You are on page 1of 39

Ref.

Ares(2021)733360 - 29/01/2021

Deliverable Reference : D5.3

Title : Devices and materials for hydrogen storage evaluation

Confidentiality Level : PU

Lead Partner : NCSRD

Abstract : Various options exist for hydrogen storage including,


pressurised gas cylinders, insulated storage tanks under
vacuum for liquid H2, and metal hydrides. This
deliverable report presents the different hydrogen
storage solutions along with detailed finite element
analysis for the designs required in each scenario for
deep underwater operation.
EC Grant N° : 824348

Project Officer EC : Renata Kadric

ENDURUNS is co-funded by the Horizon 2020


Framework Programme of the European Union
Reference : ENDURUNS_D5.3
Version : 1.0.0
Date : 27/01/2021
Page : 1

D5.3: Devices and materials for hydrogen storage evaluation

DOCUMENT APPROVAL SHEET

Name Organization Date

Prepared and cross- Theodore Sterioitis NCSRD 27/01/2020


reviewed by:
Valter Luis Jantara Junior University of Birmingham 27/01/2021
Mayorkinos Papaelias University of Birmingham 27/01/2021
Reference : ENDURUNS_D5.3
Version : 1.0.0
Date : 27/01/2021
Page : 2

D5.3: Devices and materials for hydrogen storage evaluation

DOCUMENT CHANGE RECORD

Version Date Author Changed Reason for Change / RID No


Sections or
Pages

0.1.0 5.12.2020 Theodore Steriotis All Document structure initialization and input
on various hydrogen storage technologies

Mayorkinos Papelias

0.2.0 21.1.2021 Valter Luiz Jantara Junior All Input to finite element analysis for various
hydrogen storage technologies
Mayorkinos Papaelias

1.0.0 27.1.201 Theodore Steriotis All Final review


Mayorkinos Papaelias
Reference : ENDURUNS_D5.3
Version : 1.0.0
Date : 27/01/2021
Page : 3

D5.3: Devices and materials for hydrogen storage evaluation

Executive Summary
Various options exist for hydrogen storage including, gas cylinders at pressurisation of 350, 700 and
900 bar (the latter pressure limit is currently under evaluation), insulated storage tanks under vacuum
for liquid H2, and metal hydrides. Each of these approaches offers a different weight to volume energy
density. In this project, it has been decided that the energy storage device should be mission specific
and therefore, any of them could be employed depending on the exact mission profile and its power
requirements. In the case of liquid H2 being used, this will be done without the use of further supporting
cryogenic system other than an efficient thermally-insulated storage tank under vacuum. Since liquid
H2, will evaporate with time naturally at a rate lower than 1% per day, the minimum power consumption
should be closely matched with the H2 evaporation rate. In the case of metal hydrides, various options
are available including materials that are rechargeable and hence reusable. The metal hydrides to be
used will be selected based on the mission requirements. Gas cylinders will be predominantly used for
missions where operation is limited to shorter duration. This document presents an overview of the
technical details of each hydrogen storage method along with detailed finite element analysis for the
design requirements in each scenario. For the demonstration of the ENDURUNS concept, customised
hydrogen gas cylinders will be employed for powering the AUV and an off-the-shelf solution will be
employed for the USV.
Reference : ENDURUNS_D5.3
Version : 1.0.0
Date : 27/01/2021
Page : 4

D5.3: Devices and materials for hydrogen storage evaluation

Table of Contents

1 Introduction 5
1.1 Purpose 5
1.2 Structure 5
1.3 Applicable documents 5
1.4 Reference documents 5
1.5 Acronyms 5
2 Hydrogen storage options 6
3 Analysis of pressure housings for H2 storage 10
4 Conclusion 35
5 References 35
Reference : ENDURUNS_D5.3
Version : 1.0.0
Date : 27/01/2021
Page : 5

D5.3: Devices and materials for hydrogen storage evaluation

1 Introduction

1.1 Purpose
This purpose of this deliverable report is to present the technical details of each hydrogen storage
method along with detailed finite element analysis for the design requirements in each scenario
employed during actual operation. For the demonstration of the ENDURUNS concept, the consortium
has opted for customised hydrogen gas cylinders to be employed for powering the AUV and an off-the-
shelf solution will be employed for the USV.

1.2 Structure
This document is structured as follows:

Section 1 Introduction
Section 2 Hydrogen Storage Option
Section 3 Analysis of pressure housings for H2 storage
Section 4 Conclusion
Section 5 References

1.3 Applicable documents


AD1 ENDURUNS Grant Agreement

1.4 Reference documents


RD1 D2.1: Specification document on mission requirements
RD2 D2.2: Specification document on hybrid AUV glider
RD3 D2.3: Specification document on USV

1.5 Acronyms
COTS Commercial Off-The-Shelf
AUV Autonomous Underwater Vehicle
USV Unmanned Surface Vehicle
FEA Finite Element Analysis
Reference : ENDURUNS_D5.3
Version : 1.0.0
Date : 27/01/2021
Page : 6

D5.3: Devices and materials for hydrogen storage evaluation

2 Hydrogen storage options


The development of a vehicle with an extended autonomy that is able to perform long missions requires
a particular focus on the energy storage system. Electrically powered long-range vehicles require a large
amount of energy stored, which is proportional to the cruising distance.
Increasing the cruising distances and the energy demand, if the load is supplied only by rechargeable
batteries, their size and weight is destined to rise, decreasing the manoeuvrability and energy efficiency
of the vehicle.
The use of hydrogen fuel cells allows the direct production of electricity through the chemical reaction
between hydrogen and oxygen with higher gravimetric capacity compared to batteries. Thus by
rerouting part of the energy request to fuel cells the vehicle is expected to have a lighter weight load
and/or a more durable energy supply.
The use of hydrogen as energy carrier opens up a possibility to choose between different options of
storage, commonly grouped according to the three states of matter: solid, liquid and gaseous.
The first option, involves the storage of hydrogen in a solid state and in our case refers to metallic
hydrides, where hydrogen creates a solid solution by occupying the interstices of the metal lattice.
The second, requires to lower the hydrogen temperature below its boiling point, at -252.9 °C.
Finally, the third option, requires the compression of hydrogen in its gaseous form in order to reduce
the occupied volume and reach viable storage tank sizes.
Each one of these options has characteristics that make them good candidates for deep sea and long-
term storage, but also weaknesses that have to be taken into consideration in the final choice of the
approach to be used for specific cases.
The hydrogen storage in the solid state is one of the most studied approaches for applications in vehicles.
The high safety linked to the use of these materials is important, especially considering the higher
dangers that can be experienced by a moving vehicle, even more if with human operators on board.
Hydrogen can be stored in the solid state in a wide variety of materials, since by reacting with metals
hydrogen generates compounds called generically hydrides. These materials have extremely different
behaviour depending by the position of the metal in the periodic table and are commonly classified as:
Salt-like hydrides, molecular hydrides and metallic hydrides.
Salt-like hydrides are generated by hydrogen interacting ionically with alkali or alkaline earth metals
such as (Ca, Sr and Ba). These hydrides are easily oxidized, have high temperature of decomposition
and the dehydrogenation reaction is commonly irreversible. In molecular hydrides hydrogen is
covalently linked in a complex structure. These compounds have extremely high hydrogen gravimetric
and volumetric density. However, they often lack reversibility and are characterized by slow reaction
kinetics. These materials have in fact very good characteristics but are not for the moment
technologically ready for applications.
On the other hand, metallic hydrides have a more mature history of application including on-board
systems. Metallic hydrides are generally obtained as product of reaction of hydrogen and transition
metals. The materials obtained have very different characteristics and chemical behaviour, in relation
with the position of the metal in the periodic table. Hydrogen directly reacts with the metals generating
a MHn type hydride, often deviating from the ideal stoichiometry.
The hydrogen atoms occupy the interstices among the much bigger metal atoms, and for this reason the
metallic hydrides are commonly known also as interstitial hydrides. One of the most attractive
advantages of this hydrogen storage method is safety, due to the mild working conditions and the
possibility to handle in air the metal alloys.
Reference : ENDURUNS_D5.3
Version : 1.0.0
Date : 27/01/2021
Page : 7

