as the ow condition further deviates from the design 30


condition. Further, the resulting thrust and torque exhibit smooth-
25
er variation with changing J. The result is a propeller that is, on 0 15 30 45 60 75 90
( )
o
average, more energy efcient than its rigid counterpart, requiring
less power to operate and less variation in power, which reduces 45
Standard Material
the strain and extends the fatigue life of the engine. For details 40 E 2 + 3E2
P (kW)

about the design procedure or uidstructure interaction analysis 35


E 2 - 3E2

methodology, readers should refer to [14,30,16,17]. 30


25
0 15 30 45 60 75 90
6. Parametric sensitivity for R
( )
o

45
To further simplify the model, parametric sensitivity analyses of Standard Material
the random variables XR were performed (see Figs. 3 and 4). By 40 G 12 + 3G12
P (kW)

G 12 - 3G12
taking each of the material parameters and providing them with 35
a normal distribution (i.e. with mean lXR XR;design 30
J; E1 ; E2 ; G12 ; m12 ; m21  0:66; 171:42 GPa; 9:08 GPa; 5:29 GPa; 25
0 15 30 45 60 75 90
0:32; 0:32 and standard deviation rXR 0:02l XR based on
( )
o

expected tolerances, the sensitivities to these variables can be as-


sessed. The material parameters were assigned extreme deviations Fig. 4. Effect of variations in material properties on the power requirement for the
from the design values of three standard deviations (99.7% of the self-twisting propeller.
Author's personal copy

248 Y.L. Young et al. / Composite Structures 92 (2010) 244253

realizations of a normal distribution are within three standard FEM model, and gX is the data generated from the response sur-
deviations of the distribution mean), and compared with the de- face method; R2 values closer to 1.0 represent higher accuracy.
sign condition (i.e. XR;design ). The contour maps of the tted response surfaces and data com-
As shown in Figs. 3 and 4, the effects of variations in the primary puted using the BEMFEM numerical solver are shown in Figs. 57.
bending modulus E1 and shear modulus G12 are negligible. The shaded contour values represent the tted equations, while
Bounds for extreme values of the secondary bending modulus the dashed contour lines represent the BEMFEM simulation data.
E2 , however, noticeably deviate from the design behavior, partic- The power requirement is more sensitive to J than to h. Lower
ularly for large h. This is because at low values of h, the primary values of J correspond to higher angles of attack and higher loads
stiffness is governed by E1 171:42 GPa, which is a much larger and thereby higher power demands. At higher loads, the change
value compared to E2 9:08 GPa. Hence, even at three standard in pitch caused by the uidstructure interaction is also greater,
deviations from the mean 161:13 GPa 6 E1 6 181:71 GPa, the ef- and hence the power demand is more sensitive to h at lower J val-
fects on the normalized bending deection and power are small ues. At high J values, the power demand is lower and is less sensi-
due to the high stiffness. On the other hand, as h approaches 90 , tive to h due to small changes in pitch caused by the hydrodynamic
the primary bending stiffness of the blade is governed by E2 . Be- load induced bendingtwisting deformation.
cause E2 is comparatively small with respect to E1 , the system The maximum deection is a strong function of both J and h.
behavioral effects for variations in E2 are magnied, though only This is because, as the ber orientation angle becomes larger, the
marginally. For the purposes of this paper, however, it is assumed blades are less stiff along their primary (longitudinal) axis (which,
that the material parameters, except for h, have negligible effect on at h 45 becomes oriented more as the secondary axis). As a re-
the efciency, power requirement, and tip displacement of the pro- sult, the blade tip deections have nonlinear growth with ber ori-
peller blades. As such, the random variable vector XR can be simpli- entation angle. The increasing of the tip deection with decreasing
ed to only contain the advance coefcient, J. J is also expected due to increasing longitudinal load.
The efciency is highest near the design values J J design
0:66; h hdesign 32 , which means that the original design
7. Response surface methodology

