Вы находитесь на странице: 1из 8

Integral Design for Filament-Wound Composite Pressure Vessels

Integral Design for Filament-Wound Composite Pressure Vessels


Lei Zu, Sotiris Koussios, and Adriaan Beukers
Design and Production of Composite Structures, Faculty of Aerospace Engineering, Delft University of Technology,
Kluyverweg 1, 2629 HS Delft, The Netherlands

Summary
The most important issues for the integral design of a filament-wound pressure vessel reflect on the determination
of the dome shape and applied winding patterns. The goal of this paper is to determine the meridian profiles of
continuum-based domes for pressure vessels, and to demonstrate that the utilization of non-geodesic trajectories
forms a favorable alternative to the dome design. An integral methodology for the design of such dome structures
is outlined, with emphasis on the application of the non-geodesic winding law and the classical lamination theory.
Based on the condition of equal shell strains, the governing equation for the shape of the dome meridian and the
differential equation describing non-geodesic trajectories on the dome surface are derived. The meridian profiles of
non-geodesics-based carbon-epoxy domes are obtained for various slippage coefficients; the structural efficiency
of geodesics and non-geodesics-based domes for various polar radii are then calculated and compared to each
other. The results concluded that filament-wound domes of pressure vessels designed using the non-geodesics
provide better performance than geodesics-based ones.

1. Introduction overwound pressure vessels, which dome shapes, the fiber tension and the
are governed by the condition of equal interlaminar shear stress. The optimal
A typical pressure vessel consists of shell strains, were determined in3. The design of dome shapes in helically
a cylindrical section and two quasi- shape optimization for articulated wound pressure vessels used as
spherical domes with polar openings. pressure vessels consisting of several rocket motor cases was discussed by
Since the dome regions withstand domes axially stacked on each other Fukunaga et al.10, based on two kinds
the highest stress levels and are the was conducted based on the classical of criteria: the critical failure strength
most critical locations with regard lamination theory 4. Liang et al. 5 and the maximum performance factor.
to failure of the structures, the dome evaluated the effect of the dome depth
design is one of the most important on the structural performance and The majority of previous work
issues of designing composite pressure derived the optimal dome contour by has merely considered the dome
vessels. Extensive research has been maximizing the shape factor. In6 the design based on geodesic trajectories,
focused on the determination of the optimal meridian shapes and thickness and overlooked the application of
dome shapes, fiber trajectories and distributions in a filament-wound dome non-geodesics to the design of
lamination parameters based on the closure were investigated, in order to
meridian profiles and their related
continuum theory. The optimality improve shell buckling performance
roving paths. Geodesics show great
conditions for a dome-shaped shell under static external pressure. The
stability on a curved surface and
of revolution loaded with uniform effect of the roving bandwidth on the
calculability. However, since the
internal pressure and axial forces were stability of winding patterns in a dome
obtained by maximizing the stress geodesics are entirely determined
is evaluated in 7. De Jong8 compared the
invariant1. A multi-level optimization by the initial winding conditions,
dome profiles respectively determined
strategy for pressure vessels with restricting the winding trajectories
by the netting and the continuum theory
geodesic and ellipsoidal heads was to the geodesics certainly limits
and indicated that the vessel geometry
developed using a combined FEM/ the available design space and the
and performance are dependent on the
genetic algorithm 2. The optimal elastic properties of the used materials. performance improvement of pressure
dome profiles for non-geodesically Mitkevich9 presented the equilibrium vessels11-13.