D5.3: Devices and materials for hydrogen storage evaluation

In fact, there are various systems that are able to exchange hydrogen by changing the temperature of
the system in a range of temperatures very close to ambient temperature. Moreover, the hydrogen
exchange with the metal is fast and after the activation may occur in minutes. To this we can add that
after the decomposition and the selective introduction of the hydrogen atoms during the metal
hydrogenation process, all the impurities that are commonly present in the gasses are excluded, acting
as filter for the protection of the fuel cells.
The most common intermetallic compounds are identified with the following stoichiometric
compositions: AB5, AB, AB2, AB3 and A2B. The AB5 type alloys are well known and due to the easy
substitution of the elements in the A and B positions have been studied in a wide range of compounds.
Commonly the A site is occupied by atoms belonging to the rare earth metals, calcium, zirconium or
yttrium. These commonly associate with d-metal atoms (B-sites) like nickel, copper, cobalt, iron or
platinum.
One of the most iconic members of this family of compounds is LaNi5, a compound that is particularly
interesting for due to the ability to store hydrogen reversibly and with fast kinetics at moderate pressure
and temperature conditions. Even if the storage capacity is limited to a theoretical 1.5 wt%, the safety
issues but also the low prices of the alloy, and the possibility to achieve better sorption performances
or tailor the operating pressure and temperature conditions by simple partial substitution of lanthanum
and nickel with heteroatoms make these compounds very attractive.
AB alloys are another deeply studied family of composites and known for having a cubic structure.
TiFe is probably the most eminent compound of its family. TiFe is particularly attractive due to the
higher hydrogen capacities and the much lower price compared to LaNi5. However, the high equilibrium
pressure and the difficult activation constitute two important obstacles, although as for the AB5 alloys,
the substitution may help to overcome these problems. Among the AB2 compounds we find mostly
alloys having titanium and zirconium in the A sites while the B sites are commonly occupied by d-
metals. These alloys are generally less expensive than the those containing rare earths. The hydrogen
sorption capacity can be increased by adding vanadium in the lattice, even if the use of this metal is
discouraged by its price on the market. AB3 are well known alloys with lanthanum, cerium or yttrium
commonly occupying the A sites and transition metals in the B sites. LaNi3 is the most studied example
that belongs to this group, while the hydrogen storage capacity of this family of alloys is relatively low
even if atom substitutions can improve the overall behaviour. Finally, the A2B alloys are also well
studied, with Ti2Ni particularly interesting for hydrogen storage purposes, often modified by adding
other metals, particularly zirconium, to the lattice.[1–4]
A practical example of the metal hydride storage approach is the Japanese Long-range Cruising
Autonomous Underwater Vehicle named URASHIMA built in 1998. This vehicle, powered by
hydrogen fuel cells, with hydrogen stored in metal hydrides, was able to cruise 317 km in 60 hours. In
this specific case an AB5 type alloy was used (capacity 1.5 wt%)because it is an easy to handle, safe
material withworking temperatures that are favourable for onboard operations (hydrogenation
temperature below 15 °C, dehydrogenation over 45 °C and storage temperature of 20 – 25 °C).
A similar approach was used also in the development of the Type 212 submarines recently
commissioned by the German and Italian Navy. In these cases, even if the materials have a low
gravimetric hydrogen capacity, the water environment mitigates this factor and the safer conditions for
the vessel and the crew compensate the disadvantages. Moreover, in such massive systems the excess
heat can be recycled in order to produce energy.[5, 6] The weight penalty and the need of a system for
the control of the temperature can be optimal on big vessels, more problematic in smaller one, as in our
case.
The second option we took into consideration involves liquid hydrogen. Hydrogen has critical point at
-239.96 °C and 13 bar, with critical density of 31.43 kg/m3. Liquid hydrogen has been practically
adopted for a long time beside compressed hydrogen, and some of the first examples of on-board
Reference : ENDURUNS_D5.3
Version : 1.0.0
Date : 27/01/2021
Page : 8