90
The fully-coupled boundary element method-nite element

28
method (BEMFEM) model [25,16] summarized in Section 4 is 36

24
used for the design and analysis of adaptive composite marine ro-
tors. Although the coupled BEMFEM analysis method is relatively
fast, it can still be computationally expensive to use, with wait 60
32

time requirements ranging from 5 min to 2 h for a single simula-


tion on a single processor depending on if the analysis is steady,
(o)

unsteady, with or without cavitation. For a Monte Carlo analysis

28
large enough to successfully achieve a reliable optimization, this

20
24
becomes impractical. Since the behavior of the performance 30
(power, deection, and efciency) is expected to be smooth func-
40 36 32
tions of J and h, the response surface methodology is a reasonable
analysis alternative. Data points obtained from the BEMFEM 44

model were used to predict the behavior of the self-twisting com- 48 28


0
posite propeller. By using a fully two-dimensional regression anal- 0.4 0.5 0.6 0.7 0.8 0.9
ysis, equations for the response surface of the self-twisting J
Power, P (kW)
propeller power requirement, tip deection, and efciency were Shaded Contours - Fitted
Dashed Lines - BEM-FEM 20 24 28 32 36 40 44 48 52
developed:
Fig. 5. Power demand, P (kW), contour map for the self-twisting propeller.
PST J; h 104 W5:818  0:104J  0:0909h
0:0516Jh  4:020J 2 0:000772h2
90
0:000361Jh2  0:0304J2 h  0:000330J 2 h2 8
0.03

DJ; h
0 .0 4

0:0060  0:0010J 0:0019h


D
05

 0:0013Jh  0:0021J 2  0:000006h2


0.

06
60 0.
0:000005Jh2  0:00031J2 h  0:000001J 2 h2 9
gJ; h 0:2358 2:2626J 0:0015h 0:0058Jh
(o)

 1:3411J 2  0:0001Jh2  0:0105J2 h 0:0001J 2 h2 10


0 .0 4
where J is dimensionless, h is in degrees, and W represents units of
30 0 .0 3

watts for the power surface. The goodness-of-t of the surfaces can 0 .0 2
be represented by the coefcients of determination for the power
0 .0 1
demand, blade tip deection, and efciency, which are 0.997,
0.997, and 0.988, respectively, where 0 0.0 1
0.4 0.5 0.6 0.7 0.8 0.9
Rg BEMFEM  gBEMFEM 2
J
2
R 1 11
max / D
Rg BEMFEM  gX2 Shaded Contours - Fitted
Dashed Lines - BEM-FEM 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09

where g BEMFEM is the data obtained from the numerical BEMFEM Fig. 6. Normalized blade tip deection, Dmax =D, contour map for the self-twisting
model, gBEMFEM is the mean of all data obtained from the BEM propeller.
Author's personal copy

Y.L. Young et al. / Composite Structures 92 (2010) 244253 249

90 50000

0.66

0.7 0
40000
60

0.70
(o)

P rigid (W)
30000
0 .5 4

30
0 .5 8

0 .6 2

0.
70
0 .6

20000
0 .5

6
0

0
46

0.4 0.5 0.6 0.7 0.8 0.9


J
Efficiency,
Shaded Contours - Fitted
Dashed Lines - BEM-FEM 0.46 0.50 0.54 0.58 0.62 0.66 0.70 10000
0.4 0.5 0.6 0.7 0.8 0.9
Fig. 7. Propeller efciency, g, contour map for the self-twisting propeller.
J

Fig. 9. Polynomial curve-t for rigid propeller power requirement as a function of J.

to be rigid). Using polynomial tting techniques, a second-order


curve was used to t the data R2 0:999 for the rigid propeller
(Fig. 9):
4
Prigid J 104 W38572J 2  5091J 50416 12
(o)

8. First-order reliability method

2
Eq. (1) denes the probability of unacceptable performance, and
computing this probability requires integration of the probability
density function of X over the domain of x values that would result
in unacceptable performance
Z
0 pf pgX < 0 fX x dx 13
0 15 30 45 60 75
gX60
( )
o

