Sie sind auf Seite 1von 202

Werkstofftechnische Berichte |

Reports of Materials Science and Engineering

Martin Nebe

In Situ Characterization
Methodology for the
Design and Analysis
of Composite Pressure
Vessels
Werkstofftechnische Berichte j Reports
of Materials Science and Engineering

Series Editor
Frank Walther, Materials Test Engineering (WPT), TU Dortmund University,
Dortmund, Nordrhein-Westfalen, Germany
In den Werkstofftechnischen Berichten werden Ergebnisse aus Forschungspro-
jekten veröffentlicht, die am Fachgebiet Werkstoffprüftechnik (WPT) der Tech-
nischen Universität Dortmund in den Bereichen Materialwissenschaft und Werk-
stofftechnik sowie Mess- und Prüftechnik bearbeitet wurden. Die Forschungsergeb-
nisse bilden eine zuverlässige Datenbasis für die Konstruktion, Fertigung und
Überwachung von Hochleistungsprodukten für unterschiedliche wirtschaftliche
Branchen. Die Arbeiten geben Einblick in wissenschaftliche und anwendungsori-
entierte Fragestellungen, mit dem Ziel, strukturelle Integrität durch Werkstoffver-
ständnis unter Berücksichtigung von Ressourceneffizienz zu gewährleisten.
Optimierte Analyse-, Auswerte- und Inspektionsverfahren werden als Entschei-
dungshilfe bei der Werkstoffauswahl und -charakterisierung, Qualitätskontrolle
und Bauteilüberwachung sowie Schadensanalyse genutzt. Neben der Werk-
stoffqualifizierung und Fertigungsprozessoptimierung gewinnen Maßnahmen
des Structural Health Monitorings und der Lebensdauervorhersage an Bedeu-
tung. Bewährte Techniken der Werkstoff- und Bauteilcharakterisierung werden
weiterentwickelt und ergänzt, um den hohen Ansprüchen neuentwickelter Produk-
tionsprozesse und Werkstoffsysteme gerecht zu werden.
Reports of Materials Science and Engineering aims at the publication of results
of research projects carried out at the Department of Materials Test Engineering
(WPT) at TU Dortmund University in the fields of materials science and engi-
neering as well as measurement and testing technologies. The research results
contribute to a reliable database for the design, production and monitoring of high-
performance products for different industries. The findings provide an insight to
scientific and applied issues, targeted to achieve structural integrity based on ma-
terials understanding while considering resource efficiency.
Optimized analysis, evaluation and inspection techniques serve as decision guid-
ance for material selection and characterization, quality control and component
monitoring, and damage analysis. Apart from material qualification and produc-
tion process optimization, activities concerning structural health monitoring and
service life prediction are in focus. Established techniques for material and com-
ponent characterization are aimed to be improved and completed, to match the high
demands of novel production processes and material systems.

More information about this series at http://www.springer.com/series/16102


Martin Nebe

In Situ Characterization
Methodology for the
Design and Analysis
of Composite Pressure
Vessels
Martin Nebe
Stuttgart, Germany

Publication as doctoral thesis in the Faculty of Mechanical Engineering of TU Dort-


mund University.
Date and location of defense: 22.03.2021, Dortmund
Chairman: Priv.-Doz. Dr.-Ing. Dipl.-Inform. Andreas Zabel
First examiner: Prof. Dr.-Ing. habil. Frank Walther
Second examiner: Prof. Dr. Peter Middendorf
Co-reviewer: Prof. Dr.-Ing. Prof. h.c. Dirk Biermann

ISSN 2524-4809 ISSN 2524-4817 (electronic)


Werkstofftechnische Berichte j Reports of Materials Science and Engineering
ISBN 978-3-658-35796-2 ISBN 978-3-658-35797-9 (eBook)
https://doi.org/10.1007/978-3-658-35797-9

Springer Vieweg
© The Editor(s) (if applicable) and The Author(s), under exclusive license to Springer Fachmedien
Wiesbaden GmbH, part of Springer Nature 2022
This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or
part of the material is concerned, specifically the rights of translation, reprinting, reuse of illus-
trations, recitation, broadcasting, reproduction on microfilms or in any other physical way, and
transmission or information storage and retrieval, electronic adaptation, computer software, or by
similar or dissimilar methodology now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this
publication does not imply, even in the absence of a specific statement, that such names are exempt
from the relevant protective laws and regulations and therefore free for general use.
The publisher, the authors and the editors are safe to assume that the advice and information in this
book are believed to be true and accurate at the date of publication. Neither the publisher nor the
authors or the editors give a warranty, express or implied, with respect to the material contained
herein or for any errors or omissions that may have been made. The publisher remains neutral with
regard to jurisdictional claims in published maps and institutional affiliations.

Responsible: Stefanie Eggert

This Springer Vieweg imprint is published by the registered company Springer Fachmedien Wies-
baden GmbH part of Springer Nature.
The registered company address is: Abraham-Lincoln-Str. 46, 65189 Wiesbaden, Germany
Für meinen Vater
Foreword

The research activities of the Department of Materials Test Engineering (WPT)


at TU Dortmund University in cooperation with the Fuel Cell Production Depart-
ment of Mercedes-Benz AG in Stuttgart-Untertürkheim entail the holistic analysis
of type IV composite pressure vessels used for hydrogen storage. Through the
combination of work related to material science, numerical modeling, filament
winding and advanced experimental characterization, a fundamental understanding
of the mechanical response of these components under internal pressure loading is
sought. The research activities not only provide insights into material and compo-
nent mechanics, but more importantly demonstrate the application of innovative
experimental and numerical methodologies for future structural analysis of com-
posite pressure vessels.
The current work addresses the development and implementation of an in situ
characterization methodology for composite pressure vessels that permits to as-
sess their geometry as well as their mechanical response under internal pressure
loading. Through a combination of optic and acoustic characterization methods,
insights about the deformation behavior as well as the initation and progression of
damage become accessible. This in turn not only permits to evaluate design-related
aspects, such as stacking sequence or circumferential ply drop locations, but al-
lows to validate numerical modeling strategies. The investigations are exercised
on subscale and on fullscale vessel geometries, which emphasizes the industrial
application of this work. By combining numerical modeling, filament winding and
experimental characterization, this work provides a sound foundation for future
developments in the area of the design and analysis of composite pressure vessels.

Dortmund Frank Walther


July 2021 TU Dortmund University
Department of Materials Test Engineering (WPT)
frank.walther@tu-dortmund.de
http://www.wpt-info.de
vii
Acknowledgments

The writing of this section marks the end of three exciting years of Ph.D. work
at Mercedes-Benz AG. Looking back it has been the most challenging, but at the
same time most rewarding journey for me. At this stage, I would like to express
my sincere gratitude to all the people involved in this project.
First and foremost, I would like to thank Prof. Dr.-Ing. habil. Frank Walther for
his dedicated and motivating supervision of my Ph.D. project. His genuine interest
in my work alongside with the regular and insightful discussions together with
Dr.-Ing. Daniel Hülsbusch really helped shaping this work.
I would like to express my gratitude to Prof. Dr.-Ing. Peter Middendorf for
taking on the co-supervision. His remarks on my work have given it its final touch.
Moreover, I would like to thank Prof. Dr.-Ing. Prof. h.c. Dirk Biermann and Priv.-
Doz. Dr.-Ing. Dipl.-Inform. Andreas Zabel for their serving on my committee.
I would like to thank all the people involved at Mercedes-Benz AG. Starting
with the people, who rigorously planned the facilities I was fortunate to work with,
Wolfgang Hansen, Dr.-Ing. Jürgen Güttler, Holger Stark, Timur Gebhardt and the
entire project team “CHG-Tank”. Thomas Vogel, Raffaele Monda and Axel Bauer
for being the best machine shop staff a Ph.D. student could ask for. Katja Olsen and
Dr.-Ing. Helmut Rauner for providing me with funds, supporting the project and
giving me the freedom to pursue this research. Dr.-Ing. Rolf Schaller for his gen-
uine interest in my work and for the generous amount of CT scans he provided me
with. And last but by no means least, the person with whom I spent a small infinity
of hours at work, Clemens Braun. Thank you for always being around, supporting
me and teaching me practical engineering skills. Looking back, it all started with a
phone call from Stuttgart, Germany to San Juan del Sur, Nicaragua and ended with
seemingly endless nights in the burst chamber fixing various equipment. After all,
I am glad to say I have found a very good friend.
I would like to thank Dr. ir. Julien Van Campen for following my journey from
an early stage on and for being a constant source of lively and insightful discussion
ix
x Acknowledgments

on a variety of topics. His interest in my work and his continous support throughout
helped to establish a strong bond between the TU Delft and Mercedes-Benz AG,
which arguably yielded into this work.
The greatest appreciation goes to my students, Daniel Maraite, Ana Isabel Tor-
res Guijarro, Tom Asijee, Eleonora Cesari, Benoît Porra, Alexander Kasses, An-
tonio Johman, Alejandro Soriano Sutil and Chiara Ardemani. Every single one of
you mightily contributed to the content and the results of this work. Following your
paths towards graduation has fullfilled me with the greatest passion, joy and pride
ever since.
I would like to acknowledge Gunter Sanow, Tim Winter, Stefanie Kielies and
the entire GOM support team for their continous support regarding the optical
characterization by means of digital image correlation. Establishing a multi-sensor
DIC system for the deformation analysis of CPVs became a real idea almost two
years ago. Without your help, this would not have been possible.
A special thanks goes to Adam Andrae and Jonas Velten from IDVA GmbH for
the close interaction and research exchange related to the numerical modeling of
CPVs, particularly at early stages of my work.
I would like to thank Marc Grzeschik and Miloš Drašković from the Institute
of Aircraft Design at the University of Stuttgart for the long-term cooperation over
the years and for their excellent work in material characterization.
Another big acknowledgement is directed towards my dearest friends. Dr.-Ing.
Tobias Schmack for initiating and following my journey as well as for providing
me with valuable advices on a nearly daily base. Tommy Staub for always keeping
me engaged off work – I will never forget our first Burn – and for proof-reading
my thesis. Moritz Neubauer for not letting me off the hook and for reigniting the
competitiveness within me, most of the times likely without even knowing.
I would like to express my sincerest gratitude to my beautiful partner in crime
Charlotte Katharina Nebe, who supported me throughout the entire time with
seemingly endless understanding. Her patience, kindness and affection gave me
the strength and freedom to pursue this work in the way I did. The countless
occasions, where I was cutting our time short made me realize how much I am
indebted to her. I am very much looking forward to pay this back in our common
future.
I would like to thank my family that supported me throughout the years and
for whose I always strove to do better. To my mother Katja Nebe for her love and
affection in very early and most recent years. To my sisters, Madlén Vogel-Wittig
and Michèle Nebe, for being by my side during good but especially during difficult
times. The strong bond we have developed throughout the years is something that
I am very grateful for and gives me strength whenever I need it.
Acknowledgments xi

There is no doubt that my biggest acknowledgment belongs to my father Gerd


Nebe. His endless support and guidance over the years made me the person I am
today. The rigorous care he took of me and my two sisters and the sacrifices he
made for us will always remain incomprehensible for me. In many ways he has
been my role model and taught me an endless number of life lessons, most of them
without even knowing. His accountability for us and the way he was able to provide
us with the life we have today, are things that I will never forget and will cherish
forever. Whenever I lacked motivation for this work, I thought of him knowing
that it would be my biggest goal to make him proud. In many ways, this thesis is
dedicated to him.

Stuttgart Martin Nebe


July 2021
Abstract

Recent endeavors towards the industrialization of fuel cell electric vehicles


(FCEVs) face the challenge of cost-competitive hydrogen storage systems.
For current state-of-the-art composite pressure vessels (CPVs), the raw material
price of carbon fiber contributes to the vast majority of system costs. Therefore, the
reduction of used composite material poses the most promising cost optimization
potential. Yet, in order to derive structural sound and material optimized CPV de-
signs, a comprehensive understanding of the mechanical response under internal
pressure loading is required.
Recognizing the necessity to characterize and comprehend the vessel’s me-
chanical response in its entirety is where this work begins. To this purpose, an
in situ characterization methodology is established, which permits to identify the
vessel’s meridional thickness profile, its deformation behavior and the damage evo-
lution process under internal pressure loading. The successful implementation of
the methodology yields into the development of a high fidelity finite element (FE)
model, that is subsequently contrasted against experimental data on various scales
of complexity. Through the alignment of experimental and numerical analysis,
valuable insights on the deformation behavior as well as on the initiation and pro-
gression of damage are obtained.
Subsequently, the in situ characterization methodology and the FE model are
applied to the investigation of stacking sequence influence. The substantial impact
of the stacking sequence is unveiled with regard to consolidation level, stress and
strain distribution and crack propagation through-the-thickness. Furthermore, the
high sensitivity of the strain distribution around the cylinder-dome transition zone
to the circumferential ply drop locations is highlighted as a critical aspect in regard
to the design of CPVs.
Ultimately, the derived insights are transferred onto a fullscale vessel geome-
try, representative for the use in FCEVs, where a noticeable mass saving potential
of 17% compared to the state-of-the-art is achieved. The investigation herein high-
xiii
xiv Abstract

lights the strong impact of the dome geometry on the laminate design. Furthermore,
the framework’s applicability for the structural analysis, regardless of the vessel
geometry or size, is verified.
The work presented here sets a sound foundation in regard to future experi-
mental characterization, numerical modeling and design of CPVs as it allows for
a comprehensive understanding of the mechanical response by correlating experi-
mental and numerical results on various scales of complexity.
Kurzfassung

Aktuelle Bestrebungen zur Industrialisierung von Brennstoffzellenfahrzeugen


stehen vor der Herausforderung einer kosteneffizienten Herstellung von Wasser-
stoffspeichersystemen. Bei derzeitig verwendeten Faserverbund-Druckbehältern
stellt der Materialpreis der Kohlenstofffaser den überwiegenden Teil der Sys-
temkosten dar. Somit ist die Materialreduktion der vielversprechendste Ansatz
unter den möglichen Kostenoptimierungspotentialen. Um jedoch struktur- und
materialoptimierte Druckbehälter entwickeln zu können, ist ein umfassendes
Strukturverständnis der Behälter unter Innendruckbelastung erforderlich.
Die Notwendigkeit zur vollumfänglichen Charakterisierung der Strukturant-
wort der Behälter stellt den Ausgangspunkt dieser Arbeit dar. Zu diesem Zweck
wird eine in situ Charakterisierungsmethodik entwickelt, welche es ermöglicht,
die Behälterkontur, das Verformungsverhalten und den Schädigungsentwicklungs-
prozess unter Innendruckbelastung zu charakterisieren. Die erfolgreiche Imple-
mentierung der Methodik resultiert in der Entwicklung eines hochgenauen FE-
Modells, das anschließend mit experimentellen Daten auf verschiedenen Komplex-
itätsskalen korreliert wird. Durch den Abgleich von experimenteller und nume-
rischer Analyse werden wertvolle Erkenntnisse über das Deformationsverhalten,
sowie über die Schadensinitiierung und -progression gewonnen.
Resultierend daraus wird der Einfluss der Lagenreihenfolge mit Hilfe der
beiden etablierten Methoden eingehend untersucht. Der erhebliche Einfluss der
Lagenreihenfolge wird im Hinblick auf die Kompaktierung, die Spannungs- und
Dehnungsverteilung und der Rissausbreitung aufgezeigt. Darüber hinaus wird
die hohe Empfindlichkeit der Dehnungsverteilung im Zylinder-Dom-Übergangs-
bereich gegenüber den Umfangslagenenden als ein kritischer Aspekt im Hinblick
auf die Auslegung von Druckbehältern hervorgehoben.
Abschließend werden die abgeleiteten Erkenntnisse auf eine Behältergeome-
trie übertragen, die in ihren geometrischen Abmaßen repräsentativ für die Ver-
wendung in Brennstoffzellenfahrzeugen ist. Gegenüber dem Serienbauteil kann
xv
xvi Kurzfassung

dabei eine Reduzierung der Strukturmasse von 17% aufgezeigt werden. In diesem
Zusammenhang verdeutlichen die Ergebnisse den bedeutenden Einfluss der Dom-
geometrie auf die Laminatauslegung. Darüber hinaus wird die Anwendbarkeit der
geschaffenen Methodik zur Strukturanalyse, unabhängig von Behältergeometrie
und -größe, verifiziert.
Die vorliegende Arbeit bildet eine fundamentale Grundlage im Hinblick auf die
zukünftige Charakterisierung, FE-Modellierung und Auslegung von Faserverbund-
Druckbehältern, welche durch Korrelation experimenteller und numerischer
Ergebnisse ein umfassendes Strukturverständnis ermöglicht.
Contents

1 Motivation and scope . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

2 Literature review . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2.1 Background to composite pressure vessels . . . . . . . . . . . . . . 5
2.1.1 Composites . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2.1.2 Filament winding . . . . . . . . . . . . . . . . . . . . . . . . 7
2.2 Analysis of composite pressure vessels . . . . . . . . . . . . . . . . 13
2.2.1 Analytical approaches . . . . . . . . . . . . . . . . . . . . . 13
2.2.2 Finite element analysis . . . . . . . . . . . . . . . . . . . . . 17
2.3 Experimental characterization of composite pressure vessels . . 19
2.3.1 Strain and displacement monitoring . . . . . . . . . . . . . 19
2.3.2 Acoustic emission analysis . . . . . . . . . . . . . . . . . . 21
2.4 Correlation of numerical and experimental analysis . . . . . . . . 24
2.5 Summary and research questions . . . . . . . . . . . . . . . . . . . 27

3 Material and methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29


3.1 Robot-assisted towpreg winding . . . . . . . . . . . . . . . . . . . . 29
3.1.1 Manufacturing . . . . . . . . . . . . . . . . . . . . . . . . . . 29
3.1.2 Liner geometries and layups . . . . . . . . . . . . . . . . . . 31
3.2 Experimental characterization . . . . . . . . . . . . . . . . . . . . . 33
3.2.1 Test chamber . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
3.2.2 Stripelight projection . . . . . . . . . . . . . . . . . . . . . . 35
3.2.3 Digital image correlation . . . . . . . . . . . . . . . . . . . . 36
3.2.4 Airborne acoustic emission . . . . . . . . . . . . . . . . . . 37
3.3 Determination of material properties . . . . . . . . . . . . . . . . . 39
3.3.1 Stiffness and strength . . . . . . . . . . . . . . . . . . . . . . 39
3.3.2 Fracture toughness . . . . . . . . . . . . . . . . . . . . . . . 40
3.3.3 Fiber volume fraction and porosity . . . . . . . . . . . . . . 41

xvii
xviii Contents

3.4 Structural analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43


3.4.1 Three-dimensional elasticity theory . . . . . . . . . . . . . 44
3.4.2 Finite element analysis . . . . . . . . . . . . . . . . . . . . . 44

4 In situ characterization methodology . . . . . . . . . . . . . . . . . . . 51


4.1 Evaluation of thickness build-up . . . . . . . . . . . . . . . . . . . . 52
4.1.1 Analysis of dome contour and manufacturing influence . 52
4.1.2 Correlation to thickness prediction . . . . . . . . . . . . . . 54
4.2 Characterization of deformation behavior . . . . . . . . . . . . . . 57
4.2.1 Full field strain analysis . . . . . . . . . . . . . . . . . . . . 58
4.2.2 Quantitative description of deformation . . . . . . . . . . . 64
4.3 Failure monitoring and localization . . . . . . . . . . . . . . . . . . 68
4.3.1 Analysis of interfiber failure . . . . . . . . . . . . . . . . . . 68
4.3.2 Insights on vessel failure . . . . . . . . . . . . . . . . . . . . 74
4.4 Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . 79

5 FE modeling and correlation . . . . . . . . . . . . . . . . . . . . . . . . . 81


5.1 Model definition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
5.1.1 Material characterization . . . . . . . . . . . . . . . . . . . . 82
5.1.2 Geometry definition . . . . . . . . . . . . . . . . . . . . . . . 91
5.2 Correlation of mechanical response . . . . . . . . . . . . . . . . . . 97
5.2.1 Deformation behavior . . . . . . . . . . . . . . . . . . . . . . 98
5.2.2 Damage progression and final failure . . . . . . . . . . . . 104
5.3 Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . 108

6 Influence of stacking sequence . . . . . . . . . . . . . . . . . . . . . . . . 111


6.1 Experimental design . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
6.1.1 Stress gradient through-the-thickness . . . . . . . . . . . . 113
6.1.2 Observations on vessel failure . . . . . . . . . . . . . . . . . 116
6.2 Fiber volume fraction and porosity . . . . . . . . . . . . . . . . . . 117
6.3 Deformation behavior . . . . . . . . . . . . . . . . . . . . . . . . . . 120
6.3.1 Cylinder strains . . . . . . . . . . . . . . . . . . . . . . . . . 120
6.3.2 Strain distribution in dome and cylinder-dome transition 122
6.4 Damage progression and final failure . . . . . . . . . . . . . . . . . 127
6.4.1 Interfiber failure . . . . . . . . . . . . . . . . . . . . . . . . . 127
6.4.2 Final failure . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130
6.4.3 Influence of circumferential ply drop locations . . . . . . 133
6.5 Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . 135
Contents xix

7 Application on fullscale geometry . . . . . . . . . . . . . . . . . . . . . 137


7.1 Design considerations . . . . . . . . . . . . . . . . . . . . . . . . . . 137
7.1.1 Dome contour . . . . . . . . . . . . . . . . . . . . . . . . . . 138
7.1.2 Layup design . . . . . . . . . . . . . . . . . . . . . . . . . . . 139
7.2 Structural analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140
7.2.1 Deformation behavior . . . . . . . . . . . . . . . . . . . . . . 140
7.2.2 Final failure . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142
7.3 Comparison to the state-of-the-art . . . . . . . . . . . . . . . . . . . 145
7.4 Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . 147

8 Design considerations to composite pressure vessels . . . . . . . . . . 149

9 Summary and outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157

List of publications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161

Supervised student theses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165

Curriculum vitae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167

Published volumes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171
Nomenclature

Abbreviations

AAE Airborne acoustic emission


AE Acoustic emission
ANN Artificial neural network
BD Bidirectional
BEV Battery electric vehicle
BP Burst pressure
C Circumferential layer
CC Compliance calibration
CDM Continuum damage mechanics
CFRP Carbon fiber-reinforced plastic
CLT Classical lamination theory
CNG Compressed natural gas
CPV Composite pressure vessel
CT Computed tomography
DCB Double cantilever beam
DIC Digital image correlation
ENF End notch flexure
EoL End-of-line
ESG Electrical resistance strain gauge
FBG Fiber Bragg grating
FCEV Fuel cell electric vehicle
FE Finite element
FEA Finite element analysis
FOS Fiber optic sensor
FPF First-ply-failure
FRP Fiber-reinforced plastic

xxi
xxii Nomenclature

FVF Fiber volume fraction


GTR Global technical regulation
HH High-angle helical layer
ICE Internal combustion engine
ICEV Internal combustion engine vehicle
IFF Interfiber failure
LH Low-angle helical layer
LVDT Linear variable differential transformer
MBT Modified beam theory
NDE Non-destructive evaluation
NTP Normal temperature and pressure
NWP Nominal working pressure
PA6 Polyamide 6
PTFE Polytetrafluoroethylene
SD Standard deviation
SHM Structural health monitoring
UD Unidirectional
WTW Well-to-wheel

Latin symbols

A [m2 ] Specimen cross-sectional area


ŒA [N m1 ] Extensional stiffness matrix
a [mm] Delamination length
as [s] Speed of sound in air
Amax [Pa] Maximum sound pressure amplitude
ŒB [N] Bending-extension coupling matrix
b [mm] Specimen width
ŒD [N m] Bending stiffness matrix
d1 [–] Longitudinal damage variable
d2 [–] Linear softening relation variable
E [GPa] Modulus of elasticity
E11 [GPa] Modulus of elasticity in the longitudinal direction
E22 [GPa] Modulus of elasticity in the transverse direction
eOF [–] Orthogonal vector to fiber
eON [–] Crack-normal vector to fiber
eOT [–] Transverse vector to fiber
F [–] Transverse damage activation function
F [N] Fiber tensile force
Nomenclature xxiii

FB [–] Bulk deformation gradient


FRN [–] Felicity ratio
fO [–] Time function
fb [N] Transverse force on fiber
fi [–] Recorded time functions
fn [N] Normal force on fiber
fw [N] Friction force between fiber and underlying surface
G [GPa] Shear modulus
G12 [GPa] In-plane laminate shear modulus
G1C [N mm1 ] Fibers Mode I fracture toughness
GI [N mm1 ] Laminate Mode I fracture toughness
GIc [N mm1 ] Corrected laminate Mode I fracture toughness
GII [N mm1 ] Laminate Mode II fracture toughness
GIIc [N mm1 ] Corrected laminate Mode II fracture toughness
h [mm] Specimen height
L [–] Loading function
LFR;N [MPa] Load stage of acoustic emission onset
LN 1 [MPa] Last load stage prior to acoustic emission onset
l [–] Element length
M [#] Number of sound pressure sensors
Mt [Nm] Moment resultant tensor
m [N1 mm2 ] Coefficient for fracture toughness calculation
mCFRP [kg] Laminate mass
mR [–] Count of helical subplies in the cylinder
ms [–] Softening law proportion
N [#] Number of emissions above threshold
Nt [N] Force resultant tensor
n [#] Number of specimens or discrete samples
ncycles [#] Number of loading cycles
nlayers;norm: [–] Normalized number of layers
nR [#] Number of helical subplies
ns [–] Softening law proportion
P [N] Applied load
Pmax [N] Maximum load
p [MPa] Internal pressure
pAAE [MPa] Internal pressure of airborne acoustic emission onset
pburst [MPa] Burst pressure
p burst [MPa] Average burst pressure
pk [MPa] Interpolation parameter
xxiv Nomenclature

pl [MPa m1 ] Radial pressure per unit length


ps [Pa] Sound pressure
ŒQ [MPa] Stiffness matrix
R [mm] Cylinder radius
Rcr [–] Basis coordinate system
r [mm] Radial coordinate
rb [mm] Pole radius at one tow-width from the turnaround point
r2b [mm] Pole radius at two tow-width from the turnaround point
ri [mm] Distance from sound pressure sensor to point
rs [mm] Principal meridional radius of curvature
rturn [mm] Polar opening radius
r [mm] Principal tangential radius of curvature
s [mm] Meridional coordinate
S [MPa] Piola-Kirchoff stress tensor
SL [MPa] Laminate shear strength
T [°C] Temperature
ŒT  [–] Transformation matrix
t [s] Time
t [MPa] Bulk stress tensor projection
tk [s] Time value at a discrete sample index
tl [mm] Laminate thickness
tp [mm] Ply thickness
tR [mm] Current ply thickness in cylinder
ur [mm] Radial displacement
uz [mm] Axial displacement
u˚ [mm] Tangential displacement
Vf [%] Fiber volume fraction
Vp [%] Porosity
wi [–] Shading weights
wt [mm] Tow width
XC [MPa] Longitudinal compressive strength
XT [MPa] Longitudinal tensile strength
x [mm] Length coordinate
YC [MPa] Transverse compressive strength
YT [MPa] Transverse tensile strength
y [mm] Length coordinate
zsound [mm] Axial coordinate of location with predicted thickness
zturn [mm] Axial coordinate of turnaround point
z [mm] Axial coordinate
Nomenclature xxv

Greek symbols

˛ [°] Winding angle


˛R [°] Winding angle of current ply in cylinder
 [%] Shear strain
 [–] Horizontal axis interception of cube root compliance
i [s] Sound pressure relative time delay
ı [mm] Displacement
 [–] Slippage coefficient
" [%] Normal strain
"0 [%] Mid-plane strain tensor
"p;u [%] Strain at ultimate strength
"' [%] Tangential strain
"s [%] Meridional strain
 [–] Mode-mixity ratio
 [–] Curvature tensor
[–] Coefficient of friction

[–] Poisson’s ratio

12 [–] In-plane Poisson’s ratio in the 1-2 direction

21 [–] In-plane Poisson’s ratio in the 2-1 direction

23 [–] Out-of-plane Poisson’s ratio in the 2-3 direction


[kg m3 ] Density
[MPa] Normal stress
11 [MPa] Longitudinal in-plane component of stress tensor
22 [MPa] Transverse in-plane component of stress tensor
33 [MPa] Transverse out-of-plane component of stress tensor
B [MPa] Bulk material tensor
L [MPa] Fiber stress
Res [MPa] Residual stress vector
y [MPa] Yield stress
u [MPa] Ultimate stress
[MPa] Shear stress
 [MPa] Stress vector in the cohesive interface
i [s] Absolute run time
' [°] Tangential coordinate
fiber [g] Fiber mass
List of Figures

Fig. 1.1 Mercedes-Benz GLC F-CELL with two composite pressure


vessels for an on-board storage of 4:4 kg hydrogen [3] . . . . . . 2
Fig. 1.2 Thesis structure and content of individual chapters . . . . . . . . 4
Fig. 2.1 Overview of identified pressure vessel types according to their
structural design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
Fig. 2.2 Comparison of layer types . . . . . . . . . . . . . . . . . . . . . . . 9
Fig. 2.3 Schematic representation of a tow path in the dome region,
reprinted/adapted from [18] with permission from Elsevier . . . 10
Fig. 2.4 Comparison of thickness prediction to a CT scan of a CPV,
reprinted/adapted from [18] with permission from Elsevier . . . 13
Fig. 2.5 Meridional moment distribution at the cylinder-dome transition,
adapted from [43] . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
Fig. 2.6 FE model of a CPV composed of metallic boss, plastic liner and
overwrapped CFRP, reprinted/adapted from [44] with permission
from Elsevier . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
Fig. 2.7 Embedment of a FBG sensor line during filament winding,
reprinted/adapted from [62] with permission from Elsevier . . . 20
Fig. 2.8 Display of full field strain for a CNG vessel at 10 MPa,
reprinted/adapted from [60] with permission from Elsevier . . . 21
Fig. 2.9 Average Shelby countup ratios for four impacted and two
non-impacted CPV, reprinted/adapted from [71] . . . . . . . . . . 23
Fig. 2.10 Schematic overview of correlation between experiment and
simulation with indicated literature sources . . . . . . . . . . . . . 24
Fig. 2.11 Correlation of predicted and experimental strains obtained
through integrated optic fibers, reprinted/adapted from [45] with
permission from Elsevier . . . . . . . . . . . . . . . . . . . . . . . . 25

xxvii
xxviii List of Figures

Fig. 2.12 Correlation of predicted and experimental strains at different


axial positions for an internal pressure of 3:45 MPa,
reprinted/adapted from [62] with permission from Elsevier . . . 26
Fig. 2.13 Remainders of a type III vessel after burst testing [11] . . . . . . 27
Fig. 3.1 Overview of utilized ressources and methods within this thesis 30
Fig. 3.2 Robot-assisted winding system . . . . . . . . . . . . . . . . . . . . 30
Fig. 3.3 Overview of liner geometries and coordinate system definition 31
Fig. 3.4 Overview of the investigated main layups . . . . . . . . . . . . . . 32
Fig. 3.5 Test chamber for burst and cyclic experiments of pressure vessels 34
Fig. 3.6 Overview of individual steps for the DIC postprocessing
workflow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
Fig. 3.7 Filament wound flat plates; a) unidirectional plate and b)
bidrectional plate (˙45°) . . . . . . . . . . . . . . . . . . . . . . . . 40
Fig. 3.8 Integration of a PTFE insert during filament winding for the
manufacturing of ENF and DCB specimens . . . . . . . . . . . . 41
Fig. 3.9 Porosity volume for a CPV laminate section . . . . . . . . . . . . 43
Fig. 3.10 Methodology for the geometrical description of the FE model . 46
Fig. 4.1 Experimentally obtained thickness profiles and overview of
layup composition for the Layups A, B and C . . . . . . . . . . . 52
Fig. 4.2 Influence of fiber tension variation on circumferential layer
readjustment at cylinder-dome transition . . . . . . . . . . . . . . 53
Fig. 4.3 Comparison of thickness build-up prediction made by
ComposicaD™ and experimentally obtained thickness profile . 54
Fig. 4.4 Comparison of experimentally obtained thickness profile,
nominal and scanned liner contour with a CT scan of a vessel
after manufacturing . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
Fig. 4.5 Example for liner deformation resulting from subsequently
wound low-angle helical layers with excessive fiber tension . . 57
Fig. 4.6 Example for a manufacturing defect during filament winding:
squishing of circumferential layers due to a subsequently wound
1 lagging helical pattern . . . . . . . . . . . . . . . . . . . . . . . 58
Fig. 4.7 Meridional strain fields during a burst experiment shown from
three different circumferential positions . . . . . . . . . . . . . . . 59
Fig. 4.8 Overview of test setup with indicated location of strain gauges
and surface components . . . . . . . . . . . . . . . . . . . . . . . . 61
Fig. 4.9 Meridional and tangential cylinder strains obtained by means of
strain gauges for a pressurization up to 105 MPa internal pressure 62
Fig. 4.10 Comparison of cylinder strains obtained by means of strain
gauges and DIC at an internal pressure of 25 MPa . . . . . . . . 63
List of Figures xxix

Fig. 4.11 Framework of parameters used for the description of deformation


behavior . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
Fig. 4.12 Circumferential ply drop location for the Layup B . . . . . . . . 65
Fig. 4.13 Comparison of average cylinder strains between the Layups B
and B-T1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
Fig. 4.14 Comparison of average axial boss displacement between the
Layups B and B-T1 . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
Fig. 4.15 Strains along the meridional surface path at an internal pressure
of 105 MPa for the Layups B and B-T1 . . . . . . . . . . . . . . . 67
Fig. 4.16 Channel file for a burst experiment of Layup A without prior
pressurization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
Fig. 4.17 Interfiber failure in circumferential and outer helical layers after
an internal pressure of 105 MPa . . . . . . . . . . . . . . . . . . . . 70
Fig. 4.18 Recorded maximum sound pressure amplitude Amax for a
pressurization from 0 to 105 MPa for the Layups A, A-C1 and
A-C2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
Fig. 4.19 Recorded maximum sound pressure amplitude Amax for a
pressurization from 0 to 105 MPa for the Layups A, A-S* and
A-S2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
Fig. 4.20 Circumferential view on the cylinder sections of three different
vessels of Layup A . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
Fig. 4.21 Detected hits above threshold during EoL and cyclic
pressurization from 2 to 87:5 MPa . . . . . . . . . . . . . . . . . . 74
Fig. 4.22 Comparison of emission localization on three-dimensional mesh
to the positions of fiber bundle breakages on the outermost
helical layer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
Fig. 4.23 Detected hits above threshold during burst experiment for vessels
of the Layups A, C and D . . . . . . . . . . . . . . . . . . . . . . . 76
Fig. 4.24 Experimentally obtained thickness profiles and failure pictures
of the Layups A, C and D . . . . . . . . . . . . . . . . . . . . . . . 77
Fig. 4.25 Comparison of last localized emission and resulting failure for
the Vessel D-01 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
Fig. 5.1 Overview of chapter structure including used experimental data
for the model definition and correlation of mechanical response 82
Fig. 5.2 Measurement of fiber volume fraction and porosity by means of
acid digestion tests at different axial positions of Layup A . . . 83
Fig. 5.3 Porosity through-the-thickness obtained for three laminate
sections of Layup A . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
xxx List of Figures

Fig. 5.4 Stress-strain curves for longitudinal, transverse and in-plane


shear tensile tests for the three different consolidation levels . . 86
Fig. 5.5 Load-displacement curve of DCB test procedure for Mode I
fracture toughness determination . . . . . . . . . . . . . . . . . . . 89
Fig. 5.6 Load-displacement curve of ENF test procedure for Mode II
fracture toughness determination . . . . . . . . . . . . . . . . . . . 90
Fig. 5.7 Comparison between initial thickness estimation obtained from
ComposicaD™ , scanned outer contour and adjusted simulation
geometry for the Layup A . . . . . . . . . . . . . . . . . . . . . . . 92
Fig. 5.8 FE mesh with its main modeling characteristics . . . . . . . . . . 93
Fig. 5.9 Comparison between extrapolated fiber orientation and initial
prediction made by ComposicaD™ for the outermost helical
layer of Layup A . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
Fig. 5.10 Implementation of heterogeneous material definition for the FE
model of Layup A . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
Fig. 5.11 Correlation of predicted and experimental cylinder strain
components for Layup A . . . . . . . . . . . . . . . . . . . . . . . . 98
Fig. 5.12 Deformed shape and meridional strain field of a homogeneous
orthotropic vessel at an internal pressure of 160 MPa . . . . . . . 100
Fig. 5.13 Correlation of predicted and experimental strain components
along the meridional surface path at an internal pressure of
105 MPa for Layup A . . . . . . . . . . . . . . . . . . . . . . . . . . 101
Fig. 5.14 Correlation of predicted and experimental axial displacement for
Layup A . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
Fig. 5.15 Comparison of experimental channel file, damage variable in
transverse direction at 35, 50 and 105 MPa internal pressure and
microscope images after a pressurization up to 105 MPa . . . . . 105
Fig. 5.16 Monitoring of outer surface damage on white-chalked vessel
surface during the loading up to 105 MPa internal pressure . . . 106
Fig. 5.17 Three vessel remainders of the investigated Layup A . . . . . . . 107
Fig. 6.1 Experimental design for the investigation of stacking sequence
influence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
Fig. 6.2 Distribution of longitudinal stress through-the-thickness at
an internal pressure of 160 MPa for the investigated stacking
sequence configurations . . . . . . . . . . . . . . . . . . . . . . . . 114
Fig. 6.3 Remainder and burst pressure of a single vessel for each stacking
sequence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
Fig. 6.4 Fiber volume fraction and porosity obtained by means of acid
digestion tests for Sequences A, A-S1 and A-S2 . . . . . . . . . 118
List of Figures xxxi