D5.3: Devices and materials for hydrogen storage evaluation

application are the NASA space missions. Under atmospheric conditions liquid hydrogen has an energy
density three times higher than gaseous hydrogen compressed at 350 bar.
However, due to the extremely low temperatures involved, the process of hydrogen liquefaction is
highly energy demanding while adequate thermal insulation of the storage system is pivotal. The gas
liquefaction is often performed in special units near the production site, with the possibility to integrate
the processes in order to have an overall better energetic efficiency. In general, the energy consumption
involved in the production of liquid hydrogen is about one third of the energy that liquid hydrogen itself
can provide.
The energy required to liquefy hydrogen comprises the sensible energy (i.e. the energy that is necessary
to cool down the gaseous hydrogen from ambient temperature to – 253 °C), the “latent” condensation
energy (i.e. the energy required for the gas-to-liquid phase transition) and the ortho to para conversion
energy.
The ortho and para hydrogen are the two isomeric conditions of the di-hydrogen molecule and it
depends by the nuclear spin orientations of the two hydrogen atoms in the H2 molecule. The ortho-H2
molecule has two identical atomic spins and higher energy, while para-H2 has antiparallel spins and
lower energy.
In a state of thermal equilibrium at room temperature, H2 has approximately 75 % of ortho-H2 and 25
% of para-H2. The conversion from orto to para hydrogen releases about 527 kJ/kg and this phenomenon
is one of the factors leading to H2 evaporation.
Liquid H2 consists of 99.8 % para-H2 and the reduction of the orto-para conversion helps to reduce the
hydrogen boil-off. For the large-scale production of liquid hydrogen refrigeration cycles are generally
adopted. Commonly, the cycles adopted include processes of compression and expansion and pre-
cooling.
During storage and transportation, a constant boil-off of about 0.2 – 0.3 % per day occurs. This may
happen due to heat leakage, orto-para transitions and sloshing that generates heat. The boil-off must be
kept under control in the vessels in order to avoid either the pressure build-up and the loss of hydrogen
due to evaporation.
In order to reduce these problems, the vessels may integrate refrigeration systems or cryocoolers,
combinations with metal hydrides and liquid nitrogen cooling systems in order to further insulate the
tank walls. Hydrogen powered vehicles can exploit the boil-off and the constant hydrogen evaporation
in order to produce energy, minimizing thus the waste of hydrogen. Before the use liquid hydrogen
must be brought back to the gas state. The regasified H2 has an extremely high degree of purity, making
it particularly suitable for fuel cells power generation. [7]
On the light of the large amount of energy, the various precautions and the high level of technology and
costs necessary to produce and store liquid hydrogen but mainly the constant H2 boil-off thus is
incompatible with the intermittent nature of energy usage in the AUV (e.g. glider/normal mode) we can
conclude that despite the great hydrogen density, liquid hydrogen is not a convenient hydrogen storage
method for our purposes.
Finally, the conceptually simplest method to store hydrogen involves the compression of the gas under
high pressure at room temperature. The simplicity of this approach made this method the most
commonly adopted for hydrogen storage, including the mobility sector. At ambient temperature and
atmospheric pressure, 1 kg of hydrogen is a gas that occupies 11 m3, which is completely impractical.
The large-scale transportation of compressed hydrogen in wrought iron cylinders started since the 1880s
by the British troops, in order to inflate war balloons in Asia and Africa. The wrought iron cylinders
necessitated very thick walls in order to withstand the high pressures contained, giving an overall low
gravimetric hydrogen capacity to the system.
Reference : ENDURUNS_D5.3
Version : 1.0.0
Date : 27/01/2021
Page : 9

D5.3: Devices and materials for hydrogen storage evaluation

With the development of advanced steelmaking technologies, the use of new alloys and the introduction
of Cr-Mo steel cylinders, the carrying capacity of these containers has been raised up to 0.13 m3/kg.
Tanks of this family are fabricated industrially from different sources such as plates, billets or tubes.
However, the steel cylinders subjected to multiple filling/discharging operations appeared to be prone
to failure. The reason was found in the so-called hydrogen embrittlement mechanism.
Particularly affected by this phaenomenon were the cylinders often subject to loading stress and made
of high-tensile steel. These characteristics, concurrently with a sudden change of profile of the cylinder,
eventually coupled with minor manufactory internal defects could generate local stresses and a starting
point for the cylinder failure.
The hydrogen embrittlement problem was partially contained by regulating the characteristics of the
cylinders, limiting to 950 MPa the tensile strength of the steel used for hydrogen transport cylinders,
while the maximum size of the defects on the internal surface was limited to no more than the 3% or
the cylinders’ wall thickness. A way to go around the problem of hydrogen embrittlement involves the
use of aluminium alloy cylinders.
Aluminium structure has insufficient sites where hydrogen can create stress. However, because
hydrogen is often produced from the electrolysis of an aqueous solution of sodium chloride, with
mercury in one of the electrodes, hydrogen could carry mercury impurities, and even if in very small
percentages they could attack the aluminium surface leading to a structural failure. This, with the high
prices of high-pressure aluminium cylinders discourages their use for hydrogen transportation.
The full metal cylinders are convenient for stationary applications, but they not fulfil the requirements
for on-board applications, where higher hydrogen densities, i.e. higher pressures are required. Since the
1970s a new group of cylinders has been introduced. These are characterized by the presence of a liner
wrapped with an external fibrous material as reinforcement such as glass, boron, aramid and carbon in
different resin matrices.
The use of these materials, first introduced as life-support cylinders for firefighters or underwater
applications, was initially held back by the high costs. However, the increase of the demand led to the
decrease of the prices on the market, allowing the spread also in on-board applications for hydrogen
storage.
The spreading of these new cylinders is also due to the efficiency that they can achieve, of about 0.25
m3/kg. there are two main designs used in the fabrication of composite cylinders: hoop-wound and fully
wound designs.
For both the design the fibre is applied to a thin liner that may be made of steel, aluminium or even
polymer and acts as barrier for the gas. The hoop-wound design the composite is wrapped on the central
parallel section of the cylinder and gives a significant reinforcement to the structure. The fully wound
design on the other hand, has fibres wrapped on the whole body of the cylinder sharing the stress with
the internal metal liner.
This kind of cylinders can store hydrogen under higher pressure than the other mentioned models, and
are quite common in the market with working pressures of 350 bar but reaching even pressures of 700
bar with Aluminium liner wrapped with carbon fibre. The fibre in this family of tanks is commonly
wrapped around the liner with hoop, polar or helical patterns but also with a combination of the three.
Once the liner is wrapped, the resin must be cured, usually by heating.
An enhanced family of composite cylinders involves the use of an internal liner made of high-density
polyethylene (HDPE) fully wrapped with carbon fibres. This kind of vessel can withstand pressures of
about 700 bar and higher, due to the great strength/stiffness to weight ratio and to the good resistance
to fatigue and corrosion.
Reference : ENDURUNS_D5.3
Version : 1.0.0
Date : 27/01/2021
Page : 10

D5.3: Devices and materials for hydrogen storage evaluation

These characteristics give us low weight cylinders able to carry important amounts of hydrogen and
particular attractive for on-board applications. Moreover, the cylinders with HDPE liner suffer less
stress than aluminium ones during the refuelling of the cylinder and the temperature variations.
The fatigue strength of Aluminium liner cylinders decreases at low temperatures and increases at higher
temperatures, while an opposite behaviour is observed in cylinders with plastic liner. All the different
kind of cylinders presented are commonly grouped in four main groups or “Types”.
The tanks belonging to each type are characterized by similar peculiar characteristics that increase the
capacity to withstand higher pressures going from Type I cylinders to Type IV.
Type I tanks include the first mentioned only metal cylinders commonly made of stainless-steel or
aluminium. This kind of cylinder is economical but their nature limits the maximum working pressure
between 175 and 200 bar. This means a hydrogen capacity of about 1 wt%, not suitable for on-board
applications and mostly used for stationary systems.
Type II cylinders are the tanks with hoop wrapped metal liners. In this kind of cylinders, the fibre
surrounding the walls of the tank shares with the metal part of the stress, allowing higher working
pressures, generally between 260 and 300 bar. However, the internal metal liner gives susceptibility to
internal corrosion and an important weight to the overall system.
Type III cylinders include tanks with internal metal liner fully wrapped with composite fibre. The
presence of a full cover with light fibres allows high burst pressure, limits the hydrogen permeation and
doesn’t allow galvanic corrosion between the metal liner and the fibre.
This kind of tanks can work at higher pressures, between 350 bar and 700 bar, with hydrogen capacity
respectively of 3.9 wt% and 5 wt%.
Finally, Type IV cylinders are made of a plastic liner fully wrapped with composite fibre. Due to the
absence of a metallic structure, the tanks of this group are characterized by extremely low mass if
compared with any of the previous ones and despite this can withstand extremely high pressures.
Pressures of 700 bar are easily reachable, indicating a hydrogen capacity of about 5 wt% and more. The
absence of a rigid metallic liner increases the durability of the cylinders against repeated hydrogen
charging. These characteristics, together with the refinement of the fabrication techniques are pushing
for a wide diffusion of the Type IV cylinders that includes on-board applications.[8–10]
Composite cylinders, due to their mature technology and working simplicity, compared to all the other
options appear as the best method for the storage of hydrogen on-board, especially when considering
that metal hydrides tanks are exceptionally heavy and would require a system for the control of the
temperature, while liquid hydrogen, would require a highly insulated vessel, eventually refrigerated.
High pressure cylinders will require only on tank valves, pressure reducers and safety relief valves,
facilitating the assembly of our AUV hydrogen storage system.