This integral in general involves a complex high-dimensional


Fig. 8. Change in tip pitch angle with ber orientation angle for the self-twisting failure domain and often cannot be performed analytically, as is
propeller at J Jdesign .
the case here. Thus, we turn to the rst-order reliability method
(FORM), which facilitates an approximate solution of the integral
by transforming the random variables X into variables having a
objectives are satised. Note that the efciency of the adaptive multivariate standard normal density function, linearizing the lim-
composite propeller has a strong dependence on J, which is inver- it state function in this transformed domain, and then utilizing
sely proportional to the angle of attack, but a weaker dependence analytical solutions to the transformed and linearized problem
on h. It is of note, however, that there exists a quadratic element to [31].
the behavior of the surface based on the ber orientation angle. In this application, the variables in X do not necessarily have
This curvature switches directions at J 0:66 and the local maxi- normal distributions, so we start by transforming each component
mum and minimum point is located at h hdesign , a characteristic of the vector X, denoted X i , to a corresponding standard normal
which the response surface takes into account. This change in cur- variable, denoted U i . This transformation is performed by equating
vature can be explained by combining the behavior shown in Figs. the probability of non-exceedance of any numerical values of X i
8 and 2. According to Fig. 8, for h > 32 and h < 32 , the change in and U i . This can be stated as follows:
tip pitch angle, D/, will be less than D/jhhdesign 32 . According to Uui F X i xi 14
Fig. 2, for J < J design 0:66, if D/ < D/jhdesign , the loaded pitch distri-
bution will be further away from the theoretical optimal value and where U is the standard normal cumulative distribution function
hence g < gjhdesign ; for J > J design 0:66, if D/ < D/jhdesign , the loaded (CDF) and F Xi is the CDF of the non-normal random variable X i . If
pitch distribution will be closer to the theoretical optimal value the components of X are not mutually independent, a generaliza-
and hence g > gjhdesign . tion of Eq. (14) to use a Nataf or Rosenblatt transformation would
The rigid propeller power requirement does not require re- be needed to remove that dependence from U. But the calculations
sponse surface methodology as it is only a function of J; however, here use only mutually independent random variables, so this com-
tting a curve to dene the behavior of the rigid propeller is also ponent-by-component transformation is sufcient.
faster than using the BEM model to compute the behavior at each The nonlinear limit state functions gX can be transformed into
J value (FEM analysis is not necessary since the blades are assumed a linear limit state space hU using this same transformation. We
Author's personal copy

250 Y.L. Young et al. / Composite Structures 92 (2010) 244253

20 3
g(X) < 0 h(U) < 0
FAILURE FAILURE h(U)=0
g(X) = 0
15 1.5
x* u*

U1
X2

10 0

5 -1.5
g(X) > 0 h(U) > 0
SAFE SAFE
0 -3
0 5 10 15 20 -3 -1.5 0 1.5 3
X1 U2

Fig. 10. Generalization of the transformation from the original X space (Left) into the U space (Right).

then proceed by taking advantage of the fact that pf pgX 6


0 phU 6 0: h(U) < 0
FAILURE h(U) = 0
Z Z
pf fX x dx fu u du 15
gX60 hu60 u*

This mapping of the problem from X-space to U-space is illus-

U1
trated in Fig. 10. The latter integral can still not be solved
analytically, unless the limit state function hU is linear. We thus
linearize hU at the so-called design point u , dened as
follows:
h(U) > 0
u arg minkUk jhU 0 16 SAFE