Fig. 6.5 Comparison of outer vessel surfaces between Sequences A,


A-S1 and A-S2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118
Fig. 6.6 Porosity through-the-thickness for the Sequences A, A-S1 and
A-S2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
Fig. 6.7 Meridional and tangential cylinder strains for the investigated
stacking sequences . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
Fig. 6.8 Strain components along the meridional surface path at an
internal pressure of 70 MPa for the investigated stacking
sequences . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
Fig. 6.9 Experimentally obtained thickness profiles and cross-sectional
view of circumferential ply drop locations for the Sequences A,
A-S1 and A-S2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
Fig. 6.10 Comparison of predicted and experimental strain components
along the meridional surface path for the Sequences A, A-S1 and
A-S2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126
Fig. 6.11 Comparison of channel files for a pressurization up to burst;
shown data corresponds to one vessel for each sequence; the
vessels were not pressurized prior to the experiment . . . . . . . 128
Fig. 6.12 Comparison of damage variable in transverse direction and
microscope cross-section after pressurization for the Sequences
A, A-S1 and A-S2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
Fig. 6.13 Investigation on the existence of interlaminar damage in the
cylinder-dome transition of Sequence A-S2 . . . . . . . . . . . . 132
Fig. 6.14 Strain components along the meridional surface path at an
internal pressure of 70 MPa for the Sequences A-S2 and A-S2-T1 134
Fig. 6.15 Comparison of vessel remainders and burst pressures between
Sequences A-S2 and A-S2-T1 . . . . . . . . . . . . . . . . . . . . . 134
Fig. 7.1 Comparison of dome shape between subscale and fullscale
geometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138
Fig. 7.2 Comparison of thickness build-up and dome contour during
filament winding between Layup E and F . . . . . . . . . . . . . . 140
Fig. 7.3 Strain components along the meridional surface path at an
internal pressure of 70 MPa for the Layups E and F . . . . . . . . 141
Fig. 7.4 Comparison of vessel remainder and burst pressure between
Layup E and F . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142
Fig. 7.5 Strain components along the meridional surface path at an
internal pressure of 105 MPa for the Layups F and F-C1 . . . . . 143
xxxii List of Figures

Fig. 7.6 High crack density on outer excess resin layer after surpassing
the pressure cycle test of 22,000 cycles from 2 to 87:5 MPa
internal pressure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145
Fig. 7.7 Comparison of CFRP mass and cylinder wall thickness between
state-of-the-art design and design prototype . . . . . . . . . . . . 146
Fig. 7.8 Comparison of FVF and porosity between state-of-the-art design
and design prototype . . . . . . . . . . . . . . . . . . . . . . . . . . 147
Fig. 8.1 Distribution of longitudinal stress through-the-thickness at an
internal pressure of 160 MPa for the Sequences A, A-S1, A-S2
and A-S3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150
Fig. 8.2 Comparison of CT cross-section of the Sequences A, A-S1 and
A-S2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151
Fig. 8.3 Comparison of crack growth through-the-thickness between
Sequence A and A-S1 . . . . . . . . . . . . . . . . . . . . . . . . . . 152
Fig. 8.4 Strain components along the meridional surface path at an
internal pressure of 105 MPa for the Layups B and B-S2 . . . . 154
Fig. 8.5 Variation of axial liner length over normalized number of layers
for the Layup B . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
Fig. 9.1 Overview of potential applications of this work in future research 159
List of Tables

Table 3.1 Definition of AAE signal features . . . . . . . . . . . . . . . . . . 38


Table 4.1 Obtained acoustic features during the loading with internal
pressure from 0 to 105 MPa . . . . . . . . . . . . . . . . . . . . . . 72
Table 4.2 Detected AAE from 70 MPa to burst . . . . . . . . . . . . . . . . . 78
Table 5.1 Overview of investigated configurations with different
consolidation levels . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
Table 5.2 Results of the DCB tests for the determination of Mode I fracture
toughness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
Table 5.3 Results of the ENF tests for the determination of Mode II
fracture toughness . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
Table 5.4 Properties of metallic boss and resin constituent . . . . . . . . . . 94
Table 5.5 Composite properties defined in the FE model with homogeneous
material definition . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
Table 5.6 Composite properties defined in the FE model with
heterogeneous material definition . . . . . . . . . . . . . . . . . . . 97
Table 5.7 Comparison of predicted and experimental strains and
displacements at an internal pressure of 105 MPa for the
Layup A . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
Table 5.8 Comparison of predicted and experimental burst pressures for
Layup A . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
Table 6.1 Overview of predicted outer surface strains at 105 MPa internal
pressure and final burst pressure according to three-dimensional
elasticity theory [42] . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
Table 6.2 Comparison of predicted and experimental burst pressures . . . 117
Table 6.3 Average burst pressure, CFRP mass, cylinder strains and axial
displacement for the stacking sequences with three tested
specimens . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122

xxxiii
xxxiv List of Tables

Table 6.4 Comparison of burst pressures and failure locations between


experiment and predictions made by three-dimensional elasticity
theory and FE model . . . . . . . . . . . . . . . . . . . . . . . . . . 131
Table 7.1 Comparison of normalized layer number, layup composition and
CFRP mass between Layup E and F . . . . . . . . . . . . . . . . . 139
Table 7.2 Comparison of normalized layer number, layup composition and
CFRP mass between the Layups E, F and F-C1 . . . . . . . . . . 144
Motivation and scope
1

Sustainable and emission-free transportation in all segments of todays mobility


poses a necessity to counteract the manifested human-induced climate change. In
this context, battery electric vehicles (BEVs) appear as the solution of choice for
short-distance and urban traffic due to their generally high well-to-wheel (WTW)
efficiency. Nevertheless, the low gravimetric energy density and comparably long
recharging durations inevitably pose limitations for the scope of long-distance
transportation. This in particular motivates the development and industrialization
of alternative technologies to replace internal combustion engines (ICEs).
To this purpose, fuel cell electric vehicles (FCEVs) offer a promising potential,
because of their high gravimetric energy density and fast refueling possibilities.
With regard to the storage of hydrogen, its low volumetric energy density at normal
temperature and pressure (NTP) conditions require it to be stored as a liquid under
cryogenic temperatures, highly compressed as a gas under ambient temperatures
or as combination of the aforementioned in a cryo-compressed state. Besides, the
sorbation of hydrogen in a liquid or solid carrier medium demonstrates another
viable option, with metal hydrides being the most recognized solution.
Withal, the highly compressed gaseous storage constitutes the currently most
mature technology, because of its arguably simpler system requirements and the
similitude to existent pressurized propellant systems such as compressed natural
gas (CNG). Targeting high energy densities in order to achieve similar driving
ranges than those of internal combustion engine vehicles (ICEVs) in turn leads
to high storage pressures, which in case of the transportation sector are currently
standardized to a maximum nominal working pressure (NWP) of 70 MPa [1, 2].
Within the scope of preferably lightweight vehicle architectures, the use of com-
posites is motivated by their high specific material properties in contrast to their
metallic competitors. Hence, type IV pressure vessels, which are composed of a
polymer liner, metallic boss ends and a reinforcing composite overwrapping, ren-
der the current state-of-the-art for hydrogen storage and are found in vehicles such
© The Editor(s) (if applicable) and The Author(s), under exclusive license to 1
Springer Fachmedien Wiesbaden GmbH, part of Springer Nature 2022
M. Nebe, In Situ Characterization Methodology for the Design and Analysis of
Composite Pressure Vessels, Werkstofftechnische Berichte j Reports of Materials
Science and Engineering, https://doi.org/10.1007/978-3-658-35797-9_1
2 1 Motivation and scope

Fig. 1.1 Mercedes-Benz


GLC F-CELL with two
composite pressure vessels
for an on-board hydrogen
storage of 4:4 kg, which
allows for an approximate
driving range of 430 km
without refueling [3]

as the Mercedes-Benz GLC F-CELL, shown in Fig. 1.1. Commercially available


CPVs regularly feature a geometry, which consists of a cylinder and two dome
end caps, that ultimately represents a compromise of storage volume, production
costs and manufacturing feasibility. In this context, filament winding represents the
currently most mature manufacturing technology.
With carbon fiber-reinforced plastic (CFRP) as main constituent of the over-
wrapping material, the generally strict safety regulations compel a minimum burst
pressure (BP) of 225% greater than the NWP [1, 2], which hence requires consid-
erable material use in order to ensure the needed high mechanical performance.
In combination with the elevated raw material price of carbon fiber, this leads to
the circumstance that up to 75% of the tank system costs are due to the composite
material [4]1 . Yet, the industrialization of FCEVs requires further cost reductions
in order to demonstrate a genuine alternative to conventional ICEVs. This in par-
ticular promotes the desire for structure optimization of CPVs in order to improve
material efficiency, while maintaining high safety standards.
In that sense, reliability within the design process of CPVs is sought, that is de-
termined by how well the mechanical response is understood, but more importantly
to which accuracy the final collapse can be predicted. To this purpose, analytical
and numerical analysis strategies are appealing, because of the insights they de-
liver at principally lower development costs. Yet, their representative description
of reality can only be recognized once an adequate validation of the assumptions
made, is established.
The validation through experimental testing generally features the determina-
tion of burst pressure or the number of cycles until rupture, which from an indus-

1
Considering 500,000 units per year.
1 Motivation and scope 3

trial point of view, are the most relevant parameters. Nonetheless, final failure at
high pressure stages represents the ultimate result of a complex damage evolution
process, where the exact causes and sequence of events are not always straightfor-
wardly to be determined. This motivates the acquisiton of additional experimental
data to gain further insights into the mechanical response. While the experimen-
tal characterization of composite structures has seen advances in recent years, the
testing of CPVs still remains a challenging task. The characterization requires
sufficiently safe containments, time-consuming measurement instrumentation and
finally the high energy release during burst can lead to damage and potentially loss
of the aforementioned, which reasons the high costs. As such, characterization ap-
proaches are still limited to some extent, which restricts the detailed correlation to
analysis strategies and therefore leads to uncertainties in how well the mechanical
response is understood.
The aim of this work is to contribute to an in-depth understanding of the me-
chanical response of composite pressure vessels under internal pressure loading.
To this purpose, an in situ characterization methodology is initially established,
which permits to evaluate the vessel’s meridional thickness profile, its deformation
behavior and the damage evolution process under internal pressure loading. After
the successful implementation, a high fidelity FE model is developed and vali-
dated. Consequently, insights on the mechanical response are derived by applying
in situ methodology and FE model onto a variety of stacking sequences. Finally,
the transfer of these insights onto a fullscale vessel geometry permits to achieve
a mass saving potential of 17% compared to the state-of-the-art. To this end, the
thesis structure is depicted in Fig. 1.2 and the content of each individual chapter is
explained in the following.
In Chap. 2, a review of literature is presented, where relevant aspects to the
background of CPVs are explained and the current state-of-the-art with regard
to experimental characterization, analysis strategies and the correlation of both is
discussed. Consequently, the research questions for this thesis are formulated in
Sect. 2.5.
In Chap. 3, the ressources, methods and materials utilized within the course of
this work are introduced and described.
Following the formulated research questions, an in situ characterization
methodology is established in Chap. 4, that permits the evaluation of thickness
build-up, the characterization of deformation behavior as well as the monitoring
and the localization of failure during internal pressure loading.
The acquired experimental data is used in Chap. 5 to develop and validate a high
fidelity FE model. The correlation scales on which an agreement between exper-
imental and numerical data is achieved include the vessel’s meridional thickness
4 1 Motivation and scope

In situ characterization methodology


for the design and analysis of composite pressure vessels
Chapter 1 Chapter 2 Chapter 3
Motivation and scope Literature review Material and methods
Research questions

Chapter 4 Chapter 5
In situ characterization methodology FE modeling and correlation

• Evaluation of thickness build-up • Model definition


• Characterization of deformation behavior • Correlation of mechanical response
• Failure monitoring and localization
Insights on mechanical response

Chapter 6 Chapter 7
Influence of stacking sequence Application on fullscale geometry
• Experimental design • Design considerations
• Fiber volume fraction and porosity • Structural analysis
• Deformation behavior • Comparison to the state-of-the-art
• Damage progression and final failure
Insights related to CPV design

Chapter 8 Chapter 9
Design considerations to CPVs Summary and outlook

Fig. 1.2 Thesis structure and content of individual chapters

profile, its deformation behavior, the onset of interfiber failure (IFF) and final burst
pressure.
In Chap. 6, the influence of CPV stacking sequence is studied by applying the
framework of in situ characterization methodology and FE model onto a variety
of stacking sequences. The influence is investigated with regard to consolidation
level, deformation behavior, crack propagation and final burst pressure.
In Chap. 7, the previously derived insights are transferred onto a fullscale vessel
geometry, where the influence of the dome contour to the layup composition is
outlined. The derived design is contrasted against the current state-of-the-art.
Resulting from the structural analysis conducted in the previous chapters, gen-
eral considerations to the design of CPVs are formulated in Chap. 8.
Ultimately in Chap. 9, this thesis is summarized and recommendations for fu-
ture research are given.
Literature review
2

The aim of this chapter is to provide an overview of the relevant aspects in regard
to the background, the analysis and the experimental characterization of composite
pressure vessels. Subsequently, the recent state-of-the-art in terms of numerical-
experimental correlation for CPVs is pointed out and discussed. Finally, this chap-
ter is summarized and the research questions for this thesis are formulated.

2.1 Background to composite pressure vessels

The storage of hydrogen in FCEVs is specified through the global technical reg-
ulation (GTR) No. 13, and more specifically by the regulation UNECE/R134 of
the United Nations [1, 2] with the defined objective to attain or exceed equivalent
safety levels than of those of ICEVs and establish performance-based measures
that do not restrict future technologies. In this context, the fullfillment of a min-
imum burst pressure of 225%1 of the nominal working pressure is required for
the gaseous storage, even though the reduction of this requirement is exhibited
in prospect together with a data-driven performance-linked justification. The cur-
rently most established nominal working pressure levels are 35 MPa and 70 MPa.
The choice for whether one of them depends on the application and is strongly im-
pacted by the given space constraints and/or the envisioned driving range without
refueling. Withal, the high pressures lead to considerably thick walls and in case
of full metallic pressure vessels, heavy weights. This is commonly not acceptable
for transportation purposes. As such, distinct types of pressure vessels have been
evolved, which are differentiated by their structural design. Figure 2.1 shows an
overview of the identified types. Pressure vessels of type I and II render metal-

1
Requirement is linked to the type of main reinforcing constituent, which in this case is
CFRP.

© The Editor(s) (if applicable) and The Author(s), under exclusive license to 5
Springer Fachmedien Wiesbaden GmbH, part of Springer Nature 2022
M. Nebe, In Situ Characterization Methodology for the Design and Analysis of
Composite Pressure Vessels, Werkstofftechnische Berichte j Reports of Materials
Science and Engineering, https://doi.org/10.1007/978-3-658-35797-9_2
6 2 Literature review

Type I Type II Type III Type IV Type V

Metallic vessel Metallic vessel with Metallic liner with Polymer liner with Linerless
partial composite full composite full composite composite
overwrapping overwrapping overwrapping with metallic boss
and metallic boss

Fig. 2.1 Overview of identified pressure vessel types according to their structural design

lic vessels, that in the latter case are partially reinforced with composite. Pressure
vessels of type III and IV are composed of a gas sealing liner component, which is
metallic for type III and polymer for type IV, and a full composite overwrapping.
Pressure vessels of type V are linerless, which means that the composite overwrap-
ping additionally has to ensure the sealing.
The tendencies that can be observed from the introduced vessel types is not only
that the use in composite material progressively increases, but also that the compo-
nent mass for widthstanding similar mechanical loads decreases. Next to that, the
manufacturing complexity and overall costs also increase correspondingly with the
vessel type, which justifies their different purposes. Type I and II pressure vessels
are commonly used for stationary hydrogen storage or in CNG applications. Type
III and IV currently render the most mature vessel types for the hydrogen storage
in mobile applications. Pressure vessels of type V have only found application in
the aerospace sector up to now.

2.1.1 Composites

Within the course of this work, the term composite is associated with a stack of
multiple layers of fiber-reinforced plastic (FRP). Fiber-reinforced plastics are gen-
erally composed of a resin such as epoxy or polyurethan, reinforced with fibers
such as carbon or glass, whereas only the continous type of reinforcement is re-
garded in this thesis. Unidirectional continous fiber-reinforced plastics display or-
thotropic material behavior, which means they exhibit two principal material axes
perpendicular to each other. Whilst they present superior mechanical properties in
fiber direction, their properties rapidly diminish perpendicular to it. As such, in en-
2.1 Background to composite pressure vessels 7

gineering applications layers2 of different orientations are stacked together to form


a multidirectional so-called laminate. Because of their high strength-to-weight ra-
tio, particularly for tensile dominant loading cases, composites have been identified
early on for the application in pressure vessels.

2.1.2 Filament winding

Filament winding renders one of the oldest composite processing methods. Since
its introduction in the late 1940’s, there has been a variety of filament wound struc-
tures in aerospace, military, civil and automotive applications [5–7]. The basic
principle relies on the preferably tensioned placement of dry or wet continuous
fibers on a rotating mandrel based on a prescribed geometric path. The most el-
ementary machine setup is composed of two axes of motion, where one is the
rotation of the mandrel and the other is the horizontal movement of the fiber de-
livery eyelet along the mandrel’s length axis. The introduction of additional axes
of motion, such as the delivery eyelet rotation and the yaw rotation, further en-
hances the fiber placement. After filament winding, curing is generally achieved in
an oven, in which the filament wound component is constantly rotated to prevent
resin accumulation.

Wet and towpreg winding


The two most established processes within the the filament winding industry are
wet winding and towpreg winding. Wet winding renders the currently most com-
mon process, where dry fibers are guided through a resin bath before their appli-
cation on the mandrel. The favorable aspects of wet winding are the low storage
requirements, the tailored on-demand change in resin content and the comparably
low raw material costs of fibers and resins. Yet, the exact definition of resin amount
and viscosity is difficult to maintain throughout, which inevitably leads to a larger
scatter in material properties. Moreover, the desired low resin viscosity to achieve
a sufficient wet-out inevitably leads to a lower tack, which results in a higher
probability of fiber slip during winding. Towpreg winding, which in literature is
also commonly referred to as dry winding [5], classifies the winding with pre-
impregnated fibers. The advantages are the tailorable level of tack, the homogenous
resin content, the cleaner winding setup and the generally higher delivery rates
due to the lack of excess resin. As such, the manufacturing quality resulting from
towpreg winding is superior in comparison to wet winding. The disadvantages of

2
Also denoted as laminae or plies.
8 2 Literature review

towpreg winding are the necessity to store the material in cooled conditions and
the elevated costs for the additional pre-impregnation process, which has been the
major argument for wet winding in commercial applications so far.

Process parameters of influence


The quality of the CPV’s laminate is inherently linked to the process parameters
chosen during filament winding. Among them, the tension at which the fibers are
wrapped around the mandrel, commonly referred to as fiber tension is one of the
most essential [8]. Once applied, the fiber tension translates into radial pressure
which, in case of the cylindrical section, can be described as [9]:

F sin2 ˛
pl D (2.1)
r wt
where pl denotes radial pressure per unit length, F represents the fiber tension, r
is the radius of the underlying surface, ˛ is the tow angle with respect to the longi-
tudinal mandrel axis, and w t the tow width. The applied radial pressure is counter-
balanced by the internal pressure applied to the liner. As such, equilibrium between
both forces is envisioned to prevent plastic deformation of the liner. Next to that,
their interaction not only determines the level of consolidation, but also the distri-
bution of residual stresses through-the-thickness within the laminate. Withal, the
subsequent winding of multiple layers inevitably introduces a consolidation gradi-
ent through-the-thickness, as innermost layers receive larger compaction than their
outermost counterparts. As such, fiber volume fraction (FVF) increases, whilst
porosity decreases towards the laminate inside. Thus, this leads to differences in
material properties depending on the radial position [10, 11].

Layer types
The subsequent winding of composite tows onto a rotating mandrel allows to tailor
the mechanical properties by differential positioning of consecutive tows. Within
the context of CPV filament winding, two main types of layers can be identified,
namely circumferential and helical layers, which are depicted in Fig. 2.2.
Circumferential layers3 are formed by winding consecutive tows, that closely
align with the tangential direction of the vessel’s cylinder. Depending on the band-
width, layer angles close to 90° with respect to the vessel’s longitudinal axis may be
achieved. Through the accordance of loading direction and fiber angle, these lay-
ers render an effective reinforcement to bear the tangential loads, that arise through

3
Also commonly referred to as hoop layers.
2.1 Background to composite pressure vessels 9

a) Circumferential layer b) Helical layer

Fig. 2.2 Comparison of layer types; a) circumferential layer and b) helical layer

the radial cylinder expansion. Moreover, their layer architecture corresponds clos-
est to the one of non-intertwined unidirectional (UD) composites, which is why
they present high mechanical stiffness and strength. Yet, their placement is limited
to the cylindrical section, as the meridional curvature of the domes triggers fiber
slippage. Contrarily, the winding of helical layers results in interwoven and bal-
anced ˙ angle plies. These layers are wound from dome to dome and feature a
constant ply angle in the cylindrical section, which approaches an angle of 90° at
the turnaround, where the tow is tangentially wrapped around the opening circum-
ference. The purpose of helical layers is dependent on the chosen angle4 . While
high-angle layers tangentially reinforce the cylinder and permit to tailor the tan-
gential stiffness distribution in the cylinder-dome transition due to their located
turnaround locations, low-angle layers are sought to carry the meridional loads
and ensure the structural integrity of both dome ends next to the boss.

Tow path description and thickness build-up


The filament winding of cylindrical CPVs requires the tows to drape around the
domes in order to ensure material continuity. Hence, the previously constant angle
in the cylindrical section is prompted to change while approaching the turnaround,
which can be on either geodesic or non-geodesic trajectories. The path definition
determines the change in fiber orientation in the dome which, in turn, has a direct
impact on the vessel’s mechanical performance. Therefore, several works investi-
gated the use of geodesic [12–14] and non-geodesic trajectories [15–17] for CPVs
coupled to structural analysis approaches. Geodesic trajectories connect two ar-
bitrary points on a curved surface by means of minimum arclength and therefore
ensure a stable, wrinkle-free tow application. Following Clairaut law, the geodesic

4
Here it is referred to the effective ply angle in the cylindrical section.
10 2 Literature review

Fig. 2.3 Schematic repre- z


sentation of a tow path in
the dome region, reprinted/ r0
adapted from [18] with rturn
permission from Elsevier
α

r
θ R

Meridian direction
Tow

trajectory for surfaces of revolution can be described as:

r sin ˛ D rturn (2.2)

where r denotes the radius at an arbitrarily position on the curved surface as shown
in Fig. 2.3, ˛ denotes the angle between tow path and meridian direction and rturn
represents the polar opening radius, where the tow is tangentially placed. While
geodesic trajectories present superior tow stability, they also limit the design space,
in the sense that the trajectory is solely defined by the initial ply angle in the cylin-
der and the underlying meridian profile. Contrarily, deviating from the geodesic
trajectory exploits further design opportunities, but also introduces lateral forces
onto the tow, that may lead to wrinkling or slippage. In this context, Zu et al. [15]
provided a mathematical description for non-geodesic trajectories in CPVs:
 
d˛ sin ˛ tan ˛ r 00 r 0 tan ˛
D  02
cos ˛  ; (2.3)
dz r 1Cr r

herein,  denotes the slippage coefficient between tow and underlying surface
whilst z denotes to the axial coordinate of the surface. According to De Carvalho
et al. [19] the slippage coefficient can be described as the ratio of transverse (fb )
to normal (fn ) forces acting on the tow:

fb
D : (2.4)
fn
2.1 Background to composite pressure vessels 11

The normal force acts perpendicularly to the surface and is balanced by the man-
drel’s response, while the transverse force is countered by friction. Therefore, the
tow path is stable as long as the lateral force is smaller than the friction force which
can be written as:

jfb j  jfw j D j fn j; (2.5)

where fw is the friction force and denotes the static coefficient of friction be-
tween tow and surface. Hence, stability of non-geodesic paths is ensured whenever
the slippage coefficient is smaller than the local coefficient of friction. In case of
 D 0, the Eq. 2.3 is simplified to the geodesic solution, described by Eq. 2.2. In
every other case ( ¤ 0), an iterative method such as the Runge-Kutta algorithm
[20] together with initial winding conditions are utilized to obtain the non-geodesic
trajectories. To ensure manufacturability and a smooth transition between geodesic
trajectories in the cylinder and non-geodesic trajectiories in the dome, a continuity
condition is enforced [15]:
 
 .r  rturn /
 D max cos ; (2.6)
2 .R  rturn /

where R refers to the cylinder radius, rturn is the dome radius at the turnaround
point and max is an arbitrarily chosen maximum value of the coefficient of slip-
page. In this context, the work of Wang et al. [21] showed that the resin viscosity
appears to have the most relevant influence on friction in comparison to other
variables such as roving width, winding speed and tow tension. As such, they con-
cluded that slippage coefficients for engineering purposes vary from 0:1 to 0:5
depending on the type of process (wet/dry winding). The completition of a tow
path from pole-to-pole and back is a so-called winding circuit. Repeating multiple
winding circuits leads to the generation of a so-called winding pattern, where gen-
erally full mandrel coverage is sought. The resulting pattern not only determines
the tow-overlaps in the cylinder, but also the tangential points at the polar openings.
As such, the choice defines the layer architecture and therefore its mechanical prop-
erties, but also the resulting thickness build-up at the poles [15]. As the material
build-up at the poles is inherently linked to the vessel’s mechanical performance,
thickness estimations are crucial for the determination of burst strength and have
therefore been subject to numerous works [22–24]. The challenges herein include
the consideration of bandwith, the influence of manufacturing-related thickness
smoothing and the change in mandrel shape due to previously wound layers. Leh
et al. [18] have shown the most recent progress on this matter, combining non-
geodesic trajectory descriptions [15] together with the thickness prediction based
12 2 Literature review

on cubic spline functions presented by Wang et al. [25, 26], where thickness is
predicted by:
nR mR tp h r 
turn
 r i
b
t .r/ D arccos  arccos for r2b  r  R (2.7)
 r r
and

t .r/ D A C B  r C C  r 2 C D  r 3 for rturn  r  r2b ; (2.8)

with nR and mR being the number of helical subplies and their count in the cylin-
drical section. Correspondingly, the values of rb and r2b denote the pole radius
at one and two tow-width distances from the turnaround point, rturn of every in-
dividual ply. The constants A, B, C and D are the fitting coefficients that define
the layer end thickness, which are determined from the helical ply thickness at the
turnaround point, the continuity of thickness and curvature in r2b and a constant
material volume between rturn and r2b . A helical ply is then defined as a group
of subplies that has the same winding angle (˙˛) in the cylindrical section. The
parameters nR and mR are described as:
tR
nR D (2.9)
2tp
2R cos ˛R
mR D ; (2.10)
b
where tp and b are the thickness and the width of the filament band, tR and ˛R are
the thickness and the initial winding angle of that layer in the cylindrical region.
The final thickness prediction is achieved through an iterative process, where the
underlying mandrel shape is readjusted according to the predicted thickness build-
up of each subsequently wound helical ply. Finally, a comparison between the
predicted and real thickness build-up is established through the use of a computed
tomography (CT) scan, shown in Fig. 2.4, that implies good accordance.
Altogether, the accurate estimation of thickness build-up in CPVs renders a
complex task which is dependant on a multitude of variables. Among those are the
trajectory description, the tow width consideration, the change in mandrel shape
due to previously wound layers and the number of tangential points depending
on the pattern of choice [27]. Furthermore, manufacturing related aspects such
as layer consolidation appear as additional influence. Thus, inaccurate thickness
estimations directly impact the prediction of the vessel’s final strength, which
inevitably results in uncertainty within the design process. For the industrial ap-
plication, several commercial filament winding softwares offer the convenience to
2.2 Analysis of composite pressure vessels 13

Fig. 2.4 Comparison of 70


thickness prediction (red)
derived through coupling 60
of non-geodesic trajec-
tory description and roving
thickness estimation to a 50
CT scan of a CPV, reprint-
ed/adapted from [18] with 40

r [mm]
permission from Elsevier
30

20

10

0
0 5 10 15 20 25 30 35 40 45 50
z [mm]

design trajectories, derive thickness estimations and export machine motions for
the filament winding setup. Among these are ComposicaD™ , CADWIND, Cadfil®
and Wind.

2.2 Analysis of composite pressure vessels

This section provides an overview of the structural analysis strategies to CPVs and
introduces approaches, that progressively increase in their modeling complexity
and accuracy, yet mostly accompanied by an increasing computational cost.

2.2.1 Analytical approaches

Analytical approaches provide general insights into the mechanical response of


thin- and thick-walled composites. With regard to composite pressure vessels, the
most common approaches are rooted in the so-called netting theory, the classical
lamination theory (CLT) and three-dimensional elasticity theory, which are briefly
described in the following.
14 2 Literature review

Loads in pressurized shells of revolution and netting theory


In the most fundamental way, the vessel surface can be treated as thin-walled shell
of revolution. The in-plane loads, that arise in a double curved shell of revolution
due to the loading with internal pressure, can be described as [12]:

Ns D p  r =2
N D Œ2  r =rs p  r =2
Ns D 0 (2.11)

where p is the internal pressure, rs and r refer to the principal radii of curva-
ture, whilst the subscripts s and  denote the meridional and tangential direction,
respectively. In case of a cylindrical shell (r D R; rs ! 1), Eq. 2.11 reduces to:

Ns D P  R=2
N D P  R (2.12)

which for thin-walled pressurized cylinders of isotropic materials leads to the com-
monly known hoop-to-axial stress ratio of 2:1. In contrast to isotropic materials,
unidirectional composites display orthotropic material behavior and can therefore
be used to distribute stresses more equally. Their superior mechanical strength in
fiber direction compared to their transverse and shear strength, lead to the most
fundamental analytical description for composites, the so-called netting theory
[28]. According to the netting theory, the composite structure carries its load solely
through fiber tension, whereas any contribution of the matrix constituent is effec-
tively disregarded. Furthermore, it is assumed that solely membrane loads arise
and no out-of-plane bending or shear exists. As such, an optimum design is found
from the fiber orientation, and correspondingly the surface shape, that results in
constant fiber tension throughout, which is the so-called isotensoid solution [29].
Effectively, the relationship between resultant in-plane membrane forces in a pres-
surized shell of revolution and the fiber tension in a bidirectional helical layer can
be described as follows [12]:

Ns D L tR cos2 ˛
N D L tR sin2 ˛
Ns D 0 (2.13)

where L denotes the fiber stress, tR denotes the ply thickness and ˛ refers to the
ply angle with respect to the meridional direction. Coupled with formulation of
2.2 Analysis of composite pressure vessels 15

in-plane forces in Eq. 2.11 and the filament winding path description in Sect. 2.1,
the isotensoid solution can be obtained. Altogether, netting theory renders a time-
efficient design tool for the preliminary analysis of CPVs, that has seen application
in several works throughout [28, 30, 31]. Yet, because of the established assump-
tions its validity decreases, particularly with increasing wall thickness.

Classical lamination theory and thick-walled cylinder solutions


The CLT frames the elastic constitutive formulation that considers the behavior of
orthotropic layers within a composite laminate. Hereby, it is assumed that the in-
dividual layers are perfectly bonded and that normal plane sections remain normal
and plane after deformation. Moreover, it is assumed that the layer thickness is
thin compared to their in-plane dimensions and that stress components in the out-
of-plane direction are negligible, meaning that a state of plane stress is assumed.
Force and moment resultants can be related to the deformation and curvature of
the laminate by the following expression:
( ) ( )( )
Nt A B "0
D (2.14)
Mt B D 

where N t and M t denote the force and moment resultant tensor, respectively, "0
displays the mid-plane strain tensor and  is the curvature tensor. ŒA is the ex-
tensional stiffness matrix, which relates in-plane stress resultants to the mid-plane
strains of the laminate. ŒB is the bending-extension coupling matrix, which relates
in-plane stress resultants to curvatures and moment resultants to mid-plane strains
of the laminate; and ŒD is the bending stiffness matrix, which relates moment re-
sultants to the curvatures of the laminates mid-plane. Considering the shell of a
pressure vessel with helically wound layers, each of which is formed by two ˙
subplies of the same material, the Eq. 2.14 can be expressed as [32]:
! " # !
Ns A11 A12 "s
D  (2.15)
N A21 A22 "

where the subscripts s ; denote the global component axes and 1 ;2 the local ma-
terial axes. The individual components of the extensional stiffness matrix ŒA are
calculated by:

X
n
ŒA D ŒQk  t k D t  ŒQk (2.16)
kD1
16 2 Literature review

where the superscript k indicates the kth layer, t k denotes individual layer thick-
k
ness, n is the total number of layers, tl is the total laminate thickness and ŒQ is
the reduced stiffness matrix of the kth layer in the global coordinate system, that
can be obtained from the reduced stiffness matrix ŒQ as follows:

ŒQk D .ŒT k /1  ŒQ  ŒT k (2.17)

being T k its corresponding transformation matrix from the local to the global co-
ordinate system. The components of both, transformation matrix T and stiffness
matrix ŒQ are further detailed in [33]. The stress-strain relation of the kth layer is
expressed as:

k
s; D ŒQk  s;
k
D ŒQk  s; (2.18)

therefore the stress components at each individual layer expressed in terms of the
membrane forces can be assessed by:

k
1;2 D ŒT k  ŒQk  ŒA1  Ns; : (2.19)

Coupled with the in-plane force formulation of Eq. 2.11 and the winding tra-
jectorie descriptions in Sect. 2.1, CLT enables a more refined analysis than the
netting theory by considering the orthotropic material behavior. As such, several
works demonstrated its application to CPVs regarding the optimization of dome
shapes [13, 15, 34], the prediction of burst pressure [35, 36] and as preliminary de-
sign method for the subsequent refinement through a high fidelity FE model [37].
Altogether, the use of CLT renders a versatile analysis method that comes at a rea-
sonably low computational cost. Nonetheless, the CLT, like the netting theory, is
rooted in the thin-wall assumption, which neglects the presence of normal stress
gradients through-the-thickness as a result of the thick-wall condition. As a result,
its prediction accuracy decreases with increasing wall thickness as indicated by
Parnas and Katirci [38].
To consider for these normal stress gradients through-the-thickness, further ana-
lytical analysis strategies for thick-walled composite cylinder have been developed
[39–41]. In this context, Xia et al. [42] provided a description for multi-layered
filament wound composite pipes in which the laminate is considered of cylindrical
three-dimensional anisotropy. Their work, in which they compare three different
stacking sequences in terms of deformation behavior and stress state through-
the-thickness, underlines the importance of stacking sequence with regard to the
arising normal stress gradient through-the-thickness.
2.2 Analysis of composite pressure vessels 17

Fig. 2.5 Meridional mo- π/4β


Ms
ment distribution at the
cylinder-dome transition π/4β
of an isotropic pressure
vessel with elliptical heads h -0.094 pRh
and constant wall thickness 0.0976 pRh
during internal pressure
p
loading, reprinted/adapted
from [43] R Cylinder Dome

In conclusion, analytical approaches to CPVs are manifold. The most fun-


damental description is the netting theory, which considers load transfer solely
through fiber tension and permits to establish time-efficient preliminary analy-
ses. A more refined description is found through the use of CLT, which considers
the contribution of the matrix constituent. Both analysis approaches, coupled to
geodesic and non-geodesic trajectories, allow for the detailed analysis of opti-
mum dome shapes and meridian profiles, yet limited to the meaningful predic-
tion of thickness build-up. Another contraint is the thin-wall assumption of both
approaches, which invitably decreases their validity with increasing CPV wall
thickness. As such, the use of three-dimensional elasticity approaches for the con-
sideration of through-the-thickness normal stress gradients is insightful, yet limited
to the cylindrical section of the CPV. Nevertheless, the CPV’s structural integrity
is determined by cylinder, dome and their interaction at the transition, when in-
ternally pressurized. In this regard, most common analytical approaches assume a
state of sole membrane forces. Nevertheless, the individual expansion of cylinder
and dome constrained by their rigid connection leads to the existence of bending,
as pointed out by Baličević et al. [43]. In their work, they assess the meridional
bending at the cylinder-dome transition by approximating the solutions to the dif-
ferential equations of shells of revolutions by membrane and moment theories
superposition. Figure 2.5 shows the meridional moment distribution for an inter-
nally pressurized isotropic vessel with elliptical heads.