3 Analysis of pressure housings for H2 storage


The first part of this section shows the buckling response evaluated via finite element analysis of
spherical and cylindrical pressure housings for the storage of hydrides or cryogenic liquid hydrogen for
powering the ENDURUNS AUV and USV.

Both pressure housings have similar volume. Three different alloys have been considered: Al7075 T6,
Ti-6Al-4V and stainless steel 316.

A linear buckling analysis was performed, followed by a nonlinear buckling analysis accounting for
imperfections of 2.5 mm, and a safety factor of 2.25. Different wall thicknesses were simulated, but
only vessels made out of Ti-6Al-4V would survive the intended pressure of 600 bar.
Reference : ENDURUNS_D5.3
Version : 1.0.0
Date : 27/01/2021
Page : 11

D5.3: Devices and materials for hydrogen storage evaluation

The spherical storage vessel was designed with radius of 150 mm. The cylindrical storage vessel was
designed with hemispherical domes with radius of 110 mm. The total length was 440 mm. The volume
of the spherical and cylindrical vessels is similar, 14.1 and 13.9 L respectively. The design for both
vessels can be seen in Figure 1.

Figure 1 Storage vessel drawings. Units in mm.

The simulations were conducted with the dimensions on Figure 1 being considered as the external
dimensions of the pressure vessel.

Three alloys were considered in this study, namely Al 7075 T6, Ti-6Al-4V, and stainless steel (SS) 316.
They were all simulated with a Johnson-Cook plastic constitutive rule.

The parameters used are shown in Table 1.

The true stress-strain curve obtained by using these parameters under uniaxial tensile loading conditions
can be seen in Figure 2.

Table 1 Material parameters [11-13].

Elastic parameters Johnson-Cook parameters


Young's Poisson's A B
Material n m
modulus (GPa) ratio (MPa) (MPa)
Al 7075 T6 71.7 0.33 473 210 0.3813 1
Ti-6Al-4V 109 0.34 860 683 0.47 1
SS 316 190 0.29 305 1161 0.61 1
Reference : ENDURUNS_D5.3
Version : 1.0.0
Date : 27/01/2021
Page : 12

D5.3: Devices and materials for hydrogen storage evaluation

1250

1000
True stress (MPa)

750

500

250

0
0% 5% 10% 15%
Al-7075 T6 True strain
Ti-6Al-4V SS 316

Figure 2 True stress-strain curves under uniaxial loading conditions for chosen materials.

The commercial finite element code ABAQUS 2017 was used in this research study. First, a mesh
convergence analysis was performed in order to define the final mesh to be used.

The pressure vessel was designed according to Figure 1 with full integration linear shell elements (S4).
The material employed was aluminium 7075 T6.

The number of integration points through the thickness of the vessels was fixed at 1 integration point
per mm, regardless of the wall thickness being simulated.

In the mesh convergence analysis, the thickness of the spherical vessel was 10 mm, whereas for the
cylindrical vessel it was 15 mm.

The external pressure applied was 600 bar in both cases.

Figure 3 shows both designs converge quickly, with the difference in the maximum equivalent stress
for the converged mesh and the finest mesh being less than 0.4% for both vessels.

An element size of 5 mm was then chosen as the final mesh, generating 13,336 elements for the
spherical vessel and 13,648 for the cylindrical vessel.

Figure 4 and Figure 5 show the boundary conditions (applied on symmetry planes) and equivalent stress
contour plot for converged meshes.

In all future simulations, the same boundary conditions were used.


Reference : ENDURUNS_D5.3
Version : 1.0.0
Date : 27/01/2021
Page : 13

D5.3: Devices and materials for hydrogen storage evaluation

500
Max equivalent stress (MPa) Sphere-10mm

450

400

350
600 6000 60000
Number of elements (log scale)

Figure 3 Mesh convergence analysis.

Figure 4 Boundary conditions (left), and equivalent stress contour plot for converged mesh (right).
Reference : ENDURUNS_D5.3
Version : 1.0.0
Date : 27/01/2021
Page : 14

D5.3: Devices and materials for hydrogen storage evaluation

Figure 5 Boundary conditions (left), and equivalent stress contour plot for converged mesh (right).

The vessels defined in the previous section were simulated with only external pressure acting on them,
as this is the loading condition that they will be subjected to when submerged.

Under this condition, buckling is the main failure mode, due to the high radius-to-thickness ratio of the
vessels. In order to quantify the maximum pressure that is supported by the structures, a linear
perturbation analysis was initially performed, and 3 eigenvalues were calculated.

The values for both vessels with different thicknesses and materials can be seen in Table 2.

Table 2 Eigenvalues for both spherical and cylindrical vessels with different thicknesses and materials.

Material Thickness Eigenvalue 1 Eigenvalue 2 Eigenvalue 3


Design
(mm) (bar) (bar) (bar)
Al Sphere 10 404 404 405
Ti Sphere 10 616 616 617
SS Sphere 10 1063 1063 1063
Al Sphere 15 913 913 913
Ti Sphere 15 1389 1389 1390
SS Sphere 15 2412 2412 2412
Al Sphere 20 1642 1643 1643
Ti Sphere 20 2499 2500 2500
SS Sphere 20 4342 4342 4343
Al Cylinder 15 213 213 225
Ti Cylinder 15 326 326 342
SS Cylinder 15 554 554 591
Al Cylinder 20 393 475 475
Ti Cylinder 20 598 726 726
Reference : ENDURUNS_D5.3
Version : 1.0.0
Date : 27/01/2021
Page : 15

D5.3: Devices and materials for hydrogen storage evaluation

SS Cylinder 20 1030 1235 1235


Al Cylinder 25 641 896 897
Ti Cylinder 25 978 1369 1369
SS Cylinder 25 1680 2340 2340

The linear analysis, however, does not take into account any material imperfections which will reduce
the structural integrity and strength of the component.

A nonlinear analysis was subsequently performed, as it is able to capture the effects of the geometric
imperfections generated in the linear analysis.

The nonlinear analysis uses the arc-length (modified Riks) method.

An imperfection of 2.5 mm was used in all cases.