In this paper we outline a design-


oriented methodology for the integral
design of non-geodesically overwound
dome structures subjected to uniform
©
Smithers Rapra Technology, 2011 internal pressure. With the aid of the

Polymers & Polymer Composites, Vol. 19, Nos. 4 & 5, 2011 413
Lei Zu, Sotiris Koussios, and Adriaan Beukers

optimality condition and the non-


geodesic law, the governing equations
(3) (4)
for the determination of meridian
profiles and adapted winding angle where Nφ, Nθ are the in-plane membrane The axial equilibrium of a shell of
distributions are given, while taking the forces per unit length in the meridional revolution can be formulated as
anisotropic characters of the reinforcing and circumferential directions. follows:
walls into account. The approach
assumes the dome wall as an orthotropic Note that we have sinΔ ≈ Δ for Δ → 0.
continuum and employs the classical (5)
From Eqs. (2) and (3), the relation
lamination theory to determine the for the membrane forces can thus be where r’ is the first derivative of r with
mechanical properties of the dome. The given by: respect to z. Hence:
approach is then demonstrated by three
typical composite materials, reflecting
on the most general design cases. The
calculations and comparisons of the Figure 1. Dome geometry and its parameters
performance factors for geodesically
and non-geodesically overwound
domes are carried out for various polar
opening radii, in order to evaluate the
effect of the non-geodesic winding on
the structural efficiency of composite
pressure vessels.

3. Governing Equations
3.1 Meridian Profiles
The geometry and loads of a dome,
which can be regarded as a generic shell
of revolution, are given in Figure 1.
The vector representation of a dome
structure in polar coordinates is:

(1)

where r, z, φ and θ denote the radial,


axial, meridional and circumferential
Figure 2. Geometry and loads of a surface element
coordinates, respectively.

We consider here an elementary piece


of a convex surface being part of a dome
(see Figure 2). The normal force of the
surface is given by14:

(2)

where p is the internal pressure; Rφ


and Rθ are the radii of curvatures in
the meridional and circumferential
directions, respectively.

According to the force equilibrium


in the normal direction of the surface
element, the normal force is:

414 Polymers & Polymer Composites, Vol. 19, Nos. 4 & 5, 2011
Integral Design for Filament-Wound Composite Pressure Vessels

We introduce now: where λ represents the slippage


coefficient of the roving bundles that
(6) are placed on the supporting surface.
(11)
Nθ is given by substitution of Eq. (6) Generally, the maximum allowable
where R is the radius at the equator. value of λ ranges from 0.15 to 0.6,
into (4):
Substituting Eqs. (6), (7) and (11) into depending on fiber/mandrel materials,
(8), the differential equation that links resin viscosity, fiber morphology, band
the geometry, the roving angle and the width, etc.18. The application of non-
(7) material properties of the dome shell geodesics significantly enlarges the
can be formulated by: design space for pressure vessels. The
where r” is the second derivative of r possibility appears now for modifying
with respect to z. the λ-value to provide more design
freedom for the winding paths.
We consider here dome shells made
by winding of unidirectional fibers. (12) Substitution of Eq. (12) into (13)
The classical lamination theory15 is results in:
employed in this study, that is, the role where ρ’ is the first derivative of ρ with
respect to ζ.
of the matrix is indeed considered.
As shown in1,8,16, an optimal pressure
vessel is governed by the condition of The non-geodesic law is here applied
equal shell strains or, in other words, to the design of roving trajectories. The
zero shear stress at lamina level. The differential equation for determining
the distribution of non-geodesic (14)
fiber strength is completely utilized
when the participating individual layers winding angle for a generic shell of
For the creation of C1 continuity for the
are aligned with the direction of the revolution is given by17:
roving trajectories when passing the
maximum principal stress. To satisfy equator, the derivative of the winding
the optimality condition of equal shell angle must have the same value as
strains, the ratio of the membrane forces the derivative of a geodesic at exactly
Nθ / Nφ is given by: that place. Therefore, the slippage
coefficient at that point should be equal
to zero. For this reason the following
(13) distribution for the slippage coefficient
(8) is assumed (Figure 3):
where α is the winding angle between
the directions of the roving path and Figure 3. Slippage coefficient λ as a function of ρ (μ=0.2, ρ0=0.4)
the dome meridian; k is the anisotropy
parameter, given by:

(9)

in which E1 and E2 are the elastic moduli


in respectively the longitudinal and
transverse directions of a unidirectional
layer; ν12 and ν21 are the Poisson’s ratios
satisfying the following symmetry
condition:

(10)

The specific laminate properties can be


characterized by the parameter k, Eq.
(9), which in fact expresses the degree
of orthotropy1.

