You are on page 1of 11

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 7 ( 2 0 1 2 ) 1 2 7 7 1 e1 2 7 8 1

Available online at www.sciencedirect.com

journal homepage: www.elsevier.com/locate/he

Determination of the autofrettage pressure and estimation


of material failures of a Type III hydrogen pressure vessel
by using finite element analysis
Dae-Sung Son, Jin-Ho Hong, Seung-Hwan Chang*
School of Mechanical Engineering, Chung-Ang University 221, Huksuk-Dong, Dongjak-Gu, Seoul 156-756, Republic of Korea

article info

abstract

Article history:

The autofrettage process of a Type III hydrogen pressure vessel for fuel cell vehicles with

Received 29 February 2012

preset winding pattern was simulated by finite element analysis (FEA). For a precise finite

Received in revised form

element analysis, the ply based modeling technique was used for the composite layers;

11 June 2012

a contour function was derived for the fibers at the dome part to determine the exact

Accepted 13 June 2012

winding angle; and the exact composite thickness was also considered. In order to deter-

Available online 12 July 2012

mine the most appropriate autofrettage pressure, stress analysis of the pressure vessel
according to its internal pressure was carried out with consideration of the international

Keywords:

regulations about pressure vessel design. The minimum stress ratio, the permanent

Type III hydrogen pressure vessels

volumetric expansion and the generated residual stress were investigated, and the failure

Autofrettage pressure

of the pressure vessel under minimum burst pressure was predicted by application of

Winding angle

various failure criteria of anisotropic composites.

Failure criterion

Copyright 2012, Hydrogen Energy Publications, LLC. Published by Elsevier Ltd. All rights
reserved.

1.

Introduction

Hydrogen gas produces no harmful byproducts when it burns,


so it has received much attention as a source of eco-friendly
energy [1,2]. To utilize hydrogen as a fuel source, various
storage methods have been developed. Among these
methods, the most commercialized method has been to
charge hydrogen gas in a pressure vessel at a high pressure.
One of the promising industrial fields to use hydrogen is the
automobile industry. But low energy density of hydrogen gas
has been the main reason for its low competitiveness against
existing fuels such as gasoline; therefore, high capacity
(70 MPa) pressure vessels are in demand for commercialization of hydrogen gas. To design a high capacity (70 MPa)
pressure vessel, a new material is essential, and fibrous
composites such as the carbon/epoxy composite is considered
one of the promising materials for fabricating Type III and

Type IV pressure vessels [3]. These pressure vessels are made


by the filament winding process, in which the liner is
completely wound by filaments at various winding angles. In
the filament winding process, various design factors such as
the winding pattern greatly affect the vessel performance.
Various algorithms to optimize these design factors were
studied [4e6]. Investigations on the performance of Type III
and Type IV pressure vessels according to the design factors of
the filament winding process were also carried out. Camara
et al. [7] and Biea et al. [8] investigated the fatigue behavior of
composite pressure vessels by using a statistical method and
continuum damage mechanics, and they estimated the
fatigue life of the pressure vessels for various conditions. Liu
et al. [9e11] formulated an estimation method for the property
degradation of composites according to the generated stresses
in pressure vessels and then, applied this estimating method
to FEA to predict the failure-inducing pressure. Hu et al. [12]

* Corresponding author. Tel.: 82 2 820 5354; fax: 82 2 814 9476.


E-mail address: phigs4@cau.ac.kr (S.-H. Chang).
0360-3199/$ e see front matter Copyright 2012, Hydrogen Energy Publications, LLC. Published by Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.ijhydene.2012.06.044

12772

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 7 ( 2 0 1 2 ) 1 2 7 7 1 e1 2 7 8 1