As it is unlikely that a defect of such size would exist in the component, a high value is used to take the
worst case scenario into account.

Examples of node displacement contour plots for both vessels can be seen in Figure 6 and Figure 7. The
remaining simulations not shown have deformed in the same way.

Figure 6 Node displacement contour plots for spherical vessel with 15 mm thickness.
Reference : ENDURUNS_D5.3
Version : 1.0.0
Date : 27/01/2021
Page : 16

D5.3: Devices and materials for hydrogen storage evaluation

Figure 7 Node displacement contour plots for cylindrical vessel with 15 mm thickness.

Finally, a further safety factor of 2.25 was added as precaution. The final nonlinear buckling response
can be seen in Figure 8 and Figure 9, for the spherical and cylindrical vessel, respectively. For the
spherical vessel case, only vessels made out of titanium would survive the intended pressure of 600 bar,
and they would need to be at least 15 mm thick. Titanium was also the only material that would make
the cylindrical vessel survive 600 bar, but in this case the thickness would have to be at least 20 mm.

1000

900
Pressure with safety factor (bar)

800
Al-10 mm
700 Al-15 mm
600 Al-20 mm

500 Ti-10 mm
Ti-15 mm
400
Ti-20 mm
300
SS-10 mm
200 SS-15 mm
100 SS-20 mm
0
0 5 10 15 20 25 30 35 40
Maximum displacement (mm)

Figure 8 Nonlinear buckling response for spherical pressure vessel with different materials and thicknesses,
taking into account geometry imperfections of 2.5 mm and a safety factor of 2.25.
Reference : ENDURUNS_D5.3
Version : 1.0.0
Date : 27/01/2021
Page : 17

D5.3: Devices and materials for hydrogen storage evaluation

1000

900
Pressure with safety factor (bar)

800
Al-15 mm
700
Al-20 mm
600 Al-25 mm
500 Ti-15 mm

400 Ti-20 mm
Ti-25 mm
300
SS-15 mm
200
SS-20 mm
100 SS-25 mm
0
0 5 10 15 20 25 30 35 40
Maximum displacement (mm)

Figure 9 Nonlinear buckling response for cylindrical pressure vessel with different materials and thicknesses,
taking into account geometry imperfections of 2.5 mm and a safety factor of 2.25.

The maximum buckling pressure for the sphere was taken to be at 3 mm displacement, whereas for the
cylindrical vessel that was taken at 5 mm displacement.

These were then plotted against the vessel thickness and can be seen in Figure 10.

It shows that the stronger design and material combination would be a spherical vessel made out of
titanium, followed by cylindrical aluminium, spherical aluminium, spherical stainless steel, cylindrical
aluminium, and finally cylindrical stainless steel.
Reference : ENDURUNS_D5.3
Version : 1.0.0
Date : 27/01/2021
Page : 18

D5.3: Devices and materials for hydrogen storage evaluation

1000
Maximum buckling pressure with

800
safety factor (bar)

Sphere-Al
600
Sphere-Ti
Sphere-SS
400
Cylinder-Al
Cylinder-Ti
200
Cylinder-SS

0
10 15 20 25
Vessel thickness (mm)

Figure 10 Maximum buckling pressure (considering geometrical imperfections and safety factor) for different
designs and materials as a function of the vessel thickness.

The second part of this section analyses the pressure vessel design for the hydrogen storage system of
the ENDURUNS AUV based on gas cylinders.

Three different gas cylinder types (I, II, and III) were simulated and optimised with a safety factor of 3,
for the load cases of a full cylinder (300 bar) at sea level.

A nonlinear buckling analysis accounting for imperfections was then performed on the final designs for
each type, in order to find the maximum buckling pressure that each vessel could withstand.

The type-III pressure vessel with a 5 mm liner has the best trade-off when considering the results of
both analysis and weight.

The pressure vessel was designed as a cylinder of radius 150 mm.

Hemispherical domes with the radius equal to the cylinder were chosen, as in a sphere the pressure is
equally divided across the surface. The total length of the vessel is 800 mm.
Reference : ENDURUNS_D5.3
Version : 1.0.0
Date : 27/01/2021
Page : 19

D5.3: Devices and materials for hydrogen storage evaluation

Figure 11 Pressure vessel drawing. Units in mm.

The simulations were conducted with the dimensions on Figure 1 being considered as the internal
dimensions of the pressure vessel.

This was done for two reasons: (i) keeping the internal dimensions fixed avoids the need to redraw the
vessel at every simulation; (ii) by fixing the internal dimensions, a more severe loading case is created
when compared to the option of fixing the external dimensions, hence adding an extra safety factor.

The commercial finite element code ABAQUS 2017 was used. First, a mesh convergence analysis was
first performed in order to define the final mesh to be used in the next simulations.

The pressure vessel was designed according to Figure 1 with full integration linear shell elements (S4).
The “sweep” mesh strategy was chosen. The material employed was aluminium 7075 T6 with a
Johnson-Cook plastic constitutive rule. The parameters used shown in Table 13. The wall thickness was
15 mm with 21 integration points across the thickness, the external pressure 1 bar and the internal
pressure 300 bar, which corresponds to the cylinder full at sea-level, the most severe loading case.

Table 3 Material parameters for aluminium 7075 T6.

Elastic parameters Johnson-Cook parameters [14]


Young's Poisson's
A (MPa) B (MPa) n
modulus (GPa) ratio
71.7 0.33 473 210 0.3813

The initial mesh tested had an average element size of 20x20 mm, as shown in Figure 413. Finer mesh
sizes were tested down to an element size of 1x1 mm, which resulted in a total of 468,504 elements. As
Reference : ENDURUNS_D5.3
Version : 1.0.0
Date : 27/01/2021
Page : 20

D5.3: Devices and materials for hydrogen storage evaluation

the mesh convergence analysis on


500
Sphere-10mm
Max equivalent stress (MPa)

450

400

350
600 6000 60000
Number of elements (log scale)

Figure 312 shows, however, the results converge for a mesh of 5x5 mm, with 17,848 elements. The
difference in the maximum equivalent stress for the converged mesh and the finest mesh was only
0.05%. Figure 4 shows the meshes and equivalent stress contour plot for the coarse and converged
mesh. In all future simulations, unless otherwise stated, the same boundary conditions were used, as
this corresponds to the most severe loading case possible. The only changes were the materials and wall
thickness.

270,0
Maximum equivalent stress (MPa)

Converged mesh

267,5

265,0
500 5000 50000 500000
Number of elements (log scale)

Figure 12 Mesh convergence analysis.


Reference : ENDURUNS_D5.3
Version : 1.0.0
Date : 27/01/2021
Page : 21

D5.3: Devices and materials for hydrogen storage evaluation

Figure 13 Equivalent stress contour plot for coarse mesh with 931 elements (left), and for converged mesh with
17,848 elements (right).

The first pressure vessel to be considered is the type-I pressure vessel, which is fully made of metal. In
this case, aluminium 7075 T6.