2.2.2 Finite element analysis

Given the complexity of the aforementioned and the constraints of analytical de-
scriptions, the use of FE modeling is motivated by the possibility to discretize
complex geometries, to account for non-linearities at material level, to model large
18 2 Literature review

Fig. 2.6 FE model of a


CPV composed of metal- CFRP
lic boss, plastic liner and
overwrapped CFRP where
the distinct colors indicate Liner
different material defini-
tions, reprinted/adapted
from [44] with permission Boss
from Elsevier

deformations and to reproduce contacts between components. The application to


CPVs features a large quantity of works, whereas only a selection is mentioned in
this subsection. Most commonly, the modeling of CPVs is achieved on mesoscale,
where the composite and its distinct constituents are assumed as homogeneous or-
thotropic continuum. In this context, the considered element types vary from two-
dimensional shell [27], axissymmetric shell [44–46], to three-dimensional solid
definitions [45, 47, 48]. Withal, the most accurate depiction of the thick-wall con-
dition appears to be achieved by the solid definition, yet in expense of the highest
computational cost. Furthermore, the mechanical response and final failure are
assessed by either the constitutively elastic solution [27, 49] or by considering
damage in the context of continuum damage mechanics (CDM) [45, 48, 50]. The
works by Leh et al. [18, 44, 46] present a comprehensive methodology to derive
accurate thickness estimations for non-geodesic trajectories, perform FE simula-
tions with the consideration of damage and establish an optimization framework
for the reduction of mass. Based on a pre-determined thickness profile, shown in
Fig. 2.6, axissymmetric and solid definitions are investigated.
Mechanical properties are derived through the testing of filament wound
coupons, that also consider the intertwined layer architecture of helical plies.
No boss plasticity is considered and the composite-liner interface is treated as tied.
Their findings show an overall good accordance to experimental results for both,
shell and solid definitions, even though discrepancies are assumed to exist because
of inadequate determination of strength properties, difficulties in determining ac-
curate ply thicknesses, material heterogenities due to ply relaxation, inadequate
damage variables and the absence of plasticity in the metallic bosses [46].
An alternative modeling approach applied to CPVs is the multiscale analysis,
where fiber and matrix constituent are represented separately in a so-called unit cell
2.3 Experimental characterization of composite pressure vessels 19

[51, 52]. Contrarily to mesoscale approaches, this allows for a detailed analysis of
microscopic damage (e.g. fiber-matrix debonding). Withal, the difficulty for multi-
scale approaches resides in adressing the specific circumstances related to CPVs in
terms of filament paths, dome shapes, ply drop locations and consolidation differ-
ences. These result in distinct micromechanical characteristics as they impact the
distribution of FVF and porosity, thus relevantly affecting the prediction accuracy.
In this context, the consideration of material variability in CPVs and its in-
fluence on the burst pressure prediction has also been subject to multiple works
[53–55]. While Gentilleau et al. [53] focus on the development of a damage model
that considers the statistical distribution of fiber strength in CPVs, Harada et al.
[54] demonstrate the influence of different fiber packing by considering proba-
bilistic distributions of FVF. More recently, Rafiee and Torabi [55] considered a
broader range of manufacturing-induced material variability in CPVs by applying
probabilistic functions to variables such as fiber volume fraction, winding angle
and mechanical strength, which lead to a stochastic distribution of vessel burst
pressure.

2.3 Experimental characterization of composite pressure


vessels

The high internal pressures that CPVs are generally subjected to, constitute a
potential safety risk for civil or automotive applications. As such, the vessel’s me-
chanical response needs to be well understood and its structural integrity needs to
be ensured throughout its in-service life. This motivates the acquisition of experi-
mental data to gain further insights into the on-going mechanisms when internally
pressurized, whether it may be for design validation purposes or for the monitoring
during in-service life. In this section the experimental characterization approaches
to CPV by means of strain and displacement monitoring and acoustic emission
(AE) analysis are presented and discussed.

2.3.1 Strain and displacement monitoring

The purposes for the acquisiton of strain and displacement data in CPVs presented
in literature are various, but can be generally separated in three categories: process
control during manufacturing, experimental validation of analysis stategies and
monitoring during in-service life. The measurement of discrete displacements can
be achieved through the use of linear variable differential transformers (LVDTs),
20 2 Literature review

Fig. 2.7 Embedment of


a FBG sensor line during
filament winding, reprint-
ed/adapted from [62] with
permission from Elsevier

FBG sensor line

which convert a rectilinear motion into an electric signal. Thus, boss and cylinder
displacement can be measured at principally low investment costs [45, 46]. Yet,
particularly in the case of the boss displacement, the interpretation of this value
is not straightforward as it is influenced by multiple variables including geometric
non-linearities, damage evolution in interfaces and non-linear contacts [45]. Sim-
ilarly, discrete outer surface strain information can be obtained through the use of
applied electrical resistance strain gauges (ESGs). Well-established in mechanical
testing ESGs link a change in resistance to a strain by a so-called gauge factor.
As such, they render a comparably easy method to acquire strain information for
CPVs [56–58]. Yet, their drawbacks are the limited spatial resolution of strain,
their vulnerability to outer surface damage and the sensitivity of alignment to the
principal measuring direction. A considerably more promising approach, which
is often referred to in regard of non-destructive evaluation (NDE) [59–61] is the
strain monitoring by using fiber optic sensors (FOSs) with integrated fiber Bragg
gratings (FBGs). Their principle of operation resides in the change of the reflected
wavelength due to mechanical altering of the grating’s period. Due to multiplexing,
the integration of several FBGs into a single optic fiber can be achieved. This al-
lows to obtain strain information at various discrete positions and therefore permits
an overall higher spatial resolution of strain. Thus, these sensors can be integrated
in between layers during filament winding as Fig. 2.7 shows. This not only per-
mits to monitor the manufacturing process, but also allows to measure strains at
different positions through-the-thickness [45, 58].
Nevertheless, the controllability of their exact placement presents a limitation
as the low viscosity resin prompts slippage during manufacturing and may fur-
ther cause readjustments during curing. Next to that, preliminary failure can be
2.3 Experimental characterization of composite pressure vessels 21

a) Tangential strain b) Meridional strain c) Shear strain


Implemented flaw

Fig. 2.8 Display of full field a) tangential, b) meridional and c) shear strain for a CNG vessel
at 10 MPa, reprinted/adapted from [60] with permission from Elsevier

triggered due to their brittle behavior [62]. While the spatial resolution can be sub-
stantially increased through a network of FBGs as demonstrated by Tapeinos et
al. [61], a full field strain resolution can only be achieved by means of digital im-
age correlation (DIC). Whereby, local strain information is obtained through the
comparison of subsequently acquired images. Its application to CPVs has been
demonstrated by correlating the derived strains to the results of other measuring
methods such as FBGs and ESGs [56, 57, 60]. Thereby, the detection of spatially
deviating strain fields stands out as promising feat to gain a comprehensive under-
standing of the deformation behavior. In this context, the work of Gasior et al. [60]
demonstrated the possibility of detecting locally varying strain fields by imple-
menting flaws in the vessel’s cylinder, as Fig. 2.8 shows. Nonetheless, the work’s
focus was rather on the development of a structural health monitoring (SHM)
framework than on the analysis of the vessel’s mechanical response. Among the
drawbacks of DIC and its application to CPVs are the necessity to apply a random-
ized surface pattern and the comparably more complex and cost-intense testing
setup, particularly in case of burst experiments.

2.3.2 Acoustic emission analysis

The term acoustic emission analysis generally refers to the detection and interpreta-
tion of ultrasonic waves caused by rapid displacements, whereas their most relevant
source is understood to be the formation and propagation of cracks [63]. The de-
tection of AE signals by piezoelectric sensors renders an insightful approach for
the in situ monitoring of microscopic failure within composites (e.g. fiber break-
age, matrix cracking and interfacial failure) and has seen substantial advances in
the past decades [64–66]. The recognition allows to count, localize and identify
the source of signals. Yet, during propagation in the guiding media, the signal’s
22 2 Literature review

characteristics suffer from phenomena like attenuation and dispersion, which sub-
stantially increases the sophistication for the application in large scale composite
structures. The usage of AE analysis for CPVs can be generally differentiated in
three categories: burst prediction, failure localization and mechanism differentia-
tion.
A well-established concept for the failure analysis of CPVs is achieved by
detecting AE activity during a predefined loading-unloading scheme, such as
specified by the standard ASTM E1067/E1067M-18 [67]. Here, the specimens
are subsequently loaded and unloaded, whereas the maximum load progressively
increases after each cycle and the onset of AE activity is constantly monitored
throughout. After each cycle, the loading stage at which AE activity is recognized
is put into relation to the previously applied load. This ratio is well-known as
Felicity ratio:

LFR;N
FRN D (2.20)
LN 1

where LFR;N refers to the load stage at which AE activity onset is recognized and
LN 1 denotes the previously applied load. In this context, the designated Felicity
effect describes the occurence of AE activity at loading stages lower than previ-
ously applied [68]. Metallic materials, generally display the so-called Kaiser effect,
which describes the absence of AE signals until the previously applied load is
reached [69]. However, in fiber-reinforced plastics AE signals often initiate before
reaching the previously applied load, which is the reason why the Felicity ratio be-
came an important concept to understand the progression of failure in composites.
Herein, to increase the reliability of the calculated Felicity ratio, different concepts
for the determination of AE onset have been established [70–72]. Repeating this
procedure for several loading cycles, yields into the derivation of multiple Felicity
ratios that can be assessed in dependance of the applied load. A similar measure
for the damage progression was introduced by Downs and Hamstad [71]. The so-
called Shelby ratio, which in contrast to the Felicity ratio, evaluates the AE activity
during the unloading of the component. In their work, they were able to distinguish
between four impacted and two non-impacted CPVs based on the calculated aver-
age Shelby countup ratio, which is shown in Fig. 2.9.
Concluding from the previously implied a correlation of both quantities, Fe-
licity and Shelby ratio, to the resulting burst pressure, Sause et al. established an
approach for the forward failure prediction of CPVs and coupon specimens, which
uses AE ratios in combination with artificial neural networks (ANNs) [72–74]. For
two different vessel sizes, they were able to achieve precise burst pressure pre-
2.3 Experimental characterization of composite pressure vessels 23

Fig. 2.9 Average 95


Shelby countup ratios 13500
for four impacted and 90 13000

Burst Strength [MPa]


two non-impacted CPVs,

Burst Strength [psi]


12500
reprinted/adapted from [71] 85
12000
80 11500

75 11000
10500
70 Non-impacted 10000
Impacted
65 9500
0.0 0.2 0.4 0.6 0.8 1.0
Average Shelby Countup Ratio
(Sum of All Hits From All Sensors)

dictions, which only deviated by 3% to the experimentally determined pressures


[73, 74].
Besides the determination of burst pressure, insights into spatial distribution of
emissions and the potential failure location are of interest, too. Herein, the chal-
lenges faced for the AE localization in CPVs are the large component dimensions,
which require a suffciently fine sensor grid and the multi-directional fiber orien-
tations, that cause non-uniform wave propagation, dispersion and attenuation. In
this regard, Kalafat et al. demonstrated the virtues of an ANN approach over the
classical t -based location by comparing both on the same vessel geometry [75].
According to their results, the ANN approach offered an increase in localization
accuracy by a factor of six in comparison to the t -based localization. Sause et
al. also implemented this ANN based localization into their forward prediction ap-
proach, showing that besides the forecast of failure pressure, also the location of
failure may be predicted [73, 74]. Their prediction seemed to coincide with the fi-
nal failure location and they propose a correlation to analysis approaches as a very
promising matter of future work.
Next to the prediction of burst pressure and failure location, studies further-
more aimed at differentiating failure mechanisms in CPVs [76–78]. AE has been
commonly used for differentiating failure mechanisms in composites and is par-
ticularly well-established for two-dimensional coupon specimens. Mostly, features
such as energy, amplitude and frequency (peak or centroid) are first obtained from
the signal and later partitioned by supervised or unsupervised machine learning
techniques, such as k-means. Withal, as pointed out by Chou et al. [76] the appli-
cation to CPVs places a high sophistication to the task because of the sensitivity
24 2 Literature review

of signal features to the sensor-source distance. As such, the mostly large vessel
dimensions require a sufficiently dense sensor grid to ensure an unambigious de-
termination of failure mechanisms.

2.4 Correlation of numerical and experimental analysis

The purpose of correlating numerical and experimental analyses is twofold. On


the one hand, by performing precise experimental testing, the conformity of an
analysis approach can either be validated or declined. Vice versa, through a high
fidelity numerical model, experimental results can be scrutinized regarding their
validity. In consequence, both analyses complement each other and help to gain
a comprehensive understanding to an engineering problem. The following sec-
tion introduces and discusses work related to CPVs in that matter. To this end,
an overview of the established correlations between experiment and simulation is
shown in Fig. 2.10, where the degree of complexity in experimental data acquisiton
and the degree of complexity in FE modeling is schematically indicated.
The parameter on which the majority of correlations for CPVs is established,
which at the same time is also the one of most relevance, is the final burst pressure
[11, 27, 52]. Withal, the determination of this value is influenced by a multitude of
variables, both in experiment and numerical modeling. Next to that, the complex-
ity to derive this parameter is different for both approaches. From an experimental
standpoint, the burst pressure is comparably easy to be determined as it solely
requires a pressuring device, -gauge and a safe testing environment. Yet, its nu-
merical determination stands at the end of the modeling approach, whereas it can

Fig. 2.10 Schematic overview of correlation between experiment and simulation with indi-
cated literature sources
2.4 Correlation of numerical and experimental analysis 25

25000 C4 strain
(experimental)
20000 C4 strain C17
Strain (μm/m)

(simulated)
15000 C12 strain
(experimental)
10000 C12 strain C12
(simulated)
C4
5000 C17 strain
(experimental)
0 C17 strain
0 400 800 1200 1600 (simulated) Liner Helicals Circumferentials
Pressure (bar) Optic fiber
a) Predicted and experimental optic fiber strains b) Location of optic fibers through-the-thickness

Fig. 2.11 Correlation of predicted and experimental strains obtained through integrated optic
fibers, reprinted/adapted from [45] with permission from Elsevier

be either predicted by considering first-ply-failure (FPF) or by accounting for dam-


age progression.
Likewise, displacements and strains can be directly accessed by FE model-
ing, but require further instrumentation in the experimental practice. Herein, the
more insightful the data to be acquired is, the more complex is the measurement
setup. While the acquisition of radial and axial displacement allows to correlate
the vessel’s global deformation [45, 46], its meaningfulness is limited by the dis-
crete information, which can be aggregated by a multitude of influences. In the
works of Leh et al. [46] and Berro Ramirez et al. [45] a generally better accor-
dance was achieved for the displacement in radial direction, compared to the axial
direction. This appears to be related to the number of variables that influence both
values. In radial direction, the displacement is solely influenced by the laminate
response to the cylinder expansion. The axial displacement however, is aggregated
by the overall deformation of the full vessel, which includes the transitions and
both domes. Next to that, the yielding of the boss, the boss-composite interface
and additional effects, such the presence of a composite-boss gap, can influence
the displacement [45]. This hinders to identify the exact causes for discrepancies
in correlation. Similarly to displacements, authors also correlated discrete outer
surface strains obtained from ESGs in either principal vessel axes [62, 80] or in the
local layer orientation [54]. Nonetheless, motivated by the normal stress gradient
through-the-thickness, the correlation of inner surface strain data poses a gener-
ally more relevant matter, which can further enhance the structural understanding.
In this context, Berro Ramirez et al. [45] compared the tangential strain at differ-
ent positions through-the-thickness, as depicted in Fig. 2.11. While a particular
26 2 Literature review

Analysis (Helical_between ply 1 & ply 2) 0.28


Experiment (FBG_Ch1)
0.24
Analysis (Helical_between ply 2 & ply 3)
Experiment (FBG_Ch2) 0.20
Analysis (Helical_top surface)

Strain (%)
Experiment (ESG_dome) 0.16
Analysis (Hoop_between ply 3 & ply 4)
Experiment (FBG_Ch4) 0.12
Analysis (Hoop_Hoop_top surface)
Experiment (ESG_cylinder) 0.08

0.04

0.00
-400 -300 -200 -100 0 100 200 300 400
Z-axis (mm)

Fig. 2.12 Correlation of predicted and experimental strains at different axial positions for an
internal pressure of 3:45 MPa, reprinted/adapted from [62] with permission from Elsevier

good accordance in outer hoop strain (C17) is achieved, the inner hoop strain (C4)
presents a more distinct deviation, which might be related to the gradient in mate-
rial properties, that is not being considered in the model.
In contrast to the vessel cylinder, that generally presents a uniform strain dis-
tribution along its meridional path, the strain distribution in the domes is much
more heterogeneous. As such, insights about the local strain variation contribute
to a comprehensive understanding of the deformation behavior of both subcom-
ponents, cylinder and dome. Withal, the double-curved dome surface hampers the
correct placement of ESGs and FBGs to a certain extent, which is reported by Kang
et al. [62]. In Fig. 2.12 ply-strains derived from a three-dimensional solid model
are contrasted against the discrete measured strains. While, a good accordance for
cylinder strains was achieved, differences in the dome are recognized, that were
ultimately traced back to misalingments during curing. Next to the correlation of
strains and displacements, the comparison of the damage progression process is
reported by Chou et al. [76]. In their investigation, the modeling of variability
in fiber strength and its influence on the damage evolution process is contrasted
against AE analysis conducted on UD coupons and type III/IV pressure vessels.
The results herein reveal, that clusters of broken fibers evolve steadily throughout
the loading until their density eventually rapidly increases prior to rupture, which
coincides with a sudden increase in AE activity. Yet, the high sensor-source sensi-
tivity of AE is identified as a remaining drawback for the comparably large CPV
dimensions.
2.5 Summary and research questions 27

2.5 Summary and research questions

From the presented review of literature, it can be inferred that a multitude of vari-
ables influence the structural integrity of the filament wound vessel. Among these
are the tow path description, the thickness build-up, the ply consolidation level, the
material properties and the damage evolution process that together define its final
strength. Given the aforementioned complexity, analysis strategies and in situ mon-
itoring approaches were established to gain a further comprehensive understanding
of its mechanical response under internal pressure loading. Withal, their individual
representativeness is limited without further validation. In this regard, the correla-
tion between experiment and simulation is mostly established by means of burst
pressure, which in turn, hampers to confidently pinpoint the reasons for discrep-
ancy due to the variety of influences. Next to the magnitude of burst pressure,
the location of failure in composite pressure vessels can be largely diverse. Here,
the prediction of cylinder failure is generally affiliated with a reasonably good ac-
cordance, that has been demonstrated by various authors [44, 45, 52]. Likewise,
the acquisition of discrete strains and displacements in the cylindrical section ren-
ders the current state-of-the-art. Contrarily, the considerably more heterogeneous
cylinder-dome transition and dome region impede the predictions’ certainty. Next
to that, currently available characterization approaches present themselves as lim-
ited to derive spatial information in these areas. Therefore, the identification of
causes for such failures, seen in Fig. 2.13, are not straightforwardly to be de-
termined. To ensure the reliable prediction of burst pressure commonly leads to
designs that preserve cylinder failure, which at the same time comes at the expense
of a too conservative reinforcement of the domes.

Fig. 2.13 Remainders of a type III vessel after burst testing [11]
28 2 Literature review

Nonetheless, envisioning an optimized usage of CFRP requires sound vessel


designs in which the structure is equally loaded throughout. To achieve this, a
comprehensive understanding of the vessel’s entirety under internal pressure load-
ing is sought, that can only be gained by the accordance of both, numerical analysis
and experimental validation. To this purpose, the presented thesis follows five main
research questions, which can be individually assigned to the subsequent chapters:

Research question 1:
How can composite pressure vessels and their mechanical response under internal
pressure loading be characterized in a thorough and comprehensive way?
Chap. 4

Research question 2:
How and to which extent can experimental and numerical analysis be aligned re-
garding the CPV’s mechanical response under internal pressure loading?
Chap. 5

Research question 3:
What influecing factors need to be taken into account with regard to the definition
of a CPV stacking sequence?
Chap. 6

Research question 4:
What considerations need to be made to transfer previously derived insights onto
a fullscale vessel geometry with distinct dome contour?
Chap. 7

Research question 5:
What general design considerations can be drawn from the conducted structural
analysis?
Chap. 8
Material and methods
3

Within this chapter, utilized ressources and methods for the filament winding, the
experimental characterization and the structural analysis of CPVs are introduced.
Furthermore, the determination of material properties for filament wound lami-
nates and CPV laminate sections is described. To this purpose, Fig. 3.1 gives an
overview of the methods and ressources that are detailed in the following sections.

3.1 Robot-assisted towpreg winding

The material properties found within the final vessel are inherently linked to the
manufacturing process and as such, strongly impacted by the chosen winding pa-
rameters [8, 10]. Hence, precise manufacturing is required to establish comparisons
between different sets of vessel configurations. Therefore, the entirety of vessels
presented in this work were manufactured within the same facilities of Mercedes-
Benz AG, Suttgart, Germany, that are introduced in the following.

3.1.1 Manufacturing

The vessels are filament wound using a robot-assisted winding system, shown in
Fig. 3.2, for which the software ComposicaD™ (Plastic Omnium, Levallois-Perret,
France) is used to define machine motions and winding patterns. Individual spool-
specific fiber tensions are constantly monitored and controlled through torque-

Supplementary Information The online version contains supplementary material avail-


able at https://doi.org/10.1007/978-3-658-35797-9_3.
Note: Parts of the content presented in this chapter have been previously published [81–83]
or are derived from supervised student theses [84–87].

© The Editor(s) (if applicable) and The Author(s), under exclusive license to 29
Springer Fachmedien Wiesbaden GmbH, part of Springer Nature 2022
M. Nebe, In Situ Characterization Methodology for the Design and Analysis of
Composite Pressure Vessels, Werkstofftechnische Berichte j Reports of Materials
Science and Engineering, https://doi.org/10.1007/978-3-658-35797-9_3
30 3 Material and methods

Section 3.1 Section 3.4


Robot-assisted towpreg winding Structural analysis
• Multi-axis filament winding robot • 3D Elasticity theory
• EP/ CF towpreg
eg material • Finite
F element analysis
• FW software Co
ComposicaD™
C m
mposicaD™ • CDM
C model CompDam

Section 3.2 Section 3.3


Experimental characterization Determination of material properties
• Test chamber • Coupon tensile tests
• Stripelight projection (longitudinal/ transverse/ shear)
• 3D Digital image correlation • ENF/ DCB tests
• Airborne acoustic emission/ delay-and-sum beamforming • Acid digestion/ CT scanning

Fig. 3.1 Overview of utilized ressources and methods within this thesis

a) Multi-axis filament winding robot b) Rotated eyelet at turnaround of helical layer

Fig. 3.2 Robot-assisted winding system; a) multi-axis filament winding robot and b) rotated
eyelet at turnaround of helical layer
3.1 Robot-assisted towpreg winding 31

controlled engines. The vessels are internally pressurized during the whole man-
ufacturing process. During the curing, they are also constantly rotated to prevent
resin accumulations. Manufacturing quality is assured through frequent measure-
ments of dimensions and weight.
As composite reinforcement, this work considers a towpreg system composed
of an epoxy resin and high-strength standard-modulus carbon fibers. The choice
resides in the commercial availability of the material system and its well-known ap-
plication for composite pressure vessels. Mechanical characteristics such as elastic
properties, ply strengths and fracture toughnesses are obtained through the testing
of filament wound coupons, which is further detailed in Sect. 3.3.

3.1.2 Liner geometries and layups

Two different liner geometries, depicted in Fig. 3.3, are investigated within this
work. The majority of the work is carried out on the subscale geometry to save
material and time ressources. Subsequently, the insights derived from the structural
analysis are transferred onto the fullscale geometry to investigate the influence of
the distinct dome contour and to show the applicability of this work on a geometry
that is representative for the use in FCEVs. Both, subcale and fullscale geometry
are composed of a polyamide 6 (PA6) mandrel with two aluminum bosses. The
coordinate system definition is shown in Fig. 3.3 c), where s designates the merid-
ional coordinate, ' indicates the tangential coordinate and z represents the axial

a) Subscale geometry (V = 8.6 L) c) Coordinate system definition


s
r
α
φ

z CFRP
0 200 400 585 -85 0 85
Length [mm] Radius [mm]
b) Fullscale geometry (V = 39.1 L)

0 200 400 600 800 1000 1200 1400 1595 -97 0 97


Length [mm] Radius [mm]

Fig. 3.3 Overview of liner geometries and coordinate system definition


32 3 Material and methods

a) Layups on subscale geometry

*†‡ A * B * C * D

b) Layups on fullscale geometry c) Occurence in thesis


* Chapter 4: In situ characterization methodology
† Chapter 5: FE modeling and correlation
** E ** F ‡ Chapter 6: Influence of stacking sequence
** Chapter 7: Application on fullscale geometry

d) Labeling scheme: stacking sequence change e) Labeling scheme: circumferential layer change

A ‡ A-S1
A1 A * A-C1
A1

Note: Changes in polar opening, circumferential tapering or other manufacturing parameter are separately indicated
5° ≤ α < 40° Helical layer 40° ≤ α < 88° Helical layer Circumferential layer

Fig. 3.4 Overview of the investigated main layups, their occurence in the following chapters
and the introduction into the labeling scheme of stacking sequence as well as circumferential
layer variation

coordinate. For every vessel of both geometries, the coordinate origin is located at
the center of the boss end.
The side of where the origin is located, is determined by how the vessel is
clamped inside the test chamber, seen in Fig. 3.5. The coordinate origin is always
located at the boss end, which is clamped in the locating bearing.
A brief overview of the most relevant layups in this work is given in Fig. 3.4,
where their occurence in the following chapters is indicated. Hereby, the pic-
tograms represent the layup composition qualitatively with three identified layer
categories. Derivates of these layups are investigated by either changing the stack-
ing sequence or removing circumferential layers, whereas the labeling scheme for
these is indicated in d) and e). Next to that, changes in other variables, such as polar
openings or manufacturing parameters are separately marked at the corresponding
positions. The base layup for the investigations is the Layup A, which has been
designed and tested prior to this work. The main design consideration that was fol-
lowed to derive this layup was to achieve minimum CFRP mass, whilst ensuring
the required minimum burst pressure of 157:5 MPa.
3.2 Experimental characterization 33

To acquire a detailed structural knowledge, the Layup A is first experimentally


characterized in Chap. 4, where derivates are generated by removing circumferen-
tial layers or changing the stacking sequence. In Chapter 5, Layup A is considered
to exercise a correlation between experimental and numerical results through the
development of a FE model. Following, in Chap. 6, derivatives of this layup are
considered once again to investigate the influence of stacking sequence. The Layup
B is likewise derived from Layup A and remains closely similar to it. The main dif-
ference resides in the existence of additional mid- and high-angle helical layers in
the case of Layup B. The reason for this is to reduce the sensitivity of the tangen-
tial strain distribution at the circumferential ply drop location. The influence of the
circumferential ply drop location is exercised for Layup B in Sect. 4.2. The Layups
C and D exhibit a similar stacking sequence definition compared to Layup A, but
are composed of fewer layers and show a substantially different thickness build-up
at the dome. These two layups are of particular interest due to their failure type,
which is shown in Sect. 4.3. In Chap. 7, the Layups E and F are regarded on the
fullscale geometry to demonstrate the influence of the dome geometry and its pro-
found effect on the layup composition. Layup E is similar to Layup A composed
low-angle helical and circumferential layers. Contrarily, Layup F features a vari-
ety of mid- and high-angle helical layers, which is a resultant consideration to the
distinct dome contour of the fullscale geometry.

3.2 Experimental characterization

The experimental work is carried out within a uniquely designed test chamber
located within the facilities of Mercedes-Benz AG, Stuttgart, Germany. The cham-
ber allows for burst and cyclic testing as well as for user-defined pressurizations
of CPVs. Two measurement systems are integrated within the chamber to analyze
the vessel’s contour and its mechanical response during internal pressure loading
through a combination of optic and acoustic methods.

3.2.1 Test chamber

The chamber, which is depicted in Fig. 3.5 features a robust steel containment and
additional protection cabins located on three sides. The vessels are clamped inside
the containment through a locating-floating bearing arrangement, which allows
the specimen to expand in axial and radial direction. Through the hydraulic joint,
34 3 Material and methods

a) Overview

Containment
Damper
Impact adapter
Protection cabin
incl. stereometric systems
Floating bearing

Porthole with protection glass

Pressure vessel

Sound pressure sensor

Locating bearing

Hydraulic joint

b) Inside view c) Portholes with protection glass

d) Sound pressure sensor

Fig. 3.5 Test chamber for burst and cyclic experiments of pressure vessels
3.2 Experimental characterization 35

located beneath the locating bearing, the vessels are loaded with internal pres-
sure. Water is utilized as pressurant to reduce the energy release in case of burst
and is constantly supplied through a reservoir located outside the containment.
The loading rate and profile can be flexibly varied according to corresponding test
standards, but is ultimately constrained to a maximum pressure of 200 MPa. The
modular framework allows impact adapters to be individually inserted or replaced,
which permits the testing of differently sized vessel geometries.
The chamber features two implemented measurement systems. The optic ar-
rangement consists of nine stereometric systems, which are composed of two cam-
eras and one projector. Located in protection cabins outside of the containment,
every 120° of the chamber at a bottom, center and top position, a system monitors
the fixed vessel through portholes. The setup allows for operation of two distinct
measurement methods, namely stripelight projection and three-dimensional DIC.
While outer contour scans are obtained by using stripelight projection, the vessel’s
deformation throughout the pressurization is tracked by constant image acquisition
and the use of DIC. The acoustic system, composed of 120 sound pressure sensors
distributed over all six chamber walls, allows for the record and localization of
acoustic events during internal pressure loading. The sensors are permanently fixed
as their positioning has been predetermined to minimize aliasing effects. To inter-
link the obtained information to the loading state, sound pressure channel files and
recorded images are synchronized with the internal vessel pressure, that is being
measured by four independently pressure gauges.

3.2.2 Stripelight projection

The basic measurement principle relies on the subsequent projection of black/white


stripe patterns with distinct fineness onto the vessel’s surface and the simultane-
ous image acquisition through both cameras. As the pattern is differently distorted
by the three-dimensional shape, the vessel’s surface is recognized and can be re-
constructed by triangulation [88, 89]. The setup of stereometric systems allows
to obtain 360° point clouds of the clamped vessel by using stripelight projec-
tion. Therefore, the systems perform contour scans sequentially and the obtained
point clouds are then merged together to form a three-dimensional representation.
The cumulated scan time accounts for about 240 s with a minimum number of
2,600,000 points for the subscale geometry. The purpose of the obtained point
cloud is twofold. On the one hand, it is used to evaluate the thickness build-up by
referring the acquired outer contour to the nominal liner geometry, which creates
36 3 Material and methods

a thickness profile along the meridional path. Therefore, splines at four different
positions around the vessel’s circumference are extracted and lately averaged to
represent the vessel’s contour in the two-dimensional space. On the other hand,
by using Poisson surface reconstruction [90], the point cloud is converted into a
three-dimensional mesh, that is utilized as projection body for the localization of
AE. The acquistion and surface reconstruction is realized with Final Surface (gfai
tech GmbH, Berlin, Germany).

3.2.3 Digital image correlation

The general concept of DIC is rooted in the idea to measure changes between
subsequently acquired images of an object during load application. The objective
is to obtain quantative information about the motion, rotation and deformation of
pixel groups, so-called subsets or facets. In order to achieve a unique correspon-
dance of subsets throughout an image series, a non-periodic stochastic pattern is
applied onto the object surface. The pattern fineness is hereby defined by the cam-
era resolution and the measurement volume that is being considered. In the current
case, a speckle pattern is achieved through a white base coating and randomly
distributed black top paint. Simultaneous image acquisition between the systems
is ensured for frame rates between 1–10 Hz. Each camera provides a resolution
of 1624  1234 px2 , which in combination with the regarded measurement vol-
ume leads to a pixel size of 4 µm and an average measurement reference length of
10 mm. The subset size utilized is 19  19 px2 . The postprocessing is realized with
the software GOM Correlate Professional (GOM mbH, Braunschweig, Germany).

Postprocessing workflow
A semi-automated workflow has been developed and implemented within the
GOM Scripting module to establish a reproducible and time-efficient postprocess-
ing. Figure 3.6 shows an overview of the individual steps performed during the
postprocessing.
For each system, a single project is created where the obtained images of left
and right camera are inserted. Through images taken of a reference body, the rela-
tive position of each system to the measuring volume can be reconstructed and the
single projects are merged within a multi system project. A global coordinate sys-
tem is defined accordingly to the one introduced in Fig. 3.3 to enable comparisons
between different sets of vessels. Subsequently, to quantify the vessel’s defor-
mation behavior, a framework of deformation parameters is defined that will be
3.2 Experimental characterization 37

1. Creation of 2. Creation of 3. Generation of 4. Analysis of


single system projects multi system project component features deformation

L R

Predefined and semi-automated process using GOM Scripting (~15 mins)

Fig. 3.6 Overview of individual steps for the DIC postprocessing workflow

further detailed in Sect. 4.2.2. These parameters are calculated based on so-called
component features that represent geometrical objects, such as surfaces, points or
lines within GOM Correlate Professional. Thus, the magnitude of the calculated
deformation parameter is dependent on the alignment and the dimension of the cor-
responding component feature. For example, the calculated average cylinder strain
depends on the position and the dimension of the defined surface in the cylinder.
Hence, to ensure comparibility between different measurements, the generation of
component features is uniformly throughout by predefined dimensions and loca-
tions within the workflow. Finally, through the established framework the vessel’s
deformation behavior can be analyzed and quantified.

3.2.4 Airborne acoustic emission

The array of sound pressure sensors allows for the record of acoustic events
throughout the internal pressure loading. Contrarily to the commonly used AE
technique, which uses structure borne noise, this is achieved by means of air-
borne noise. As such, to indicate the difference, the term airborne acoustic
emission (AAE) is used within this thesis. During the pressurization occuring
AAE are continously monitored by the sound pressure sensor array, which allows
the record with three distinct sampling frequencies of 48, 96 and 192 kHz for each
individual sensor.
38 3 Material and methods

Table 3.1 Definition of Feature Definition Unit


AAE signal features used Amax Maximum amplitude [Pa]
within this thesis
N 2 Pa Number of hits above 2 Pa threshold [#]
pAAE;onset Internal pressure for onset of threshold [MPa]
crossing

Feature extraction
Signal features are defined in order to extract and isolate single information from
the obtained channel files. Table 3.1 summarizes the signal features used within
this thesis. A threshold of 2 Pa was chosen to prevent any influences resulting from
noise development of the hydraulic system.

Delay-and-sum beamforming
The distribution of sensors furthermore allows to localize detected AAE signals.
This is achieved by using a time domain delay-and-sum beamforming algorithm
implemented in the software NoiseImage (gfai tech GmbH, Berlin, Germany). The
general operating principle roots in the delays of an acoustic signal travelling to the
individual sound pressure sensors. As introduced by Doebler and Heilmann [91],
the reconstruction of the time function fO for a point x in the three-dimensional
space for an acoustic event can be described as follows:

1 X
M
fO.x; t / D wi fi .t  i / (3.1)
M i D1

Where t denotes time, M indicates the number of sound pressure sensors in the
array, wi are shading weights, fi are the recorded time functions of the individual
sound pressure sensors and i denotes the appropriate relative time delays. The
relative time delays are calculated from the different absolute run times i to the
sensors:

i D i  min. i / (3.2)

Where the absolute run times are determined by:

jri j
i D (3.3)
as

Where as is the speed of sound in air and jri j denotes the geometrical distance
from the sensor i to the point x. Finally, the effective sound pressure at the point x
3.3 Determination of material properties 39

can be calculated:
v
u n1
u1 X
pOeff .x/ D t fO2 .x; tk / (3.4)
n
kD0

where n indicates the total number of discrete samples considered for the effec-
tive value calculation, fO denotes the reconstructed time function of the sound
pressure at point x and tk is the time value at a discrete sample index k. The
three-dimensional mesh, obtained by means of stripelight projection, is utilized
as projection body for each vessel individually. The triangle length for the mesh is
chosen to be 1 mm to enable a sufficient local resolution. An average beamforming
time interval of 125 µs with respect to the maximum absolute amplitude is cho-
sen, which provided most consistent results. To this extent, further information is
provided in the Appendix A.2 in the Electronic Supplementary Material (ESM).

3.3 Determination of material properties

To accurately depict the vessel’s mechanical response, its material properties need
to be determined adequately. Hereby, the consideration of manufacturing-related
influences stands out as a necessary feat to realistically reproduce the mechanical
properties found within a CPV. Therefore, elastic modulus, final strength and frac-
ture toughness are determined based on filament wound flat coupons. To relate the
mechanical properties determined from coupon level to the final filament wound
CPV, laminate sections are investigated. Thereby, the determination of fiber vol-
ume fraction and porosity allows to access the material heterogeneities found in
dependence of the position through-the-thickness and along the meridional path.

3.3.1 Stiffness and strength

The mechanical properties of the composite reinforcement are determined through


the testing of filament wound flat coupons. The manufacturing of testing plates
is realized through plate tools compatible within the same robot-assisted winding
system used for CPV manufacturing, which enables to recreate process-induced
influences, such as the tensioned application of tows, adequately. A mold case is
applied after the filament winding, to ensure similar surface quality on both plate
sides, but also allows to adjust the layer consolidation depending on the applied
40 3 Material and methods

Fig. 3.7 Filament wound


flat plates; a) unidirectional
plate and b) bidrectional
plate (˙45°)

a) Unidirectional plate b) Bidirectional plate (± 45°)

compression. Figure 3.7 shows the two derived plate types, that ultimately repre-
sent the distinct layer architectures found in CPVs, being a) UD circumferential
layers and b) intertwined bidirectional (BD) helical layers. Especially the recre-
ation of intertwined helical pattern was deemed essential to accurately reproduce
both, mechanical properties and damage evolution process of helical layers. Tensile
properties in longitudinal and transverse direction are obtained through the test-
ing of UD coupons according to standard DIN EN ISO 527-4:1997 [92], whereas
shear characteristics are determined by testing the BD coupons according to stan-
dard DIN EN ISO 14129:1997 [93]. Strain measurement is achieved through strain
gauges in case of the UD specimens and optic extensometer for the BD specimens.
For the determination of average and scatter, eight samples per configuration are
considered. The manufacturing quality and consolidation level of each configura-
tion is assessed through the measurement of FVF and porosity by means of acid
digestion tests. The tests are carried out at the Institute for Aircraft Design, Uni-
versity of Stuttgart, Germany.