investigated the failure of pressure vessels caused by cracks in


the liner with consideration of the autofrettage process.
Tomioka et al. [13] investigated the failure strength of a pressure vessel according to the environmental temperature by an
experimental approach. Unlike the Type IV pressure vessels,
the Type III pressure vessel undergoes the autofrettage
process to enhance its fatigue characteristics [14]. The autofrettage process generates tensile residual stress in the
composite layers and compressive residual stress in the liner
because the excessive internal pressure (autofrettage pressure) makes the liner yield. Therefore, when service pressure
is applied to the vessel, a low level tensile stress is generated
in the material, which enhances fatigue strength. The
performance of the autofrettage process is affected by the
radius of the vessel, the composite thickness and modulus,
plastic behavior of the liner, the magnitude of the autofrettage
pressure, and so on [15]. Therefore, the autofrettage pressure
should be determined by considering the shape of the pressure vessel and the winding pattern of the carbon fibers.
In this paper, vessel performance according to the autofrettage pressure was investigated with consideration of
international regulations to determine the appropriate autofrettage pressure for a Type III hydrogen pressure vessel with
a pre-set winding pattern. The important factors for accurate
finite element analysis of the autofrettage process are the
consideration of the anisotropic property of the composites
and the information about the liners plastic behavior. Even
a slight variation of these factors may generate a big difference in the analysis result because of the involvement of nonlinear plastic deformation and the interaction of stress
distributions in the two materials. Therefore, the composite
layer has to be modeled precisely, and the exact non-linear
stressestrain relationship of the liner is needed. In this
paper, the ply based modeling technique [16], which can
provide the exact stress distribution for the composite layer,
and the exact stressestrain relationship for an aluminum
liner were used for the finite element analysis. With this
modeling technique, the stresses in orthotropic directions in
every single composite layer were calculated accurately. The
variable winding angles and composite thickness at the dome
part were also considered through a contour function of the
fibers. To determine the most appropriate autofrettage pressure, four major conditions from the international regulations
on the Type III pressure vessel design were applied.

with a solid object and a tie condition was imposed on the


interface of the boss and the liner. A simple contact condition was imposed on the interface between the liner and the
composite layers because these two parts are combined by
filaments in tension and bonded by excessive resins, which
is expected to be very weak [16]. The surface contact function provided by ABAQUS 6.9-1 was used to define contact
surfaces and impose friction coefficient on the interface. In
normal direction at the interface a hard contact, which is
a general contact condition, was applied and the friction
coefficient of 0.4 [17] was imposed on the interface in
tangential direction. For a precise stress analysis with
reduced computation time, only the small axisymmetric
part (5 ) of the pressure vessel was modeled by using surface
symmetric condition in the hoop direction. An 8 node 3D
rectangular element (C3R8R) was used for the modeling, and
a layup tool supplied by ABAQUS 6.9-1 was used for the ply
based modeling [16] of the composite laminate. The port and
the liner were made of aluminum (6061-T6), and the orthotropic property of the carbon/epoxy composite laminate
(MR60H-24K/epoxy, Grafil Inc., USA) was used for the
modeling of the composite layers. This prepreg had the
same type of carbon fiber and fiber volume fraction (65%) as
the filament winding structure. The material properties
used in the finite element analysis are listed in Table 1.
As mentioned, one of the most important factors for estimating the appropriate autofrettage pressure was the plastic
behavior of the liner. Therefore, the non-linear stressestrain
relationship of an aluminum 6061-T6 [18] was used in the
analysis. The composite laminate was composed of 132
windings (plies) with various winding angles (a hoop winding
(90 ), high helical windings (80 , 75 , 65 ) and helical windings
(30 , 25 , 23 , 21 , 18 , 15 , 13 )). The details of the winding
patterns were not specified for company confidentiality.
Because of the nature of the filament winding process, the
winding angle (J) contains simultaneously the positive (J)
and the negative (J) ply angles with respect to the axial axis,
so the balanced stacking sequence was used in the analysis.
Material degradation of the composite layers due to
temperature rise was not considered in this paper because the
generated temperature during fueling was expected not to
exceed 100  C [19] and the composite layers maintain its
original material property in this temperature condition [20].

2.2.

2.

Modeling

2.1.
vessel

Geometry and material properties of the pressure

The Type III hydrogen pressure vessel with capacities of


70 MPa and 113 L was investigated and its shape and
geometric details are shown in Fig. 1. It is composed of
a port, an aluminum liner and composite layers. The
aluminum liner was fabricated by deep drawing and ironing
processes so it has a single port, as shown in Fig. 1. At the
opposite part of the port, a knob was formed to support
filaments during the filament winding process. To simulate
the actual service condition of the vessel, the port was filled

Composite modeling

If a composite laminate is thin enough, the ply based


modeling and the laminate based modeling techniques give
almost the same results [21]. But for thick composite structures such as a Type III hydrogen pressure vessel (70 MPa
class), the ply based modeling technique provides much more
precise stress distribution [16]; therefore, the ply based
modeling technique was used for the accurate stress analysis
of the autofrettage process in this study. In the ply based
modeling technique, every single winding layer (ply) of the
pressure vessel was modeled by using the composite layup
tool provided by ABAQUS 6.9-1, and the variations of the
winding angle and thickness at the dome part were also
considered. To apply the variation of the winding angle of the
carbon filaments at the dome part to the finite element model,

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 7 ( 2 0 1 2 ) 1 2 7 7 1 e1 2 7 8 1

12773

Fig. 1 e Finite element model of a Type III hydrogen pressure vessel.