By this denition, u is the point on the limit state surface U2


hU 0 closest to the origin in U-space, and thus has the highest
Fig. 11. Generalization of the determination of u and b.
probability density of all points in the failure domain hU 6 0. This
high probability content means that linearizations of hU around
u should produce pf estimates close to the pf estimates obtained 9. Design example
without this linearization. Once u is known and the limit state
surface is linearized at this point, the reliability index b can be 9.1. Parameter denitions
computed as
A design example is presented based on the reliability and re-
b au 17 sponse surface methodology explained above. The rst step be-
yond the response surface methodology involves determining
where a is the unit normal vector perpendicular to the limit state how to dene the random distribution of the variables. It is typical
surface at u and pointing into the failure domain for a manufacturer to provide a ber orientation tolerance around
23 in the construction of the laminates for propeller or turbine
rhu blades, with a condence level of 95%. With this as a reference
a 18
krhu k point, it is reasonable to assume that the ber orientation angle
has a Gaussian distribution with a mean value of hdesign . A tolerance
This reliability index is directly related to the probability of fail-
of 3 with 95% condence can be approximated by a standard devi-
ure by the equation
ation of 1:5 (for a normal distribution, 95% of values are within
pf Ub 19 two standard deviations of the mean).
Determining an appropriate distribution for the advance coef-
This pf is the FORM approximation of the pf associated with the cient is slightly less intuitive. It can be assumed that the propeller
original non-normal random variables and nonlinear limit state will operate near the design advance coefcient J design 0:66 un-
function. The variables u ; a and b are illustrated graphically in der most operating conditions and that this would be an appropri-
Fig. 11. Several well-studied numerical algorithms for nding these ate mean or mode value. A standard deviation of 0.10 (for normal
variables are available [31]. distribution) will provide a realistic range of operating conditions.
In addition to the probability of failure estimate, this FORM cal- A normal distribution, however, may not be appropriate here.
culation provides several other informative outputs. The elements Operating conditions are sensitive to many variables, including
of the vector a provide information about the relative importance currents, waves, ship acceleration and deceleration, turning and
of the random variables in U (and, after a simple transformation, towing. A log-normal distribution can be used to simulate a ran-
provides the same information about the original random variables dom variable that is the product of many random variables. The
X). The design point u can also be transformed (using the inverse log-normal distribution tends more toward a positive skew (the
of the operation in Eq. (14)), to nd a corresponding x : the values right tail is larger). However, surface vessels tend to favor a nega-
of X that have the highest probability of causing failure of the sys- tive skew (the left tail is larger). More often than not, deviations
tem of interest. will occur at values lower than J design because of towing or high
Author's personal copy

Y.L. Young et al. / Composite Structures 92 (2010) 244253 251

0.06 Normal
stresses. Further, the natural frequency limit is approximately
Lognormal 60 Hz, or 3600 rpm, which is about 4.6 times higher than the de-
0.04
PDF

Gumbel sign propeller rotational frequency at n = 780 rpm.


0.02 The minimum target efciency for all loading conditions is set
0 at g 0:60 to ensure satisfactory performance. The serviceability
0 0.5 1 1.5 limit state is set at pf 1 0:50 to maximize the probability that the
J
self-twisting propeller yields better performance (lower power de-
1 mand), on average, than its rigid counterpart. The safety limit state
is set at pf 2 0:001 to limit the deections to avoid strength or sta-
CDF

0.5 bility failure while allowing enough exibility to enable perfor-


mance improvement through passive bendtwist coupling
induced by the hydrodynamic loads.
0
0 0.5 1 1.5
J 9.2. Validation studies

Fig. 12. Comparison of the normal, log-normal, and gumbel distribution of J.


The FORM calculations of the previous section require an
approximation in that they linearize the limit state surface after
wave resistance. As a result, the lower limit of J deviates more from transforming the random variables affecting the structure perfor-
the mode than the higher limit, resulting in a negative skew. This mance. This linearization typically has only a minor impact on
can be modeled using a Gumbel distribution, which can be gener- the computed failure probabilities, especially when the true failure
ated by xing the mode (the most frequent value) at J design and lim- probability is small.
iting the minimum value to J 0 (i.e. forward operations only). To verify that FORM did not introduce any signicant errors in
This ensures that the most common value, the mode, of J is that this application, FORM and comparable Monte Carlo results are
of the design condition. As shown in Fig. 12, all three distributions compared. Monte Carlo pf estimates can be obtained by simulating
shown as both a probability density function (PDF) and cumula- realizations of X having appropriate probability distributions, and
tive density function (CDF) share the same mode, while the dis- then counting the fraction of realizations for which the structure
tribution away from the mode differs. has unacceptable performance. This approach produces pf esti-
A second initial step is to dene an acceptable maximum tip mates that approach the true pf when the number of simulations
deection and minimum target efciency. An inherent problem is large, because no restrictions are required on the permissible
of self-twisting propellers is that they can be subject to strength- distribution of X or on the form of the limit state surface. The
based failures, as well as hydroelastic instabilities and resonance one approximation retained here is to use the response surfaces
issues. As described in Section 3, by limiting the tip deection rather than the fully coupled BEMFEM solver to represent the
these issues can be avoided or minimized. Hence, the value structural performance. Although Monte Carlo simulations can
Dmax =D 0:05 is selected for this design example. Extending this produce estimates that converge to the true pf value and are more
value too high can lead to static divergence (during deceleration widely applicable, FORM is still preferred here because of its lower
and backing, although not considered here), increased stresses, computational cost and its additional diagnostic tools such as the a
and higher susceptibility to resonance. As seen in Fig. 13, vector dened in Eq. (18).
Dmax =D 0:05 provides a bound which limits the maximum von Fig. 14 shows the results of the FORM methodology as com-
Mises stress to approximately 500 MPa, where the tensile and pared to a 1,000,000 simulation Monte Carlo analysis based on
compressive ber failure stresses for this material are 2300 and the response surface technology described above. Note the near-
1200 MPa, respectively. While matrix failure stresses may be an or-
der of magnitude less than ber failure stresses, matrix failures are 1
generally shear based failures and the internal shear stresses tend
p(gobj <0)