The maximum stress calculated both at the cylinder and at the dome of the pressure vessel are shown
in Figure 14.

The safety factor (SF) is calculated as the ration between the yield stress and the maximum stress. The
safety factor for both the dome and cylinder area are shown in Figure 15.
Reference : ENDURUNS_D5.3
Version : 1.0.0
Date : 27/01/2021
Page : 22

D5.3: Devices and materials for hydrogen storage evaluation

500 YIELD STRESS (SF=1)

400
Equivalent stress (MPa)

66% YIELD STRESS (SF=1.5)


300

200
33% YIELD STRESS (SF=3)

100
Max cylinder stress
Max dome stress
0
5 10 15 20 25 30
Thickness (mm)

Figure 14 Maximum equivalent stress at the cylinder and at the dome of a type-I pressure vessel, as a function of
thickness.

5
Cylinder
Dome
4
Safety factor

0
5 10 15 20 25 30
Thickness (mm)

Figure 15 Safety factors for the cylinder and at the dome of a type-I pressure vessel, as a function of thickness.

It can be seen that a pressure vessel thinner than 9 mm would yield at the cylinder area. A thickness of
17 mm would give a safety factor of 2, whereas a thickness of 26 mm would be needed for a safety
factor of 3, which is usually the requirement for these components. The stress contour plot for this vessel
can be seen in Figure 16.
Reference : ENDURUNS_D5.3
Version : 1.0.0
Date : 27/01/2021
Page : 23

D5.3: Devices and materials for hydrogen storage evaluation

Figure 16 Equivalent stress contour plot for type-I pressure vessel with 26 mm thickness.

A type-II pressure vessel consists of a metal liner reinforced in the cylinder area by fibres. The liner
was made of the same aluminium 7075 T6, while the reinforcing fibre material employed in this section
was T700 fibre/epoxy composite. Table 44 and Table 55 show the material parameters used.

Table 4 Elastic material parameters for T700 fibre/epoxy [2].

Young's modulus (GPa) Shear modulus (GPa) Poisson's ratio


E1 E2 E3 G12 G13 G23 ν12 ν 13 ν 23
141 11.4 11.4 7.1 7.1 7.1 0.28 0.28 0.4

Table 5 Strength parameters for T700 fibre/epoxy [2].

Strength parameters (MPa)


Fibre tensile Fibre compressive Matrix tensile Matrix compressive Shear
strength (XT) strength (XC) strength (YT) strength (YC) strength
2080 1250 60 290 110

The safety factor (SF) for the aluminium liner is calculated in the same way as type-I. The Tsai-Hill
failure parameter was chosen to analyse the composite layer as this is the most conservative approach
in comparison to Tsai-Wu and maximum stress failure theory [14], as well as the Hashin failure criteria,
as is shown in Figure 17. The Tsai-Hill [15], as well as all other damage criteria range from 0 for the
structurally sound material to 1 for a completely failed material. The safety factor for the composite
Reference : ENDURUNS_D5.3
Version : 1.0.0
Date : 27/01/2021
Page : 24

D5.3: Devices and materials for hydrogen storage evaluation

layer will be calculated as the inverse of maximum damage value obtained in each simulation, similarly
to ref. [16].

0,6
Maximum damage parameter

0,3

0
Tsai-Hill Maximum stress Tsai-Wu Hashin

Figure 17 Comparison of maximum damage parameters of a type-II pressure vessel with an Al liner of 17 mm
and a T700 hoop layer of 7 mm.

The first analysis performed aimed to maintain a thickness of 25 mm around the cylinder, while varying
the aluminium and composite thicknesses.

For a composite thickness of 10 mm, for example, the liner thickness would then be 15 mm. The results
of this preliminary analysis can be seen in Figure 18.

The safety factors for the simulations can be seen in Figure 19. In this graph, we then identify an area
of interest (highlighted in blue) with the aluminium liner thickness varying from 15 to 19 mm.
Reference : ENDURUNS_D5.3
Version : 1.0.0
Date : 27/01/2021
Page : 25

D5.3: Devices and materials for hydrogen storage evaluation

350

300
Equivalent stress (MPa)

250

200

150

100 Max composite stress


Max liner cylinder stress
Max liner dome stress
50
0 2 4 6 8 10
Composite layer thickness (mm)

Figure 18 Equivalent stress as a function of composite layer thickness. The total thickness at the hoop is kept
constant at 25 mm, and the ratio between liner and composite thickness is changed.

4
Composit
e
Safety factor

2
0 2 4 6 8 10
Composite layer thickness (mm)

Figure 19 Safety factor as a function of composite layer thickness. The total thickness at the hoop is kept
constant at 25 mm, and the ratio between liner and composite thickness is changed.

A sensitivity study was then performed and is shown in


Reference : ENDURUNS_D5.3
Version : 1.0.0
Date : 27/01/2021
Page : 26

D5.3: Devices and materials for hydrogen storage evaluation

Figure 20 Safety factor as a function of composite layer thickness for a range of fixed liner thickness.

Figure . For all the different liner thicknesses, there is a point in which the addition of more composite
layers does not increase the safety factor of the pressure vessel. This is due to the fact that the dome,
which is not reinforced by composite in type-II pressure vessels, becomes the weakest area of the vessel.

3,5
Liner
19 mm
thickness
18 mm
17 mm
16 mm
15 mm
Safety factor

2,5
3 4 5 6 7 8 9 10 11 12 13 14 15
Composite layer thickness (mm)

Figure 20 Safety factor as a function of composite layer thickness for a range of fixed liner thickness.

Figure 20 shows the safety factor as a function of composite layer thickness for a range of fixed liner
thicknesses. Circle markers indicate that the lowest safety factor is at the liner dome, squares at the liner
cylinder and triangles at the composite. Based on

Figure 20 Safety factor as a function of composite layer thickness for a range of fixed liner thickness.

Figure , the thinnest vessel able to obtain a safety factor of 3 would have an aluminium liner of 17 mm
and a composite layer 7 mm thick. The equivalent stress and Tsai-Hill contour plots for these
dimensions can be seen in Figure 21.
Reference : ENDURUNS_D5.3
Version : 1.0.0
Date : 27/01/2021
Page : 27

D5.3: Devices and materials for hydrogen storage evaluation

Figure 21 Equivalent stress contour plot (left), and Tsai-Hill failure criteria contour plot (right). The liner is 17
mm and the composite layers are 7 mm thick in total.

As shown in

Figure 20 Safety factor as a function of composite layer thickness for a range of fixed liner thickness.

Figure , the dome is the limiting factor in making the liner of type-II vessels thinner. Type-III pressure
vessels address this problem by completely wrapped the liner with reinforcing fibre material, including
the domes. This is done by employing helical wrapping, which is able to wrap the dome and cylinder
area of the vessel. If additional reinforcement is needed, hoop layers can be added, similarly to the type-
II vessel. Three different liner thicknesses were chosen to be simulated and be optimised until a safety
factor of 3 was achieved, namely 15 mm, 10 mm and 5 mm.