the fiber angle (a) was defined as the angle between the
meridian line and the fiber, which varies with the angular
position (F) and winding angle (J), as shown in Fig. 2a. It was
assumed that fibers did not slip and they traveled the shortest
path during the winding process. When the distance from an
arbitrary point D on the surface of the dome to the central axis
is ri the position of the point D can be expressed by (ricos A, risin
A, f(ri)) where the function f(ri) is defined by the dome geom!0
etry. The fiber angle (a) between a fiber direction F and the
!0
derivative of the meridian line V can be expressed by eq. (1).
And the angular position (F) and the length of ri have the
relationship of eq. (2);
ri f 0 ri  f ri
a tan1 qq
2
2
1 ff 0 ri g ri cot2 J  ff ri g
tan F

f ri
ri

(1)

(2)

By combining the eqs. (1) and (2), the fiber angle (a) can be
expressed as a function of the angular position (F) [16].
Therefore, once the winding angle (J) is known, the fiber angle
(a) at an arbitrary angular position (F) can be identified. The
variation of the fiber angle of helical windings (80 , 75 , 65 , 30 ,
25 , 23 , 21 , 18 , 15 , 13 ) was calculated, as shown in Fig. 2b.
The hoop winding (90 ) was not carried out at the dome part. To
consider the thickness variation along the angular position,
a prototype of the pressure vessel was cut and the actual
composite thickness of the dome part was measured, as shown
in Fig. 3a. At the dome part, there was no hoop winding, which
occupied a large portion of the winding angles at the cylinder
part, and the fiber density increased along the angular position
(see Fig. 3b); therefore, the total composite thickness at the
dome part increased overall along the angular position. The

initial decrease of the composite thickness was caused by the


absence of hoop winding and the small areal coverage of the
high helical winding fibers (80 , 75 , 65 ), which covered around
20 of the angular position, as shown in Fig. 2b.
To estimate a single ply thickness, it was assumed that the
composite thickness variation at the dome part did not
correlate with the winding angle. After considering the
coverage area of each fiber for different winding angles, the
single ply thickness was calculated, as shown in Fig. 3a. To
apply this thickness variation to the finite element model,
additional balanced plies were added to the existing
composite laminate, as shown in Fig. 3b, and this method
helped to simulate the actual winding pattern and the shape
of the dome part. And for realizing a more accurate winding
angle, the local coordinate system of each finite element was
used to align the fiber direction along the actual winding
direction. The additional modeling details are described in the
preliminary study [16].

3.
Determination of the autofrettage
pressure
3.1.

Conditions for determining autofrettage pressure

Among many international regulations on the design of


pressure vessels, the compressed gaseous hydrogen regulation [22], which was proposed by European Integrated
Hydrogen Project e Phase II (EIHP-II), and the ISO 15869:2009
[23] were considered in the determination of the appropriate
autofrettage pressure of a Type III hydrogen pressure vessel.
The important conditions for determining autofrettage pressure are as follows;

12774

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 7 ( 2 0 1 2 ) 1 2 7 7 1 e1 2 7 8 1

Table 1 e Material properties of aluminum (Al6061-T6)


and carbon/epoxy composite (HR60H-24K/epoxy).
Aluminum
E
v
syield
sultimate (true stress)

Carbon/epoxy composite
E1
E 2, E 3
G12, G13
G23
v12, v13
v23
Xt
Xc
Yt
S

Value
68.26 GPa
0.33
275.79 MPa
333.72 MPa

Value
165 GPa
8.56 GPa
4.39 GPa
2.7 GPa
0.326
0.43
3190 MPa
1440 MPa
82 MPa
141 MPa
0.019333
0.009579
0.032118

t
31
t
32

directions (fiber, in-plane transverse, and in-plain shear


directions) reach their corresponding material strengths
(Xt, Yt, S ) or failure strains 3 t1 ; 3 t2 ; G, material failure can be
expected.
s1
 1;
Xt

s2
 1;
Yt

31

32

t
31

 1;

t
32

 1;

s6
1
S
36

(3)

1

(4)

where the subscripts 1, 2, 6 and a superscript t represent the


fiber direction, in-plane transverse direction, in-plane shear
direction and tensile stress, respectively. Only tensile stress
was considered because the internal pressure in a pressure
vessel simply generates tensile stress in the material. But
these criteria do not consider the interactions of stresses
among the orthogonal principal directions, so material failure
may be predicted inaccurately for the multi-axial loading
condition (see Fig. 5). For a more reliable failure estimation
with consideration of the inter-stress effect on material
failure, the generalized Tsai-Wu failure criterion [24] was
considered, as shown in eq. (5).
Fi si Fij si sj  1 i; j 1; 2; .; 6