3.3.2 Fracture toughness

In the context of a CDM approach used to describe the damage evolution process
within the composite, its fracture toughness needs to be determined. Within this
thesis, fracture toughnesses in Mode I and II are obtained through double can-
tilever beam (DCB) and end notch flexure (ENF) tests according to the standards
ASTM D5528-13 [94] and ASTM D7905/D7905M-19e1 [95]. Accordingly to the
3.3 Determination of material properties 41

Fig. 3.8 Integration of a


PTFE insert during filament
winding for the manufac-
turing of ENF and DCB
specimens

specimens for the determination of stiffness and strength, the testing plate was
filament wound and a polytetrafluoroethylene (PTFE) insert with a thickness of
50 µm was integrated at the UD plate’s midplane during winding, which is shown
in Fig. 3.8. The manufactured specimens are first tested for the DCB test procedure,
where the delamination growth is constantly monitored through frequent image
acquisition. Afterwards, the specimens are trimmed and re-used for the ENF test
method. Hence, the delamination tip of the ENF specimens does not correspond
to the blunt end of the manufactured insert, but to the cracked end from the DCB
procedure. A limitation of the presented experimental characterization resides in
the fact, that only the fracture toughness of a UD composite is determined. While
this is representative for the circumferential layers, the intertwined helical layer
architecture inherently leads to differences in crack propagation. Nevertheless, no
current standards are available for the fracture toughness determination of inter-
twined composites, which thus remains as a matter of future work. The tests are
carried out at the Institute for Aircraft Design, University of Stuttgart, Germany.

3.3.3 Fiber volume fraction and porosity

To relate the determined mechanical properties on coupon level to the material


properties found within the final component, insights about magnitude and dis-
tribution of fiber volume fraction and porosity are of crucial importance. More-
over, as the filament winding process inevitably introduces material heterogeneities
through-the-thickness and along the meridional path, the quantification of these in
42 3 Material and methods

dependence of the position appears to be insightful, too. As such, laminate sections


are prepared out of CPVs and analyzed regarding their FVF and porosity by using
acid digestion tests as well as CT scans.

Determination of fiber volume fraction and porosity


The determination of FVF is conducted through acid digestion method according
to the standard DIN EN 2564:2019-08 [96]. Samples are taken from either entire
laminate sections or partially stacked layer groups. The sample density sample is
first determined through hydrostatic weighing of the sample in water and in air,
and the relationship described in Eq. 3.5:

msample;air
sample D  . water  air / C air (3.5)
msample;air  msample;water

Hereinafter, the composite is exposed to an acid environment, which leads to the


decomposition of the matrix constituent. The remaining fiber mass fiber and the
fibers’ density fiber are used to determine the FVF Vf by using Eq. 3.6:

sample
Vf D  fiber (3.6)
fiber

Correspondingly, the porosity Vp is calculated by using the matrix’ density matrix


and the relation described by Eq. 3.7:
 
sample sample
Vp D 100  fiber  C .100  fiber /  (3.7)
fiber matrix

For the determination of average and scatter, five samples per configuration are
considered. The sample weight is specified between 0.5 and 0:65 g to ensure re-
producible results. The measurements are carried out at the Institute for Aircraft
Design, University of Stuttgart, Germany.

Porosity through-the-thickness
Acid digestion is a well-known and efficient method to reliably determine FVF
and porosity within composites. However, insight into the distribution within the
sample is not acessible. For a thick-walled CPV laminate, a gradient in material
properties through-the-thickness is recognized, that is inevitably induced by the
filament winding process. To assess this gradient, cut-out laminate samples with
a cross section of 10  10 mm2 are analyzed by CT scans with a scan resolution
3.4 Structural analysis 43

Fig. 3.9 Porosity volume Volume [mm3]


for a CPV laminate section 0.72

0.58

0.43

0.29

0.14

0.00

of 20 µm. The porosity volume through-the-thickness is postprocessed by using


the porosity/inclusion analysis of VGStudio® (Virtual Graphics GmbH, Heidel-
berg, Germany). An example for the porosity volume of a CPV laminate section is
shown in Fig. 3.9. Average and scatter are determined by considering three sam-
ples per configuration. The scans are carried out at the Department of Materials
Test Engineering, TU Dortmund University, Germany.

3.4 Structural analysis

Within the context of this thesis, the structural analysis is composed of two dif-
ferent approaches, being an analytical description for thick-walled cylinders and a
high fidelity FE model. The choice of both individual analysis methods is reasoned
by their different application envisioned. Through the three-dimensional elastic-
ity approach the varying stress distribution throughout the vessel’s cylinder wall is
investigated. The suitability is reasoned by the consideration of normal stress gra-
dients through-the-thickness and the computational efficiency, which allows for
the analysis of multiple stacking sequences equally. Yet, the approach is limited
to the cylindrical section, which introduces an inevitable drawback if failure is not
preserved in the cylinder. Correspondingly, a solid FE model is developed to accu-
rately reproduce the vessel’s deformation behavior, the damage evolution process
and finally the loss of structural integrity.

Note: Parts of the content presented in this section have been previously published [82, 83]
or are derived from supervised student theses [84, 86, 97].
44 3 Material and methods

3.4.1 Three-dimensional elasticity theory

The analysis approach for multi-layered filament wound pipes described by Xia et
al. [42] is utilized and implemented for the analysis of the vessel’s cylinder. To per-
form stress and strain analysis, the approach uses three-dimensional stress-strain
and strain-displacement relations together with equilibrium equations formulated
in cylindrical coordinates. The main set of assumptions is provided in Eq. 3.8:

ur D ur .r/I u' D u' .r; z/I uz D uz .z/ (3.8)

where ur , u˚ , uz denote the displacements in radial, tangential and axial direction.


Through this it is possible to determine the radial and tangential displacements by
imposing a set of boundary conditions. These boundary conditions account for con-
tinuity of radial displacements and stresses in between layer interfaces. Moreover,
the radial pressure at the inner and outermost interface must match the internal and
external pressure. Finally, the cylinder is assumed to have closed ends and zero
torsion. The resulting strain and stress can be obtained from the radial and tangen-
tial cylinder displacements by using strain-displacement and strain-stress relations,
respectively. Further detailed information is provided by Xia et al. [42]. Following,
burst is predicted by considering FPF based on a maximum stress criterion in lon-
gitudinal direction, shown in Eq. 3.9:
11
D1 (3.9)
XT
where 11 represents the longitudinal component of the stress tensor and X T the
tensile strength of the composite material. Being one of the most fundamental fail-
ure criterion for composites, the drawback of this criterion is the non-interactivity
of the different stress components. Additionally, any degradation in the matrix con-
stituent is effectively neglected. Yet, from an engineering perspective this approach
renders a computationally efficient method to estimate CPV cylinder strength,
which is ultimately lost by exceeding the maximum strength in fiber direction.

3.4.2 Finite element analysis

The numerical analysis is approached by creating a three-dimensional solid FE


model for its implementation within the explicit module of Abaqus (Dassault Sys-
tèmes, Vélizy-Villacoublay, France). The vessel geometry definition is achieved
through a methodology on which its initial input the filament winding software
ComposicaD™ stands. The model features independent layer meshes, whose are
3.4 Structural analysis 45

derived from the individual ply discretization and permit to adequately reproduce
the load distribution along the meridional direction. Material properties are adopted
from the testing of flat filament wound coupons, as described in Sect. 3.3. For the
prediction of burst, a constitutively elastic solution as well as CDM approach are
considered.

Geometry definition
The geometry definition is achieved through the adjustment of the thickness profile
generated by the filament winding software ComposicaD™ . While the software
permits a time-efficient CPV pre-design and the derivation of machine motions,
the thickness prediction appears to be particularly imprecise in the cylinder-dome
transition and at the poles as it can be inferred from Fig. 4.3. Thus, the prediction
is revised and validated through the correlation to experimental data by means of
contour and CT scans of manufactured vessels. This is achieved by following the
methodology, shown in Fig. 3.10, which is composed of four main steps:

1. Circumferential ply drop readjustment

Full mandrel coverage at circumferential ply ends is enforced through the def-
inition of additional dwell angles. These ensure that in case the axial position
of the ply boundary is reached, regardless of the mandrel’s angular position, the
ply covers the entire circumference. This definition leads to a nominal increase in
thickness at the ply ends and a sudden decrease evolves afterwards. However, in
reality ply readjustments appear through the subsequent winding of plies which
induces friction forces onto the previously wound surface. In the particular case of
circumferential ply ends, subsequently wound low-angle helicals lead to the cre-
ation of meridional fricton forces that smooth out the circumferential ply ends. To
account for this effect, the excess material arising from the additional dwell defi-
nition is calculated based on the local thickness increase with respect to nominal
ply thickness and added in a new tapered ply drop end. To achieve this in a physi-
cally sound manner, the new ply drop end is defined by the exsiting taper solution
and extended by the excess material, while taking into account the nominal ply
thickness.

2. Inner contour verification and adjustment

To maintain the liner’s original shape during filament winding, force equilibrium
between the sustained internal liner pressure and the applied fiber tension is en-
visioned. Withal, in reality this unfolds as a rather complex tasks as the local
deformation of the liner is influenced by numerous variables such as tow tension,
46 3 Material and methods

1. Circumferential Tow- 2. Inner Contour 3. Layer Confinement


Drop Readjustment Verification and Adaption Correction
1.1. Identification of plies 2.1. Deviation profile 3.1. Correlation of stacking
of increased tow-drop definition sequence to layer group
thickness Boss Cylinder Helical layers

Deviation
Circumferential
layers
1.2. Calculation of material
excess Helical layers
Axial Coordinate
2.2. Offset thickness contour 3.2. Ply effective thickness
Cylinder definition
Offset
1.3. Addition of excess
material in a tappered ply Coordinate Mandrel
ss
drop end
Radial

Bo

Axial Coordinate

4. Boss Ply End Profile Export Contour End Extrapolation


Boundary Slope Match until Boss
Extrapolation Component
4.1. Fit of a parabolic
equation at the end of the
predicted dome profile

Fig. 3.10 Methodology for the geometrical description of the FE model

stacking sequence definiton, resin viscosity and the liner’s stiffness. Readjusting
the thickness profile based on the scanned outer contour roots in the assumption
that the inner laminate contour equals the nominal liner contour. Therefore, the
accordance is evaluated through the use of CT scans. To this purpose, a deviation
profile is created that considers the radial offset between nominal liner geometry
to final inner composite contour in dependence of the axial coordinate.

3. Layer confinement correction

The radial pressure applied by the internal liner pressure and tensioned tow appli-
cation leads to the consolidation of the uncured viscous material. The resin outflow
prompted by the material following the Terzaghi principle, can be described ac-
cording to Darcys law and therefore an estimation of the resulting FVF could
potentially be established. Yet, additional difficulties are introduced through the
influence of the curing process on the resin viscosity, hence additionally impact-
ing the outflow. Effectively, the consolidation leads to a ply thickness decrease and
a FVF increase depending on the position through-the-thickness, which needs to
3.4 Structural analysis 47

be accounted for in order to reallistically represent the stiffness and strength dis-
tribution. As such, the layer consolidation levels are correlated with CT scans of
laminate sections.

4. Boss ply end extrapolation

Lastly, the thickness build-up at the boss neck is corrected for. At the neck, where
the mandrel shapes changes from a convex to a concave slope, the filament winding
software experiences difficulties to reproduce a physically sound thickness profile.
This resides in the fact, that the software predictes the thickness normal to the seg-
ment’s surface, which in case of the investigated neck geometry inevitably causes
intersections in the thickness profile due to the change in slope. This limitation is
overcome by extrapolating parabolic ply ends from a sound vessel thickness pre-
diction location in the form of:

r.z/ D A0 C A1  z C A2  z 2 for zturn  z  zsound (3.10)

where r and z refer to the coordinates in radial and axial direction, respectively.
zturn denotes the axial coordinate at which the helical ply achieves its turnaround
and also represents the coordinate at which the extrapolated profile intersects with
the mandrel shape. zsound is an arbitrarily chosen axial coordinate at which a physi-
cally sound thickness from the initial prediction is ensured and serves as boundary
condition for the contour definition. A0 , A1 and A2 are coefficients which are de-
fined by the thickness and the slope continuity of the thickness profile. In the
presented case however, continuity in the curvature is not enforced and a shape
parameter is defined instead.
Resulting from this geometry adjustment methodology, a noticeably more real-
istic vessel geometry is achieved, which makes use of the initial thickness predic-
tion made by ComposicaD™ , experimental data and phenomenological knowledge
gained through the production of various vessels.

Puck criterion for longitudinal first-ply-failure


For the prediction of burst, the Puck failure criterion in the longitudinal direction
is considered. In contrast to the maximum stress criterion, it takes into account the
Poisson effect of the transverse stress components under combined loading. Failure
is predicted whenever the composite’s longitudinal strength is exceeded in either
compression or tension and can be expressed as:
   
1 E11
11 
21 
m
21;f f . 22 C 33 / D1 (3.11)
˙X T;C E11;f
48 3 Material and methods

where 11 refers to the longitudinal stress state and


21 denotes the Poisson effect
in the longitudinal direction because of transverse loading. X T and X C correspond
to the tensile and compression strength, respectively. E11 and E22 denote the longi-
tudinal and transverse stiffness, which allow to obtain
12 through the orthotropic
relation
12 =E22 D
21 =E11 . The subscripts f refer to variables applicable to the
fibers alone. 22 and 33 denote the transverse stress components in the in-plane
and out-of-plane directions. An arbitrarily factor m f is introduced to account for
differences in transverse stresses between fiber and matrix on a micromechanical
level [98]. In case of CFRP, this factor is proposed to be 1:1.

Intralaminar damage model CompDam


Besides the consideration of first-ply-failure, the influence of material degrada-
tion during internal pressure loading is examined through the implementation of
the CDM model CompDam developed by Leone et al. [99]. This assumes a linear
elastic orthotropic behavior before the onset of damage, which is then evaluated
independently in transverse and longitudinal direction. In the following, brief ex-
planations for the description of transverse and longitudinal tensile damage are
given. For further detailed information it is referred to Leone et al. [99].

Transverse damage
Transverse damage is considered through a smeared crack surface representation
by means of a cohesive law embedded within the continuum element. Displace-
ment jumps are calculated from a deformation gradient tensor decomposition F in
elastic FB and cracked components D [100]. The magnitude of each of these com-
ponents is determined from the configuration that minimizes the residual difference
between the cohesive surface stress vector and the bulk material stress projection
in that plane:
Res D   t  Rcr
T
(3.12)
where Res is the residual stress vector and denotes the stress vector in the co-
hesive interface. t  Rcr T is the bulk material stress tensor projection in the crack
surface basis coordinate system Rcr D ŒeOF eON eOT . The basis coordinate sys-
tem Rcr is formed in the cohesive surface by the orthogonal vectors in the fiber
eOF , crack-normal eOF and transverse directions eOT . This basis is determined at the
onset of damage from the angle that maximizes the result of the Benzeggagh and
Kenane cohesive failure criterion [101]. The transverse damage activation function
F is then defined from the crack displacement magnitude ı, the loading function
L.ı/ and the linear softening relation d2 as shown by Turon et al. [102]:

F .ı; d2 / D L.ı/  d2  0; (3.13)


3.4 Structural analysis 49

The cohesive surface stress vector  is composed of an elastic  coh and a friction
 f term. The first stems from the evaluation of the calculated crack opening vector
with arbitrarily chosen cohesive penalty stiffness. These need to be large enough
to ensure elastic displacement continuity while preventing numerical instabilities
because of ill-conditioned matrices [100]. The second term accounts for the fric-
tion that is being generated in fractured but contacting surfaces. By introducing
a Coulomb friction law within the cracked surface of the continuum it is possi-
ble to account for the tangential stress that is induced by the applied normal load
as described by Turon et al. [103]. During crack propagation, the effect of friction
effectively increases the stress at the cohesive interface. Once damage has fully de-
veloped, a purely Coulombian friction relationship is considered at the interface.
Due to this, it also increases the cohesive surface shear strength in dependence of
the surface normal loads [103]. Concerning the bulk stress tensor projection t, this
is calculated from the Cauchy stress tensor as t D B  eON . Accordingly, the bulk
material tensor is calculated from the bulk deformation gradient FB and the second
Piola-Kirchhoff stress tensor S :

B D FB SFBT jF j1 (3.14)

Longitudinal tensile damage


The growth of longitudinal tensile damage entails the existence of fiber-matrix
debonding and fiber pullout. The fracture toughnesses for these modes cannot be
easily calculated from standard procedures [104], what derives in the CompDam
implementation to consider longitudinal damage as the non-interactive constitutive
model of Maimí et al. [105, 106]. Damage is accounted through the longitudinal
damage variable d1 in the lamina compliance tensor:
2 1
12 3
 0
6 .1  d1 /E1 E1 7
6 7
6
12 1 7
H D6
6 E 0 77: (3.15)
6 1 E2 7
4 1 5
0 0
G12

The onset of damage is predicted by a non-interactive maximum strain criterion.


After the initation of failure, damage evolution is then represented by a bilinear
relation generated through the superposition of fiber-matrix debonding and fiber
pullout failure mechanisms. As shown by Dávila et al. [107], these values of frac-
ture toughness can be obtained from the resistance curve analysis of crack growth
50 3 Material and methods

perpendicular to the fiber direction. This permits to define the softening law pro-
portions ms and ns that characterize damage progression [107]. To prevent local
snap-backs in the stress-strain relations, the damage variable function needs to be
normalized with respect to the typical element length l  [108]. This leads to the
definition of a maximum element size:

2E11 G1C
l  (3.16)
XT 2

where E11 is the elastic modulus in the longitudinal direction, G1C refers to the
fiber tensile fracture toughness and X T denotes the longitudinal tensile strength.
In situ characterization methodology
4

As pointed out in the presented literature review in Chap. 2, the current approaches
followed for the experimental characterization of composite pressure vessels are
still limited to some extent. This in turn hampers to which degree numerical and
experimental analysis can be aligned. In consequence, uncertainties arise when-
ever discrepancies between prediction and experiment are existent. To overcome
these limitations, a characterization methodology is sought that permits to capture
the vessel’s mechanical response under internal pressure loading. In light of the
aforementioned, this chapter aims to answer the first research question.

Research question 1:
How can composite pressure vessels and their mechanical response under internal
pressure loading be characterized in a thorough and comprehensive way?

To this purpose, the presented methods in Chap. 3 are used to establish an in situ
characterization methodology, that adresses the following key issues:

 Evaluation of thickness build-up


 Characterization of deformation behavior
 Failure monitoring and localization

Supplementary Information The online version contains supplementary material avail-


able at https://doi.org/10.1007/978-3-658-35797-9_4.
Note: Parts of the results presented in this chapter have been previously published [81–
83, 109, 110] or are derived from supervised student theses [84, 85, 97, 111].

© The Editor(s) (if applicable) and The Author(s), under exclusive license to 51
Springer Fachmedien Wiesbaden GmbH, part of Springer Nature 2022
M. Nebe, In Situ Characterization Methodology for the Design and Analysis of
Composite Pressure Vessels, Werkstofftechnische Berichte j Reports of Materials
Science and Engineering, https://doi.org/10.1007/978-3-658-35797-9_4
52 4 In situ characterization methodology

4.1 Evaluation of thickness build-up


The prediction of thickness, particularly at the poles and in the cylinder-dome tran-
sition, remains a challenging but crucial task to accurately reproduce the vessel’s
mechanical response [18, 112, 113]. Hence, inprecise thickness predictions e.g. re-
ceived by commercially available filament winding softwares result in uncertainty
to which degree the vessel’s mechanical response is understood. Next to this, the
influence of the manufacuturing process and its parameters (e.g. fiber tension, resin
viscosity and curing cycle) lead to phenomena that further influence the fiber place-
ment and layer consolidation, which is hardly to account for within the predictions.
In consequence, quantative information about the final wall thickness distribution
is sought. To this extent, the following section introduces the derivation of experi-
mental thickness profiles through the juxtaposition of outer vessel contour obtained
by means of stripelight projection and nominal liner geometry.

4.1.1 Analysis of dome contour and manufacturing influence

The derivation of meridional thickness profiles enables the quantification of ma-


terial build-up in dependance of the chosen layup composition. This permits to
compare various layups for the same liner geometry and allows to identify at which
positions sudden changes in thickness are encountered. To this extent, Fig. 4.1

100
1 LH*
Normalized No. of Layers [-]
Radial Coordinate r [mm]

HH
80 C

60

Liner
40 Layup A
Layup B
Layup C
20 0
80 100 120 140 160 180 A B C
Axial Coordinate z [mm] Layup
a) Thickness profile b) Layup composition
*LH - Low-angle helical (5 ≤ α < 40°) | HH - High-angle helical (40 ≤ α < 88°) | C - Circumferential

Fig. 4.1 a) Experimentally obtained thickness profiles and b) overview of layup composition
for the Layups A, B and C
4.1 Evaluation of thickness build-up 53

shows the obtained thickness profiles for three differently composed layups, wound
with identical manufacturing settings.
The thickness build-up at the poles is a function of the number, the patterns
and the polar opening definitions of low-angle helicals. Next to that, the slope
at the cylinder-dome transition is determined by the amount of circumferential
layers, their tapering and the presence or absence of mid- to high-angle helical
layers. In the case of Layup C, a comparably low thickness build-up at the poles
is recognized, which is reasoned in the small number of low-angle helical layers.
Furthermore, the presence high-angle helical layers contributes to a gradual change
in wall thickness and thus results in a comparably milder slope from cylinder to
dome. Contrarily, Layup A exhibits no further high-angle helical layers that, to-
gether with the higher number of circumferential layers, contribute to an overall
steeper transition than recognized for the Layup C. Due to the overall higher num-
ber of layers, Layup B presents the largest thickness distribution in cylinder and
dome, where respective to Layup A, a noticeable difference is seen from 100 to
120 mm due to the presence of high-angle helical layers.
Besides the evaluation of differently composed layups, the influence of the man-
ufacturing process can be assessed by comparing the identical layup manufactured
with different parameters. Therefore, Fig. 4.2 illustrates an example of two vessels
with the identical layup, where solely the fiber tension of the circumferential layers
was varied. The increase in fiber tension for the Vessel FT-C2, not only leads to a

100
Radial Coordinate r [mm]

80

60

40 Liner
Layup A | FT-C1
Layup A | FT-C2
20
80 100 120 140 160 180 a) FT-
T C1 b) FT-
T C2
Axial Coordinate z [mm]
a) Thickness profile b) Detail view during winding
FT-C - Fiber tension of circumferential layers, where FT-C2 > FT-C1

Fig. 4.2 Influence of fiber tension variation on circumferential layer readjustment at


cylinder-dome transition
54 4 In situ characterization methodology

slightly higher compaction, which is recognized by the reduced wall thickness in


the cylinder, but also results in a readjustment of plies towards the cylinder-dome
transition. Due to the radial compaction applied by the subsequent winding of cir-
cumferential layers, the viscous tow material experiences transverse expansion due
to the Poisson effect. Thus, the stacking of multiple circumferential layers leads to
smearing out ply ends that move towards the domes. As such, mildly different
slopes at the transition are noticed for the Vessels FT-C1 and FT-C2, reasoned in
the different magnitude of applied fiber tension. Besides the changes in geometry,
the readjustments of plies also leads to changes in the local stiffness distribution,
which in turn is expected to impact the vessel’s mechanical response, too.

4.1.2 Correlation to thickness prediction

The thickness profile generation furthermore permits to correlate, validate and


scrutinize thickness predictions made by commercial filament winding software.
This in turn, allows to establish a much more realistic wall thickness distribution
as input for the finite element analysis (FEA) of CPVs. Figure 4.3 shows the com-
parison of predicted thickness profile derived from the filament winding software
ComposicaD™ and the outer contour scan for Layup A. Here, the comparison
reveals discrepancies particularly at the boss neck (1) and at the cylinder-dome
transition (2). The mismatch at the boss neck is anticipated in the fact, that the
filament winding software estimates wall thicknesses normal to the underlying sur-

100 2 100 2
Radial Coordinate r [mm]

80
80
120 140 160
60 60 1
1

40 Liner 40
Layup A | Prediction
Layup A | Experiment
20
80 100 120 140 160 180 80 100
Axial Coordinate z [mm]

Fig. 4.3 Comparison of thickness build-up prediction made by ComposicaD™ and experi-
mentally obtained thickness profile
4.1 Evaluation of thickness build-up 55

face. As the curvature of the liner-boss-assembly changes from convex to concave,


intersections are generated, which hence lead to an unrealistic thickness prediction.
The deviation at the cylinder-dome transition can be attributed to a redistribution
of the viscous towpreg material during winding. For every circumferential layer,
additional dwell angles are defined in order to ensure full mandrel coverage at the
end of each winding circuit. This leads to a local thickness increase at the ply
drop location. Nevertheless, this is smoothed out in reality through the subsequent
winding of helical layers, which create meridional friction forces on the underly-
ing geometry and therefore lead to a readjustment of the stacked circumferential
layers below. In addition to the aforementioned two main discrepancies, minor de-
viations are also recognized throughout cylinder and dome, which are anticipated
in the manufacturing influence that the filament winding process inevitably intro-
duces. Depending on the roving bandwidth and chosen layer pattern, the thickness
build-up varies. The sustained fiber tension applies radial pressure onto the under-
lying laminate, which achieves layer consolidation accompanied by a decreasing
porosity and wall thickness, and an increasing FVF, respectively. Depending on
the fiber tension, the stacking sequence definition and in general the overall choice
of angles, the consolidation level greatly varies, which is currently not accounted
for within the utilized software. As a result, discrepancies in wall thickness arise
between prediction and reality, thus justifying the need to capture and assess the
final filament wound contour.

Consideration of liner deformation


Yet, the thickness profile generation is rooted in a general assumption, which is
that the nominal liner geometry does not alter and remains the same throughout
the manufacturing process. This assumption holds whenever force equilibrium be-
tween internal pressure and fiber tension is achieved and the liner does not deform
plastically. As consequence, the inner laminate contour corresponds to the nominal
outer liner contour, which translates into a reasonable conformity of the approach
presented. Nonetheless, discrepancies may occur in case the liner contour deviates
from the nominal or if the filament winding process induces a plastic liner defor-
mation. For this reason, Fig. 4.4 shows the nominal and experimental liner contour,
the measured laminate contour together with a CT scan for the case of Layup B.
Overall, a reasonable agreement between nominal liner contour and inner lam-
inate contour is recognized with minor deviations in the dome. Between nominal
and experimental liner contour a constant offset is noticed, which accounts for
about 0:7 mm in the cylinder. The deviation is anticipated in a variety of influ-
ences, whose include manufacturing tolerances in liner production, the humidity
and temperature sensitivity of the PA6 material and not least the fact that the liner
56 4 In situ characterization methodology

100
Radial Coordinate r [mm]

80

Laminate
60
Liner
* unpressurized
40 Boss Liner | Nominal
Liner | Experiment*
Layup B | Experiment
20
80 100 120 140 160 180 200 220 240
Axial Coordinate z [mm]

Fig. 4.4 Comparison of experimentally obtained thickness profile, nominal and scanned
liner contour with a CT scan of a vessel after manufacturing

has been scanned unpressurized. Once mildly pressurized during manufacturing,


the difference is likely to diminish, which is confirmed by the good agreement of
inner laminate contour and nominal liner contour.
Notwithstanding, the equilibrium of internal and external forces on the liner
mandrel is not always ensured in reality. Excessive fiber tension may potentially
trigger liner shrinkage or denting in either radial or axial direction, which is im-
pacted by layer type, angle and its position through-the-thickness. Contrarily, too
high internal pressures potentially result in a bulging liner cylinder. In addition, the
likelihood of plastic liner deformation is sensitively impacted by the liner stiffness
itself and the chosen sequence of subsequently wound layers. Hereby, axial liner
shrinkage represents a particularly severe case, which is potrayed in Fig. 4.5. Here,
the deviation between nominal liner geometry to liner geometry after manufactur-
ing is depicted through the use of a CT scan.
The high lateral forces applied at the turnaround of low-angle helicals lead to
a denting of the dome towards the cylinder. At the tip of the boss, where stiff-
ness is rapidly decreasing, the liner deforms, which leads to a length decrease and
a diameter increase, respectively. The implications of this geometry change are
manifold. The reduced length increases the thickness build-up at the poles for sub-
sequent layers in comparison to the undeformed scenario. Next to that, the change
in geometry inevitably impacts the structural loading. When internally pressurized,
an increased cylinder diameter translates into higher tangential loads compared
to the undeformed shape. Moreover, the deformed dome shape leads to changing
4.2 Characterization of deformation behavior 57

Fig. 4.5 Example for liner


deformation resulting from 120
subsequently wound low- Laminate

Radial Coordinate r [mm]


angle helical layers with 100
excessive fiber tension
Force applied Bulging liner
80 by low-angle cylinder
helical layers
60 Liner

Boss
40
Liner | Nominal
20
100 120 140 160 180 200 220
Axial Coordinate z [mm]

tangential and meridional loading components of the internal pressure, that may
not longer match the intended local stiffness distribution of the laminate. Further-
more, the shape of the dome also dictates the severity of meridional bending at
the cylinder-dome transition, which is another circumstance to consider. Resulting
from this severe change in liner geometry, a realistic thickness profile cannot be
obtained from outer contour scans, which thus necessitates the additional use of
CT scans in order to quantify the thickness distribution. Yet, in this context more
importantly is the consideration of liner deformation within the virtual design pro-
cess, in order to ensure its stability in dependence of manufacturing influences
(e.g. fiber tension, internal liner pressure and stacking sequence definition). Re-
cent progress in this matter is demonstrated by Zhang et al. [50], where a buckling
analysis of the liner during winding is conducted.

4.2 Characterization of deformation behavior


Discrete strain and displacement measurements represent the current state-of-the-
art for the experimental validation of analysis approaches. In this sense, the ac-
quisition of cylinder strain and displacement is appealing, because it allows to
obtain general insights about the cylinder deformation along with comparably low
investment costs (e.g. ESGs, LVDTs). Yet, contrarily to the cylinder, the domes
represent areas of inherent material heterogeneity, which further translates into a
more heterogeneous response. Withal, the applicability of discrete measurements
to the double-curved surface is impeded by the sensitive sensor application and
58 4 In situ characterization methodology

the limited correlation due to the uncertainty in meridional positions. Nonetheless,


in light of the aforementioned constraints in the description of material thickness,
insights into the resulting strain distribution of the dome region are of vital impor-
tance. To this purpose, this section introduces the use of three-dimensional DIC for
the characterization of the vessel’s deformation behavior.

4.2.1 Full field strain analysis

The use of two- and three-dimensional DIC has been applied to CPVs prior to
this work, yet primarily limited to obtain cylinder strain information for a certain
angular distance of the circumference [56, 57, 60]. In this context, the assump-
tion of axissymmetry generally holds for CPVs. Notwidthstanding, the influence
of manufacturing-related defects, such as fiber slippage, tow overlaps, gaps or fiber
damage, can trigger the appearance of non-axissymmetric phenomenas. To this ex-
tent, Fig. 4.6 shows an example of such a phenomena, where circumferential layers
were locally squished by a subsequently wound helical layer with a -1 lagging pat-
tern. The layer angle together with the minor tow overlapping pattern, introduced
tangential friction forces onto the circumferential layers, which further lead to their
displacement and the local accumulation of wrinkled material. In consequence,
these defects may potentially trigger a preliminary vessel failure as they constitute
areas of local decreased mechanical strength and therefore falsify the final burst
pressure.
This in turn motivates the acquisition of strain data from different circumferen-
tial positions in order to capture these defects, understand their implications and

Fig. 4.6 Example for Local squishing of circumferential layers


a manufacturing defect
during filament winding:
squishing of circumferential
layers due to a subsequently
wound 1 lagging helical
pattern

-1 Lagging helical pattern


4.2 Characterization of deformation behavior 59

p = 0 MPa 70 MPa 105 MPa 125 MPa Burst (126.9 MPa)


[%]
1.50
Bottom system 3

1.00

0.50

0.00

[%]
1.50
Bottom system 2

1.00

0.50

0.00

[%]
Bottom system 1

1.50

1.00

0.50

0.00

Fig. 4.7 Meridional strain fields during a burst experiment shown from three different cir-
cumferential positions

interprete the vessel failure more thoroughly. To this purpose, Fig. 4.7 shows the
meridional strain fields for a vessel from three different circumferential positions
at four distinct pressure stages. Starting at a comparably low internal pressure of
70 MPa, two regions of high local strain are identified at the bottom cylinder-dome
transition. Consequently, both areas progressively grow during the internal pres-
sure loading, whilst the occuring region at the bottom system 1 appears to be more
severe in terms of size and strain magnitude. Ultimately, at an internal pressure
of 125 MPa, the strain magnitude of the area in the bottom system 1 accounts for
about 3%, followed by the sudden collapse of the vessel in the bottom dome. Un-
like the documented manufacturing defect, depicted in Fig. 4.7, the exact nature of
this failure cannot exactly be pinpointed to a noticed defect during filament wind-
ing. Yet, the failure is also anticipated to be in relation with the circumferential ply
drop location as position and magnitude of these local high strain areas indicate a
60 4 In situ characterization methodology

sudden change in local stiffness at the transition. In this regard, ply readjustment
due to the increased resin flow during curing poses a possible explanation, yet this
is hardly to trace back.