Polymers & Polymer Composites, Vol. 19, Nos. 4 & 5, 2011 415
Lei Zu, Sotiris Koussios, and Adriaan Beukers

given by:

(15)

where μ is the maximum allowable


(17)
coefficient of friction between the
roving bundle and the supporting where γ is the specific weight of the used composite material, and zp is the height
surface, or between the roving bundle of the dome. The dimensionless performance factor and thickness are defined as:
and the previously wound material.

Substituting Eq. (15) into (14) gives:


(18)

where YT is the tensile strength, transverse to fibers.

Substitution of Eq. (18) into (17) yields:

(16)

Simultaneous solution of the system of


(19)
differential equations (12) and (16) will
finally provide the continuum-based Two assumptions support the calculation of the shell thickness in the dome part:
meridian profiles and related non- the fiber volume fraction is maintained consistently and the number of fibers in a
geodesic winding angle distributions cross-section is always constant19. With these assumptions, the laminate thickness
along the coordinate ζ. The initial set of along the meridional direction can be formulated as:
values consists of ζ0=0, ρeq=1, ρeq’=0.
Considering the manufacturability of
filament-wound pressure vessels, the
winding angle at the polar opening (20)
should rapidly approach the value Substitution of Eq. (20) into (19) leads to the final form of the objective function:
π/2. This condition is essential for
proceeding to the next wound circuit.
For the numerical solution, however,
a slightly reduced initial value for α is
rather desirable (herein α=0.4999π),
(21)
in order to avoid the infinities during
the solution procedure of equations
3.3. Determination of Laminate Thickness
(12) and (16). Under a given polar
radius ρ0, the resulting meridian profile The shell is assumed to consist of helically wound layers, each of which forms an
will strongly depend on the value for angle-ply configuration of ±α layers, as shown in Figure 4. The stress components
µ and the initial winding angle αeq. referred to the material axes in an individual layer are given by:
The {µ, αeq}-parameter set is able to
completely determine the meridian
shapes of non-geodesically overwound
domes, subjected to the predetermined
distribution of slippage coefficient.

3.2 Structural Performance


(22)
The performance factor I=PV/W is
used as an index for rating pressure where mij (i,j=1,2,3) are functions of the fiber angles and the material properties:
vessels, where P, V and W are the burst
pressure, internal volume and vessel
weight, respectively. Accordingly,
the performance factor of a dome is

416 Polymers & Polymer Composites, Vol. 19, Nos. 4 & 5, 2011
Integral Design for Filament-Wound Composite Pressure Vessels

(23)

in which c=cosα, s=sinα and α is the winding angle of the fiber path. Qij (i, j=1,2,6) are the components of the reduced
stiffness matrix, which are related to the laminate stiffness matrix [Q]:

(24)

in which the components of the laminate stiffness matrix [Q] are stated as:

(25)

where E1, E2, G12 are the elastic moduli in respectively the longitudinal and transverse directions and the shear modulus
of a unidirectional layer.

The failure of a composite pressure vessel includes generally two main steps: firstly, cracks appear in the matrix, and then
the pressure is taken up by the fibers until they fail20. In a commercial storage vessel, a leak-before-break safety assessment
plays a vital role for avoiding pressure loss and fluid leakage of pressure vessels. The matrix failure thus becomes a major
issue for the safety of a pressure vessel. In this study, the Tsai-Wu failure criterion21 is used:

(26)

in which the strength parameters F11, F22, F12, F66, F1, F2 and F6 are given by:

(27)

where XT, XC, YT, YC are the tensile and compressive strengths of the unidirectional layer in the fiber and transverse directions,
and S is the in-plane shear strength.

Substitution of Eq. (22) into (26) leads to a quadratic failure criterion in terms of the dimensionless thickness:

Polymers & Polymer Composites, Vol. 19, Nos. 4 & 5, 2011 417
Lei Zu, Sotiris Koussios, and Adriaan Beukers

Figure 4. Coordinate systems and loads of a symmetric lamination an interpolator between two extreme
cases: the sphere and the netting
solution. The results also show that
for the same material anisotropy,
the non-geodesics-based dome has
a slightly smaller volume and lower
aspect ratio (the ratio of the height
to the width) than the geodesics-
based one.