[Condition 1] Test pressure should not exceed autofrettage


pressure in an internal pressure test and it should be 1.5 times
the service pressure.
[Condition 2] The permanent volumetric expansion should
not exceed 5% under the test pressure condition.
[Condition 3] Minimum stress ratio defined as (the maximum
fiber stress at the minimum burst pressure)/(the maximum
fiber stress at service pressure) should be larger than 2.35.
[Condition 4] Failure should not occur below the minimum
burst pressure, which is to be 2.35 times the service pressure.
The most conservative condition from the related regulations was selected for the same test condition. But some
conditions on fatigue behavior and environmental tests,
which are beyond the scope of this current research, were
excluded. Based on [Condition 1], the autofrettage pressure
should be larger than 105 MPa. To determine the permanent
volumetric expansion, a series of loads should be imposed on
the pressure vessel after the autofrettage process, as shown in
Fig. 4a. To check [Condition 3] and [Condition 4], load history
(see Fig. 4b) was imposed on the pressure vessel and fiber
stresses were calculated for the service condition and the
minimum burst pressure condition, respectively. In these
calculations, the possibility of composite failure was inspected for the minimum burst pressure condition by applying
various failure criteria, as shown in Fig. 5. After examining all
the conditions above, the autofrettage pressure was
determined.

3.2.

Failure criteria for composite laminates

To investigate the composite failure of the pressure vessel


studied in this research, several failure criteria were introduced. The maximum stress and strain criteria expressed in
eqs. (3) and (4) can be simply applied for evaluating material
failure. Once the generated stresses or strains in orthotropic

(5)

where si and sj are ply stresses along the ply axes and Fi and Fij
are the 2nd and 4th order strength tensors, respectively,
which are determined by uniaxial and biaxial tests using
tensile, compressive and shear strengths. For the plane stress
condition (s3 s4 s5 0), this relation can be expressed in
a simple form as follows;
F11 s21 F22 s22 F66 s26 F1 s1 F2 s2 2F12 s1 s2  1

(6)

To determine the 2nd and 4th strength tensors, Hoffman


(eq. (7)) and Tsai-Wu (eq. (8)) suggestions, which are the same
relation except F12, were used;
1
1
1
; F22 t c ; F66 2 ;
Xt Xc
YY 
S

1
1
1
1
F2 t  c ; F12 
Y
Y
2 Xt Xc
F11

F1

1
1
 ;
Xt Xc

1
1
1
1
1
; F22 t c ; F66 2 ; F1 t  c ;
Xt Xc
YY
S
X
X
1
1
1 p
F11 F22  F12  0
F2 t  c ; 
Y
Y
2

(7)

F11

(8)

Because the failure envelop is affected greatly by the interstress strength tensor F12,as shown in Fig. 5, various biaxial
test methods to determine the appropriate strength tensor
F12 were studied to obtain a more reliable failure estimation
[25,26]. The first quadrant (s1 > 0, s2 > 0) of the failure
envelop is important because this region dominates brittle
failure, which leads to the failure of composite pressure
vessels. With the Tsai-Wu criterion, failure envelop varies
greatly according to the value of the strength tensor F12 (see
the yellow region in Fig. 5). High level of F12 represents a high
level of interaction between the stresses in the fiber direction
and those in the in-plane transverse direction under the
biaxial stress condition. But for a unidirectional composite,
the level of stress interaction may be small because the
modulus and strength in those two directions (fiber and inplane
transverse
directions)
are
quite
different.

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 7 ( 2 0 1 2 ) 1 2 7 7 1 e1 2 7 8 1

12775

Fig. 2 e Determination of fiber angle and areal coverage according to the angular position; (a) schematic diagram for defining
the fiber angle according to dome geometry, (b) variations of fiber angle according to the angular position and helical
winding angle.
Narayanaswami et al. [27] proved that the most accurate
failure prediction can be obtained when the inter-stress
strength tensor F12 becomes zero or the value of Hoffmans
suggestion. Because the value of F12 in Hoffmans suggestion
depends on the fiber directional strength (Xt, Xc), the value of
F12 becomes very low, almost zero, as shown in Fig. 5. In this
paper, the material failure of the composite part was estimated for the four failure criteria (maximum stress and
strain criteria, Tsai-Wu failure criterion, Hoffman failure
criterion), and the results were compared.

4.

Analysis results

4.1.

Permanent volumetric expansion

The permanent volume expansion (r) presented in [Condition


2] in the previous chapter is defined as the ratio of the
permanent volume change to the total volume change under
the test pressure condition after the autofrettage process, and
is expressed by the following relation;

12776

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 7 ( 2 0 1 2 ) 1 2 7 7 1 e1 2 7 8 1

Fig. 3 e Winding details at a dome part; (a) thickness changes in whole composite layers and a single ply at a dome part, (b)
fiber density and thickness variations according to the angular position.