Correlation to strain gauges


To further confirm the validity of the strains measured, a reference measurement
is conducted to contrast strains obtained by means of DIC against strains derived
from electrical resistance strain gauges. The motivation for this experiment resides
in the different measuring principles of both methods that can further help to scru-
tinize each other. Withal, the comparison is subject to various influences, where
the reasons for deviation are manifold. While strain gauges render one of the most
established methods to determine surface strains, they are also subject to well-
known drawbacks. Among these are, the sensitivity of the application angle with
respect to the envisioned measurement direction, the vulnerability to non-planar
and porous surfaces and eventual preliminary failure due to overstraining and the
existence of damage [63, 114]. Likewise, strain measurements by means of DIC
are inevitably influenced by errors, whose most recognized are, the influence of
the image noise level, the focal length of the lense and the quality of the applied
speckle pattern with respect to the subset size [63, 115]. Additionally to the afore-
mentioned, the protection glasses inside the portholes of the chamber, shown in
Fig. 3.5, represent another influencing factor because of the distortion of the glass.
In this regard, a clear determination of the potential error sources is not envisioned
in this work. Instead, it is sought to correlate the obtained values and identify dis-
crepancies with regard to measured average and scatter. To this purpose, Fig. 4.8
shows an overview of the test setup.
At three different circumferential positions, two strain gauges are applied to ob-
tain meridional and tangential strain information around the vessel’s longitudinal
midpoint. The utilized strain gauges are of type L4 (Hottinger Baldwin Messtech-
nik GmbH, Darmstadt, Germany) with a nominal resistance of 120  and a mea-
suring grid length of 10 mm. The gauges are bond onto the surface by using the
cyanoacrylate adhesive Z70 and covered by the silicon rubber SG250. Prior to the
application, the surface is gently sandblasted to ensure a genuine bond over the en-
tire grid surface. After the application, a speckle pattern is applied to the surface,
which is depicted in Fig. 4.8 b). The strains obtained by means of strain gauges
are correlated against the strains derived from two differently sized surface com-
ponents, being a regular cylinder surface of 200  67 mm2 and a trimmed cylinder
surface of 20  35 mm2 , which solely covers the area of both strain gauges. The
choice of the different surface sizes resides in two distinct objectives envisioned.
On the one hand, by comparing strains obtained from the trimmed surface compo-
4.2 Characterization of deformation behavior 61

meridional 1. Application and


tangential sealing of strain
gauges on vessel
3 Surface components surface
200 x 67 mm²

3 Surface components 2. Application


20 x 35 mm² of speckle pattern

6 Strain gauges
Measuring grid: 10 mm

a) Overview of test setup b) Detail view strain gauges

Fig. 4.8 a) Overview of test setup with indicated location of strain gauges and surface com-
ponents and b) detailed view of applied strain gauges

nent (20  35 mm2 ) with the strains obtained from the strain gauges an indication
about the mangnitude and scatter of measured local strain is targeted. In this re-
gard a further trimmed surface size equal to the grid of the strain gauges was not
viable due to the used average measurement reference length of 10 mm. On the
other hand, both local strain information are correlated to the strains obtained from
the regular cylinder surfaces (200  67 mm2 ) to assess the sensitivity of surface
size to obtained average and scatter of cylinder strains. For the correlation, a pres-
surization up to 105 MPa was conducted. To begin with, the strains obtained by
the strain gauges are introduced and discussed, while subsequently the correlation
to both DIC surfaces is established. Figure 4.9 shows the strains in meridional and
tangential direction derived from each individual strain gauges.
In mild pressure stages up to 25 MPa, an almost linear increase in strain with
minor deviations between the different gauges can be recognized. Starting at an
internal pressure of 26 MPa first spikes are recognized within the recorded merid-
ional strains, whilst shortly after 70 MPa two tangentially aligned strain gauges
present unreasonably high strains, that emphasize a malfunction. In this context, a
relationship between the initiation and progression of IFF and the recorded strain
gauge signals is anticipated. For the tested Layup A, the onset of IFF takes place
from 25 to 30 MPa internal pressure, signified by the occurence of AAE and fur-
ther discussed in Sect. 4.3. In this case, a stack of circumferential layers fails due
to transverse tension, which is causing sudden displacements in the meridional di-
rection. Yet, as the stack is located in between helical layers, the generated fracture
62 4 In situ characterization methodology

Fig. 4.9 Meridional and 1.2


Tangential Strain
tangential cylinder strains
Meridional Strain
obtained by means of strain
gauges for a pressurization 0.9
up to 105 MPa internal

Strain ε [%]
pressure
0.6

0.3
p = 1 MPa s-1
T = 23 °C
0
0 30 60 90 120
Internal Pressure p [MPa]

is stopped at the layer boundary and thus does not detrementally harm the inter-
face adhesive nor the strain gauges’ measuring grid. Withal, IFF also initiates in
latter pressure stages on the outermost helical layer, which exhibits a low-angle
with respect to the cylindrical section. Furthermore, the surface is characterized by
a top resin layer, whose outflow was prompted during manufacturing through the
applied compaction by the cirumferential layers. Due to the high tangential loads
and the low matrix strength, fracture surfaces are generated on the outer surface
that align parallel to the vessel’s longitudinal axis, thus causing displacements in
the tangential direction. Contrarily to the prior, these fractures are in direct contact
with the adhesive, which eventually causes a sudden rupture of the measuring grid
and falsified strain values. Given the perpendicular alignment of the strain gauges
with respect to the longitudinal vessel axis, there is a higher probability of pre-
liminary failure compared to the ones, whose alignment is parallel to the vessel’s
longitudinal axis. For this reason, a correlation of measured strains is established
in Fig. 4.10 for an internal pressure at which apparent material degradation has not
taken place yet, whilst strains for latter pressure stages are further summarized in
the Appendix A.1 in the Electronic Supplementary Material. Figure 4.10 shows
the determined average strains with indicated error bars at an internal pressure of
25 MPa. A reasonable accordance in strain can be recognized with the maximum
deviation being between the meridional strain of trimmed and regular defined DIC
surface. Although the average strains of the trimmed surface are subject to a notice-
ably high scatter, which resides in the considered surface size with respect to the
used subset size of 19  19 px2 . The small dimensions of the surface lead to a low
number of subsets considered for the determination of average strain. Hence, lo-
4.2 Characterization of deformation behavior 63

Fig. 4.10 Comparison of 0.3


cylinder strains obtained by Meridional Strain Tangential Strain
means of strain gauges and
DIC at an internal pressure
of 25 MPa 0.2

Strain ε [%]
0.1

p = 25 MPa
T = 23 °C
0
Strain Gauge DIC Surface DIC Surface
(20 x 35 mm²) (200 x 67 mm²)

cal strain variations more substantially influence the measurement, that effectively
leads to a higher scatter obtained.
Contrarily, because of the larger quantity of subsets considered, the regular
cylinder surface strains show reasonably lower scatter. Compared to the strains ob-
tained by strain gauges, the regular DIC surface strains also assure similar average
strain magnitudes, which may be interpreted as a fact, that despite their large size
solely the cylinder expansion is captured and no influences of the cylinder-dome
transition arise. This in turn, implies a much more representative measurement of
average cylinder strain compared to the trimmed surface due to the larger data set
considered. Altogether, the deviations in obtained DIC and strain gauge strains ap-
pear to be reasonable, although this needs to be further confirmed by subsequent
measurements to prove its statistical significance. Nonetheless, the measurement
revealed certain insights about the application of both measuring methods to CPVs.
The initiation and progression of outer surface damage appeared to be critical for
the strain gauges, which in some cases led to preliminary failure. Thus, the choice
of strain gauge type, measuring grid and surface adhesive needs to be re-evaluated
for the application to CPVs. Contrarily, for the determination of average strains by
means of DIC, the relation of subset size to surface size didactes the vulnerability
to strain singularities. As such, larger surfaces are more favorable for the deter-
mination of average strains in homogeneous strain fields, due to the larger data
set considered, whilst the depiction of local effects necessitates smaller surfaces,
which comes in expense of a higher scatter.
64 4 In situ characterization methodology

4.2.2 Quantitative description of deformation

Besides the display of full field strain, the vessel’s deformation behavior needs to
be captured in a quantative manner in order to contrast different layups against each
other and furthermore enable correlations between predicted and experimental de-
formation. To this purpose, a framework of deformation parameters is established,
which is exemplary shown for the subscale vessel geometry in Fig. 4.11.
The definition of these parameters is based on component features created
within the software GOM Correlate Professional. To assure the replicability and
comparability of strain data, the creation of these component features has been au-
tomated through the postprocessing methodology described in Sect. 3.2.3. In total
three surface components are created in the cylinder and three surface components
are created in each dome. The components have a circumferential coverage of 40°
and an axial span width of 200 mm in case of the cylinder and 60 mm in case of the
domes. Moreover, through an average of 15 point components, the displacement
at the top boss is tracked. Ultimately, to consider strain variation depending on the
axial position, five section lines along the meridional surface path are generated
per side, separated by an angular distance of 5°.
To illustrate the applicability of the established framework, a comparison be-
tween two sets of vessels is introduced. To this purpose, the Layup B and its

Boss displacement
15 Point components y Section 2
(180°) 130-170°
Section 3
250-290°
Strain along meridional path x
z
15 Section lines
(90°)
Section 1
Cylinder strain 10-50°
3 Surface components
200 mm x 40°
Surface component
40°
Dome strain
z
6 Surface components
60 mm x 40°

Section lines
x y

Fig. 4.11 Framework of parameters used for the description of deformation behavior
4.2 Characterization of deformation behavior 65

Fig. 4.12 Circumferential


Circumferential
ply drop location for the
ply drop location
Layup B
Layup B

12 mm Ply drop
retraction

Circumferential Helical layers Circumferential layers


ply drop location
Layup B-T1 40 mm

derivative B-T1 are considered. The sole difference between both resides in their
circumferential ply drop locations, shown in Fig. 4.12, which in case of B-T1
were retracted by 12 mm compared to B. The choice for this comparison resides
in the triggering of very local phenomena, which is sought to be detected by the
established framework. The area around the cylinder-dome transition, where cir-
cumferential layers are subsequently dropped, represents a critical area for the
structural integrity of CPVs. By retracting the circumferential ply drop further into
the cylinder, it is anticipated that the tangential strain at the cylinder-dome transi-
tion will increase correspondingly to the introduced loss of tangential stiffness. In
the following, a selection of deformation parameters is further discussed and their
application exercised by comparing both layups.

Cylinder strain
The vessel’s cylindrical section represents a region with constant ply angles, ply
thicknesses and material properties along the meridional path. The axissymmet-
ric loading with internal pressure leads to a prevailing cylinder expansion with a
generally uniform strain distribution. Figure 4.13 shows the comparison of aver-
age tangential and meridional cylinder strains. No differences in tangential strain
exist between B and B-T1, which is expected given the fact, that tangential stiff-
ness was exclusively reduced at the cylinder-dome transition. Minor differences in
meridional strain are recognizeable, which are due to a different meridional bend-
ing response at the transition that seems to minorly affect the determined cylinder
surface.
66 4 In situ characterization methodology

Fig. 4.13 Comparison of 1.6


average cylinder strains Layup B
between the Layups B Layup B-T1
and B-T1 1.2 Tangential Strain
Meridional Strain

Strain ε [%]
0.8

0.4
p = 1 MPa s-1
T = 23 °C
0
0 30 60 90 120 150 180
Internal Pressure p [MPa]

Boss displacement
Contrarily to the cylinder strains, the axial boss displacement is influenced by the
deformation of the vessel’s entirety. Aside to the cylinder expansion, the strain dis-
tribution in dome and transition determine its overall magnitude. Furthermore, the
displacement is affected by the boss-composite interface, the damage introduced
within the composite and the yielding of the boss itself. In turn, the aggregation
of effects hampers to always identify the causes for the displacing magnitude. In
the case of the B/B-T1 comparison however, the differences seen in Fig. 4.14, are
attributed to the changes in tangential stiffness at the cylinder-dome transition. By
removing circumferential material at the transition, the tangential expansion in-
creases correspondingly. This leads to a decreasing axial displacement in the case
of B-T1.

Fig. 4.14 Comparison of 8


average axial boss displace- Layup B
Axial Displacement uz [mm]

ment between the Layups B Layup B-T1


and B-T1 6

2
p = 1 MPa s-1
T = 23 °C
0
0 30 60 90 120 150 180
Internal Pressure p [MPa]
4.2 Characterization of deformation behavior 67

Dome and cylinder-dome transition


The most insightful information is derived through the display of strains along the
meridional surface path. Here, the heterogeneous strain distribution in the dome, as
well as the response at the cylinder-dome transition can be assessed. To interlink
geometry with resulting strain magnitude, the obtained vessel thickness profiles
are aligned within the same coordinate system as the derived strains. Figure 4.15
shows the meridional and tangential strains as a function of the axial coordinate
together with the obtained thickness profiles for both layups. As anticipated, no
changes in meridional and tangential strains are noticeable in the cylinder. Towards
the transition a distinct behavior is noticed. In general, the individual expansion of
dome and cylinder leads to the existence of bending near the transition, due to
their rigid connection. Considering a vessel of isotropic material, the individual
expansion leads to an inward bending cylinder and a matching outward bending
dome. With regard to the outer surface meridional strain distribution, this means
that meridional strains increase at the cylinder end, while the matching slope leads
to a decrease in the dome. For both layups an inward bending cylinder is indicated,
by the increasing meridional strain at the cylinder end. Yet, the decreased tangential
stiffness not only leads to an increase in tangential strain at the transition, but also
shifts the position at which the cylinder is bending inward. This appears as an

Layup B Meridional Strain Contour


Layup B-T1 Tangential Strain SD
1.5
p = 105 MPa Nominal liner geometry B B-T1
100
Measured contours Radial Coordinate r [mm]

1
Strain ε [%]

0.5
50

Boss end Geometrical cylinder-dome transition


0
100 150 200
Axial Coordinate z [mm]

Fig. 4.15 Strains along the meridional surface path at an internal pressure of 105 MPa for
the Layups B and B-T1
68 4 In situ characterization methodology

aggregated effect resulting from the change in geometry, as well as the tangential
swelling at this position.
In conclusion, the use of three-dimensional DIC not only permits to derive full
field strain information, but also allows to quantitatively compare different sets of
layups. Through the established framework the vessel’s deformation can be ana-
lyzed on various scales, ranging from cylinder strains to axial displacement. Most
valuable is the display of strains along the meridional surface path, that in combina-
tion with the derived thickness profile, delivers comprehensive information about
the overall vessel deformation.

4.3 Failure monitoring and localization

Failure in composites progresses in stages, which resides in the different proper-


ties that fiber and matrix constituent exhibit. As such, the use of acoustic emission
analysis has been commonly applied to composite pressure vessels [73, 75, 76] in
order to detect, differentiate and localize microscopic damage prior to final fail-
ure. Withal, the sensitivity of the signal characteristics require sufficiently dense
sensor grids in order to ensure a meaningful interpretation of the results obtained.
Therefore, the high sensitivity comes at the expense of a time-intense instrumen-
tation and the potential loss of measurement equipment in case of burst. Likewise,
the integrated distribution of sound pressure sensors in the test chamber permits to
record acoustic emissions, yet by means of airborne sound. Therefore the term
AAE is being used within this thesis. Compared to the conventional AE tech-
nique, the sensitivity at which failure can be detected is reasonably lower as the
wave is required to alter from the structural medium to air and furthermore suffers
from its propagation in air. Yet, given the fixed and secured distribution of sensors
permits for data acquisition without time-intense instrumentation and furthermore
reduces the likelihood of equipment damage. Next to that, considering the delay
shifts of the acoustic wave traveling to the individual sensors allows to localize the
sources of radiation. In the following section, the use of AAE for the monitoring
and localization is demonstrated, whereas it is differentiated between the analysis
of interfiber failure and insights on the vessel failure.

4.3.1 Analysis of interfiber failure

The loading with internal pressure leads to a predominantly tensile stressed com-
posite, which causes interfiber failure to develop at lower pressure stages due to
4.3 Failure monitoring and localization 69

Internal Pressure p [MPa]


0 30 60 90 120 150
50
High amplitude events Burst
Sound Pressure ps [Pa]

25

-25
p = 1 MPa s-1
A T = 23°C
-50
0 25 50 75 100 125 150
Time t [s]

Fig. 4.16 Channel file for a burst experiment of Layup A without prior pressurization

the minor matrix strength. It is anticipated that IFF arises in the circumferential
layers because of transverse tension, whilst the helical layers additionally experi-
ence shear. In this context, Fig. 4.16 shows the channel file of a burst experiment
of Layup A, which has not been pressurized prior to the test. During the pressure
stages from 20 to 50 MPa numerous high amplitude events are recognized, whereas
final failure takes place at an internal pressure of 164:3 MPa.
Interfiber failure develops in both layer types, circumferential and helical layers
throughout the entire vessel. Yet, their distinct layer architectures lead to differ-
ences in crack propagation. The intertwined character of helical layers causes crack
growth to stop due to the continous change in fiber orientation and tow overlap.
Contrarily, fracture surfaces generated in circumferential layers are free to prop-
agate in fiber direction around the circumference. Moreover, within a stack of
circumferential layers, the cracks are also not confined to propagate through-the-
thickness until they are stopped by a boundary with different fiber orientation. For
the investigated Layup A, this creates large fracture surfaces to be generated due to
the large number of circumferential layers stacked together. To further emphasize
the difference in crack propagation between both layer types, Fig. 4.17 shows a
laminate section of Layup A after an end-of-line (EoL) pressurization to 105 MPa
internal pressure.
The relationship between fracture surface and released energy [63, 116] implies
substantial differences in the resulting amplitudes between both of the fractures,
due to their different size. As such, the fractures within the circumferential stack
are determined to generate the maximum amplitudes during internal pressures from
70 4 In situ characterization methodology

Outer helical layers 1 2


Interfiber failure 2
Interfiber
1 failure
Layer boundary
Void
Interfiber failure

Circumferential layers
Interfiber
Void failure

Inner helical layers


A p = 105 MPa 1 mm 2 mm 2 mm
a) Circumferential view b) Top view
Crack propagation through-the-thickness Crack propagation in-plane

Fig. 4.17 Interfiber failure in circumferential and outer helical layers after an internal pres-
sure of 105 MPa

20 to 50 MPa seen in Fig. 4.16. To further confirm this, two variables are investi-
gated, that are anticipated to directly impact the recorded amplitudes, namely the
fracture surface and its position through-the-thickness.

Fracture surface
To demonstrate the relationship between fracture surface and sound pressure am-
plitude, derivates of the Layup A are considered. Herein, solely the number of
layers in the circumferential stack was varied, whilst the stacking sequence and
the number of helical layers remained identical. Figure 4.18 shows the recorded
maximum sound pressure amplitudes for three layups with different circumferen-
tial stack size. The considered pressure range was reduced from 0 to 105 MPa to
exclude the interference of other mechanisms at latter pressure stages.
With decreasing circumferential stack size, the maximum sound pressure am-
plitude decreases accordingly. Yet, the relative amplitude decrease appears to be
larger than the relative decrease in stack size, which is reasoned in the twofold im-
pact that the layer reduction has. By removing layers, the laminate thickness and
the circumferential length are reduced, which hence leads to a decreasing crack
size in both, in-plane and through-the-thickness simultaneously. With regard to the
onset of failure, no apparent influence of the so-called in situ effect [117] was ex-
amined based on the pressure levels for first threshold crossings. To this extent,
Table 4.1 summarizes the internal pressure for first threshold crossing pAE;onset;2 Pa
and the number of detected hits over set threshold N2 Pa .
4.3 Failure monitoring and localization 71

Maximum SP-Amplitude Amax [Pa]


Fig. 4.18 Recorded 30 1.5
Maximum SP-Amplitude

Normalized No. of Layers [-]


maximum sound pres- No. of Circumferential Layers
sure amplitude Amax for 25
a pressurization from 0 to p = 0-105 MPa
p = 1 MPa s-1 1
105 MPa for the Layups A, 20
T = 23 °C
A-C1 and A-C2
15

10 0.5

0 0
n=7 n=7 n=1
A A-C1 A-C2

Position through-the-thickness
Besides the size of the fracture, its position through-the-thickness furthermore im-
pacts the recorded maximum amplitude as the signal suffers from propagation,
attenuation and dispersion. To this purpose, derivates of the Layup A were con-
sidered, in which solely the through-the-thickness position of the circumferential
stack was varied, while the total number of layers remained identical. Figure 4.19
shows the comparison of maximum sound pressure amplitude in dependence of
the circumferential stack position through-the-thickness, where 0 corresponds to
the position at the liner interface and 1 to the laminate outside.
Maximum SP-Amplitude Amax [Pa]

Position Through-The-Thickness [-]


Fig. 4.19 Recorded 140 1.5
Maximum SP-Amplitude
maximum sound pres- Circumferential Stack Position
sure amplitude Amax for 120
p = 0-105 MPa
a pressurization from 0 to
100 p = 3 MPa s-1 1
105 MPa for the Layups A, T = 23 °C
A-S* and A-S2 80

60
0.5
40

20

0 0
n=1 n=7 n=7
A-S* A A-S2
72 4 In situ characterization methodology

Table 4.1 Obtained acoustic features during the loading with internal pressure from 0 to
105 MPa
Layup n [#] pAAE;onset;2 Pa [MPa] Amax [Pa] N 2 Pa [#]
A 7 19.71 ˙ 1.87 23.21 ˙ 3.42 59 ˙ 5

A-C1 7 23.72 ˙ 1.11 11.54 ˙ 1.31 48 ˙ 5


A-S2 7 21.33 ˙ 2.93 4.65 ˙ 1.17 19 ˙ 7

A-C2 1 22.78 6.62 11


A-S* 1 18.01 96.81 72

With further placement of the circumferential stack towards the liner interface,
the maximum sound pressure amplitude decreases accordingly. Yet, the relative
large decrease from configuration A-S* to A is reasoned in another effect that
arises due to the specific stacking sequence of A-S*. As the circumferential stack
is exposed to a free edge condition, the overall energy release is higher as the
crack is not constrained by any boundary. This effect is also known as free bound-
ary effect and is described by Talreja and Veer Singh [118]. As such, quantitive
information about the relative amplitude decrease in dependence of the position
through-the-thickness can only be ascertained by the comparison of A to A-S2,
which nonetheless further confirms the trend. From the aforementioned, it can be
confirmed that a relationship between the size and the position of the circumferen-
tial layer stack and the resulting sound pressure amplitude exists. In fact, for the
investigated layups the maximum amplitudes for pressure stages until 105 MPa are
dictated by the stack size and its position. Herein, large groups of circumferential
layers towards the laminate outside result in generally higher sound pressure am-
plitudes. Contrarily, low amplitudes are associated to layups with circumferential
layers towards the liner interface or generally smaller sized circumferential stacks.
In this context, it is also recognized that for a representative number of vessels, the
layup-specific maximum amplitudes follow a reproducible pattern (e.g. A, A-C1,
A-S2). Besides, this trend can also be observed for the internal pressure of first
threshold crossing pAAE;onset;2 Pa and the number of detected emissions above set
threshold N 2 Pa , which are summarized in Table 4.1.
Withal, the previously established correlation was based on the first loading of
the CPV with internal pressure, which in the commercial application is generally
represented by the EoL pressurization. Hence, this constitutes the moment at which
4.3 Failure monitoring and localization 73

Outer helical layers IFF Void IFF

Circumferential layers IFF

IFF

Inner helical layers


A IFF 2 mm
a) After manufacturing b) After EoL pressurization c) After EoL and cyclic
(105 MPa) pressurization (6700 cycles)

Fig. 4.20 Circumferential view on the cylinder sections of three different vessels of Layup
A after a) manufacturing, b) EoL pressurization and c) EoL and cyclic pressurization (6700
cycles from 2 to 87:5 MPa internal pressure)

IFF is introduced within the composite on a mesoscopic scale and these fractures
are generated. After this, it is anticipated that further loading leads to friction be-
tween the fractured but contacting surfaces, existing fractures eventually grow and
new fractures are introduced within the composite, which leads to an overall in-
crease in matrix crack density and thus progressing material degradation, [77]. In
this context, Fig. 4.20 shows the circumferential view on three cylinder laminate
sections of different vessels of Layup A with distinct loading histories.
Compared to the laminate section after the EoL pressurization b), a moderately
higher crack density in the circumferential layer stack as well as in the innermost
and outermost helical layers is noticed for the laminate section, which has been fur-
ther cyclic loaded from 2 to 87:5 MPa for 6700 cycles, c). Given the fact that not
the identical vessel is compared, this trend was furthermore affirmed by observing
multiple additional samples, not presented herein. As such, the re-distribution of
load after local failure, appears to trigger new cracks to grow and thus leads to a
progressively developing damage. Nontheless, the gradual increase in crack den-
sity appears to not release comparable energy levels than those encountered within
the first loading cycle, which can be observed in Fig. 4.21.
Whilst numerous hits are recognized during the EoL pressurization, no further
acoustic activity above the set threshold is detected until the pressurizations were
aborted because of liner leakage. This pattern in acoustic activity appears to be in
accordance to findings reported by other authors [71, 77].
74 4 In situ characterization methodology

Fig. 4.21 Detected hits A-01 A-02 A-03 A-04


above set threshold of 2 Pa 50
during EoL and cyclic

Number of Hits N2Pa


pressurization from 2 to 40
87:5 MPa End of test due
30 liner leakage

20
p = 2-87.5 MPa
10 p = 10 MPa s-1
T = 23 °C
0
10 0 10 1 10 2 10 3 10 4
(EoL) Number of Cycles ncycles

4.3.2 Insights on vessel failure

Besides its monitoring, failure can furthermore be localized by considering the


signals’ delay shift to each individual sound pressure sensor. To this purpose,
an implemented time domain delay-and-sum beamforming algorithm within the
software NoiseImage is used and detected hits are localized on to previously es-
tablished three-dimensional vessel mesh. Withal, it is herein considered that the
identified source of radiation equals the position of crack initation. To this ex-
tent, obtained results are scrutinized by using in situ camera frames, CT scans
and microscope observations. In order to determine the localization accuracy of
the experimental setup, a reference measurement was conducted. A selection of
piezo signal generators was applied at different locations on the vessel’s outer sur-
face and were subsequently triggered. Consequently, the generator midpoints were
contrasted against the determined sources of radiation for which a maximum de-
viation of 5:83 mm was obtained. The measurement setup and results are further
detailed in the Appendix A.2 in the Electronic Supplementary Material. Withal,
the signal characteristics of the utilized piezo signal generators deviate from those
of real failure mechanisms, which thus impedes a direct transfer of these results for
the detection of failure in CPVs. Therefore, a selection of pressurizations was per-
formed and stopped prior to failure, in order to establish a correlation of localized
AAE to real failure mechanisms. Hereof, an example is given in Fig. 4.22, where
the locations of two identified events are contrasted against the positions of fiber
bundle breakages occuring in the outermost helical layer.
The causes for this particular type of failure are manifold and ultimately need to
be determined through the correlation to a detailed FE model. In general however,
it is understood that the area close to the cylinder-dome transition renders a critical
4.3 Failure monitoring and localization 75

Fig. 4.22 Comparison of emission localization on three-dimensional mesh to the positions


of fiber bundle breakages on the outermost helical layer

area as the geometrical configuration inevitably leads the existence of meridional


bending. This in turn, results in overall higher stress states for the helical layers on
both, innermost and outermost surface. Nonetheless, the established comparison
reveals a generally good accordance between detected position and location of
failure. This was furthermore confirmed for the localization of other mechanisms,
such as IFF in helical and circumferential layers in a likewise manner.

Failure analysis
Due to the numerous design variables, the type and the location of final CPV fail-
ure can be largely diverse and is furthermore impacted by the influence of the
manufacturing process as well as the evolution of damage throughout the internal
pressure loading. In general, two distinct types of CPV failure can be differenti-
ated, failure within or outside the cylinder. The latter can further be specified into
a failure that is triggered in the cylinder-dome transition or in the dome proper.
Depending on the structural design and the type of final failure, it is anticipated
that the sequence and characteristics of mechanisms prior to failure differ as well.
Therefore, Fig. 4.23 shows the detected hits for burst experiments of three different
layups, namely A, C and D.
76 4 In situ characterization methodology

100

Burst Layup D

Burst Layup C

Burst Layup A
Layup A (n=7)
Layup C (n=4)
Layup D (n=4)
Sound Pressure ps [Pa]

75
p = 3 MPa s-1 High amplitude
T = 23 °C events
Individual vessels
50
Interfiber failure Mid amplitude
within circumferential layer stack events
25
Low amplitude
events
0
0 250 500 750 1000 1250 1500 1750
Internal Pressure p [MPa]

Fig. 4.23 Detected hits above threshold during burst experiment for vessels of the Layups
A, C and D

For all three layups, a similar dense accumulation of hits is recognized within
low pressure stages from 20 to 75 MPa, designated to the development of interfiber
failure. This apparent similarity in acoustic behavior is reasoned in their compa-
rable stacking sequence, which exhibit a large circumferential stack towards the
laminate outside. Nonetheless, distinct behaviors are recognized for latter pressure
stages until final failure takes place. For the Layup A, little to non acoustic activity
is noticed until low amplitude events arise shortly prior to failure. Contrarily, for
the Layups C and D numerous mid to high amplitude events take place, starting
from 75 MPa to final failure. In this context, it is worthwile to interlink the acous-
tic activity to the obtained thickness profiles and the resulting failure types, whose
are depicted in Fig. 4.24. Discernible differences in the type of failure are noticed
between the Layup A, which shows a primarily cylindrical failure to the Layups C
and D, that display dome dominated failures.
Besides their different layup composition and total layer number, the most rec-
ognizable difference is encountered for the thickness build-up at the boss neck. In
this region, Layup A presents an increased thickness that appears to sufficiently
reinforce the boss neck, thus preserving a cylinder failure. Contrarily, the Layup
C and D show comparably low thicknesses that may prompt a preliminary fail-
ure. This is further affirmed by the localization of the emissions prior to failure,
Fig. 4.25 depicts the comparison of the last localized emission before final fail-
ure and the resulting failure of the Vessel D-01. The recognized sound pressure
4.3 Failure monitoring and localization 77

100
Radial Coordinate r [mm]

80

60

Liner
40 Layup A
Layup C
Layup D
20
80 100 120 140 160 180
Axial Coordinate z [mm] Layup A Layup C Layup D
a) Thickness profiles b) Failure pictures

Fig. 4.24 a) Experimentally obtained thickness profiles and b) failure pictures of the Layups
A, C and D

Fig. 4.25 Comparison of a) last localized emission and b) resulting failure for the Vessel
D-01

amplitude of 105:51 Pa seems relatable to those obtained by the identified fiber


bundle breakages, shown in Fig. 4.22. This tendency is furthermore preserved for
all vessels of this layup, implying that this effect appears to be reproducible. To
this extent, Table 4.2 shows a specimen-wise overview of the number of emissions
above threshold from after 70 MPa to final failure, their relative share of posi-
tion distinguished between dome and cylinder, and lastly the observed location of
failure.
78 4 In situ characterization methodology

Table 4.2 Detected AAE from 70 MPa to burst


Layup Vessel N 2 Pa N 2 Pa;cyl: N 2 Pa;domes Amax pburst Failure
No. [#] [#] [#] [Pa] [MPa] locationa
A 01 3 2 1 2.83 162.41 Cylinder
02 0 0 0 – 166.12 Cylinder
03 0 0 0 – 164.31 Cylinder
04 0 0 0 – 168.34 Cylinder
05 1 0 1 3.69 168.81 Cylinder
06 1 0 1 5.72 165.33 Cylinder
07 2 2 0 6.14 168.12 Cylinder
C 01 10 1 9 16.62 129.48 Dome
02 8 1 7 49.51 123.97 Dome
03 9 0 9 48.49 125.41 Dome
04 8 0 8 30.87 125.69 Dome
D 01 2 0 2 105.51 88.61 Dome
02 5 0 5 89.58 94.96 Dome
03 3 0 3 88.91 86.83 Dome
04 4 0 4 88.12 86.21 Dome
N 2 Pa;cyl: – number of emissions above threshold, located in the cylinder
N 2 Pa;domes – number of emissions above threshold, located in the domes
a
determined through visual observation

Likewise, the emission locations after 75 MPa for Layup C exhibit a similar
trend. In fact, the vast majority of detected emissions are located in the dome re-
gions, where accordingly the final failure takes place. Yet, in contrast to Layup D,
numerous emissions with lower amplitude events between 5 and 50 Pa are recog-
nized prior to failure. The differences in the acoustic behavior between both layups
are herein anticipated in their distinct dome layup. Besides a slightly larger thick-
ness build-up at the boss neck, Layup C furthermore features mid- to high-angle
helical layers that contribute to the structural integrity of the dome in the axial
positions from 90 to 130 mm seen in Fig. 4.24. The overall larger thickness and
correspondingly reinforcement appear to result in a more progressively increasing
damage evolution process, that exhibits numerous low amplitudes, contrarily to a
sudden failure event seen for the Layup D. Nonetheless, the various differences in
winding trajectories do not permit a clear identification of these relationships with-
out further detailed experimental work combined with a sound structural analysis
strategy, which is not further envisioned here.
Instead, a general trend between the acoustic behavior and the type of CPV
failure is aimed to be pointed out. In this sense, the cylindrical failure of Layup A
4.4 Concluding remarks 79

appears to be related with generally small number of emissions above set threshold
before burst. Considering the normal stress gradient through-the-thickness, failure
within the cylinder is triggered in the inner layers that exhibit high stiffness in the
tangential direction. In consequence, failure occurs whenever the stress exceeds
the longitudinal strength of the inner plies, which then triggers the development of
fiber breakage. Single fiber breakages render small fracture surfaces, that despite
their high initial stiffness, exhibit relatively low energy. As such, the progressively
increasing failure might not be captured as these emission fall below the current de-
tection threshold. Contrarily, failure triggered within the domes appears to coincide
with the continous presence of airborne acoustic emissions until final failure takes
place. Unlike in the cylinder, where the circumferential layers are closely aligned
to the governing tangential load, the helical layers in the dome are more prone to
the development of damage due to their alignment with respect to the applied load.
This leads to the development of interfiber failure at relatively low pressure stages,
which may continously grows in latter pressure stages and further exhibit airborne
acoustic emission prior to burst. Here, the validation of these hypotheses remains
as a matter of future work.

4.4 Concluding remarks

Research question 1:
How can composite pressure vessels and their mechanical response under internal
pressure loading be characterized in a thorough and comprehensive way?

The established in situ characterization methodology allows to capture the ves-


sel’s meridional thickness profile as well as its mechanical response under internal
pressure loading through a combination of optic and acoustic methods. The use
of stripelight projection permits to derive a three-dimensional model of the ves-
sel, that in combination with the nominal liner geometry, enables to generate an
experimental thickness profile.
Through the use of the multi camera system and three-dimensional DIC the
vessel’s deformation behavior can be accessed holistically from different circum-
ferential and axial positions. Local phenomena, arising from the filament winding
process, can be traced back through the display of full field strain. An established
framework of deformation parameters allows to quantify and compare the vessel’s
deformation. Together with the obtained thickness profile, the geometry and its
resulting deformation pattern can be analyzed along the entire meridional coordi-
80 4 In situ characterization methodology

nate, which provides valuable insights for design critical regions such as the dome
or cylinder-dome transition.
The record of AAE through the distributed sound pressure sensors and the local-
ization by means of delay-and-sum beamforming enable the analysis of the damage
evolution process during internal pressure loading. While the onset and progression
of IFF can be identified at mild pressure stages, the monitoring and localization of
events at latter pressure stages permits to further interprete the resulting vessel
failure.
FE modeling and correlation
5

The established in situ characterization methodology allows for a detailed descrip-


tion of the vessel’s mechanical response under internal pressure loading. Given the
complexity of multi-layered and thick-walled CPVs however, an explanation for
certain phenomena can not always be provided. To this purpose, FE simulations
emerge as a bridging tool that complements the structural understanding, where
experimental work is limited. Nonetheless, the representativeness of these models,
can only be acknowledged once a validation of the established assumptions was
realized. Therefore, this chapter aims to answer the second formulated research
question.

Research question 2:
How and to which extent can experimental and numerical analysis be aligned with
regard to the CPV’s mechanical response under internal pressure loading?

Figure 5.1 gives an overview of the chapter structure and the experimental data that
is derived through the previously established in situ characterization methodology.
This chapter can be separated into two main parts, the model definition and the
correlation of the mechanical response. The model definition entails the determi-
nation of the composite’s material properties and the creation of a representative
vessel geometry. To this purpose, the experimentally obtained thickness profile is
compared against the predicted thickness profile. Further, the numerical results on
the mechanical response are contrasted against experimental results on different
scales of complexity. Step by step, results are compared with regard to the defor-

Supplementary Information The online version contains supplementary material avail-


able at https://doi.org/10.1007/978-3-658-35797-9_5.
Note: Parts of the content presented in this chapter have been previously published [82, 83]
or are derived from supervised student theses [84–87, 97].

© The Editor(s) (if applicable) and The Author(s), under exclusive license to 81
Springer Fachmedien Wiesbaden GmbH, part of Springer Nature 2022
M. Nebe, In Situ Characterization Methodology for the Design and Analysis of
Composite Pressure Vessels, Werkstofftechnische Berichte j Reports of Materials
Science and Engineering, https://doi.org/10.1007/978-3-658-35797-9_5
82 5 FE modeling and correlation

Chapter 4 Chapter 5
In situ characterization methodology FE modeling and correlation
Section 5.1
datadata:

• Thickness profile Model definition


• Material characterization
Input

• Geometry definition
Input

surface
• Outer surfa
f ce strains/ displacements Section 5.2
Correlation of mechanical response
Airbo
r rne acoustic emission
• Airborne • Deformation behavior
• Burst pressure • Failure and damage progression

Fig. 5.1 Overview of chapter structure including used experimental data for the model defi-
nition and correlation of mechanical response

mation behavior, the onset and progression of damage and the final failure. The
correlation in this chapter is exercised for the case of Layup A, where three vessels
are investigated in detail.

5.1 Model definition


The model definition is divided into two subsections. First, the composite’s me-
chanical properties are determined through a testing campaign and measurements
on the final CPV laminate. Subsequently, a realistic vessel geometry is established
through the use of experimentally determined thickness profiles.

5.1.1 Material characterization

This subsection entails the determination of FVF and porosity found within a CPV,
a material testing campaign of filament wound coupons where longitudinal, trans-
verse and shear tensile properties are identified in dependence of the consolidation
level, and lastly the characterization of fracture toughnesses in Mode I and II.

Determination of fiber volume fraction and porosity


Filament winding inevitably leads to a gradient in the consolidation level of plies
through-the-thickness [8, 11]. This generally translates into high fiber volume frac-
tions and low porosities found within the innermost layers, whilst the opposite is
true for the outermost layers. Next to that, the applied radial pressure component
of the fiber tension is dependent on the layer angle of incidence. This means that
5.1 Model definition 83

70

65 64.16 63.94
60.68 60.91 59.86
FVF Vf [%]

60 57.48 58.92 57.55 57.99

55

50 Circumferential
Helical
45
90.5 190.5 290.5 390.5 490.5
Axial Coordinate z [mm]

3 2.19
1.87 1.97
1.6 1.55
Porosity Vp [%]

2 1.48 1.38
1.23
1
0.30
0
Circumferential
Helical

90.5 190.5 290.5 390.5 490.5


Axial Coordinate z [mm]
Laminate Section Helical Layer Group Circumferential Layer Group

Fig. 5.2 Measurement of fiber volume fraction and porosity by means of acid digestion tests
at different axial positions of Layup A

circumferential layers compact particularly well in the cylindrical section. Con-


trarily, low-angle helical layers apply comparably low compaction in the cylinder,
while they achieve the highest compaction at their turnaround. As such, CPVs are
subject to material heterogeneities along the axial coordinate as well as through-
the-thickness. To relate the mechanical properties obtained from the testing of flat
coupons, the consolidation level found within the final vessel needs to be de-
termined. To this purpose, the FVF and porosity of CPV laminate sections are
determined by means of acid digestion tests, as detailed in Sect. 3.3.3. Grouped se-
quences of helical and circumferential layers are separately accessed by grinding
the remainder of the laminate section. Figure 5.2 shows an overview of FVFs and
porosities at different axial locations.
84 5 FE modeling and correlation

18

Helical layers
Helical layers
Circumferential layers
A-01 A-03
15 A-02 2 mm
Porosity Vp [%]

A-03
12
9
6
CFRP
3 Void
0 1
0
0 0.2 0.4 0.6 0.8 1
Normalized Radial Coordinate r/ra [-]

Fig. 5.3 Porosity through-the-thickness obtained for three laminate sections of Layup A

Moderate differences in FVF and porosity are noticed between the investi-
gated positions. At the boss neck, where the low-angle helical layers achieve their
turnaround, FVF is determined to be highest while porosity tends be lowest. Be-
tween the results of cylinder and dome, no real trend can be identified based on the
conducted measurements as the scatter bars do overlap. The layer group sections
of helical and circumferential layers show noticeably higher FVF values than de-
termined in the laminate sections at similar axial positions. This resides in the
fact that the laminate sections generally feature a layer of accumulated excess
resin on the outer surface that lowers the effectively measured FVF. Therefore,
the average FVF found within the Layup A is expected to be similar to the values
obtained for the helical and circumferential layer groups, yet locally decreasing
towards the laminate outside. Differences in porosity are recognized between the
circumferential and helical layer group, which are supposed to be the result of their
different layer architecture and the existing gradient in consolidation level through-
the-thickness. To qualitatively access this gradient, Fig. 5.3 shows the distribution
of porosity through-the-thickness for three laminate sections obtained by means of
CT scans. To derive this, the postprocessing methodology described in Sect. 3.3.3
was followed.
The specific stacking sequence of Layup A with the circumferential layers posi-
tioned towards the laminate outside provides an effective compaction for the inner
helical layers, recognized in low porosities. Towards the outside a gradual increase
in porosity is noticed until at the layer boundary to the outermost helicals a major
peak in porosity is seen. This sudden increase is due to the accumulation of resin at
the end of the circumferential layer group. The compaction that the circumferen-
tial layers apply, prompts the outflow of resin through-the-thickness. Aligned in the
similar principal fiber direction, the resin is allowed to flow through the circumfer-
5.1 Model definition 85

ential stack until the next layer boundary where it gets trapped by the tow-overlap
and the deviating fiber orientation of a helical layer.