0.75
to also be an order of magnitude smaller than the primary axis
0.5
0.25
0
600 180 0 30 60 90
design

160 1
500
0.75
p(gf1 <0)

max
v
140 0.5
400
0.25
(MPa)

120 Serviceability
n (Hz)

0
300 0 30 60 90
design
max

100
v

200 1
Safety
80 0.75
p(gf2 <0)

100 n 0.5
60
FORM
0.25
MonteCarlo
0 0
0.02 0.04 0.06 0.08 0 30 60 90
max / D design

Fig. 13. Maximum von mises stress rmax


v and fundamental frequency xn of the Fig. 14. FORM vs. Monte carlo methodology comparison. h is assumed to follow a
blade as a function of the normalized blade tip deection, Dmax =D. normal distribution and J is assumed to follow a gumbel distribution.
Author's personal copy

252 Y.L. Young et al. / Composite Structures 92 (2010) 244253

1 random variations in the material stiffness parameters. The distri-


Optimal Design Range
bution of the random variable J was examined with normal, log-
p(gobj < 0)

0.75
0.5 normal, and Gumbel distributions. The Gumbel distribution was
considered to be more appropriate physically, and thus was used
0.25
as the advance coefcient distribution.
0
0 30 60 90 First-order reliability methods (FORM) were shown to be an
design adequate design tool instead of the more time consuming Monte
1 Carlo simulations for probabilistic propeller optimization. Through
0.75 Serviceable Operating Range FORM analysis, it was shown that the optimal ber orientation an-
p(gf1 < 0)

0.5
gle for the adaptive propeller is 29 6 h 6 31 , which will yield a
93% probability of acceptable performance based on three criteria:
0.25
a serviceability limit state based on propeller power requirement, a
0 safety limit state based on blade tip deection, and an objective
0 30 60 90
design function that ensures the maximum overall energy efciency.
1
The serviceability limit state and constraint functions are designed
such that, on average, the power requirement of the adaptive pro-
0.75 Safe Operating Range
p(gf2 < 0)