Vzero2  Vzero1
 100%
Vtest  Vzero1

(9)

where Vzeroe1, Vzeroe2 and Vtest are the internal volumes under
the conditions of zero pressure-1 and zero pressure-2, and
under the test pressure condition in Fig. 4a, respectively. A low
level of permanent volumetric expansion represents high
reliability of gas storage capacity and low level of plastic
deformation of an aluminum liner, meaning that if the liner
deforms only in the elastic region under 1.5 times the service
pressure condition, it will be safer under the service condition
and even under the fatigue loading condition. The internal

volume at each loading condition shown in Fig. 4a is


summarized in Fig. 6a. As shown in the figure, the permanent
volumetric expansion under the test pressure decreased as
the autofrettage pressure increased and saturated after the
value of 107 MPa. As a result, autofrettage pressures larger
than or equal to 105 MPa satisfied the maximum allowable
permanent volumetric expansion of 5%.

4.2.

Minimum stress ratio

Minimum stress ratio is defined as the maximum fiber stress


at the minimum burst pressure divided by the maximum fiber

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 7 ( 2 0 1 2 ) 1 2 7 7 1 e1 2 7 8 1

a
Pressure

Test
Pressure

Autofrettage
Pressure

(105MPa)

Service
Pressure
(70MPa)

Zero
Pressure-1

Zero
Pressure-2

(0MPa)

(0MPa)

Step

Minimum Burst
Pressure

Pressure

(164.5MPa)

Test
Pressure
(105MPa)

Service
Pressure
(70MPa)

Zero
Pressure

stress at the service pressure. The minimum burst pressure


was 164.5 MPa, which was 2.35 times the service pressure
(70 MPa) as defined in EIHP-II [22]. The stress ratios decreased
as the autofrettage pressure increased, as shown in Fig. 6b.
This decrease was due to the fact that the fiber stress under
service pressure increased with the autofrettage pressure
because higher autofrettage pressures generated higher levels
of tensile residual stress in the composite layers. But the fiber
stress under the minimum burst pressure was seldom
affected by the level of autofrettage pressure because of the
relatively low level of the residual stress compared to the total
fiber stress. When the autofrettage pressures are in the ranges
of 105e107 MPa, the minimum stress ratio becomes higher
than 2.35, which satisfies [Condition 3].

4.3.
Failure prediction under the minimum burst
pressure

Autofrettage
Pressure

12777

Step

(0MPa)

Fig. 4 e Loading histories for determining (a) permanent


volumetric expansion, (b) the minimum stress ratio and
material failure.

The Type III hydrogen pressure vessel should be safe under


the minimum burst pressure (164.5 MPa) based on the regulation ([Condition 4]). Therefore, the possibility of any failures
both in the aluminum liner and the composite layers was
investigated by using finite element analysis. The four failure
criteria were used to estimate any material failures in the
composite layers. First, the stresses of the aluminum liner
under various pressure conditions were plotted against the
autofrettage pressure in Fig. 6b. The aluminum liner stresses
decreased as the autofrettage pressure increased under the
service pressure condition but they showed little change
under the minimum burst pressure condition. The maximum
stress (314.9 MPa) generated in the aluminum liner was lower
than the ultimate strength (333.72 MPa) of the aluminum liner,
so the liner will not fracture.
To estimate the failure of the composite layers of the
pressure vessel, the case of 107 MPa autofrettage pressure
amongst 105 MPa, 106 MPa and 107 MPa, which satisfied the

Fig. 5 e Comparison of various failure criteria.

12778

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 7 ( 2 0 1 2 ) 1 2 7 7 1 e1 2 7 8 1

Fig. 6 e Results of stress analysis for determining (a) the


permanent volumetric expansion, (b) the minimum stress
ratio at composite layer.

[Condition 3], was investigated. Once the pressure vessel that


experienced 107 MPa autofrettage pressure remained safe, the
other cases were considered to be theoretically safe because
higher autofrettage pressures generate higher residual
stresses in the composite layer of the pressure vessel. But the
generated fiber stress according to the autofrettage pressure
was almost uniform, as shown in Fig. 6b, so the failure
behavior of the composite layers was expected to be the same
for the other autofrettage pressure conditions.
To estimate the failure of the composite layers, the
maximum stress criterion was applied to the stress analysis
result. The failure index, defined as the ratio of the generated
stress to the material strength in each orthogonal direction, as
expressed in eq. (3), was plotted against characterized length
(x/L), as shown in Fig. 7. The characterized length (x/L) is
defined as the ratio of the contour path (x) of the pressure
vessel surface to the axial perimeter (L), as shown in Fig. 1. The
yellow surfaces in Figs. 7e9 represent failure surfaces. The
calculation result showed that the generated fiber stress and
in-plane shear stress were much lower levels than their corresponding strengths, so material failure was not predicted.
On the other hand, it showed almost a failure index of unity in
the transverse direction, which means the hoop winding and
helical windings of the composite laminate may fail by the
transverse tensile stress if the internal pressure increases due
to an abnormal situation such as a sudden temperature rise
during fuel injection [19]. The maximum strain criterion