As a starting point, these vessels were simulated with only helical layers. The total thickness was kept
constant to 24 mm in all three cases, meaning that the composite thicknesses were 9 mm, 14 mm and
19 mm, respectively. Each composite layer was 0.5 mm thick, and the stacking sequence was (+α,-α)n,
where α is the helical angle, and n is half of the total number of layers. As the dome is the limiting
design factor, it is needed to identify the optimal helical angle f increasing the dome safety factor. This
analysis is shown in Figure 22. An angle range between 45° and 60° chosen to be iterated and simulated.
Reference : ENDURUNS_D5.3
Version : 1.0.0
Date : 27/01/2021
Page : 28

D5.3: Devices and materials for hydrogen storage evaluation

4,50
Liner 15 mm
Liner 10 mm
3,50 Liner 5 mm
Dome Safety factor

2,50

1,50

0,50
15 25 35 45 55 65 75
Helical angle (°)

Figure 22 Dome safety factor as a function of the helical angle for type-III pressure vessels with no hoop
reinforcing layers. The total thickness of all vessels is 24 mm.

An iterative process of analysis was then carried out for all three cases. The idea is to find the thinnest
helical layer able to bring the dome safety factor as close as possible to 3. When this is achieved, the
cylinder will need to be reinforced with hoop layers, as its safety factor will be below 3. Adding the
hoop layer also reinforces the end areas of the dome, slightly increasing the safety factor of the dome.
The helical thickness can then be decreased or the helical angle can be changed – and the whole process
is repeated – until it can no longer be improved. This is then the optimised design. Table 66 shows a
list of different designs that were simulated until the optimal design was obtained.

Table 6 List of designs simulated in order to obtain optimal design for a type-III pressure vessel.

Helical
Hoop layer
Helical layer
Liner (mm) thickness Cylinder SF Dome SF
angle (°) thickness
(mm)
(mm)
15 60 9 0 2.7 3.5
15 60 7 0 2.5 3.2
15 60 7 5 3.0 3.1
15 60 5 5 2.9 2.9
15 60 6.5 5 3.0 3.0
15 60 5 5 2.9 2.9
10 60 14 0 2.5 3.1
10 60 14 5 2.8 2.7
10 60 17 5 3.1 3.0
Reference : ENDURUNS_D5.3
Version : 1.0.0
Date : 27/01/2021
Page : 29

D5.3: Devices and materials for hydrogen storage evaluation

10 60 18 4 3.1 3.2
10 60 17.5 4 3.0 3.1
10 60 16.5 4.5 3.0 3.0
10 52 16.5 4.5 3.3 3.4
10 52 13 5 3.0 3.0
10 60 16 4 3.0 2.9
10 60 12.5 6 2.8 2.5
5 45 19 0 1.9 2.7
5 45 25 0 2.4 3.3
5 45 25 7 3.7 3.8
5 45 20 7 3.2 3.2
5 45 18 7 3.1 3.0
5 45 19 6.5 3.1 3.1
5 45 18.5 6.5 3.0 3.0
5 52 18.5 6.5 3.0 2.7
5 45 15 10 3.1 2.6
Figure 23 shows a summary of optimal designs for the three cases. As the liner thickness decreases, so
does the pressure vessel overall weight. The overall thickness, however, will increase. All three pressure
vessels have the same safety factor of 3, and the choice should be based on whether the designer wants
to optimise the weight or volume of the vessel.
Reference : ENDURUNS_D5.3
Version : 1.0.0
Date : 27/01/2021
Page : 30

D5.3: Devices and materials for hydrogen storage evaluation

40 65
Liner Composite Helical Composite Hoop Weight

60
30
Thickness (mm)

55

Weight (kg)
20
50

10
45

0 40
Cylinder Dome Cylinder Dome Cylinder Dome
Type-III 15 mm (α=60°) Type-III 10 mm (α=52.5°) Type-III 5 mm (α=45°)

Figure 23 Optimal design for type-III pressure vessels with different liner thicknesses able to withstand an
internal pressure of 300 bar with a safety factor of 3.

Figure 24 Equivalent stress contour plot (left), and Tsai-Hill failure criteria contour plot (right), for optimised
the type-III 5 mm liner case.

The five pressure vessel designs defined in the previous sections were simulated with only external
pressure acting on them, to reproduce the boundary conditions of submerged pressure vessels. Under
Reference : ENDURUNS_D5.3
Version : 1.0.0
Date : 27/01/2021
Page : 31

D5.3: Devices and materials for hydrogen storage evaluation

this conditions, buckling is the main failure mode, due to the high radius-to-thickness ratio of the vessel.
In order to quantify the maximum pressure that is supported by the structure, a linear perturbation
analysis was initially performed, and 3 eigenvalues were calculated. The values for each one of the
structures can be seen in Table 27.

Table 7 Maximum buckling pressure for different pressure vessels at different eigenvalues.

Maximum linear buckling pressure (bar)


Type-I Type-II Type-III 15 Type-III 10 Type-III 05
Eigenvalue 1 1693 1198 1404 1448 1445

Eigenvalue 2 1693 1198 1404 1448 1445

Eigenvalue 3 2627 2198 2656 2689 2659

Failure pressure 1693 1198 1404 1448 1445

The linear analysis, however, does not take into account any material imperfections which will reduce
the maximum pressure allowed. A nonlinear analysis was subsequently performed, as it is able to
capture the effects of the geometric imperfections generated in the linear analysis. The nonlinear
analysis uses the arc-length (modified Riks) method. The results for each one of the designs can be seen
in Figure 25.

1500

1250
External pressure (bar)

1000

750

500

250

TI TII TIII-15 TIII-10 TIII-05


0
0 5 10
Arc length (modified Riks method)

Figure 25 Nonlinear buckling response of five different cylinder designs subjected to external pressure.

The simulation concerning type-II pressure vessel encountered too many numerical instabilities and
always failed on the onset of buckling. The pressure obtained before the end of the simulation will be
Reference : ENDURUNS_D5.3
Version : 1.0.0
Date : 27/01/2021
Page : 32

D5.3: Devices and materials for hydrogen storage evaluation

considered as the maximum buckling pressure. The node displacement contour plots for all five designs

can be seen in

Figure 6 Node displacement contour plots for spherical vessel with 15 mm thickness.

to Figure 30. As discussed, no deformation is visible for the type-II pressure vessel (

Figure 7). All the remaining designs have buckled on the hoop section.
Reference : ENDURUNS_D5.3
Version : 1.0.0
Date : 27/01/2021
Page : 33

D5.3: Devices and materials for hydrogen storage evaluation

Figure 26 Node displacement contour plots for type-I pressure vessel.

Figure 27 Node displacement contour plots for type-II pressure vessel.


Reference : ENDURUNS_D5.3
Version : 1.0.0
Date : 27/01/2021
Page : 34

D5.3: Devices and materials for hydrogen storage evaluation

Figure 28 Node displacement contour plots for type-III pressure vessel with 15 mm Al liner.