peller is less than its rigid counterpart. The safety limit state and
0.5
constraint functions are designed to limit the blade tip deection
0.25 to a specied value to prevent excessive deections, stresses, and
0 to reduce the susceptibility to hydroelastic instability failures. Fi-
0 30 60 90
design nally, the objective function was used to determine the optimal -
ber orientation angle that will maximize the average energy
Fig. 15. Results of the FORM methodology for the probability of failure of the efciency of the self-twisting propeller for all forward operating
objective and limit state functions. conditions. With the knowledge of the optimized equivalent single
layer ber orientation, a series of possible layup sequences can be
developed that will provide an equivalent optimal loaddeforma-
perfect agreement between the two methods, suggesting that tion behavior of the blade.
FORM is valid for this example, where h is assumed to follow a nor- The results show that a probabilistic approach is more appropri-
mal distribution and J is assumed to follow a Gumbel distribution. ate than a deterministic approach for the design and optimization
of adaptive composite structures that rely on uidstructure inter-
9.3. Results action. This is because such structures are inherently more sensi-
tive to random variations in material properties, geometric
The results of the objective function and limit states are shown congurations, and loading conditions. It is important to note that,
in Fig. 15. Again, h is assumed to follow a normal distribution while while the self-twisting propeller can be optimized for efciency, it
J is assumed to follow a Gumbel distribution. According to the top is possible that the limit state functions would provide no viable
gure, the optimal ber orientation angle in terms of the objective design option that would be both serviceable and safe. Should a de-
function is about 59 ; however, the objective function showed lit- sign singularity occur, changes in the blade geometry, material
tle variation in the failure probability across the entire range of h. properties, and material congurations may be necessary. Addi-
The limit states play a very important role beyond the objective tional work is needed to assess the effect of material, geometry,
function. The constraint functions each have denitive boundaries and load uncertainties on the initiation and evolution of failure
for acceptable performance. The serviceability constraint is that modes. This is more complex due to the need to consider the many
the power requirement of the self-twisting propeller is, on average, layers of laminates and the many possible modes of failure, as well
lower than that of the rigid propeller, which can only be satised if as uncertainties in the failure modeling of CFRP. It should be
29 6 h 6 83 . Second, the safety constraint limits the ber orien- emphasized here that although the methodologies presented here-
tation angle to h < 31 . Hence, what seemed initially to be a wide in focus on adaptive composite marine propellers, the framework
range of viable options for the design variable based on the objec- is also generally applicable to other exible structures that under-
tive function is limited to a small range of 29 6 h 6 31 . This go uidstructure interactions, including wind or tidal turbines.
places the optimal design based on FORM methodology to be very
close to the optimal deterministic design. In this case, the probabil-
ity of failure of the objective function ranges between 6.6% and Acknowledgements
6.9%, which represents approximately 93% condence that the
self-twisting propeller will exhibit safe and improved performance The authors are grateful to the Ofce of Naval Research (O N R)
over the rigid propeller for a realistic range of operating conditions. and Dr. Ki-Han Kim (program manager) for their nancial support
through grant numbers N00014-07-1-0491 and N00014-08-1-
10. Conclusions 0475.

The objective of this research is to develop a reliability-based


design and optimization methodology to improve the energy ef- References
ciency of self-adapting composite marine rotors while minimizing
the power requirement and susceptibility to blade failure with [1] Khan AM. Flexible composite propeller design using constrained optimization
techniques. Ph.D. thesis, Iowa State University; 1997.
consideration for material and loading uncertainties. It was shown
[2] Friedmann PP, Venkatesan C, Yuan K. Development of a structural optimization
that the uncertainties in material stiffness parameters, considered capability for the aeroelastic tailoring of composite rotor blades with straight
as random variables, have a marginal effect on the hydroelastic and swept tips. In: 4th AIAA symposium on multidisciplinary analysis and
behavior of the self-twisting propeller. This due to the optimiza- optimization, Cleveland, OH; 1992.
[3] Ganguli R, Chopra I. Aeroelastic optimization of a composite helicopter rotor.
tion of the bendingtwisting coupling that produces a system In: 4th AIAA symposium on multidisciplinary analysis and optimization,
which is more sensitive to variations in h and J than to expected Cleveland, OH; 1992.
Author's personal copy