Fig. 7 e The maximum stress criterion under the loading


condition of 107 MPa; (a) fiber direction, (b) in-plane
transverse direction, (c) in-plane shear.

produced the similar result to the maximum stress criterion


(see Fig. 8) but it predicted a safer condition, especially in the
transverse direction. This safe condition was due to the
Poissons ratio, which correlate the strains in other directions;
as a result, the Poissons ratio effect reduced the strains in the
transverse direction, which is more realistic.
In order to estimate the material failure more precisely,
second order failure criteria were also applied. For the case of
the Tsai-Wu criterion, the inter-stress strength tensor F12 was
set to be zero for the precise prediction of brittle failure under
the tensile stress condition [27]. The failure index expressed
by the left hand side of eq. (6) was calculated as shown in
Fig. 9a. The result was similar to the failure index distribution
of the in-plane transverse stress for the maximum stress
criterion, and this similarity explains why the transverse
tensile stress is the most harmful factor that leads to material
failure of this composite structure. The maximum failure
index of 0.786 is predicted at the outermost center part of the
cylinder of the helical windings. This result was similar to the
results of Hoffmans criterion (Fig. 9b), because in this criterion the inter-stress strength tensor is mainly related to the
very high fiber directional strengths, which make the strength

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 7 ( 2 0 1 2 ) 1 2 7 7 1 e1 2 7 8 1

12779

Fig. 9 e The second order failure criteria of orthotropic


materials; (a) Tsai-Wu criterion, (b) Hoffman criterion.

Fig. 8 e The maximum strain criterion under the loading


condition of 107 MPa; (a) fiber direction, (b) in-plane
transverse direction, (c) in-plane shear.

tensor very low, comparable to zero. Both of the two criteria


predict a sharp rise of stresses at the border between the dome
and the cylinder parts as shown in Fig. 9. This stress distribution may cause problems under a fatigue loading condition;
therefore, a special design consideration such as winding
pattern changes needs to be taken into account.
The failure analysis of the Type III hydrogen pressure
vessel with an autofrettage pressure of 107 MPa proved the
safety of the vessel without any material failures. It was also
proved that this autofrettage pressure satisfied all the conditions of a safe design of a Type III hydrogen pressure vessel.

5.

Discussions

In order to determine the autofrettage pressure of a Type III


pressure vessel, the four design conditions should be satisfied
and the generated stresses need to be low for high fatigue life.
First, the range of appropriate autofrettage pressures which
satisfied the four design conditions for the current pressure
vessel with a pre-set winding pattern were estimated
(105 MPae107 MPa). The most appropriate autofrettage pressure should be determined in this range by using the stress

analysis results. First, to enhance the fatigue characteristics,


a low rate of permanent volume expansion was preferable
based on [Condition 2]. Second, as the minimum stress ratio
increased, the composite layers became safer, but it raised
stresses in the liner. The preferable autofrettage pressure
should make the minimum stress ratio exceed the allowable
value of 2.35 slightly, and should minimize the liners stress
based on [Condition 3] because fibrous composites have good
fatigue characteristics. As a result, the autofrettage pressure
that guaranteed structural safety and the lowest permanent
volume expansion (see Fig. 6a), made the minimum stress
ratio exceed 2.35, and generated the lowest stress in the liner
was found to be 107 MPa, as shown in Fig. 6b.
A simple contact surface condition rather than a perfect
bonding condition at the interface was used to estimate the
stress distribution of a pressure vessel during autofrettage
process.
Based on the measurement of the bonding strength of
a carbon/epoxy composite-aluminum single lap joint [28] and
preliminary experimental investigation the level of bonding
strength of a carbon/epoxy composite-aluminum single lap
joint is below 15 MPa but the analyzed bonding stress of the
Type III hydrogen pressure vessel at the interface is higher
than 17 MPa for the case of perfect bonding condition, which
represents the bonding layer fails during the autofrettage
process. Therefore, the perfect bonding condition is not able
to represent the actual behavior of the interface during the
autofrettage process.
In order to check the effect of the friction coefficient at the
interface the lower level of friction coefficient (0.25) was also
applied and the result was compared with the case of the
higher friction coefficient (0.4) which was used for estimating
the stress distribution of the pressure vessel. The lower level
of friction coefficient (0.25) generated almost the same stress