Figure 29 Node displacement contour plots for type-III pressure vessel with 10 mm Al liner.
Reference : ENDURUNS_D5.3
Version : 1.0.0
Date : 27/01/2021
Page : 35

D5.3: Devices and materials for hydrogen storage evaluation

Figure 30 Node displacement contour plots for type-III pressure vessel with 5 mm Al liner.

The maximum pressure allowed for each one of the vessels can be seen in Table 88. The best
performance is obtained by the type-III pressure vessel with 5 mm liner, which is able to withstand 894
bar of external pressure before buckling.

Table 8 Nonlinear buckling pressure for five different pressure vessel designs.

Nonlinear buckling pressure (bar)


Type-I Type-II Type-III 15 Type-III 10 Type-III 05
761 700 828 853 894
Different pressure vessel types were simulated to operate with internal pressure of 300 bar and a safety
factor of 3.

All pressure vessels in Figure 31 are viable options, however, a type-II vessel with an aluminium liner
of 17 mm and a 7 mm composite hoop layer would achieve the best trade-off between weight and
thickness.
Reference : ENDURUNS_D5.3
Version : 1.0.0
Date : 27/01/2021
Page : 36

D5.3: Devices and materials for hydrogen storage evaluation

40 65
Liner Composite Helical Composite Hoop Weight
60
30
Thickness (mm)

Weight (kg)
55
20
50

10
45

0 40
Cylinder Dome Cylinder Dome Cylinder Dome Cylinder Dome Cylinder Dome
Type-I Type-II Type-III 15 mm Type-III 10 mm Type-III 5 mm
(α=60°) (α=52.5°) (α=45°)

Figure 31 Optimal design for different pressure vessel types able to withstand an internal pressure of 300 bar
with a safety factor of 3.

However, when considering the buckling analysis, the type-III pressure vessel with 5 mm liner is able
to withstand the highest pressure to almost 900 bar, as seen in Figure 32.

900
Maximum external pressure (bar)

850

800

750

700

650

600
Type-I Type-II Type-III 15Type-III 10Type-III 05

Figure 32 Nonlinear buckling pressure for five different pressure vessel designs.
Reference : ENDURUNS_D5.3
Version : 1.0.0
Date : 27/01/2021
Page : 37

D5.3: Devices and materials for hydrogen storage evaluation

4 Conclusion
Two different pressure housing designs and three different alloys were assessed regarding their
structural integrity under buckling, for hydrogen storage based on LH2 storage tank and metal hydrides
in the ENDURUNS AUV. After carrying out a non-linear analysis with imperfections of 2.5 mm and a
safety factor of 2.25, it was found that the only vessels made out of Ti-6Al-4V would survive the
intended pressure of 600 bar. For a cylindrical design, the vessel would have to be at least 20 mm thick,
or at least 15 mm for a spherical design.

Also different types of gas cylinders for hydrogen storage were analysed with respect to both internal
and external pressure and the results are presented in summarised form. Based on the results obtained
it has been decided that for the demonstrator a customised type III design will be employed for the
AUV.

5 References
1. Züttel A., Materials for hydrogen storage, Mater. Today 6 (2003) pp. 24–33
2. Møller K. T.; T. R. Jensen; E. Akiba; and H. wen Li, Hydrogen - A sustainable energy carrier,
Prog. Nat. Sci. Mater. Int. 27 (2017) pp. 34–40
3. Principi G.; F. Agresti; A. Maddalena; and S. Lo Russo, The problem of solid state hydrogen
storage, Energy 34 (2009) pp. 2087–2091
4. Hirscher M., Handbook of Hydrogen Storage, Wiley-VCH Verlag GmbH & Co. KGaA 2010doi:
10.1002/9783527629800
5. Hyakudome T.; H. Yoshida; S. Tsukioka; T. Sawa; S. Ishibashi; T. Aoki; T. Iwamoto; Y.
Kawaharazaki; A. Muto; T. Oda; and Y. Fujita, High Efficiency Hydrogen and Oxygen Storage
System Development for Underwater Platforms Powered by Fuel Cell, ECS Transations 26
(2010) pp. 465–474
6. Baricco M.; C. Buckley; G. Capurso; N. Gallandat; D. M. Grant; M. N. Guzik; I. Jacob; E. H.
Jensen; T. Jensen; J. Jepsen; T. Klassen; M. V Lototskyy; K. Manickam; A. Montone; J.
Puszkiel; S. Sartori; D. A. Sheppard; A. Stuart; G. Walker; C. J. Webb; H. Yang; V. Yartys; and
A. Zuttel, Application of hydrides in hydrogen storage and compression: Achievements, outlook
and perspectives, Int. J. Hydrogen Energy 44 (2019) pp. 7780–7808
7. Wijayanta A. T.; T. Oda; C. W. Purnomo; T. Kashiwagi; and M. Aziz, Liquid hydrogen,
methylcyclohexane, and ammonia as potential hydrogen storage: Comparison review, Int. J.
Hydrogen Energy 44 (2019) pp. 15026–15044
8. Irani R. S., Hydrogen Storage: High-Pressure Gas, MRS Bull. (2002) pp. 680–682
9. Barthélémy H., Hydrogen storage - Industrial prospectives, Int. J. Hydrogen Energy 37 (2012)
pp. 17364–17372
10. Li M.; Y. Bai; C. Zhang; Y. Song; S. Jiang; D. Grouset; and M. Zhang, Review on the research
of hydrogen storage system fast refueling in fuel cell vehicle, Int. J. Hydrogen Energy 44 (2019)
pp. 10677–10693
11. Zhang, D.-N., et al., A modified Johnson–Cook model of dynamic tensile behaviors for 7075-
T6 aluminum alloy. Journal of Alloys and Compounds, 2015. 619: p. 186-194.
Reference : ENDURUNS_D5.3
Version : 1.0.0
Date : 27/01/2021
Page : 38

D5.3: Devices and materials for hydrogen storage evaluation

12. Karkalos, N.E. and A.P. Markopoulos, Determination of Johnson-Cook material model
parameters by an optimization approach using the fireworks algorithm. Procedia
Manufacturing, 2018. 22: p. 107-113.
13. Chen, G., et al., Finite element simulation of high-speed machining of titanium alloy (Ti–
6Al–4V) based on ductile failure model. The International Journal of Advanced
Manufacturing Technology, 2011. 56(9-12): p. 1027-1038.
14. Liao, B.-b., et al., Continuum damage modeling and progressive failure analysis of a Type III
composite vessel by considering the effect of autofrettage. Journal of Zhejiang University-
SCIENCE A, 2019. 20(1): p. 36-49.
15. Tsai, S.W., Strength theories of filamentary structure. Fundamental aspects of fiber reinforced
plastic composites, 1968.
16. Alcantar, V., et al., Optimization of type III pressure vessels using genetic algorithm and
simulated annealing. International Journal of Hydrogen Energy, 2017. 42(31): p. 20125-
20132.

You might also like