Determination of material stiffness and strength


Stiffness and strength properties of the towpreg material are derived through the
manufacturing and testing of flat filament wound coupons. For the derivation of
longitudinal and transverse tensile properties, unidirectional coupons are consid-
ered, whilst shear properties are obtained through the testing of intertwined bidi-
rectional coupons. The tests are performed according to the standards described in
Sect. 3.3.1. Three distinct consolidation levels per testing direction are considered
to investigate the influence of consolidation level on the mechanical properties. To
ensure the comparability of results, the material batch was kept the same through-
out the testing series and solely the compaction force onto the flat filament wound
plate was varied. Table 5.1 summarizes the three different consolidation levels per
testing direction, where the values for fiber volume fraction and porosity were de-
termined by means of acid digestion tests.
Fig. 5.4 illustrates the obtained stress-strain curves for the tests, where single
specimen curves are slightly shaded and the series average curve is indicated in
bold. The series average strength and strain values are represented by the marker,
accompanied by error bars that represent the standard deviation. Here, the dif-
ferent colors identify the different consolidation levels. A general trend that can
be observed for both fiber and matrix dominated composite properties, is that
higher consolidation levels generally yield higher stiffnesses and strengths. This
appears reasonable given the fact that the fiber volume fraction increases, which
predominantly influences the longitudinal properties and porosity decreases, which

Table 5.1 Overview of investigated configurations with different consolidation levels


Stacking Consolidation n Œ# Vf [%] Vp [%] sample
Sequence Level Œg cm3 
Œ09 High 5 62:24 ˙ 0:36 0:36 ˙ 0:04 1:57 ˙ 0:00
Medium 5 59:17 ˙ 0:70 1:55 ˙ 0:66 1:54 ˙ 0:01
Low 5 56:55 ˙ 0:50 7:46 ˙ 0:82 1:45 ˙ 0:01
Œ909 High 5 62:94 ˙ 0:57 1:07 ˙ 0:53 1:56 ˙ 0:00
Medium 5 60:02 ˙ 0:36 2:60 ˙ 0:43 1:53 ˙ 0:01
Low 5 56:77 ˙ 0:64 6:81 ˙ 0:31 1:46 ˙ 0:00
Œ˙455 High 5 62:90 ˙ 1:31 1:31 ˙ 0:72 1:56 ˙ 0:01
Medium 5 60:17 ˙ 3:07 3:65 ˙ 1:26 1:52 ˙ 0:03
Low 5 55:01 ˙ 5:23 5:65 ˙ 1:93 1:46 ˙ 0:04
86 5 FE modeling and correlation

a) Longitudinal 3000
stress-strain curve

Tensile Stress σ [MPa]


Low Consolidation 2250
Vf = 56.55 ± 0.50%
Vp= 7.46 ± 0.82%
Medium Consolidation 1500
Vf = 59.17 ± 0.57%
Vp= 1.55 ± 0.66%
750 Single Specimen
High Consolidation
Series Average
Vf = 62.24 ± 0.36% Average Failure
Vp= 0.36 ± 0.04%
0
0 0.4 0.8 1.2 1.6 2
Strain ε [%]
b) Transverse 50
stress-strain curve
Tensile Stress σ [MPa]

Low Consolidation 37.5


Vf = 56.67 ± 0.64%
Vp= 6.81 ± 0.31%
Medium Consolidation 25
Vf = 60.02 ± 0.36%
Vp= 2.60 ± 0.43%
12.5 Single Specimen
High Consolidation
Series Average
Vf = 62.94 ± 0.57% Average Failure
Vp= 1.07 ± 0.53%
0
0 0.12 0.24 0.36 0.48 0.6
Strain ε [%]

c) Shear 60
stress-strain curve

45
Shear Stress τ [MPa]

Low Consolidation
Vf = 55.01 ± 5.23%
Vp= 5.65 ± 1.93%
Medium Consolidation 30
Vf = 60.17 ± 3.07%
Vp= 3.65 ± 1.26%
High Consolidation 15 Single Specimen
Series Average
Vf = 62.90 ± 1.31%
Average Failure
Vp= 1.31 ± 0.72%
0
0 1.1 2.2 3.3 4.4 5.5
Shear Strain γ [%]

Fig. 5.4 Stress-strain curves for longitudinal, transverse and in-plane shear tensile tests for
the three different consolidation levels
5.1 Model definition 87

arguably effects the transverse and shear properties. Furthermore, from a testing
perspective the calculated stresses are a function of the specimen’s cross-section.
As the overall material amount remains the same, but the specimen’s cross-section
decreases, equal load levels lead to higher calculated stiffnesses and strengths for
the higher consolidated specimens. A detailed overview of maximum applied load,
specimen cross-section and derived stiffnesses and strengths is further summarized
within the Appendix B.1 in the Electronic Supplementary Material. Withal, certain
phenomena related to the strength properties are noticed, which are discussed in the
following. The results for the longitudinal tensile properties, depicted in Fig. 5.4
a), show a relatively higher strength increase for the series with high consolidation,
compared to the low and medium consolidation series. When loaded longitudinally,
fibers carry the load coherently by equating their longitudinal strain by means of
shear loading through the matrix. As the amount of voids increases, the effective
area of resin between the bundles decreases, compromising the load-carrying ca-
pabilities of the laminate and decreasing its stiffness. This may potentially lead to
rapid failure and lower material strength. In general, a low level of porosity and an
adequate amount of resin facilitate gradual failure. Contrarily, when the amount of
resin is not appropriate or the porosity level is too elevated, shear load transfer be-
tween the bundles becomes inefficient and the elastic energy released with single
fiber failure stays within the neighboring bundles, causing them to fail prematurely.
The influence of the consolidation level on the transverse properties is shown in
Fig. 5.4 b). Here, the porosity has a more significant impact on transverse tensile
stiffness and strength than in the longitudinal direction. In fact, the composite’s
transverse properties are dictated by the matrix. Increasing porosity reduces the
matrix cross-sectional area, which carries the transverse load. This not only leads
to a larger strain component, translating in premature failure when the material
limit strain is reached, but also to higher local stress states. Effectively, cracks in
the matrix initiate in areas of high stress concentrations in proximity of the material
imperfections, where the load path through the laminate is interrupted. Thus, it is
clear that high levels of porosity translate into premature failure in the composite’s
transverse direction.
The in-plane shear responses are shown in Fig. 5.4 c). Shear stresses are com-
puted as the 45° rotation of the axial stresses generated during tensile testing,
whereas shear strains are computed as the difference between axial and transverse
normal strains. As premises, it should be mentioned that the shear strength results
should be used carefully when obtained from tensile testing. As shown by Kellas
and Morton [119], the outer plies undergo significant in-plane transverse stress,
in addition to in-plane shear stress, due to their unconstrained location. This ef-
fectively lowers the actual strength of the material. Furthermore, the impact of
88 5 FE modeling and correlation

interlaminar free ply edge is neglected, although it is expected to have a substan-


tial effect on the shear strength results. The obtained stress-strain curves clearly
show a non-linear relation. This is explained by the matrix ductility during degra-
dation and by the scissoring of fibers with increasing axial deformation. The latter
is reported to be in the range of 1° per 3.5% shear strain by Kellas and Morton
[119]. Therefore, a shear strain limit of 5% is set to ensure that the assumed ˙ 45°
fiber orientation is valid. As for longitudinal and transverse case, the stiffness is
influenced by the amount of voids, which in turn influences the compliance of the
matrix and therefore the shear load transferring capabilities between the fiber bun-
dles. Furthermore, it is observed that for low porosity specimens, the stress-strain
curve flattens after the non-linear deviation. This is explained by the uniform load
distribution typical of the low porosity specimen, which results in longitudinal
fiber failure to occur only after matrix degradation. On the other hand, the slope
of the curve in the non-linear regime increases significantly for high porosity lev-
els. Here, delaminations initiate at relatively low loading stages on the outer plies
and further progress to the inner undamaged plies. This causes matrix failure to
develop throughout the laminate, and load redistribution to progress rapidly, until
the specimen finally fails.
Whilst the results of this material characterization provide necessary input data
for the FE model creation and deliver insights about general trends concerning
the relationship of consolidation level and composite properties, further detailed
investigations are necessary to gain a more comprehensive understanding of the
on-going mechanisms on micromechanical level. In this sense, the aforementioned
descriptions concerning the damage evolution and failure process need to be val-
idated through additional characterization methods such as AE analysis or CT
scans.

Determination of Mode I and II fracture toughness


The composite’s damage tolerance is accessed through the determination of Mode
I and Mode II fracture toughnesses by means of double cantilever beam and
end notch flexure tests. The specimens first undergo the DCB test procedure
and are used afterwards in the ENF test. Further details about the manufacturing
and the preparation of these specimens are summarized in Sect. 3.3.2. The load-
displacement curve derived from the DCB test is depicted in Fig. 5.5, where the
individual specimens are lighly shaded and the average curve is indicated in bold.
Two specific marks are emphasized. The first one, marked in red, indicates
the point at which the load-diplacement curve becomes non-linear and the sec-
ond one, marked in green, at which the delamination is visually detected on the
specimens’ edge. The corresponding fracture toughness GI and the corrected frac-
5.1 Model definition 89

Fig. 5.5 Load- 48


displacement curve of DCB Single specimen
test procedure for Mode I Average series
36
fracture toughness determi-

Load P [N]
nation
24

12 Delamination Onset
Delamination Propagation
Delamination Measurement
0
0 15 30 45 60
Displacement d [mm]

ture toughness GIc are then calculated by using the applied load P , the opening
displacement ı, the specimen width b and the delamination length a according to
standard ASTM D5528-13 [94], which is further detailed in the Appendix B.2
in the Electronic Supplementary Material. The results from this procedure are
presented in Table 5.2. Here, two sets are summarized. The first set labeled NL
considers the parameters corresponding to the point where the load-displacement
curve becomes non-linear. The second set labeled VIS utilizes the parameters col-
lected when the delamination is first visually observed on the specimen.
As shown by the respective standard deviations, the scatter in applied force P ,
delamination length a and opening displacement ı is quite noticeable, due to the
fluctuations at the crack tip. Furthermore, it is observed that identifying delami-
nation by determining the non-linearity in the curve leads to a more conservative
result than approaching the analysis visually. This is in line with the expectation
that the delamination grows faster in the inner region of the specimen, therefore
causing the curve to deviate from its linear path before it becomes observable on
the outer surfaces. For this reason, the correction applied on the fracture toughness
value GI effectively has a small impact on the results. The increase in load shown
in Fig. 5.5 underlines that the laminate can less efficiently resist delamination onset
than delamination propagation. In fact, the latter requires an increasing load which

Table 5.2 Results of the DCB tests for the determination of Mode I fracture toughness
Stack. Type n b h a0 ı P GI GIc
[#] [mm] [mm] [mm] [mm] [N] [N mm1 ] [N mm1 ]
Œ010 NL 6 24:99˙0:02 2:29 ˙ 0:08 50:16 ˙ 0:41 6:67 ˙ 0:40 29:33 ˙ 4:08 0:23 ˙ 0:04 0:21 ˙ 0:03
VIS 6 24:99˙0:02 2:29 ˙ 0:08 55:42 ˙ 0:89 9:84 ˙ 0:77 32:68 ˙ 4:02 0:35 ˙ 0:06 0:32 ˙ 0:06
90 5 FE modeling and correlation

Fig. 5.6 Load-displace- 600


ment curve of ENF test Single specimen
procedure for Mode II
fracture toughness deter- 450
mination

Load P [N]
Average series
300
CC20
150 CC40
Fracture Test
Max. Load
0
0 1.5 3 4.5 6
Displacement d [mm]

however reaches a maximum due to the increasing distance between the point of
load application and the delamination front, or in other words, due to the increasing
moment arm of the applied load.
Mode II fracture toughness is obtained by means of ENF tests of pre-cracked
specimens. In this case, compliance calibration (CC) is used for the calculation of
the fracture toughness GII . First, the compliance of the specimen is determined by
subsequently loading and unloading the laminate to reach delamination lengths of
20 and 40 mm, as well as full fracture. Then, the fracture toughness GII is computed
as follows:
2
3mPmax a02
GII D (5.1)
2b
where Pmax is the load carried by the specimen until fracture, m is the ratio between
the slope of the compliance measurements and the delamination length cubed, a0
is the original crack length and b is the specimen width. The value of GII is ac-
cepted only when it falls between 15% and 35% of the associated toughness value.
Therefore, the percent fracture toughness is computed as:
 
100.Pj aj /
%GII;j D where j D 20; 40 mm (5.2)
.Pmax a0 /2

This procedure requires an initial guess of the maximum load until fracture, which
is used to determine which result to deem as acceptable. This renders a limita-
tion of the method, as this estimation may alter the scatter of the results. The
load-displacement curves for the three cases are presented in Fig. 5.6, where the
individual specimens are lightly shaded and the average curve is indicated in bold.
5.1 Model definition 91

Table 5.3 Results of the ENF tests for the determination of Mode II fracture toughness
Spec. h P 20 P 40 Pmax m GII %GII;20 %GII ;40 Accepted?
[mm] [N] [N] [N] [N1 mm2 ] [N mm1 ] [%] [%]
1 2:15 500 250 367:36 1:08  107 0:79 82:3 82:3 No
2 2:33 350 150 423:56 8:01  108 0:77 30:3 22:2 Yes
3 2:36 350 150 402:76 8:67  108 0:76 33:5 24:6 Yes
4 2:35 359 150 401:88 8:24  108 0:72 33:7 24:7 Yes
5 2:31 350 150 422:00 8:23  108 0:79 30:5 24:4 Yes
6 2:21 350 150 371:68 1:05  107 0:78 39:4 28:9 Yes
Average 2:29 ˙ 0:08 – – 404:38 ˙ 18:74 .8:72 ˙ 0:90/ 0:76 ˙ 0:03 33:5 ˙ 3:26 24:6 ˙ 2:40 –
accepted  108

The results are presented in Table 5.3. It is observed that lower specimen thick-
ness leads to lower maximum fracture load. This is related the lower bending stiff-
ness associated to the reduced thickness, which effectively translates into higher
levels of compliance. Therefore, the obtained fracture toughness of the specimens
with the lowest thickness is unacceptable, as it lies above the threshold of 35%.

5.1.2 Geometry definition

Following the determination of material properties, the vessel geometry for the
investigated Layup A is defined. To this purpose, the geometry correction method-
ology described in Sect. 3.4.2 is used, at which initially the thickness estimation
of the filament winding software ComposicaD™ stands. As previously seen in
Fig. 4.3, discrepancies in estimated and experimental thickness arise particularly at
the boss neck and the cylinder-dome transition. These inprecisions are accounted
for by the readjustment of circumferential ply drop locations, the verfication of the
inner laminate contour, the consideration of layer consolidation and the interpola-
tion of ply ends at the boss neck. To achieve this, experimentally obtained thickness
profiles, CT scans and microscope sections are used to refine the geometry for the
simulation input. A comparison of initial thickness estimation, outer contour scan
and final simulation geometry is shown in Fig. 5.7.
For the analysis, a three-dimensional FE model is generated. To reproduce the
through-the-thickness gradient of in-plane stresses, a solid geometry model is con-
sidered. Symmetry conditions are applied to the vessel geometry to reduce the size
of the model to one eighth of the original CPV. The model is composed of the com-
posite reinforcement and the metallic boss. The liner is not considered due to its
low stiffness and strength. The geometry of the composite reinforcement is gener-
92 5 FE modeling and correlation

Radial Coordinate r [mm] 100

80

60

40 Liner
Thickness Estimation
Outer Contour Scan
Simulation Geometry
20
80 100 120 140 160 180 200 220 240
Axial Coordinate z [mm]

Fig. 5.7 Comparison between initial thickness estimation obtained from ComposicaD™ ,
scanned outer contour and adjusted simulation geometry for the Layup A

ated by revolving the corrected cross-sectional area presented in Fig. 5.7, where a
mesh composed of solid three-dimensional reduced integration elements (C3D8R)
is generated per ply. The meshed geometry is depicted in Fig. 5.8.
Reduced integration solid elements are used because of their ability to retrieve
complete stress tensors, that reproduce the coupled interaction in all principal el-
ement directions without compromising on computational tractability. This is of
particular interest considering the present tangential, meridional and radial stress
components found within the CPV’s wall. The single integration point per element
also facilitates the following implementation of the damage progression model,
where the degradation of stiffness is associated to each element integration point.
On the other hand, several limitations of this approach are worth noting. First,
the presence of several through-the-thickness elements makes the model suscep-
tible to develop zero energy hourglass deformation modes. This is particularly
prompted at interfaces at which sudden changes in stiffness are recognized (e.g.
helical and circumferential layer interface). Controlling hourglass deformation re-
quires the definition of arbitrary artificial stiffness factors and the consideration of
the influence of viscous damping during loading. Another drawback of using solid
elements is that relatively high aspect ratios in the thickness direction are needed
to reduce computational cost, which also challenges the numerical regularity of the
solution at times. Mass scaling factors are defined to counteract unmanageable low
stable increments resulting from relatively thin ply definitions. These scaling fac-
tors are assigned according to a variable description until stable increments in the
5.1 Model definition 93

Circumferential layer definition


Fiber orientation of approx. 90 deg

Helical layer definition


Constant fiber orientation in cylinder
Non-geodesic filament path in dome

Filament wound reinforcement


Transversely isotropic composite with structured mesh. Reduced-
integration solid elements (C3D8R). Intralaminar damage model.

Layer interface Resin gaps


Perfectly tied contact condition Homogeneous isotropic with structured mesh. Solid elements
(C3D6). Elastic constitutive behavior

Metallic boss
Homogeneous isotropic with free
mesh. Solid elements (C3D6). Elasto-
plastic constitutive behavior

Fig. 5.8 FE mesh with its main modeling characteristics

order of 1  107 s are reached. This corresponds to a percentage change in total


mass in the magnitude order of 1  103 . Because of the influence of the different
parameters, a sensitivity analysis of the numerical configuration is provided in the
Appendix B.3 in the Electronic Supplementary Material.
In reality helical layers present an intertwined [C˛/˛] structure of subplies,
which in the model is accounted for as configuration of [C˛/˛/˛/C˛] sub-
plies. This effectively eliminates the shear-extensional coupling response of the
ply, while keeping the bending-twisting interaction about two orders of magnitude
below the bending response. The varying fiber orientation in the dome is defined
for every element according to the filament winding path description generated
by ComposicaD™ . In the vicinity of the boss, where helical layers achieve their
turnaround, the software struggles to define a physically sound layup. Therefore,
the fiber orientations are extrapolated following a least-squares polynomial of sixth
degree on the initially predicted fiber orientations:

p.s/ D p0 C p1  s C    C pk  s k (5.3)
94 5 FE modeling and correlation

100
ComposicaD™
Polynomial Fit
Fiber Angle α [°]

75

50

25

0
40 90 140 190 240 290
Path Coordinate s [mm]

Fig. 5.9 Comparison between extrapolated fiber orientation and initial prediction made by
ComposicaD™ for the outermost helical layer of Layup A

where pk is the interpolation parameter and p.s/ is the approximate fiber ori-
entation of the kth, extracted at the arc length coordinate s. The result of this
extrapolation is exemplary shown in Fig. 5.9 for the outermost helical layer of
Layup A.
In areas where the blunt composite layers end, resin-rich gaps are detailed to
meet the geometry obtained from the geometry adjustment methodology presented
in Sect. 3.4.2. To this purpose, solid elements with reduced integration (C3D6R
and C3D8R) are used, where the properties presented in Table 5.4 are assumed
to be purely elastic, isotropic and homogeneous. The metallic boss is allowed a
free solid element mesh (C3D8R) with an elasto-plastic constitutive behavior of an
Al-6061-T6 aluminum. Table 5.4 summarizes the boss material properties, where

Table 5.4 Properties of metallic boss and resin constituent


Al-6061-T6 Boss
Elastic Properties Plasticity Properties
E D 68.90 GPa; G D 25.94 GPa;
D 0.33 y D 241.30 MPa; u D 290.00 MPa;
"p;u D 0.12
Misc.
D 2:70 g cm3 ; D 0.3

Resin Constituent
Elastic Properties Misc.
E D 2.83 GPa; G D 1.05 GPa;
D 0.35 D 1:18 g cm3 ; D 0.3
5.1 Model definition 95

the non-linear stress-strain curve is defined from the Military Handbook [120].
The load application is defined by a smooth amplitude curve definition built-in
in Abaqus until an internal pressure of 160 MPa is reached. However, numerical
difficulties arising from the damage progression implementation may lead to a
premature termination of the analysis.
The composite material properties for the model with homogeneous material
definition are summarized in Table 5.5. Due to limited experimental testing, sev-
eral material properties are assumed from sources of literature. The softening law
proportions m and n and the fiber fracture toughness G1C are assumed from a
IM7-8552 material system [107, 121]. These values are considered because of the
comparable longitudinal strength of the IM7-8552 material and the investigated
material system. The mode-mixity ratio  is also assumed from a IM7-8552 sys-
tem, whereas the out-of-plane Poisson’s ratio
23 is obtained from a T300-914
system [122]. Since the matrix properties of the investigated material noticeably
differ from those of the IM7-8552 and T300-914 combinations, the applicability
of these values is questionable. Nonetheless, the limited availability in literature
justifies their usage in the presented investigation. In future work, these assumed
values may be verified or adjusted through a comprehensive experimental testing
campaign.
Lastly, the influence of material heterogeneity is evaluated by accounting for
different material properties along the meridional path as well as through-the-
thickness. Therefore, the model is partitioned in five distinct meridional sections

Table 5.5 Composite properties defined in the FE model with homogeneous material defi-
nition
Composite Reinforcement
Elastic Properties Fracture Toughness Circumferential Layers
E11 D 135.65 GPa; E22 D 8.03 GPa; G1C D 133:30 N mm1c ; GIc D 0:21 N mm1 ;
G12 D 3.85 GPa;
12 D 0.34;
23 D 0.52a GIIc D 0:76 N mm1 ;  D 1.634a
Ply Strengths Fracture Toughness Helical Layers
X T D 2524.20 MPa; Y T D 31.47 MPa; G1C D 133:30 N mm1c ; GIc D 80:00 N mm1 ;
Y C D 225.80 MPa; S L D 49.56 MPa; GIIc D 160:00 N mm1 ;  D 1.634a
ms D 0.375b ; ns D 0.75b
Misc.
D 1:54 g cm3 ; D 0.3
a
assumed from Camanho et al. [122];
b
assumed from Dávila et al. [107];
c
assumed from Catalanotti et al. [121]
96 5 FE modeling and correlation

CT-Scan FE-model

4 3 2 1
5

a) Variation of helical layer properties b) Variation of circumferential layer


along the meridional path properties through-the-thickness

Fig. 5.10 Implementation of heterogeneous material definition for the FE model of Layup A

and the material properties are scaled according to the measured porosity val-
ues depicted in Fig. 5.2. The mechanical properties are interpolated accordingly
to the material characterization results shown in Fig. 5.4 and follow the relation-
ships that are further detailed in the Appendix B.1 in the Electronic Supplementary
Material. Material properties of the helical layers are varied along the meridional
path, whilst circumferential layer properties are varied depending on their position
through-the-thickness. This gradient is motivated by the observation of porosity
through-the-thickness in the CT scans, seen in Fig. 5.3. Yet, as the average poros-
ity values determined by means of CT scans are highly sensitive with respect to
the chosen greyvalue threshold, the variation through-the-thickness is scaled cor-
respondingly to the porosity values determined by means of acid digestion tests.
The partioned model is shown in Fig. 5.10, whereas the different material prop-
erties are summarized in Table 5.6. Resulting from the consideration of material
heterogeneity, particularly changes in the transverse properties are recognized. In
turn, it is anticipated that this will primarily affect the onset and progression of IFF,
leading to a more realistic depiction of the damage evolution process. Moreover,
the decreased consolidation levels of helical layers are expected to further promote
a higher compliance in the dome compared to the homogeneous solution.
Altogether, through the aforementioned steps, a realistic three-dimensional FE
model is established, where its elastic characteristics, ply strengths and fracture
toughnesses are obtained from the testing of flat filament wound coupons. In the
following, the results of three distinct model configurations are contrasted against
experimental data.
5.2 Correlation of mechanical response 97

Table 5.6 Composite properties defined in the FE model with heterogeneous material defi-
nition
Group k Ref. Vf Ref. Vp E11 XT E22 YT G12 SL
[%] [%] [GPa] [MPa] [GPa] [MPa] [GPa] [MPa]
Hel. 1 60.68 0.30 133.92 2477.85 8.96 47.08 4.08 51.78
2 59.50 1.00 131.71 2328.23 8.72 44.74 4.07 51.64
3 58.00 1.70 128.89 2138.04 8.49 42.39 4.05 51.50
4 58.67 1.49 130.15 2222.99 8.56 43.10 4.06 51.54
5 64.05 1.35 140.24 2905.14 8.61 43.57 4.06 51.53

Group Layer Ref. Vf Ref. Vp E11 XT E22 YT G12 SL


[%] [%] [GPa] [MPa] [GPa] [MPa] [GPa] [MPa]
Circ. 10–12 60.91 1.80 134.35 2507.01 8.46 42.00 4.03 51.44
13–15 60.91 2.00 134.35 2507.01 8.39 41.33 4.02 51.38
16–18 60.91 2.19 134.35 2507.01 8.32 40.75 4.02 51.25
19–21 60.41 2.60 133.41 2443.61 8.19 39.38 3.98 50.85
22–23 59.91 3.40 132.47 2380.22 7.92 36.64 3.79 48.13
24 59.91 8.00 132.47 2380.22 6.37 21.24 3.10 38.30

5.2 Correlation of mechanical response

In this section various numerical and experimental results are correlated with each
other. First, the deformation behavior is compared with regard to cylinder strains,
strain components along the meridional surface path and the vessel’s axial dis-
placement. Following, the predicted onset and progression of IFF is correlated with
detected AAE events. Finally, the predicted and experimental burst pressures are
compared. To this purpose, three different model variants are investigated:

 Constitutively elastic with homogeneous material definition


In the following labeled as ’Simulation | w/o damage | homo’
 Damage progression with homogeneous material definition
In the following labeled as ’Simulation | w/ damage | homo’
 Damage progression with heterogeneous material definition
In the following labeled as ’Simulation | w/ damage | hetero’
98 5 FE modeling and correlation

5.2.1 Deformation behavior

The vessel’s deformation behavior is analyzed by correlating the deformation pa-


rameters introduced in Sect. 4.2. Successively cylinder strains, strains along the
meridional surface path and the vessel’s axial displacement are compared and dis-
crepancies between the distinct models are pointed out.

Cylinder strains
The vessel’s cylindrical section features constant ply thickness, ply angle and
material properties along the meridional path, whilst the successive winding of
layers causes differences in ply thicknesses and material properties through-the-
thickness. The loading with internal pressure leads to a predominant cylinder
expansion that induces tensile tangential and meridional stresses as well as radial
compressive stresses. The corresponding strain magnitude shows to decrease to-
wards the outside in the radial and tangential directions because of the thick-wall
condition, whilst this turns to be constant in the meridional direction through-
the-thickness. Figure 5.11 shows the comparison of predicted and experimental
cylinder strain components. A tendency that can be observed is the nearly elastic
tangential expansion, which is closely reproduced by all three model variants.
Nonetheless, discrepancies between the results of the constitutively elastic pre-

1.8
Experiment
Simulation | w/o damage | homo
1.5 Simulation | w/ damage | homo*
Simulation | w/ damage | hetero*
Tangential Strain
1.2 Meridional Strain
Strain ε [%]

*values extrapolated after 120 MPa


0.9

0.6

0.3
p = 1 MPa s-1
T = 23 °C
0
0 30 60 90 120 150 180
Internal Pressure p [MPa]

Fig. 5.11 Correlation of predicted and experimental cylinder strain components for Layup A
5.2 Correlation of mechanical response 99

diction and the experiment are encountered in the meridional direction. This
appears to be related with the initiation of damage. Interfiber failure develops at
mild pressure stages in both layer types, circumferential and helical layers. This
has comparably minor implications on the tangential expansion as it is mostly
controlled by the fibers of the circumferential layers. In the meridional direction
however, the influence of the matrix constituent on the strain component is larger,
since the fiber orientations of the helical layers do not exactly match the direction
of the meridional loading component. Thus, developing interfiber failure has a
larger repercussion on the meridional strain component.
At an internal pressure of about 35 MPa an increasing compliance in the merid-
ional direction is noticed. This coincides with the onset of AAE activity, which is
attributed to developing IFF and will be further detailed in Sect. 5.2.2. Altogether,
a reasonable accordance in cylinder strain components is recognized, where the
larger discrepancy is found in the meridional strain component. Here, the consid-
eration of damage shows to more realistically reproduce the vessel’s deformation.
Between the homogeneous and heterogeneous material definition only marginal
differences in cylinder strains are noticed.

Strain distribution in dome and cylinder-dome transition


In contrast to the cylinder, the domes feature greater material heterogenities. The
deformation behavior is not only impacted by local changes in stiffness along the
meridional direction, but also by changes in the ply consolidation level and the
thickness build-up. When loaded with internal pressure, the cylinder-dome transi-
tion ensures displacement and strain continuity between the distinct responses of
dome and cylinder. This leads to the creation of meridional bending close to the
transition and is influenced by the specific geometrical configuration of cylinder
and dome as well as their respective deformation behavior [43].
To further comprehend the mechanical response at the cylinder-dome transi-
tion, a simplified FE model of a homogeneous orthotropic vessel with constant
wall thickness is considered, where its equivalent engineering constants are derived
from Table 5.5. These results are used to analyze the responses of the individual
components, cylinder and dome, to the loading with internal pressure. Based on
this, the insights are transferred to the deformation behavior of the Layup A under
investigation.
In Fig. 5.12 the meridional strain field of the orthotropic vessel is depicted at
an internal pressure of 160 MPa. Comparatetively, its mechanical response is pre-
sented for a) detached dome and cylinder and b) as full component. In Fig. 5.12
a), where the interaction face is assumed as simply supported, it is observed that
100 5 FE modeling and correlation

a) Detached dome and cylinder with a simple support at the face of their connection

Inward meridional bending


Outward meridional bending

b) Full homogeneous orthotropic vessel

Fig. 5.12 Deformed shape and meridional strain field of a homogeneous orthotropic vessel
at an internal pressure of 160 MPa; the deformed shape (colored) is shown alongside its
undeformed configuration (gray)

the dome mainly expands as a membrane without creating any relevant merid-
ional bending. Contrarily, the cylinder expands and rotates inwards to match the
radial displacement constraint at the boundary, which creates meridional bending.
This tendency is furthermore observed in the full component b), where the axial
component expansion induces a radial constraint at the transition, thus leading to
an inward bending cylinder. Correspondingly, this produces a meridional outward
bending dome that ensures the prescribed strain compatibility condition.
5.2 Correlation of mechanical response 101

Experiment
Simulation | w/o damage | homo Contour p = 105 MPa
Simulation | w/ damage | homo Meridional Strain p = 1 MPa s-1
Simulation | w/ damage | hetero Tangential Strain T = 23 °C
1.5
Nominal liner geometry
Measured contour 100

Radial Coordinate r [mm]


1
Strain ε [%]

0.5
50

Boss end Geometrical cylinder-dome transition

0
100 150 200
Axial Coordinate z [mm]

Fig. 5.13 Correlation of predicted and experimental strain components along the meridional
surface path at an internal pressure of 105 MPa for Layup A

For a CPV, the deformation behavior at the transition is additionally influenced


by a multitude of aspects. These entail ply drop locations of circumferential lay-
ers, the bending and extensional-bending coupling response of the laminate and
the dome thickness profile, which results from the choice of ply angles and polar
opening radii. Therefore, this renders a much more complex mechanical response.
Figure 5.13 shows the outer surface strain components along the meridional path
for the investigated layup at an internal pressure of 105 MPa. Accordingly to the
orthotropic case, a local maximum in meridional strain is noticed before the transi-
tion, that indicates an inward meridional bending cylinder. Contrarily, a minimum
in meridional strain is recognized, signifying an outward bending dome. In the
dome proper, a gradual increase in strains is recognized, which is anticipated to
the decreasing reinforcement as the circumferential plies gradually taper. At the
boss tip, the stiffness increases rapidly, as the load path is effectively altered by the
contact of the composite reinforcement with the aluminium boss.
Overall, all three model variants depict the general deformation behavior with
reasonable accuracy. Minor discrepancies in both tangential and meridional strains
are recognized at the cylinder-dome transition, which potentially reside in the
definition of circumferential ply drop locations as well as the helical ply angle def-
102 5 FE modeling and correlation

inition. As previously inferred, the consideration of damage particularly affects the


meridional strain component, which in turn also leads to a better representation of
the meridional bending at the transition. The consideration of heterogeneous ma-
terial properties appears to minorly affect the cylinder strains, while the increased
compliance in the dome is palpable. This leads to a slightly better accordance to
the experimentally measured strains.

Axial displacement
The vessel’s axial displacement is influenced by a variety of effects. These include
the cylinder deformation, the deformation of dome and cylinder-dome transition,
the progression of damage within the composite, the interface between boss and
composite and at last the yielding of the boss itself. The sum of these effects is
anticipated to yield into larger discrepancies between simulation and experiment.
Figure 5.14 shows the comparison of predicted and experimental axial displace-
ment.
For the experimental displacement, an initial non-linearity is recognized. This
appears to be related with the premature failure of the resin accumulated at the ply
tips around the boss neck at the early stages of pressurization, which physically
allows the boss to bridge the gap between the liner and composite reinforcement
and engage fully with the laminate. Once this happens, the curve stabilizes in a
quasi-linear behavior at low pressure stages. Nonetheless, the experimental dis-

9
Experiment
Simulation | w/o damage | homo
7.5 Simulation | w/ damage | homo*
Axial Displacement uz [mm]

Simulation | w/ damage | hetero*


*values after 120 MPa extrapolated
6

4.5

3
Initial non-linearity

1.5
p = 1 MPa s-1
T = 23 °C
0
0 30 60 90 120 150 180
Internal Pressure p [MPa]

Fig. 5.14 Correlation of predicted and experimental axial displacement for Layup A
5.2 Correlation of mechanical response 103

Table 5.7 Comparison of predicted and experimental strains and displacements at an inter-
nal pressure of 105 MPa for the Layup A
Experiment Simulation Simulation Simulation
w/o damage w/ damage w/ damage
homo homo hetero
"' ;cyl [%] 1:02 ˙ 0:01 1.00 (1.96%) 1.00 (1.96%) 1.00 (1.96%)
"s ;cyl [%] 0:77 ˙ 0:00 0.69 (9.61%) 0.75 (1.75%) 0.75 (1.75%)
"' ;dome [%] 0:66 ˙ 0:02 0.65 (2.01%) 0.65 (2.01%) 0.66 (0.00%)
"s ;dome [%] 0:69 ˙ 0:01 0.65 (5.34%) 0.68 (0.97%) 0.69 (0.00%)
uz [mm] 4:69 ˙ 0:12 3.47 (26.01%) 4.06 (13.43%) 4.22 (10.02%)

placement shows a noticeably higher compliance than the predictions. Particularly


the constitutively elastic model shows a substantially underestimated displacement.
By considering damage, the predicted axial displacement increases for pressure
stages after 30 MPa due to the developing IFF in the composite reinforcement.
However, still a noticeable discrepancy to the experimental displacement exists,
which does not decrease substantially, when heterogeneous material properties are
considered. The discrepancy is argued in the difference between predicted and real
damage that is existent within the composite reinforcement. The used CDM model
solely considers the development of intralaminar damage. In reality however, inter-
laminar stresses further may lead to the development and growth of delaminations
that hamper the load transfer between the plies and increase the vessel’s compliance.
Moreover, through the tied constraint between composite and boss defined in the FE
model, interfacial damage is neglected in the prediction, which further contributes
to an underestimation of the axial displacement.
A comparison of the individual deformation parameters at an internal pressure
of 105 MPa is provided in Table 5.7. Generally, tangential cylinder strains present
the lowest deviations, even without the consideration of damage. Meridional cylin-
der strains appear to be more affected by the development of interfiber failure
throughout the pressurization. To achieve a comparable accordance like in the tan-
gential direction, the consideration of damage is necessary. The axial displacement
stands out as a parameter of largest discrepancy, which resides in the multitude
of variables it is affected by. Here, the sole consideration of intralaminar damage
appears not to be sufficient to represent the damage process within the compos-
ite, thus leading to lower predicted displacements than experimentally measured.
Overall, it can be concluded that the heterogeneous solution with the consideration
of damage yields the highest accordance to the experimentally measured values.
Yet, the differences to its homogeneous counterpart are comparably minor.
104 5 FE modeling and correlation

5.2.2 Damage progression and final failure

After the correlation of the deformation behavior, the damage evolution process
and the final failure are compared between simulation and experiment. To this
purpose the predicted onset and progression of IFF is contrasted against the exper-
imental observations by means of AAE and in situ camera frames. Ultimately, the
predicted and experimental burst pressures are compared.