12780

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 7 ( 2 0 1 2 ) 1 2 7 7 1 e1 2 7 8 1

distribution as the case of 0.4 friction coefficient in all the


materials under the loading history.
It seemed a friction coefficient within moderate range
seldom affected the stress distribution.
On the other hand, for the case of the perfect bonding
condition, it generated higher stress in the liner and the
composite layers under the service condition after the autofrettage process. This was caused by the fact that this condition did not allow any relative motion between the composite
layers and the aluminum liner at the interface. As a result, the
perfect bonding condition generated 15 MPa and 20 MPa
higher stresses in the cylinder part and in the dome part of the
liner, respectively than those of the case of surface contact
condition and this may affect the determination of the autofrettage pressure.
For more reliable design of the pressure vessel some more
design consideration needs to be investigated.
The temperature dependence on vessel mechanical
performance was not considered in this paper. Based on the
analysis result of temperature rise during the hydrogen
fueling process [19] it was expected that the temperature
would rise to around 100  C when the pressure vessel was
filling from empty to 70 MPa. The mechanical properties of the
carbon/epoxy composite used in this paper maintains its
original value until the service temperature of 100  C [20],
which represent the calculation results of the material
behavior and the autofrettage pressure are reliable. But the
temperature may rise over 100  C for unexpected circumstances, therefore it is needed to investigate the temperature
effect on the behavior of the pressure vessel including the
determination of the autofrettage pressure, which is a part of
our future work.
The stressestrain relationship of the aluminum liner
affects much the determination of the autofrettage pressure
because of the non-linear behavior of the aluminum liner. If
the yield stress increases by 5% with Youngs modulus
unchanged the liner stress under the service condition after
the autofrettage process will increase by 3.3%. But, if the yield
stress decreases by 5% the liner stress decreases by 10.9%
under the same condition. This represents the exact nonlinear relationship of the stressestrain curve is essential to
determine the appropriate autofrettage pressure. On the other
hand, the variation of Youngs modulus does not affect much
the stress distribution in the liner, which represents the
moderate changes of Youngs modulus does not contribute to
the changes of the autofrettage pressure.

6.

Conclusions

In this paper the appropriate autofrettage pressure for a Type


III hydrogen pressure vessel for fuel cell vehicles with a preset
winding pattern was determined by stress analysis with
consideration of the related international regulations.
Because the effect of the level of autofrettage pressure on
stress generation in materials and service life of a vessel is
great, the stress was accurately calculated by finite element
analysis. To calculate the stress distribution accurately in the
composite layers and the aluminum liner according to the
level of autofrettage pressure, the ply based modeling

technique for the composite layers was carried out to model


the pressure vessel. To estimate the changes in stress distributions in the materials of the pressure vessel such as those of
the generated residual stresses after the autofrettage process,
which induces yielding of the liner, a non-linear stressestrain
relationship was considered. Moreover, to calculate the exact
winding angle at the dome parts, a contour function was
derived for the fibers, and by using this, the exact areal
coverage of a fiber at a certain winding angle was accurately
estimated. And the thickness variation of the composite
layers at the dome part was also considered by measuring the
actual composite thickness after cutting the vessel prototype.
From the analysis results, it was found that as the autofrettage
pressure increased, the permanent volumetric expansion
decreased and the compressive residual stress in a liner
increased, which enhanced the structural performance of the
pressure vessel. To determine the most appropriate autofrettage pressure of the Type III hydrogen pressure vessel, the
four conditions ([Condition 1]e[Condition 4]) proposed by
European Integrated Hydrogen Project e Phase II and ISO
15869:2009 were considered. The appropriate autofrettage
pressure was determined to be 107 MPa.
The modeling technique and the process to determine the
autofrettage pressure presented in this paper can be utilized
in the optimal design of a Type III pressure vessel to determine
the optimal winding pattern and liner shape.

Acknowledgments
This research was supported by the Chung-Ang University
Excellent Student Scholarship and partly by the research
program of Estimation of the material property for FCV Type
III Hydrogen Storage Vessel (70 MPa) through Ministry of
Knowledge and Economy of Korea.