4.2 Evaluation of Structural


Performance
The structural performance is
evaluated by considering composite
materials with various fibers and
matrices: glass-epoxy (k=0.2645),
carbon-epoxy (k=0.0977) and aramid-
(28) epoxy (k=0.0706). Typical values of
the mechanical properties for these
where the dimensionless shell forces are given by:
composites are given in Table 115.
It should be noted here that the non-
geodesic trajectories is determined
using the coefficient of friction μ=0.2.
(29) Figure 6 shows the dimensionless
performance factors for geodesics
When a meridian shape and related non-geodesic trajectory are determined, the and non-geodesics-based domes,
minimally required strength-dominated thickness t (ρ) at each point assigned with polar opening radii (typical
on the dome is evaluated by solving Eq. (28). Then, the equatorial thicknesses t range [0-0.5]). Figure 7 illustrates the
eq
(ρ) corresponding to each t (ρ) are calculated using the “geometric” equation rates of performance improvement,
(20), and their maximum value t eqmax is selected as the final shell thickness at triggered by application of non-
the equator: geodesic trajectories. I G and I NG
are the performance factors of
geodesically and non-geodesically
overwound domes, respectively.
(30) It is revealed that the structural
performance of the dome improves by
Once t eqmax is obtained, the dimensionless performance factor I can be computed using the non-geodesic trajectories.
by Eq. (21) according to the Gaussian quadrature rule. The results also indicate that the
rate of performance improvement
can further be improved by using
4. Results and Discussion composite materials of which the
4.1 Meridian Shapes orthotropic character tends to the
netting configuration (k=0).
In the numerical solution procedure, the goal is now to provide a pair of
design variables {µ, αeq} that ensures a winding angle of 90° at exactly the
polar opening which should be achievable. Figure 5 shows the resulting 5. Conclusions
non-geodesics-based meridian profiles of carbon-epoxy polymeric domes This paper has presented an integral
(ρ0=0.4), with slippage coefficients range [0-1.5]. The continuous lines denote design method for the determination
the non-geodesics-based meridian shapes and the dashed lines stand for the of the continuum-based dome
geodesics-based meridians (i.e. µ≡0). The meridian profile is additionally shapes and related structural
governed by the material anisotropy parameter k. Such a dome would have a efficiency for non-geodesically
spherical shape in the case of isotropic materials (k=1). In general, due to the overwound pressure vessels, and
anisotropic character of the reinforced wall the resulting meridian shape is evaluated the effect of non-geodesic
oblate-spherical. For a vessel loaded solely by internal pressure, the “flatness” trajectories on the geometry and
of the dome is entirely dependent on k. The parameter k can be regarded as

418 Polymers & Polymer Composites, Vol. 19, Nos. 4 & 5, 2011
Integral Design for Filament-Wound Composite Pressure Vessels

Figure 5. Non-geodesically overwound carbon-epoxy dome profiles for various μ structural performance of domes.
The governing equations that relates
the meridian shape of the dome
and the trajectories of the rovings,
has been derived with the material
anisotropy parameter k. A specific
function has been chosen to describe
the distribution of the slippage
coefficient along the coordinate ρ for
the desired non-geodesic trajectories,
in order to ensure C1 continuity of
the roving paths when passing the
dome-cylinder conjunction. The
meridian shapes of non-geodesics-
based domes have been outlined for
various anisotropy parameters k. The
dimensionless performance factors
for geodesics and non-geodesics-
based domes have been respectively
calculated, in order to demonstrate
the gain in structural performance
Figure 6. Dimensionless performance factors of the geodesics and non-geodesics- that the non-geodesics can result in.
based domes (Carbon-epoxy, μ=0.2) It is concluded that the structural
efficiency of continuum-based
filament-wound pressure vessels
can be improved by the application
of the non-geodesic winding and the
optimality condition of equal shell
strains.