Interfiber failure
As previously seen in Sect. 4.3.1, cracks in the matrix constituent develop at al-
ready mild pressure stages during the pressurization. The high tensile in-plane
loads trigger a transverse failure in the circumferential layers, whereas interfiber
failure initiates in the helical layers due to the shear load they experience. The
evolution of damage is accessed through the CDM model CompDam. Figure 5.15
shows the transverse damage variable of the FE models with homogeneous and
heterogeneous composite material definition at three different pressure stages to-
gether with an acoustic channel file and microscope images after a pressurization
up to 105 MPa. Based on these results, an attempt is made to reconstruct the ma-
terial degradation process that takes place during the pressurization. Therefore,
additional experimental data such as in situ camera frames seen in Fig. 5.16 are
used and correlated against the numerical results.
In Fig. 5.15 b) Material degradation occurs first in the lower dome region at the
interface between the composite reinforcement and the aluminium boss, due to the
rigid constraints imposed by the contact. Here, the stress concentration is enhanced
by the high meridional and through-thickness shear stresses. Subsequently, dam-
age develops in the cylinder-dome transition region, where meridional bending is
experienced in proximity of the varying stiffness at the ply drop. At later stages,
transverse damage is triggered in the outer plies of the cylinder. This may initially
seem counterintuitive, as for thick laminates the longitudinal and radial stresses
progressively decrease towards the outside of the laminate. However, due to the
Poisson effect, radial compression reduces the in-plane tensile stress state. When
this is associated to low transverse stresses, the likelihood of interfiber failure in
the inner plies decreases. Contrarily, as the radial stress component is minimum in
the outer plies, this effectively decreases this counteracting effect and results in an
early predicted transverse failure. This damage propagation sequence appears to
be represented in both model variants.
Experimentally, the triggering of IFF is recognized by the recording of AAE ac-
tivity at low pressure stages up to 50 MPa. While the onset in Fig. 5.15 is noticed
at about 18 MPa, the vast majority of emissions above detection threshold take
place from 20 to 50 MPa signifying the development of IFF within the circum-
5.2 Correlation of mechanical response 105

a) Experimental channel file Internal Pressure p [MPa]


0 30 60 90 120 150
50
Sound Pressure ps [Pa]

A B C
25

-25
p = 1 MPa s-1
AA T = 23°C
-50
0 25 50 75 100 125 150
Time t [s]
b) Transverse damage variable
A 35 MPa Homogeneous Heterogeneous

B 50 MPa Homogeneous Heterogeneous

C 105 MPa Homogeneous Heterogeneous

c) Microscope images

p = 105 MPa 5 mm p = 105 MPa 5 mm

Fig. 5.15 Comparison of experimental channel file, damage variable in transverse direction
at 35, 50 and 105 MPa internal pressure and microscope images after a pressurization up to
105 MPa
106 5 FE modeling and correlation

a) 0 MPa b) 50 MPa c) 88.5 MPa d) 105 MPa

Fig. 5.16 Monitoring of outer surface damage on white-chalked vessel surface during the
loading up to 105 MPa internal pressure

ferential stack. This crack propagation pattern within the circumferential stack is
accurately reproduced in both models as the comparison to the microscope image
in Fig. 5.15 c) shows. Nonetheless its onset appears to be delayed in the simulation.
Whilst numerous emissions take place until an internal pressure of 35 MPa, both
model variants predict no IFF in the circumferential stack at 35 MPa. The reason
for this discrepancy is anticipated in the locally increased porosity within the resin
layer formed on top of the circumferential stack, seen in Fig. 5.3. The minimal
transverse strength of the porous resin layers triggers preliminary crack initiation,
that once developed, propagates through-the-thickness. This leads to a comparably
earlier failure in reality than predicted, even when reduced composite properties for
outer layers are consirered like in the heterogeneous model variant. Next to that,
the onset of IFF on the outer surface appears overestimated. In both model variants,
the outer surface damage is predicted for pressure stages from 35 to 50 MPa. Yet,
in the in situ camera frames in Fig. 5.16, rifts in the white-chalked vessel surface
are only seen at pressure stages after 50 MPa. The difference is expected in the
distinct depiction of helical layer architecture in simulation and reality. Whilst he-
lical layers are numerically depicted as a pair of two unidirectional subplies, they
represent in reality an intertwined layer structure. The constant tow overlap is an-
ticipated to further resist constant crack propagation and therefore delays the onset
and progression of interfiber failure compared to the numerical representation.
5.2 Correlation of mechanical response 107

Burst
With increasing pressure, discernible low amplitude AAEs are recognized in the
channel file in Fig. 5.15, signifying a further degradation within the composite
reinforcement until final failure takes place. The unambigious determination and
localization of these emissions prior to burst were not achieved in this work and
thus remain as a matter of future research. Certainly, a further detailed investigation
into the analysis of AAE will yield valuable insights into the progression of dam-
age that leads to the collapse of the component. The final failure of Layup A takes
place in the cylindrical section, which is emphasized by the three vessel remainders
depicted in Fig. 5.17. Due to numerical instabilities, it is not possible to reproduce
the vessel response at very high internal pressures for the model variants with im-
plemented damage progression subroutine. Therefore, the estimation of burst is
achieved by evaluating the corresponding longitudinal failure criterion with degra-
dated material properties. Table 5.8 summarizes predicted and experimental burst
pressures of Layup A. A very reasonable accordance in predicted and experimen-
tal burst pressures can be noticed, whereas minor differences between the three
model variants exist. Here, the constitutively elastic solution appears to slightly
overestimate the mechanical performance, whilst this gets further refined by the
consideration of damage and by accounting for heterogeneous material properties.

Fig. 5.17 Three vessel re-


mainders of the investigated
Layup A

Table 5.8 Comparison of predicted and experimental burst pressures for Layup A
Experiment Simulation Simulation Simulation
w/o damage w/ damage w/ damage
homo homo hetero
pburst [MPa] 166:2 ˙ 2:2 169.5 (C2.0%) 165.8 (0.2%) 165.2 (0.1%)
108 5 FE modeling and correlation

5.3 Concluding remarks

Research question 2:
How and to which extent can experimental and numerical analysis be aligned with
regard to the CPV’s mechanical response under internal pressure loading?

A high fidelity FE model was established through the recreation of the vessel’s
geometry by using CT and outer contour scans, the determination of mechanical
properties through a material characterization campaign and an adequate modeling
of the damage progression process in filament wound materials. The modeling of a
solid geometry with individual ply discretization not only permitted to account for
the stress gradient in the through-the-thickness direction, but also to reallistically
recreate ply group boundaries and ply drop locations. In total three model vari-
ants were inspected to assess considerations to the FE modeling. These entailed a
constitutively elastic solution with homogeneous material definition and the mod-
eling of damage progression with either homogeneous or heterogeneous material
definition.
The predicted deformation behavior was correlated to experimental cylinder
strains, strains along the meridional path and axial displacement. While cylinder
strains were found to be reproduced fairly accurate, the axial displacement showed
the largest discrepancy. The tangential cylinder strains showed a very good ac-
cordance by all three model variants (1.96%). In the meridional direction, the
influence of developing IFF on the strain magnitude was more noticeable. Whilst
for the constitutively elastic solution a certain discrepancy existed (9.61%), the
consideration of damage improved the accordance to the experimentally measured
strains (1.75%). The prediction of the axial displacement showed the largest
discrepancy for the constitutively elastic solution (-26.01%) and the lowest dis-
crepancy for the modeling of damage progression with heterogeneous material
definition (10.02%). The remaining discrepancy is reasoned in the aggregation
of multiple phenomena that influence the axial displacement. Amongst these, the
damage within the composite-boss interface and the development of interlaminar
damage were not accounted for by the present modeling approach.
The predicted onset and progression of IFF was contrasted against the occurence
of AAE. A good accordance was found in the recreation of the through-the-thickness
crack propagation pattern within the circumferential layer stack. Discrepancies be-
tween prediction and experiment were found concerning the onset and sequence of
events. Numerically, the onset of IFF within the circumferential layer stack was pre-
dicted around 50 MPa internal pressure, whilst numerous mid-amplitude AAE al-
ready took place between 20 to 35 MPa in the experiment. The reason for this dis-
crepancy is argued in the locally resin-rich areas that form at the layer interfaces
5.3 Concluding remarks 109

between helical and circumferential layers. These regions exhibit minor transverse
strength and trigger a premature development of IFF, that further propagates through
the remaining circumferential layers. Moreover, the recognized experimental onset
of IFF in the outer helical layers appears to take place at noticeably later pressure
stages than predicted. A possible explanation for this discrepancy is the distinct de-
piction of the helical layer architecture in simulation and reality. Whilst a helical
layer is formed by unidirectional subplies in the FE model, they represent an inter-
twined structure in reality. Hence, this difference is anticipated to influence the crack
initation and propagation pattern, which impacts the results.
The final failure by means of burst pressure was predicted by all three vari-
ants fairly accurate. While the constitutively elastic solution predicted a slightly
overestimated burst pressure (C2.0%), the results were further refined by the con-
sideration of damage (0.2%) and heterogeneous material properties (0.1%).
The reason for this generally good agreement is believed in the preserved cylin-
der failure. In all three cases, the predicted failure was determined by the innermost
circumferential layer, which exhibited the highest longitudinal stress. The resulting
vessel remainders further emphasized this failure mode. Due to the higher hetero-
geneities in the dome and cylinder-dome transition region, it is expected that a
failure in these locations will have a larger repercussion on the burst prediction of
the different model variants.
In sum, the exercised correlation of numerical and experimental results pro-
vided a very reasonable accordance on various degrees of detail. Whilst for the
most part a good or excellent agreement (e.g. for the prediction of burst) was
achieved, certain discrepancies in the deformation behavior and the damage se-
quence remained. Ultimately, these discrepancies can be argued in both, exper-
imental and numerical setup and require further investigations. The correlation
showed that an accurate depiction of the vessel’s geometry and material properties
are of vital importance for reproducing the mechanical response. The consideration
of damage progression and material heterogeneity further refined the model’s pre-
diction accuracy. Nonetheless, it should be mentioned that this modeling approach
required a variety of experimental data to reproduce the measured mechanical
response. The availability of this experimental data is seldomly the case in the
engineering reality. Thus, more refinements to the modeling approach need to be
applied to ensure a fast forward prediction, that requires less experimental data.
Moreover, the observed trends in this chapter hold true for the investigated Layup
A. Differences in the layup composition, the manufacturing procedure or the fail-
ure location may lead to other phenomena and conclusions. To this purpose, the
modeling approach is extended in Chap. 6, where it is projected onto various layups
and failure scenarios. This in turn, permits a more generic conclusion about its pre-
diction accuracy and the modeling considerations with regard to CPVs.
Influence of stacking sequence
6

The established framework of in situ characterization methodology and FE mod-


eling approach permits to investigate structural aspects with regard to the design
of composite pressure vessels in-depth. In this chapter the influence of the CPV
stacking sequence is studied thoroughly on the subscale geometry by taking into
account its effect on the laminate consolidation level, deformation behavior, dam-
age propagation and final failure. This is exercised by changing the sequence and
grouping of layers for a known layup, while remaining its initial composition of
plies. The purpose of this is twofold. On the one hand, further insights about the
mechanical response of CPVs are derived by considering a variety of stacking se-
quences. On the other hand, the established methods can further be scrutinized
regarding their validity through the application to various stacking sequences. The
formulated research question for this chapter is as follows.

Research question 3:
What influencing factors need to be taken into account with regard to the definition
of a CPV stacking sequence?

As previously inferred, two main types of layer can be identified for the wind-
ing of a CPV, being circumferential and helical layers. In a broad sense, helical
layers ensure the structural integrity of the domes, whilst circumferential layers
reinforce the cylinder. This also directly translates to the loads which they primar-
ily carry. While circumferential layers carry the tangential loads, the helical layers
mainly bear the axial loads. In this regard, the tangential stress gradient through-

Supplementary Information The online version contains supplementary material avail-


able at https://doi.org/10.1007/978-3-658-35797-9_6.
Note: Parts of the content presented in this chapter have been previously published [82, 83]
or are derived from supervised student theses [84–87, 97].

© The Editor(s) (if applicable) and The Author(s), under exclusive license to 111
Springer Fachmedien Wiesbaden GmbH, part of Springer Nature 2022
M. Nebe, In Situ Characterization Methodology for the Design and Analysis of
Composite Pressure Vessels, Werkstofftechnische Berichte j Reports of Materials
Science and Engineering, https://doi.org/10.1007/978-3-658-35797-9_6
112 6 Influence of stacking sequence

the-thickness, that arises from the thick-wall condition, poses a relevant structural
aspect to the choice of stacking sequence [42].
With regard to the manufacturing, the applied radial pressure of each wound
layer is not only a function of the fiber tension, but also depends on the angle of
incidence of application. Thus, circumferential layers provide substantially higher
compaction in the cylinder than for example low-angle helical layers. This in turn,
has direct implications on the fiber volume fraction and porosity found within the
CPV laminate. Higher compaction generally leads to higher fiber volume frac-
tions and lower porosities, which overall yields better mechanical performance and
decreases the likelihood of damage initiation. Therefore, changing the stacking se-
quence not only changes the consolidation level through-the-thickness, but also the
mechanical strength. Besides, the sequence of wound layers also dictates the num-
ber of necessary connector plies and the applied load onto the liner mandrel, which
needs to be accounted for from a manufacturing perspective.
For a layup composed of only two different layer types, being circumferential
and helical layers, one can consider two extreme cases. One, where two main
groups of layers are indentified and one, where the stacking sequence is divided
into several equally sized groups of circumferential and helical layers. The former
may allows to achieve higher compaction if the circumferential stack is placed
on the laminate outside. Yet, the grouping also introduces a single interface,
where a large stress gradient between circumferential and helical layer exists. Fur-
thermore, the grouping of circumferential layers leads to the unobstructed crack
growth through-the-thickness, which has been observed in the case of Layup A in
Fig. 4.17. Contrarily, frequently alternating layer groups appears to mitigate crack
growth through-the-thickness as the fracture is stopped at every layer boundary
with an alternating fiber orientation.
The following chapter outlines the results of a comprehensive experimental
set aimed at investigating the impact of stacking sequence on several aspects,
including the final mechanical performance. Furthermore, by applying the prior
developed FE model to various sequences, further insights into the mechanical re-
sponse of different stacking sequences is sought.

6.1 Experimental design

Based on the well-known Layup A where an extensive database in terms of burst


pressure p burst D 166:19 MPa (SD D 2:19 MPa, n D 7) is established, the influence
of stacking sequence on CPVs is investigated. Besides the large database, another
reason for the choice of this layup resided in the preserved cylinder failure seen
6.1 Experimental design 113

Positioning

3 A-S1
A1 1 A-S3 3 A 3 A-S4 3 A-S2

Grouping

5° ≤ α < 40° Helical layer Circumferential layer Number of vessels

Fig. 6.1 Experimental design for the investigation of stacking sequence influence

in Fig. 5.17. This is initially considered to prevent any influences triggered from
the dome to cause preliminary failure. Therefore, the layer sequence and the layer
grouping of this layup is altered, while remaining the total number of layers and
their orientations the same. Particular care was given in the design of the experi-
mental set to ensure that the considered configurations would represent a variation
in both, the positioning and grouping of layers. Figure 6.1 shows an overview of
the experimental design. Thereby, the resultant Sequences1 A, A-S1 and A-S2 rep-
resent the most distinct configurations, that are primarily regarded.
To ensure the comparability of the results, structural parameters such as pole
openings, tapering length of circumferential layers and manufacturing parameters
such as fiber tension, curing cycle and internal pressure were kept the same. The in-
fluence of a changing helical winding pattern, because of their different positioning
through-the-thickness, was considered small and is therefore neglected. To prevent
any influences from material heterogeneity, all specimens were manufactured with
the same material batch.

6.1.1 Stress gradient through-the-thickness

The configurations in this experimental set are first analyzed using the three-
dimensional elasticity approach described by Xia et al. [42] in order to assess the
through-the-thickness gradient in the governing in-plane stresses in the cylindrical
section of the vessel. Based on the premise that the polar openings and the helical

1
In this chapter the term sequence is used instead of layup to refer to different derivates of
the same set of angles.
114 6 Influence of stacking sequence

Fig. 6.2 Distribution of A A-S1 A-S2 A-S3 A-S4


longitudinal stress through-
the-thickness at an internal

Longitudinal Stress σ11 [MPa]


2400 CFRP
pressure of 160 MPa for
the investigated stacking 0 1
2200
sequence configurations
2000

1800

1600

1400
0 0.2 0.4 0.6 0.8 1
Normalized Radial Coordinate r/ra [-]

layer angles remained the same for all sequences, an influence of the dome or
cylinder-dome transition is initially neglected. For the prediction, the material
properties summarized in Table 5.5 are used. Figure 6.2 shows an overview of the
configurations and their respective distribution of longitudinal stress through-the-
thickness at an internal pressure of 160 MPa. Worth noting in this context is that
the used approach does not consider the different levels of consolidation that arise
through the different stacking.
Within the Fig. 6.2, the sections of high longitudinal stresses represent the cir-
cumferential layers that, due to their high tangential stiffness, exhibit considerably
higher longitudinal stresses than the low-angle helical layers. The most noticeable
differences are recognized between Sequences A, A-S1 and A-S2. The alternation
of layer groups in the Sequence A-S1 leads to a fragmentation in the number of
interfaces between circumferential and helical layers. This appears to increase the
longitudinal stresses within the innermost group of circumferential layers, where
due to the geometrical setup, the tangential loads are highest. Contrarily, the group-
ing of circumferential layers promotes a decrease in longitudinal stress within the
innermost circumferential layer. Here, the influence of the different positioning of
the circumferential stack between Sequence A and A-S2 appears minor. Likewise
to the distinct distribution of stresses through-the-thickness, the sequences show
a different distribution of strains through-the-thickness, which also influences the
general cylinder deformation. In this regard, placing the load carrying circumfer-
ential layers towards the outside, like in Sequence A, inevitably leads to a larger
cylinder expansion as the load needs to be transferred from the inner helical layers
6.1 Experimental design 115

Table 6.1 Overview of predicted outer surface strains at 105 MPa internal pressure and final
burst pressure according to three-dimensional elasticity theory [42]
A A-S1 A-S2 A-S3 A-S4

"s ;cyl;3d [%] 1.00 0.96 0.92 0.98 0.95


"' ;cyl;3d [%] 0.65 0.63 0.62 0.64 0.63
pburst;3d [MPa] 171.17 164.29 171.52 167.10 163.96

to the outside. Vice versa, placing the circumferential layer stack on the inside, like
in the case of Sequence A-S2, results in a comparably low cylinder expansion as
the circumferential layers directly bear the high loads on the inside.
Given the magnitude of longitudinal stress through-the-thickness and the previ-
ously determined ply strengths, allows to obtain predictions regarding the cylinder
strength of the distinct sequences. In this case, a basic maximum stress failure
criterion is applied to predict burst. Correspondingly, Table 6.1 summarizes outer
meridional and tangential strains as well as predicted burst pressures for the five
sequences.
Accordingly to the stress distribution shown in Fig. 6.2, the predicted burst
pressures show the same trend. In fact, the longitudinal stress in the innermost cir-
cumferential layer determines the magnitude of the burst pressure. Moreover, the
aforementioned trends with regard to outer surface strains are preserved. While
placing layers with high tangential stiffness on the laminate inside prevails mi-
nor cylinder expansion (e.g. Sequence A-S2), the positioning towards the outside
comes at the expense of a larger cylinder expansion (e.g. Sequence A). With regard
to its mechanical performance it can be observed that the grouping of circumferen-
tial layers (e.g. Sequences A/A-S1) has a notiably higher impact on the predicted
burst pressure than the positioning (e.g. Sequences A/A-S2). Indeed, staggering
multiple groups of circumferential layers leads to an increasing stress gradient
from inner to outermost circumferential layer. While the stress in the innermost
circumferential layer increases, the outermost cannot contribute as much, which
thus leads to a lower predicted burst pressure. Contrarily, by grouping circumfer-
ential layers, the stack can bear the tangential loads in collective. This in turn,
leads to a smaller longitudinal stress gradient within the different circumferential
layers and promotes a higher predicted burst pressure. Here, the different position-
ing of the stack appears to only minorly affect the maximum stresses, as seen by
the comparison of predicted burst pressures of Sequences A and A-S2.
116 6 Influence of stacking sequence

6.1.2 Observations on vessel failure

Following the observations by using the three-dimensional elasticity theory, the


experimental burst pressures and vessel remainders are discussed. Therefore, ini-
tially one vessel per configuration is analyzed, whilst the average values for the
configurations with multiple vessels are indicated in the following subsections.
Figure 6.3 shows the remainder of a selected vessel and its burst pressure for each
tested stacking sequence.
Noticeable differences in the remainders are observed between the Sequences
A, A-S2 and A-S4. While for Sequence A, a cylinder failure can be assured, the
failure of Sequences A-S2 and A-S4 seems to be driven by the cylinder-dome
transition or the dome proper. The failures of Sequences A-S1 and A-S3 remain
ambiguous to determine as partially cylinder and dome components are missing.
Besides the differences in failure, the experimentally determined burst pressures
also show major differences. Table 6.2 summarizes the experimental burst pres-
sures and the predictions made by the three-dimensional elasticity approach.
The error in burst prediction appears to be related with the location of failure.
For a preserved cylinder failure, such as in the case of Sequence A, the deviation
between predicted and experimental burst pressure only accounts for about 4.18%.
In contrast, a failure that is not preserved in the cylinder, such as in the case of
Sequence A-S2, leads to a considerable discrepancy between prediction and ex-

A A-S1 A-S2 A-S3 A-S4


pburst= 164.31 MPa 156.08 MPa 98.31 MPa 163.13 MPa 136.63 MPa

Fig. 6.3 Remainder and burst pressure of a single vessel for each stacking sequence
6.2 Fiber volume fraction and porosity 117

Table 6.2 Comparison of predicted and experimental burst pressures


A A-S1 A-S2 A-S3 A-S4

pburst;3d [MPa] 171.17 164.29 171.52 167.10 163.96


pburst;exp [MPa] 164.31 156.08 98.31 163.13 136.63
Rel. Diff. [%] 4.18 5.26 74.47 2.43 20.00

periment of about 74.47%. Resulting from these distinct behaviors, the following
chapter is assembled to successively unveil the effects of stacking sequence and to
pinpoint the causes for the different failures, seen in Fig. 6.3.

6.2 Fiber volume fraction and porosity

As previously demonstrated in Chap. 5, the filament winding process inherently in-


troduces a gradient in consolidation level through-the-thickness. Besides the fiber
tension, the angle at which the tow is applied onto the winding surface determines
the radial pressure and therefore the achieved compaction. For layers whose angle
in the cylindrical section is mostly aligned with the meridional direction of the ves-
sel, such as low-angle helical layers, compaction is generally higher at the vessel
poles and rather low in the cylinder. In contrast, for layers whose angle is closer
aligned with the tangential direction of the vessel, such as circumferential layers,
maximum compaction is achieved in the cylinder. In consequence, changes in the
stacking sequence lead to differences in the consolidation level and finally affect
the distribution of fiber volume fraction and porosity through-the-thickness. To this
purpose, the distribution of FVF and porosity within the cylindrical section of the
Sequences A, A-S1 and A-S2 is investigated in detail. First, general magnitudes are
obtained through acid digestion tests, while the gradient in porosity is observed by
means of CT scans. To this purpose, the methodologies described in Sect. 3.3.3
are used. Figure 6.4 shows the determined fiber volume fractions and porosities
for the Sequences A, A-S1 and A-S2 obtained by means of acid digestion tests.
Here, each bar represents the average of five samples taken from the center of the
vessel’s cylindrical section. In addition to laminate sections, the grouped sections
of circumferential and helical layers in the Sequences A and A-S2 were separately
accessed by grinding the remaining laminate sections away.
No substantial differences in the laminate FVF can be observed between the
sequences, which is anticipated due to the fact that there is no resin loss during
118 6 Influence of stacking sequence

Fig. 6.4 Fiber volume frac- 70 *includes excess 10


Vf - Laminate*

Fiber Volume Fraction Vf [%]


tion and porosity obtained Vf - Circumferential resin on outer
by means of acid digestion 65 Vf - Helical surface 8
tests for Sequences A, A-S1 Vp - Sample

Porosity Vp [%]
and A-S2
60 6

55 4

50 2

45 0

A A-S1 A-S2

manufacturing. Nonetheless, the distribution through-the-thickness underlies no-


ticeable differences as the comparison of FVF within the helical layer stack in
Sequences A and A-S2 indicates. In Sequence A, the circumferential layer stack
effectively compacts the helical layer stack beneath, which translates into a high
FVF and low porosity. Contrarily, positioned on top of the circumferential layer
stack, the helical layer stack in Sequence A-S2 are missing compaction which
yields noticeably lower FVF and higher porosity values. These differences in com-
paction are furthermore visible on the outer vessel surface. Figure 6.5 shows the
outer vessel surfaces of Sequences A, A-S1 and A-S2. The high compacting effect
of the circumferential layer stack in sequence A prompts the resin to flow out from
the beneath located helical layer stack towards the laminate outside. In contrast, the
missing compaction in the case of Sequence A-S2 leaves the outer surface almost
dry. In consequence, by considering the entire laminate section together with the
outer surface resin layer for the Sequences A and A-S1 suggests lower FVF values
than in reality found in the laminate.

a) b) c)

A A-S1 A-S2

Fig. 6.5 Comparison of outer vessel surfaces between Sequences A, A-S1 and A-S2
6.2 Fiber volume fraction and porosity 119

Fig. 6.6 Porosity through- 12


A CFRP
the-thickness for the 10 A-S1

Porosity Vp [%]
Sequences A, A-S1 and 0 1
8 A-S2
A-S2
6
4
2
0
0 0.2 0.4 0.6 0.8 1
Normalized Radial Coordinate r/ra [-]

While the results obtained from acid digestion tests give insights about the mag-
nitude of FVF and porosity, the resolution with respect to the position through-the-
thickness is rather low. Therefore, CT scans are used to obtain information about
the gradient in porosity through-the-thickness. Figure 6.6 illustrates the porosity
through-the-thickness for the Sequences A, A-S1 and A-S2. Here, each data point
represents the average of three samples, whereas the standard deviation is indicated
by the shading.
As anticipated, the porosity increases from the innermost to the outermost lay-
ers for all three sequences. While this trend is preserved for all sequences, the
gradient and the magnitude of average porosity noticeably differ between them.
The alternating change of layer groups in Sequence A-S1 leads to a gradual in-
crease in porosity. Every subsequent group of circumferential layers compacts the
helical layer group below, which results in an average porosity of 2.20%. In the
case of Sequence A, the presence of the circumferential layer group on the lami-
nate outside reduces the porosity especially for the inner helical layer group, which
is in accordance to the results, obtained from the acid digestion tests shown in
Fig. 6.4. Overall, the Sequence A shows the lowest average porosity with 1.85%.
Contrarily, the Sequence A-S2 shows the highest porosity with an average value
of 3.41%. Due to the fact that the compacting effect of the circumferential layers
cannot be exploited by placing them on the laminate inside, the porosity notice-
ably increases towards the outside. Furthermore, similarly to the sudden increase
in porosity for Sequence A, the Sequence A-S2 shows a high porous area at the
transition from circumferential layer stack to helical layer stack. Likewise to the
explanation given in Chap. 5, this sudden increase resides in the accumulation of
resin after the compacting circumferential stack.
Concluding from this, it becomes evident that the stacking sequence consider-
ably influences the distribution of FVF and porosity through-the-thickness within
120 6 Influence of stacking sequence

CPVs. Whereas the high compaction force applied through circumferential layers
generally results in high fiber volume fractions and low porosities for the pre-
viously wound layers, the compacting effect of low-angle helical layers in the
cylindrical section is rather small. At layer group transitions, there is a tendency
of increasing porosity due to changing fiber direction. This effect is amplified at
transitions after large stacks of circumferential layers. The stacking of multiple cir-
cumferentials results in an increased resin flow through-the-thickness towards the
transition where it accumulates and increases the probability of air inclusion.

6.3 Deformation behavior

The influence of stacking sequence on the deformation behavior is evaluated with


regard to two aspects, the change in cylinder strains and the strain distribution in the
dome/cylinder-dome transition. As previously inferred, the change in the location
of circumferential layers directly impacts the cylinder expansion. In turn, notice-
able differences between Sequences A and A-S2 are anticipated as the through-the-
thickness position of load carrying circumferential layers is completely opposite.
In fact, a much lower cylinder expansion is expected for the Sequence A-S2 due
to the circumferential layers being positioned as innermost layers. Contrarily, it
is anticipated that the Sequence A will show the largest cylinder expansion due
to the placement of circumferential layers on the outside. Furthermore, informa-
tion about the strain distribution within the dome and cylinder-dome transition is
sought to investigate the observed failure of the Sequences A-S2 and A-S4 seen in
Fig. 6.3.

6.3.1 Cylinder strains

A comparison of meridional and tangential cylinder strains for the five investi-
gated sequences is shown in Fig. 6.7. Here, one vessel per stacking sequence is
illustrated, whilst the average values for the sequences with more than one tested
vessel are summarized in Table 6.3. Following the trends predicted by the three-
dimensional elasticity theory, the most recognizeable difference is observed in the
tangential cylinder strains between the Sequences A and A-S2. The placement of
circumferential layers towards the laminate outside promotes a noticeably larger
cylinder expansion for Sequence A at the same pressure stage than in compar-
ison to the other sequences. Vice versa, Sequence A-S2 with the vast majority
of circumferential layers positioned on the laminate inside translates into a much
6.3 Deformation behavior 121

1.8 1.8
A A
A-S1 A-S1
Meridional Strain εs [%]

A-S2 A-S2

Tangential Strain εφ [%]


1.2 A-S3 1.2 A-S3
A-S4 A-S4

0.6 0.6

p = 1 MPa s-1 p = 1 MPa s-1


T = 23 °C T = 23 °C
0 0
0 60 120 180 0 60 120 180
Internal Pressure p [MPa] Internal Pressure p [MPa]
a) Meridional cylinder strain b) Tangential cylinder strain

Fig. 6.7 Meridional and tangential cylinder strains for the investigated stacking sequences

lower cylinder expansion at the same pressure stage. The difference between both
is likely to further increase with higher pressure stages as the larger cylinder ex-
pansion in Sequence A inevitably leads to more damage within the helical layers
as they are continously required to transfer the load to the stiffer circumferen-
tial layers on the laminate outside. Contrarily, in Sequence A-S2 due to the inner
positioning, the circumferential layer stack can directly bear the high tangential
loads, which limits the cylinder expansion and therefore the damage created within
the helical layers. Yet, this direct comparison at higher pressure stages cannot be
shown as Sequence A-S2 preliminary fails at 98:31 MPa. Following the observa-
tions with regard to the cylinder expansion and circumferential layer placement, it
is observed that the tangential strains of the Sequences A-S1, A-S3 and A-S4 fall
between the two extreme configurations and further assure this trend.
Concerning the meridional strains, a similar trend to the tangential strains is
observed although the differences between the sequences appear smaller. Non-
theless, as previously demonstrated in Chap. 5, the meridional cylinder strain is
particularly affected by the creation of damage within the composite. In this case,
two effects can be attributed to the higher meridional cylinder strain of Sequence
A. First, the larger tangential expansion promotes a higher probability for damage
within the low-angle helical layers, that in turn increases their compliance in the
meridional direction, too. Second, the recognized unobstructed transverse failure
122 6 Influence of stacking sequence

Table 6.3 Average burst pressure, CFRP mass, cylinder strains and axial displacement for
the stacking sequences with three tested specimens
Stacking pburst mCFRP "' ;70 MPa "s ;70 MPa uz ;70 MPa
sequence [MPa] [kg] [%] [%] [mm]
A 164.29 ˙ 1.49 3.57 ˙ 0.01 0.68 ˙ 0.01 0.51 ˙ 0.00 3.15 ˙ 0.09

A-S1 151.57 ˙ 3.21 3.65 ˙ 0.03 0.61 ˙ 0.01 0.47 ˙ 0.01 2.45 ˙ 0.06

A-S2 97.17 ˙ 4.18 3.64 ˙ 0.02 0.61 ˙ 0.00 0.47 ˙ 0.00 2.35 ˙ 0.02

A-S4 130.40 ˙ 1.26 3.61 ˙ 0.01 0.63 ˙ 0.00 0.48 ˙ 0.01 2.68 ˙ 0.17

of the circumferential layer stack leads to decreasing meridional stiffness as well.


As such, from 20 to 30 MPa internal pressure a minor increase in meridional strain
for Sequence A is recognized. Although comparably small, this effect appears to be
further reproducible as the meridional strains for all tested specimens in Table 6.3
indicate.
Altogether, the observed differences in cylinder deformation follow the trends
predicted by the three-dimensional elasticity approach and show no apparent phe-
nomena that can provide an explanation for the distinct burst pressures and failure
pictures. Therefore, as anticipated in prior, special attention is paid to the strain
distribution in the dome and cylinder-dome transition.

6.3.2 Strain distribution in dome and cylinder-dome transition

The dome and cylinder-dome transition can confidently be considered as critical


design regions for a CPV. Within this experimental setup, the Layup A was ini-
tially chosen for the investigation of the stacking sequence influence due to its
preserved cylinder failure. Yet, the failures of the Sequences A-S2 and A-S4 appear
to be driven outside of the cylinder. In this regard it is worth mentioning the mul-
titude of effects that potentially affect the behavior in the cylinder-dome transition
before the evaluation of the distinct strain responses. The end of the cylindrical sec-
tion also marks the area in which circumferential layers are subsequently dropped.
Thus, sudden changes in stiffness are recognized at the boundary between cylinder
and dome. Next to that, as previously discussed in Chap. 5, the geometrical setup
inevitably leads to the existence of bending close to the transition of cylinder to
dome. Here, the change in stacking sequence not only influences the laminate’s
6.3 Deformation behavior 123

1.5 1.5
A A
A-S1 A-S1
A-S2 A-S2
Meridional Strain εs [%]

Tangential Strain εφ [%]


A-S3 A-S3
1 A-S4 1 A-S4

0.5 0.5

p = 70 MPa p = 70 MPa
0 0
100 150 200 250 100 150 200 250
Axial Coordinate z [mm] Axial Coordinate z [mm]

a) Meridional strain b) Tangential strain

Fig. 6.8 Strain components along the meridional surface path at an internal pressure of
70 MPa for the investigated stacking sequences

bending stiffness, but can also lead to the existence of extensional-bending cou-
pling, that either promotes or counteracts the already existing meridional bending.
Considering these aspects allows to better evaluate the strain distributions seen
in Fig. 6.8. For all five sequences, both meridional and tangential strains are dis-
played along the meridional surface path at an internal pressure of 70 MPa.
Here, the dashed vertical line represents the geometrical cylinder-dome transi-
tion of the liner geometry. Vastly distinct responses at the cylinder-dome transition
region from 100 to 150 mm are recognized. The largest difference is noticed be-
tween Sequence A and A-S2. In fact, both sequences show contrary strain distri-
butions. For Sequence A, a minimum in meridional and tangential strain is seen
after the transition. Contrarily, Sequence A-S2 shows a maximum in meridional
and tangential strains at these positions. In principle, both strain components can
be separately accessed to obtain different informations. While the meridional strain
component allows to assess the bending response at the cylinder end, the tangential
strain component gives insights about the varying tangential stiffness distribution.
Yet, as both strains are conditioned by each other, a clear determination of the
causes for the strain magnitudes cannot always be given.
124 6 Influence of stacking sequence

100 A
Radial Coordinate r [mm]

80

A-S1
60

A
40 A-S2
A-S1
A-S2
Liner 20 mm
20
80 100 120 140 160
Axial Coordinate z [mm]
a) Experimental thickness profiles b) Cross-section cylinder-dome transition

Fig. 6.9 a) Experimentally obtained thickness profiles and b) cross-sectional view of cir-
cumferential ply drop locations at the cylinder-dome transition for the Sequences A, A-S1
and A-S2

Considering the geometrical induced bending, which has been depicted in


Fig. 5.12 in case of a homogeneous orthotropic vessel, the trend in meridional
strain for Sequence A-S2 appears counterintuitive. While the stacking sequence
asymmetry may lead to extensional-bending coupling that partially impacts the
bending response, the observed strain distribution appears as an aggregation of
additional effects. This is further assured by considering the response of Se-
quence A-S1. Through its closely symmetric stacking sequence, the influence of
extensional-bending coupling is effectively non-existent. Still, a similar trend in
meridional strain is recognized at the transition. Instead, the high tangential strain
after the transition may indicate missing tangential stiffness in this region, that
further contributes to the distinct meridional strain distribution. To this purpose,
the thickness build-up and the circumferential ply drop locations are investigated.
Figure 6.9 shows the experimentally obtained thickness profiles and the cross-
sections of the cylinder-dome transition region of one vessel for the Sequences A,
A-S1 and A-S2.
From the experimental thickness profiles it can be observed that the thickness
build-up at the boss neck is almost identical between the sequences. Yet, slight
differences in the cylinder-dome transition region are noticed. By observing the
cross-sections of the cylinder-dome transition region, it can be seen that the ex-
6.3 Deformation behavior 125

tent to which the circumferential layers reach into the dome proper varies. This
effect appears to be manufacturing-related as the nominal circumferential ply drop
locations remained identical within the filament winding program. In fact, it is an-
ticipated that two main effects lead to the discrepancies in circumferential ply drop
locations, which are briefly described here. First, a minor shrinkage in liner length
of about 3 mm is recognized within the filament winding process. This arises due to
the high-tensioned application of low-angle helical layers. In the case of Sequence
A, these helical layers are first wound, which leads to a very early decrease in
length. After the liner shrank, the circumferential layers are wound, which virtually
moves their ply drop locations towards the domes. Second, the different underlying
surface and curvature due to the different stacking might affect the ply readjust-
ments during winding and curing, too. The implications of these differences in
circumferential ply drop locations hamper to exactly pinpoint which effects seen
in the mechanical response arise due to the change in stacking sequence. The re-
tracted circumferential ply drop locations in Sequences A-S1 and A-S2 lead to
locally missing tangential stiffness in these regions. Thus, in these areas an appar-
ent expansion is noticed by high tangential strains. This in turn, appears to further
drive the preliminary failure of the Sequence A-S2. Yet, this cannot confidently be
assured through sole experimental observations.
To this purpose, the established FE model is applied onto the Sequences A-S1
and A-S2 in order to derive further insights into their distinct mechanical response.
Figure 6.10 shows the comparison of predicted and experimental strains along the
meridional surface path for the Sequences A, A-S1 and A-S2. Due to the pre-
liminary failure of Sequence A-S2, the comparison is established at an internal
pressure of 70 MPa, whereas the other two sequences are compared at 105 MPa
internal pressure.
For all three stacking sequences, the predicted deformation behavior closely
resembles the experiment along the entire meridional surface path. Minor dis-
crepancies in strain magnitude are recognized, which are argued to arise from
deviations in the predicted and experimental circumferential ply drop locations.
This furthermore highlights the sensitivity of this effect on both, experimental and
numerical scale. On the one hand, the position of circumferential tapering highly
impacts the tangential strain distribution for the given layups. On the other hand,
the minor mismatch in the predicted tapering causes noticeable deviations to the
experiment.
126 6 Influence of stacking sequence

Tangential Strain Meridional Strain Contour


1.5
p = 105 MPa Measured contour A
100
Nominal liner geomety

Radial Coordinate r [mm]


1
Strain ε [%]

0.5
50
Avg. Experiment
SD Experiment
Simulation
0
1.5
p = 105 MPa A-S1
100

Radial Coordinate r [mm]


1
Strain ε [%]

0.5
50
Avg. Experiment
SD Experiment
Simulation
0
1.5
p = 70 MPa A-S2
100
Radial Coordinate r [mm]

1
Strain ε [%]

0.5
50
Avg. Experiment
SD Experiment
Simulation
0
100 150 200
Axial Coordinate z [mm]

Fig. 6.10 Comparison of predicted and experimental strain components along the merid-
ional surface path for the Sequences A, A-S1 and A-S2
6.4 Damage progression and final failure 127

6.4 Damage progression and final failure

The stacking sequence further impacts the initiation and progression of damage. As
seen in Fig. 6.2, the stacking sequence definition determines the stress distribution
through-the-thickness. Placing circumferential layers towards the laminate outside,
such as in the case of Sequence A, requires the low-angle helical layers to transfer
the high tangential loads to the load carrying circumferential layers positioned on
the outside. Vice versa, placing circumferential layers closer to the liner surface,
as in the case of Sequence A-S2, does not necessitate a load transfer from helical
to circumferential layers. This does not only affect the cylinder expansion, as seen
in Fig. 6.7, but also the amount of damage that is created in the different layer
types. Once a crack in the matrix constituent is initiated within a ply, the layer
architecture and the grouping of likewise aligned layers determine how far the
crack is propagating, which was demonstrated in Fig. 4.17. Besides the propagation
of IFF in the cylinder, it is expected that the protruding shapes at the cylinder-dome
transition of Sequences A-S1 and A-S2 are likely to further promote the existence
of interlaminar damage. In the following, the onset and progression of IFF as well
as the final failure of the Sequences A, A-S1 and A-S2 is investigated in detail by
accessing the transverse damage variable, experimental microscope cross-sections,
predicted and experimental burst pressures and CT scans.