references

[1] Neelis ML, van der Kooi HJ, Geerlings JJC. Exergetic life cycle
analysis of hydrogen production and storage systems for
automotive applications. Int J Hydrogen Energy 2004;29:
537e45.
[2] Aceves SM, Berry GD, Martinez-Frias J, Espinosa-Loza F.
Vehicular storage of hydrogen in insulated pressure vessels.
Int J Hydrogen Energy 2006;31:2274e83.
[3] Zheng J, Liu X, Xu P, Liu P, Zhao Y, Yang J. Development of
high pressure gaseous hydrogen storage technologies. Int J
Hydrogen Energy 2012;37:1048e57.
[4] Xu P, Zheng J, Chen H, Liu P. Optimal design of high pressure
hydrogen storage vessel using an adaptive genetic algorithm.
Int J Hydrogen Energy 2010;35:2840e6.
[5] Kim CU, Hong CS, Kim CG, Kim JY. Optimal design of
filament wound type 3 tanks under internal pressure using
a modified genetic algorithm. Comp Struct 2005;71:16e25.
[6] Liu P, Xu P, Zheng J. Artificial immune system for optimal
design of composite hydrogen storage vessel. Comput Mat
Sci 2009;47:261e7.
[7] Camara S, Bunsell AR, Thionnet A, Allen DH. Determination
of lifetime probabilities of carbon fibre composite plates and

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 7 ( 2 0 1 2 ) 1 2 7 7 1 e1 2 7 8 1

[8]

[9]

[10]

[11]

[12]

[13]

[14]

[15]

[16]

pressure vessels for hydrogen storage. Int J Hydrogen Energy


2011;36:6031e8.
Bie H, Li X, Liu P, Liu Y, Xu P. Fatigue life evaluation of high
pressure hydrogen storage vessel. Int J Hydrogen Energy
2010;35:2633e6.
Liu PF, Zheng JY. Strength reliability analysis of aluminumcarbon fiber/epoxy composite laminates. J Loss Prev Process
Indust 2010;23:231e7.
Liu PF, Zheng JY. Recent developments on damage modeling
and finite element analysis for composite laminates:
a review. Mater Des 2010;31:3825e34.
Liu PF, Zheng JY. Progressive failure analysis of carbon fiber/
epoxy composite laminates using continuum damage
mechanics. Mater Sci Eng A 2008;485:711e7.
Hu J, Chandrashekhara K. Fracture analysis of hydrogen
storage composite cylinders with liner crack accounting for
autofrettage effect. Int J Hydrogen Energy 2009;34:3425e35.
Tomioka J, Kiguchi K, Tamura Y, Mitsuishi H. Influence of
temperature on the fatigue strength of compressedhydrogen tanks for vehicles. Int J Hydrogen Energy 2011;36:
2513e9.
Adibi-Asl R, Livieri P. Analytical approach in autofrettaged
spherical pressure vessels considering the Bauschinger
effect. J Press Vess-T 2007;129:411e9.
Jahromi BH, Ajdari A, Nayeb-Hashemi H, Vaziri A.
Autofrettage of layered and functionally graded
metaleceramic composite vessels. Comp Struct 2010;92:
1813e22.
Son DS, Chang SH. Evaluation of modeling techniques for
a type III hydrogen pressure vessel (70 MPa) made of an
aluminum liner and a thick carbon/epoxy composite for fuel
cell vehicles. Int J Hydrogen Energy 2012;37:2353e69.

12781

[17] Schon J. Coefficient of friction for aluminum in contact with


a carbon fiber epoxy composite. Tribol Int 2004;37(5):
395e404.
[18] Military handbook; December 1998. MIL-HDBK-5H.
[19] Kim SC, Lee SH, Yoon KB. Thermal characteristics during
hydrogen fueling process of type IV cylinder. Int J Hydrogen
Energy 2010;35:6830e5.
[20] Kang DH, Park SW, Kim SH, Kim CG, Hong CS. Measurement
of material properties of composites under high temperature
using fiber bragg grating sensors. J Korean Soc Compos Mater
2003;16:41e7 [Korean].
[21] Kim CU, Kang JH, Hong CS, Kim CG. Optimal design of
filament wound structures under internal pressure based on
the semi-geodesic path algorithm. Comp Struct 2005;67:
443e52.
[22] GRPE information group: hydrogen/fuel cell vehicles, draft
ECE compressed gaseous hydrogen regulation. EIHP II; 2003.
[23] ISO 15869. Gaseous hydrogen and hydrogen blends d land
vehicle fuel tanks; 2009.
[24] Lee DG, Suh NP. Axiomatic design and fabrication of
composite structures. New York: Oxford University Press;
2006.
[25] Pipes RB, Cole BW. On the off-axis strength test for
anisotropic materials. J Comp Mater 1973;7:245e56.
[26] Collins BR, Crane RL. A graphical representation of the
failure surface of a composite. J Comp Mater 1971;5:408e13.
[27] Narayanaswami R, Adelman HM. Evaluation of the tensor
polynomial and Hoffman strength theories for composite
materials. J Comp Mater 1977;11:366e77.
[28] Park SW, Kim HS, Lee DG. Optimum design of the co-cured
double lap joint composed of aluminum and carbon epoxy
composite. Comp Struct 2006;75:289e97.

You might also like