Although the shape differences in


the optimal meridian profiles of
geodesically and non-geodesically
overwound pressure vessels are
relatively small, the available design
space is sufficiently enlarged. This
particularly reflects on improved
structural performance, while still
being able to satisfy the conditions
of the winding process.

Table 1. Typical properties of the unidirectional composites15


Property Glass-epoxy Carbon-epoxy Aramid-epoxy
Longitudinal modulus (GPa) 60 140 95
Transverse modulus (GPa) 13 11 5.1
Shear modulus (GPa) 3.4 5.5 1.8
Poisson’s ratio ν12 0.3 0.27 0.34
Longitudinal tensile strength (MPa) 1800 2000 2500
Transverse tensile strength (MPa) 40 50 30
Longitudinal compressive strength (MPa) 650 1200 300
Transverse compressive strength (MPa) 90 170 130
In-plane shear strength (MPa) 50 70 30

Polymers & Polymer Composites, Vol. 19, Nos. 4 & 5, 2011 419
Lei Zu, Sotiris Koussios, and Adriaan Beukers

Figure 7. Vessel performance improvement by means of the non-geodesics (μ=0.2) 11. Zu L., Koussios S. and Beukers A.,
Int. J. Hydrogen Energy, 35 (2010)
660-670.
12. Koussios S., Beukers A. and
Stathis P.T., Proceedings of the
16th International Conference on
Composite Materials (ICCM/16),
Kyoto, Japan (2007).
13. Koussios S., Filament Winding: a
Unified Approach, Delft University
Press, Delft, The Netherlands
(2004).
14. de Jong Th., Koussios S. and
Bergsma O.K., Proceedings of the
14th International Conference on
Composite Materials, San Diego,
USA (2002).
15. Vasiliev V.V. and Morozov E.V.,
Advanced mechanics of composite
materials, Elsevier Ltd, UK (2007).
16. Wasiutynski Z. and Brandt A., Appl.
Mech. Rev., 16 (1963) 341-350.
17. Zu L., Koussios S. and Beukers A.,
Compos. Struct., 92 (2010) 2307-
References 6. Blachut J., Comput. Struct., 48 2313.
(1993) 153-160. 18. Koussios S. and Bergsma O.K., J.
1. Vasiliev V.V. and Krikanov A.A.,
7. Teng T.L., Yu C.M. and Wu Y.Y., Thermoplast. Compos., 19 (2006)
Compos. Struct., 62 (2003) 449-
Mech. Compo. Mater., 41 (2005) 5-34.
459.
333-340. 19. Park J.S., Hong C.S., Kim C.G.
2. Vafaeesefat A. and Khani A., Appl.
8. de Jong Th., A theory of filament and Kim C.U., Compos. Struct., 55
Compos. Mater., 14 (2007) 379-
wound pressure vessels, Report (2002) 63-71.
391.
LR-379, Structures and materials 20. Vasiliev V.V., Composite Pressure
3. Zu L., Koussios S. and Beukers A., laboratory, Faculty of Aerospace Vessels – Analysis, Design, and
Compos. Part A: Appl. Sci. Manuf., Engineering, Delft University of Manufacturing, Bull Ridge
41 (2010) 1312-1320. Technology, Delft, The Netherlands Publishing, Blacksburg, Virginia,
4. Zu L., Koussios S. and Beukers (1983). USA (2009).
A., Compos. Struct., 92 (2010) 9. Mitkevich A.B., Mech. Compos. 21. Tsai S.W. and Wu E.M., J. Compos.
339-346. Mater., 41 (2005) 497-504. Mater., 5 (1971) 58-80.
5. Liang C.C., Chen H.W. and Wang 10. Fukunaga H. and Uemura M.,
C.H., Compos. Struct., 58 (2002) Compos. Struct., 1 (1983) 31-49.
469-482.

420 Polymers & Polymer Composites, Vol. 19, Nos. 4 & 5, 2011

Вам также может понравиться