Y.L. Young et al. / Composite Structures 92 (2010) 244253 253

[4] Ganguli R, Chopra I. Aeroelastic tailoring of composite couplings and blade [18] Chen B, Neely S, Michael T, Gowing S, Szwerc R, Buchler D, et al. Design,
geometry of a helicopter rotor using optimization methods. J Am Helicopter fabrication and testing of pitch-adapting (exible) composite propellers. In:
Soc 1997;42(3):21828. The SNAME propeller/shafting symposium, Williamsburg, VA; 2006.
[5] Soykasap O, Hodges DH. Performance enhancement of a composite tilt-rotor [19] Nicholls-Lee RF, Turnock SR. Enhancing performance of a horizontal axis tidal
using aeroelastic tailoring. J Aircraft 2000;37:8508. turbine using adaptive blades. In: OCEANS 2007 Europe, Aberdeen, Scotland;
[6] Glaz B, Friedmann P, Lu L. Helicopter vibration reduction throughout the entire 2007.
ight envelope using surrogate-based optimization. J Am Helicopter Soc [20] Frangopol DM, Maute K. Reliability-based optimization of civil and aerospace
2009;54(1). systems. Engineering design reliability handbook. CRC Press; 2005. 24-124-
[7] Yang Y. Structural analysis and multidisciplinary design of exible uid loaded 32 [chapter 24].
composite canard. Ph.D. thesis, Iowa State University; 2005. [21] Allen M, Maute K. Reliability-based shape optimization of structures
[8] Lee AT, Flay RGJ. Compliant blades for passive power control of wind turbines. undergoing uidstructure interaction phenomena. Comput Method Appl M
Wind Eng 2000;24:311. 2005;194:347295.
[9] Lobitz DW, Veers PS, Eisler GR, Laino DJ, Migliore PG, Bir G. The use of twist- [22] Chamis CC. Design of smart composite structures in the presence of
coupled blades to enhance the performance of horizontal axis wind turbines. uncertainties. J Chin Inst Eng 2004;27:77181.
Tech rep, Sandia National Laboratories; 2000. [23] Lekou DJ, Philippidis TP. Mechanical property variability in FRP laminates and
[10] Lobitz DW, Veers PS. Load mitigation with bending/twist-coupled blades on its effect on failure prediction. Compos Part B Eng 2008;39:125674.
rotors using modern control strategies. Wind Energy 2003;6:10517. [24] Rais-Rohani M, Singh MN. Comparison of global and local response surface
[11] Gowing S, Cofn P, Dai C. Hydrofoil cavitation improvements with elastically techniques in reliability-based optimization of composite structures. Struct
coupled composite materials. In: Proceedings of 25th American towing tank Multidiscip O 2004;26:33345.
conference, Iowa City, IA; 1998. [25] Young YL. Time-dependent hydroelastic analysis of cavitating propulsors. J
[12] Lee Y, Lin C. Optimized design of composite propeller. Mech Adv Mater Struct Fluid Struct 2007;23:26995.
2004;11:1730. [26] Huang TT, Groves NC. Effective wake: theory and experiment. Tech rep, David
[13] Lin C, Lee Y. Stacking sequence optimization of laminated composite Taylor Naval Ship Research and Development Center, Bethesda, MD;
structures using genetic algorithm with local improvement. Compos Struct 1981.
2004;63:33945. [27] Choi JK, Kinnas SA. Prediction of non-axisymmetric effective wake by a three-
[14] Young YL, Michael TJ, Seaver M, Trickey ST. Numerical and experimental dimensional Euler solver. J Ship Res 2001;45:1333.
investigations of composite marine propellers. In: 26th symposium on naval [28] ABAQUS. ABAQUS version 6.5 documentation. ABAQUS, Inc.; 2005.
hydrodynamics, Rome, Italy; 2006. [29] Young YL, Liu Z, Motley MR. Inuence of material anisotropy on the
[15] Young YL, Liu Z. Hydroelastic tailoring of composite naval propulsors. In: 26th hydroelastic behaviors of composite marine propellers. In: Proceedings of
International conference on offshore mechanics and arctic engineering, San the 27th symposium on naval hydrodynamics, Seoul, Korea; 2008.
Diego, CA; 2007. [30] Plucinski M, Young YL, Liu Z. Optimization of a self-twisting composite marine
[16] Young YL. Fluidstructure interaction analysis of exible composite marine propeller using a genetic algorithm. In: Proceedings of 16th international
propellers. J Fluid Struct 2008;24:799818. conference on composite materials, Kyoto, Japan; 2007.
[17] Motley MR, Liu Z, Young YL. Utilizing uidstructure interactions to improve [31] Melchers RE. Structural reliability analysis and prediction. 2nd ed. Chichester
energy efciency of composite marine propellers in spatially varying wake. (NY): John Wiley; 1999.
Compos Struct 2009;90(3):30413.

You might also like