6.4.1 Interfiber failure

From the moment of first loading with internal pressure on, IFF develops in CPVs
at already mild pressure stages, which was shown in Sect. 4.3.1. Between the
two commonly used layer types, circumferential and helical layers, an intrinsic
difference in crack propagation exists. While the intertwined helical layer archi-
tecture hampers crack growth in-plane and through-the-thickness, the closely uni-
directional circumferential layer architecture allows cracks to propagate around
the circumference and through-the-thickness. As such, the stacking of multiple
circumferential layers inevitably leads to large fracture surfaces, that, when posi-
tioned towards the outer laminate surface, result in high amplitude AAE events.
Contrarily, alternating the layer orientation frequently and grouping small num-
bers of circumferential layers together leads to smaller fracture surfaces and in
turn lower AAE event amplitudes. The preceding description can be observed for
the Sequences A, A-S1 and A-S2 in Fig. 6.11, where a comparison of channel files
is seen. The channel files correspond to pressurizations up to burst for one vessel
128 6 Influence of stacking sequence

Sound Pressure [Pa] Sound Pressure [Pa] Sound Pressure [Pa]


25

0
a) A

-25
0 25 50 75 100 125 150
Time t [s]
25

b) A-S1 0

-25
0 25 50 75 100 125 150
Time t [s]
25

c) A-S2 0

-25
0 25 50 75 100 125 150
Time t [s]

Fig. 6.11 Comparison of channel files for a pressurization up to burst; shown data cor-
responds to one vessel for each sequence; the vessels were not pressurized prior to the
experiment

of each sequence. The vessels were not pressurized prior to the experiment. In all
three cases t D 0 corresponds to the beginning of the pressurization.
Within early stages of the pressurization (< 50 s), the highest AAE amplitudes
are recorded for the Sequence A, signifying IFF in the circumferential stack lo-
cated on the laminate outside. Contrarily, for Sequence A-S2 noticeably lower
amplitudes are recognized due to the muffling effect that the subsequently over-
wound helical layers have. Sequence A-S1 shows the lowest AAE amplitudes,
which follows the previous discussion about the relationship of fracture surface
and AAE amplitude. With regard to the development and progression of damage,
Fig. 6.12 a) shows the transverse damage variable for all three stacking sequences
at internal pressures of 50 and 105 MPa and b) a microscopic view on the laminate
cross-section in the cylindrical section for each sequence. To obtain these images,
one vessel per sequence was loaded and unloaded with an internal pressure and a
cross-section was cut out of the cylindrical section afterwards. While for the Se-
6.4 Damage progression and final failure 129

a) Simulation | Transverse damage variable b) Experiment | Microscope cross-section


(Circumferential view on cylinder)

p = 50 MPa

p = 105 MPa

p = 105 MPa 5 mm
A

p = 50 MPa

p = 105 MPa

A-S1 p = 105 MPa 5 mm

p = 50 MPa

p = 70 MPa

A-S2 p = 70 MPa 5 mm

Fig. 6.12 Comparison of damage variable in transverse direction and microscope cross-
section after pressurization for the Sequences A, A-S1 and A-S2

quences A and A-S1, the vessels were loaded to 105 MPa internal pressure, the
vessel of Sequence A-S2 was only loaded with 70 MPa due to the expected pre-
liminary failure.
In all three cases, IFF is visible within the laminate. As anticipated, for Se-
quence A-S1 the frequent change from circumferential to helical layer groups
prevents the development of large fracture surfaces, such as in the case of Se-
quences A and A-S2. Besides the distinct crack propagation in the cylindrical
section, differences in the amount of damage are recognized by the transverse
130 6 Influence of stacking sequence

damage variable in dome and cylinder-dome transition. The large expansion in the
cylinder-dome transition for the Sequences A-S1 and A-S2 leads to a substantially
higher damage diffusion in this region at the same pressure stage of 50 MPa. The
developed IFF causes a redistribution of load to the undamaged neighbouring plies,
which in turn increases their stresses and causes them to also fail progressively. As
such, damage in these areas is evolving step by step, which is not only limited to
intralaminar but also interlaminar damage, that eventually prompts a local failure.

6.4.2 Final failure

The vessel remainders seen in Fig. 6.3 indicate a very distinct failure of the Se-
quences A, A-S1 and A-S2. Depending on the remainders, either failure in the
cylinder or outside the cylinder can be concluded. Nonetheless, a certain ambigu-
ity remains, for which the preceding observations can help argue for the reasons
and location of final failure. For the Sequences A-S1 and A-S2 a particularly large
tangential expansion is noticed in proximity to the cylinder-dome transition, that
appears to be impacted by the retracted circumferential ply drop locations seen
in Fig. 6.9. The protruding shape at the transition further promotes the develop-
ment of damage in this area as Fig. 6.12 shows. The question arises whether this
diffusion of damage further impacts the final failure, particularly for Sequences
A-S1 and A-S2. Therefore, the prediction of burst pressure and failure location are
considered and compared with experimental observations. Table 6.4 provides an
overview of predicted and experimental burst pressures and failure locations. The
overview includes the prediction made by the three-dimensional elasticity theory
as well as the predictions made by the FE model with and without considering
damage. The predicted failure location was determined based on the axial location
of the highest failure index.
A very good accordance between all three predicted and experimental burst
pressures is achieved for Sequence A, where failure is not only predicted but also
experimentally observed in the cylinder. Depending on the used prediction method,
the discrepancy to the experiment accounts between 0.92 and C4.19%. Slightly
more deviation is noticed for Sequence A-S1, where the prediction accuracy varies
between 1.96 and C8.39%. Particularly the consideration of damage in the FE
model appears to refine the prediction accuracy from C7.28 to 1.96%. Based on
the vessel remainders, the failure location in the experiment cannot be determined
without ambiguity for this sequence. The predicted failure location however re-
mains in the cylinder even when damage is considered in the FE model. This in
turn would suggest, that even though a large diffusion of damage is seen in the
6.4 Damage progression and final failure 131

Table 6.4 Comparison of burst pressures and failure locations between experiment and
predictions made by three-dimensional elasticity theory and FE model
Stacking Experiment 3D elasticity FEA FEA
sequence theory w/o damage homo w/ damage homo
Failure Burst Failure Burst Failure Burst Failure Burst
locationa pressure locationb pressure locationb pressure locationb pressure
[MPa] [MPa] [MPa] [MPa]
A CS 164.29 CS 171.17 CS 169.50 CS 165.80
˙1.49 (C4.19%) (C3.17%) (0.92%)
A-S1 CNBD 151.57 CS 164.29 CS 162.60 CS 148.60
˙3.21 (C8.39%) (C7.28%) (1.96%)
A-S2 OCS 97.17 CS 171.52 OCS 132.60 OCS 116.40
˙4.18 (C76.51%) (C36.46%) (C19.79%)

CS – Cylindrical section; OCS – Out of cylindrical section; CNBD – Cannot be determined


a
Failure location was determined based on visual inspection of the vessel remainders
b
Failure location was determined based on the axial location of highest failure index
() Indicates the relative difference to the experimentally obtained burst pressure

cylinder-dome transition of this sequence, it does not trigger a preliminary failure.


Contrarily, for Sequence A-S2 not only a failure outside the cylinder is seen in the
experiment but is also predicted by both FE model variants (w/o and w/ damage).
While the discrepancy to the experiment accounts C36.46% for the constitutively
elastic solution, the consideration of damage noticeably refines the prediction ac-
curacy, which leads to a difference of C19.79% to the experiment. The particular
strong influence of the consideration of damage on the predicted burst pressure
of Sequence A-S2 is reasoned in the large diffusion of damage in the cylinder-
dome transition region. The large tangential expansion in this area appears to not
only increase the development of damage but also seems to drive a preliminary
failure. Still with considering damage, a noticeable difference between predicted
and experimental burst pressure of about C19.79% exists. The discrepancy is ar-
gued in the sole depiction of intralaminar damage, which does not fully represent
the damage created within the composite. Besides the development of intralam-
inar damage, interlaminar damage is further growing between layer interfaces.
Promoted through interlaminar stresses, which are particularly present at ply drop
locations at the cylinder-dome transition due to the so-called free edge condition
[123], delaminations progressively evolve and impair the load transfer between the
different layers.
To investigate the existence of interlaminar damage in the cylinder-dome tran-
sition region of Sequence A-S2, an additional vessel of this sequence is loaded to a
pressure stage close to final failure and unloaded afterwards. The transition is then
132 6 Influence of stacking sequence

Internal Pressure [%]


Helical layers
DIM +74.511 MPa 2.00
1.80 Helical layer Helical layers
1.60 Circumferential Delamination
Liner
1.40 layers
1.20 6 mm
High strain area Circumferential
layers
1.00
0.80
0.60
0.40
A-S2 Delamination
0.20
3 mm 35 mm
0.00
a) High strain area at 75 MPa b) CT scan after pressurization c) Vessel failure

Fig. 6.13 Investigation on the existence of interlaminar damage in the cylinder-dome tran-
sition of Sequence A-S2

observed by using CT scans. Since the magnification to this resolution requires


cutting of the dome and cylinder, this vessel cannot be burst. Therefore, one of the
burst vessels is also CT scanned to compare the failure with the observed trends in
the cylinder-dome transition region. Fig. 6.13 a) shows a localized high strain area
found within the cylinder-dome transition of Sequence A-S2 at an internal pressure
of 75 MPa, b) a developed delamination between the layer groups at the tip of the
circumferential layer group, which corresponds to the identified high strain area
and c) the resulting vessel failure of Sequence A-S2.
The growing delamination seen in Fig. 6.13 b) is argued to be the result of the
excessive tangential expansion seen for Sequence A-S2. Due to the retracted cir-
cumferential ply drop locations, the compliant outer helical layer group is exposed
to high tangential loads. This causes a large expansion at the tip of the circum-
ferential layer group, which induces tensile stresses within the interface of helical
and circumferential layers. As such, the force translation at the tip gets impaired
causing the loading path to deviate and further promoting stress concentrations
within the circumferential layer stack that are likely to trigger the seen failure in
Fig. 6.13 c). Nonetheless, the observation of these CT images does not allow to un-
ambigously pinpoint the cause and sequence of events to assure this hypothesis. To
further enhance the structural understanding in this area, the implementation of co-
hesive interfaces between the layer groups may be beneficial and could potentially
further increase the prediction accuracy of the FE model.
6.4 Damage progression and final failure 133

6.4.3 Influence of circumferential ply drop locations

The multitude of effects that are present at the cylinder-dome transition hinders
to confidently pinpoint the influence of each individual phenomenon. As seen in
Fig. 6.10, the deformation behavior as well as the development of damage in this
area noticeably vary between the Sequences A and A-S2. This appears to be not
only related to the distinct stacking sequence but also impacted by the differences
in circumferential ply drop locations, seen in Fig. 6.9. To this purpose, the influ-
ence of varying circumferential ply drop locations is investigated more in-depth
for the Sequence A-S2. The influence is assessed in a two-stage process. First, the
FE model is used and the impact of circumferential ply drop locations on the de-
formation behavior is evaluated. Thereby, the circumferential ply drop locations
are virtually extended or retracted and the influence on the strain distribution in the
cylinder-dome transition region is analyzed. The results imply a high sensitivity
of the strain distribution on the circumferential ply drop locations and are further
detailed in the Appendix C.1 in the Electronic Supplementary Material. Second,
a modified version of the Sequence A-S2 is manufactured and tested, where the
circumferential ply drop locations are effectively extended by 6 mm towards the
dome. This sequence is labeled as A-S2-T1. The resulting change in outer contour
and strain distribution is shown in Fig. 6.14. While the change in outer contour is
barely noticeable between both configurations, the changes in the strain distribu-
tion at the cylinder-dome transition can be undoubtly identified.
The farther circumferential ply drop location for Sequence A-S2-T1 ensures
sufficient tangential stiffness to prevent a swelling expansion in the cylinder-dome
transition, which is seen in the case of Sequence A-S2. The prevention of this ex-
cessive deformation furthermore impacts the resulting failure and burst pressure,
as Fig. 6.15 shows. With a determined burst pressure of 148:60 MPa, a relative
increase in burst pressure by C52.93% compared to the Sequence A-S2 is recog-
nized. Furthermore, differences in the vessel remainders are recognized. While for
Sequence A-S2 large sections of the vessels remain intact, the higher failure pres-
sure for Sequence A-S2-T1 leads to noticeably less remainders. Interesting in this
regard is the liner rupture in the upper dome section for Sequence A-S2-T1, which
potentially implies that even though the final failure was delayed, the failure may
still be triggered in the dome/cylinder-dome transition. In light of this, a further
detailed analysis on the development of interlaminar damage at the tip of the cir-
cumferential layer stack, seen in Fig. 6.13, poses an interesting subject for future
research.
134 6 Influence of stacking sequence

Layup A-S2 Meridional Strain Contour


Layup A-S2-T1 Tangential Strain SD
1.5
p = 70 MPa Nominal liner geometry A-S2 A-S2-T1
100
Measured contours

Radial Coordinate r [mm]


1
Strain ε [%]

0.5
50

0
100 150 200
Axial Coordinate z [mm]

Fig. 6.14 Strain components along the meridional surface path at an internal pressure of
70 MPa for the Sequences A-S2 and A-S2-T1

Fig. 6.15 Comparison


of vessel remainders and
burst pressures between Se-
quences A-S2 and A-S2-T1

A-S2 A-S2-T1
pburst = 97.17 ± 4.18 MPa pburst = 148.60 MPa
(+52.93%)
6.5 Concluding remarks 135

6.5 Concluding remarks

Research question 3:
What influencing factors need to be taken into account with regard to the definition
of a CPV stacking sequence?

Changing the stacking sequence of a CPV inevitably influences a multitude of as-


pects. Among these aspects is the consolidation level of individual plies and of the
general laminate, the stress-strain distribution through-the-thickness, the deforma-
tion behavior, the initiation and progression of interfiber failure as well as the final
failure.
Beginning with the manufacturing, the stacking sequence strongly impacts the
applied compaction during filament winding and therefore the consolidation level
through-the-thickness. The applied compaction of a layer is dependent on its orien-
tation relative to the longitudinal vessel axis. While circumferential layers are par-
ticularly effective in compacting the laminate in the cylindrical section, low-angle
helical layers contribute little to the overall compaction in the cylinder. Hence, in
this investigation the sequences where circumferential layers were placed as outer-
most layers (e.g. Sequence A) showed a higher consolidation level in the cylinder
than sequences, where the circumferential layers were located as innermost (e.g.
Sequence A-S2).
The grouping and positioning of different layers determines the stress-strain dis-
tribution through-the-thickness and by that it also affects the general deformation
behavior of the vessel. Placing layers with high tangential stiffness (e.g. circum-
ferential layers) as innermost layers in the cylinder leads to generally higher fiber
stresses in these layers as they carry the high tangential loads on the inside di-
rectly. Yet, it also causes a much smaller expansion in the cylindrical section (e.g.
Sequence A-S2). Contrarily, placing layers with high tangential stiffness as out-
ermost plies, generally decreases the fiber stresses in these layers in expense of a
larger cylinder expansion (e.g. Sequence A)
The grouping of layers furthermore affects the development and progression
of interfiber failure. Grouping circumferential layers together in a stack (e.g. Se-
quences A/A-S2) inevitably leads to the generation of larger fracture surfaces as
cracks can propagate through-the-thickness unhindered. In contrast, fragmenting
the stacking sequence into multiple stacks of circumferential and helical layers
prevents crack growth through-the-thickness due to numerous layer interfaces (e.g.
Sequence A-S1).
136 6 Influence of stacking sequence

In this investigation, the most significant differences were encountered in re-


gard to the strain distribution in the cylinder-dome transition, which impacted the
final failure too. The strain distribution in the cylinder-dome transition is affected
by a multitude of variables. Depending on the individual expansion of cylinder
and dome, the meridional bending at the transition varies. As the cylinder expan-
sion is heavily linked to the definition of the stacking sequence, the magnitude
of bending experienced at the transition is too. Moreover, the stacking sequence
not only determines the laminate’s bending stiffness but can also lead to the ex-
istence of extensional-bending coupling in case of highly asymmetric sequences
(e.g. Sequences A/A-S2). An additional effect, that appeared to majorly impact the
strain distribution in the cylinder-dome transition is the location of circumferen-
tial ply drops. Due to the different stacking sequences, the axial location of the
circumferential ply drops varied. This in turn, seemed to heavily impact the tan-
gential strain distribution in this region. For sequences, where the circumferential
ply drop locations were retracted towards the cylinder (e.g. Sequences A-S1/A-
S2) the growth of swelling areas in the cylinder-dome transition was recognized.
In case of Sequence A-S2, this not only led to the development of interlaminar
damage, but also to preliminary failure. Further numerical and experimental inves-
tigations highlighted the sensitivity of the deformation behavior and final failure
to the circumferential ply drop locations. In fact, for Sequence A-S2 a relative in-
crease in burst pressure by C52.93% compared to the baseline configuration was
noticed once the circumferential ply drop locations were extended by only 6 mm
towards the dome. Therefore, the results related to the strain distribution in the
cylinder-dome transition and the final burst pressures should be taken with cau-
tion. The differences in circumferential ply drop locations between the sequences
sensitively impact these two aspects, which overlays the investigation of stacking
sequence.
Finally, it should be considered that the observed effects in this chapter are valid
for the investigated selection of angles. Changing the layup composition may lead
to different observations. In this regard, the absence of high-angle helical layers
particularly increased the sensitivity of the deformation behavior and burst pres-
sure to the circumferential ply drop locations. Introducing high-angle helical layers
to the laminate will likely decrease the criticality of circumferential ply drop loca-
tions, which in turn would enable a more representative investigation of the sole
influence of stacking sequence.
Application on fullscale geometry
7

Finally, the derived insights are applied onto a fullscale vessel geometry, which
permits to expand the existing knowledge by considering a different dome geom-
etry and considerably larger geometrical dimensions. This chapter encompasses
an overview of the design considerations, the manufacturing and the testing of a
fullscale vessel geometry representative for the use in fuel cell electric vehicles.
The dimensions of this geometry are depicted in Fig. 3.3. The aim of this chapter
is to derive a design that fullfills the regular certification requirements with regard
to burst pressure and cyclic loading, while preserving a lower structural mass com-
pared to the state-of-the-art. In light of the aforementioned the formulated research
question assigned to this chapter is as follows.

Research question 4:
What considerations need to be made to transfer previously derived insights onto
a fullscale vessel geometry with distinct dome contour?

7.1 Design considerations

With the variety of effects interpreted from Chaps. 5 and 6, the observations are
still limited in the sense that solely one particular liner geometry has been consid-
ered. Yet, dome shape and cylinder diameter determine the resultant forces under
internal pressure loading. As such, the design of the composite reinforcement is
inherently linked to geometrical definition of the liner. Naturally, with increas-
ing diameter the required tangential reinforcement rises correspondingly. Here, the
relative increase in necessary tangential reinforcement to the relative diameter in-
crease is becoming larger for big diameters, because of the progressively increasing
normal stress gradient through-the-thickness. Yet, more importantly to the laminate
design of a CPV, is its dome shape that further dictates the number, the angles and
© The Editor(s) (if applicable) and The Author(s), under exclusive license to 137
Springer Fachmedien Wiesbaden GmbH, part of Springer Nature 2022
M. Nebe, In Situ Characterization Methodology for the Design and Analysis of
Composite Pressure Vessels, Werkstofftechnische Berichte j Reports of Materials
Science and Engineering, https://doi.org/10.1007/978-3-658-35797-9_7
138 7 Application on fullscale geometry

the trajectories of helical layers. In the following, the dome contour of the fullscale
geometry is first analyzed and the laminate design is discussed. To this purpose,
two differently composed layups are introduced and compared.

7.1.1 Dome contour

A comparison of subscale and fullscale liner geometry is shown in Fig. 7.1. The
shapes present noticeable differences in the variation of curvature. The subscale
geometry shows a rather steep dome contour, which closely resembles the shape
of an isotensoid solution [28]. Contrarily, the fullscale geometry exhibits a con-
siderably milder slope from cylinder to pole. Besides their implications on other
aspects (e.g. storage density or application of impact protectors), this difference in
shape has a profound impact on the load distribution in this region.
In case of the subscale geometry, the rapid decline from cylinder radius to polar
opening leads to a quick change in the predominant loading components of the in-
ternal pressure. From predominantly radial to axial loading, the laminate stiffness
is required to change equally fast while approaching the boss. While the sharp ge-
ometry change implies a rapid decrease in required tangential reinforcement, the
necessary reinforcement to bear the loads in the axial direction rises correspond-
ingly. The fast change in geometry also leads to a more pronounced constraint at
the cylinder end. This further increases the existing bending at the rigid connection
between cylinder and dome, when both sub-components expand under internal

100
Radial Coordinate r [mm]

50

Subscale Geometry
Fullscale Geometry

100 150 200 250 300


Axial Coordinate z [mm]

Fig. 7.1 Comparison of dome shape between subscale and fullscale geometry
7.1 Design considerations 139

pressure. Contrarily, the mild dome slope for the fullscale geometry requires a
gradual change in stiffness along the meridional path. The moderate change in di-
ameter lowers the required axial reinforcement and further narrows the existing
bending at the transition. Yet, sufficient tangential reinforcement needs to be en-
sured while approaching the boss neck.

7.1.2 Layup design

Following to the discussion concerning the geometry and laminate design, two
differently composed layups are exemplary applied to the fullscale geometry and
their deformation behavior as well as their mechanical performance is analyzed.
The layups were chosen based on a preliminary design study and were aimed to
achieve the burst certification requirement of 157:5 MPa according to the regula-
tion UNECE/R134 [2] while preserving a minimum amount of CFRP mass. Two
design philosophies were followed and are here compared. In one scenario the
choice of angles was strongly oriented on the Layup A of the subscale vessel ge-
ometry. This layup is labeled as Layup E in the following. On the other side, an
alternative approach was followed by introducing a multitude of different sets of
angles, here labeled as Layup F. The stacking sequence for both layups was cho-
sen according to Layup A, to exclude possible influences arising from the stacking
sequence definition. Their layup composition as well as the resulting masses are
summarized in Table 7.1. While the overall number of layers is the same, the layup
composition is noticeably different. Layup E exclusively features low-angle helical
and circumferential layers, whereas Layup F is composed of a variety of low-angle
and high-angle helical layers. Despite their identical number of layers, minor dif-
ferences in CFRP mass are recognized which is due to their different composition
in layer types.

Table 7.1 Comparison of E F


normalized layer number,
layup composition and
CFRP mass between Layup nlayers;norm: [–] 1 1
E and F
LH : HH : C a 0.46 : 0 : 0.54 0.37 : 0.18 : 0.44
mCFRP [kg] 19.17 19.44
a
LH – Low-angle helical layer (5°  ˛ < 40°); HH – High-
angle helical layer (40°  ˛ < 88°); C – Circumferential layer
140 7 Application on fullscale geometry

Fig. 7.2 Comparison of


thickness build-up and
dome contour during fil-
ament winding between
Layup E and F (similar
number of layers wound)

E F

Fig. 7.2 shows a comparison of both dome shapes during filament winding,
where the impact of their different layup composition can be unambigously iden-
tified. Layup E presents a comparably steep transition from circumferential to
low-angle helical layers. This not only results in a sudden change in wall thick-
ness, but also in tangential stiffness. At the tip, where the circumferential layers are
dropped off, the low-angle helical layers and their high compliance in the tangen-
tial direction are exposed to the radial loading component of the internal pressure.
Contrarily, a more gradual change in tangential reinforcement is ensured by the
layer sequence of Layup F. After the circumferential ply drop, a stepwise descent
by the chosen polar openings of high-angle helical layers is recognized. This in
turn, is anticipated to not only provide a more gradual distribution of tangential
stiffness while approaching the boss neck, but also to prevent locally increased
thickness build-up.

7.2 Structural analysis

Given the aforementioned discussions about dome geometry and layup composi-
tion, it is anticipated that the response at the cylinder-dome transition will notice-
ably differ between the two layups. In this section, the mechanical response of both
layups is analyzed with regard to the deformation behavior and final failure.

7.2.1 Deformation behavior

The comparably sudden change in tangential stiffness for Layup E is expected to


yield noticeably higher strains in the cylinder-dome transition. Contrarily, a much
7.2 Structural analysis 141

Layup E Meridional Strain Contour


Layup F Tangential Strain
1.5
p = 70 MPa
E F
Nominal liner geometry

Radial Coordinate r [mm]


100

1
Strain ε [%]

0.5
50

0
100 150 200 250
Axial Coordinate z [mm]

Fig. 7.3 Strain components along the meridional surface path at an internal pressure of
70 MPa for the Layups E and F

more gradual strain distribution is anticipated for Layup F, where the existence of
high-angle helical layers mitigates the still moderate radial loading component of
the internal pressure. Figure 7.3 shows the strain components along the meridional
surface path for the both layups at an internal pressure of 70 MPa.
The strain responses are plotted without the experimentally obtained thickness
profiles. After manufacturing, a light liner deformation for both layups was recog-
nized through the use of CT scans. The comparably low liner stiffness led to a local
denting in both axial and radial direction. Since the nominal liner geometry is not
the same between both layups, a comparison by means of experimental thickness
profiles would be misleading. Therefore, only the nominal liner geometry is shown
for reference. The differences are most noticeable in the cylinder-dome transition.
While both layups show similar cylinder strains, they exhibit distinct responses
when approaching the boss neck. Particularly recognizable is the high tangential
strain for Layup E, that positionwise coincides with the circumferential ply drop lo-
cation, seen in Fig. 7.3. The peak indicates an insufficient tangential reinforcement
at this position, which leads to a growing bulge. Comparing the magnitude of mea-
sured strains, it is likely to conclude that the swelling determines the maximum
vessel strength. In contrast, Layup F shows a more distributed strain response.
In fact, magnitude-wise the cylinder shows the highest strains. The existence of
142 7 Application on fullscale geometry

high-angle helical layers appears to sufficiently reinforce the cylinder-dome tran-


sition/dome region and therefore prevents an excessive expansion.

7.2.2 Final failure

Likewise to their distinct deformation behavior, both layups present apparent dif-
ferences in failure type and burst pressure. Figure 7.4 provides a comparison of the
vessel remainder and the burst pressure of the Layups E and F. The high tangen-
tial expansion in the case of Layup E, led to a preliminary failure located within
the upper dome/cylinder-dome transition. After the initial failure, the vessel was
accelerated against the bottom bearing which crushed the bottom dome, leaving
the cylindrical section intact. Contrarily, the layer composition of Layup E caused
a very clear shift in failure location. The predominant cylinder expansion, seen
in Fig. 7.3, resulted in a failure preserved in the cylindrical section. This is con-
cluded by the remainder of the domes and the longitudinal aligned liner rupture
in the cylinder, which can be interpreted as failure triggered by the governing tan-
gential stresses. The burst pressure of 155:93 MPa presents a relative increase of

E pburst = 93.91 MPa F pburst = 155.93 MPa

Fig. 7.4 Comparison of vessel remainder and burst pressure between Layup E and F
7.2 Structural analysis 143

Layup F Meridional Strain Contour


Layup F-C1 Tangential Strain SD
1.5
p = 105 MPa
F F-C1
Nominal liner geometry

Radial Coordinate r [mm]


100

1
Strain ε [%]

0.5
50

0
100 150 200 250
Axial Coordinate z [mm]

Fig. 7.5 Strain components along the meridional surface path at an internal pressure of
105 MPa for the Layups F and F-C1

66% in comparison to Layup E, while preserving similar mass. Based on this trend
in deformation behavior and burst pressure performance, the Layup F is modified
in order to achieve the minimum burst pressure requirement of 157:5 MPa. Given
the recorded strains, the measures to adjust the layup were limited to reinforcing
the cylinder with additional circumferential layers. Therefore, the derivative of the
Layup F is found in the Layup F-C1 with three additional circumferential layers
and an overall CFRP mass of 20:29 kg. The comparison of measured strains at an
internal pressure of 105 MPa between the Layup F and F-C1 is depicted in Fig. 7.5.
With increasing tangential stiffness, a decrease in tangential strain is recognized
for Layup F-C1 relative to Layup F. Whilst no substantial changes in the merid-
ional strain distribution is recognized in the cylinder, slight deviations are noticed
in the transition towards the dome. This change is anticipated to be in relation
to the distinct tangential cylinder expansion. With the cylinder expanding more,
the meridional bending at the transition increases correspondingly. Thus, smaller
cylinder expansion in the case of Layup F-C1 leads to a lower meridional strain at
the transition. In this regard, it is furthermore noted that the additional tangential
reinforcement does not lead to a shift in the location of maximum strain. While
the relative difference between highest strain in the dome and in the cylinder de-
creases, the cylinder still exhibits the overall higher strain. Correspondingly, the
144 7 Application on fullscale geometry

Table 7.2 Comparison of normalized layer number, layup composition and CFRP mass
between the Layups E, F and F-C1
E F F-C1

nlayers;norm: [–] 1 1 1.05


LH : HH : C a 0.46 : 0 : 0.54 0.37 : 0.18 : 0.44 0.35 : 0.18 : 0.47
mCFRP [kg] 19.17 19.44 20.29 ˙ 0.07
pburst [MPa] 93.91 155.93 167.83 ˙ 1.36
a
LH – Low-angle helical layer (5°  ˛ < 40°); HH – High-angle helical layer (40°  ˛ < 88°);
C – Circumferential layer

Layup F-C1 still preserves a failure in the cylinder. To assure this result, two addi-
tional vessels are tested virginally and another vessel is cyclic loaded.
The resultant average burst pressure of Layup F-C1 is p burst D 167:83 MPa
(SD D 1:36 MPa, n D 3). This confidently surpasses the minimum burst pressure
requirement, whilst allowing a certain span for manufacturing and material related
scatter. A final comparison of Layup F-C1 to the previously tested Layups E and F
is provided in Table 7.2. After surpassing the burst pressure requirement, an addi-
tional vessel of the Layup F-C1 is hydraulically pressure cycled to verify the layups
ability to widthstand cyclic pressure loading. The cyclic pressure loading test is of
vital importance as it demonstrates the vessel’s applicability for the on-road use
with regard to fatigue performance. As observed in Fig. 4.20, with progressively in-
creasing number of loading cycles, the crack density within the laminate increases.
Existing crack may further grow and connect, leading to delaminations in between
layer interfaces. Altogether, these various damage mechanisms prompt a local re-
distribution of load to the stiffer fibers, which in turn increases their longitudinal
stress and causes prematurely failure.
In this sense, the robustness of a design is defined by how much the virgin burst
pressure is decreased in comparison to the burst pressure determined after cyclic
testing. According to the regulation UNECE/R134 [2], the vessel is cycled for a
total number of 22,000 cycles from 2 to 87:5 MPa internal pressure. Afterwards,
the vessel is pressurized until burst to determine its residual burst strength and to
obtain insights into the material degradation that has taken place. Figure 7.6 shows
an image of the outer vessel surface after surpassing 22,000 loading cycles, where
an increased crack density in the outer excess resin layer is noticed.
The progressively increasing crack density is anticipated to trigger the redis-
tribution of load to the stiffer fibers, that in turn will exhibit higher stresses and
7.3 Comparison to the state-of-the-art 145

Fig. 7.6 High crack density


on outer excess resin layer
after surpassing the pres-
sure cycle test of 22,000
cycles from 2 to 87:5 MPa
internal pressure

are therefore likely to fail earlier. The determined residual burst strength accounts
138:51 MPa, which signifies a relative decrease of 17% compared to the previously
determined average. No apparent change in resultant failure location was observed,
which means that the cylinder still appeared to be driving the failure. The regulation
UNECE/R134 does not require to determine a certain residual burst strength after
the initial pressure cycle life test. Yet, the determined decrease in burst pressure pro-
vides valuable insights for other test sequences within the regulation, in which the
vessel is required to demonstrate a residual burst strength after being cyclic loaded
and exposed to other influences, such as chemicals or temperature changes [2].

7.3 Comparison to the state-of-the-art

Finally, the derived Layup F-C1 is contrasted against the current state-of-the-art
design on this liner geometry. In this context, it is worth mentioning that this com-
parison is not only impacted by a different layup definition, but also by another
manufacturing procedure, filament winding equipment and lastly material system.
The current state-of-the-art design is manufactured by using wet winding on a
regular portal filament winding system. In contrast, the design prototype is man-
ufactured by using robot-assisted towpreg winding. Moreover, the state-of-the-art
renders a design that has surpassed all requirements related to the certification of a
CPV [2], whilst the design prototype was only tested with regard to its performance
in burst and cyclic experiments. Nonetheless, fibers with very similar longitudinal
strength properties were used for both designs, which allows for a quantative com-
parison in terms of used CFRP mass and wall thickness. Figure 7.7 shows the
comparison of used CFRP mass and cylinder wall thickness.
In comparison to the current state-of-the-art design, Layup F-C1 reveals a mass
saving potential of 17%, while maintaining the regulation requirements with regard
146 7 Application on fullscale geometry

Fig. 7.7 Comparison of 30


CFRP Mass Wall Thickness

Cylinder Wall Thickness t [mm]


CFRP mass and cylinder
wall thickness between 25 20
state-of-the-art design and

CFRP Mass mCFRP


-17%
design prototype 20 -30%
15

15
10
10

5
5
24.6 19.4 20.3 13.6
0 0
State-of-the-Art Design Prototype
Wet Winding Towpreg Winding

to burst pressure and cyclic loading. Furthermore, an even larger reduction in wall
thickness of 30% is recognized. The larger relative difference in wall thickness
is anticipated in the distinct laminate quality, which is impacted by the towpreg
material, the used fiber tension and the chosen stacking sequence definition. The
towpreg material allows for a predefined and constant resin content, which not only
translates into more uniform material properties, but also enables a certain tack of
the material. This allows for considerably higher fiber tensions during winding, as
the friction created by the material prevents fiber slippage. Lastly, the chosen stack-
ing sequence definition promotes a good consolidation level of the laminate due to
the placement of high-angle helical and circumferential layers as outermost plies.
To verify these differences in laminate quality, FVF and porosity are measured by
means of acid digestion tests. Therefore, for each laminate five samples were cut
out of the cylindrical section of an intact but pressurized vessel. Figure 7.8 shows
the comparison of FVF and porosity in the cylindrical section.
Resulting from the aforementioned differences in material and manufacturing,
the design prototype shows superior a laminate quality compared to the state-of-
the-art laminate. The higher consolidation level leads to relative increase in FVF by
3% and a relative decrease in porosity by 66%. Besides the noticeably smaller wall
thickness, which allows for more storage volume within the same design space, the
considerably higher laminate consolidation is anticipated to positively impact the
fatigue performance of the design prototype.
7.4 Concluding remarks 147

Fig. 7.8 Comparison of 65


FVF and porosity between FVF Porosity 6

Fiber Volume Fraction Vf [%]


state-of-the-art design and 60
design prototype 5
+3%

Porosity Vp [%]
55 4

-66% 3
50
2
45
1
57.1 4.5 59.0 1.5
40 0
State-of-the-Art Design Prototype
Wet Winding Towpreg Winding

7.4 Concluding remarks


Research question 4:
What considerations need to be made to transfer previously derived insights onto
a fullscale vessel geometry with distinct dome contour?

The layup design of a CPV is strongly linked to its dome shape. For the fullscale
vessel geometry, the mild dome slope necessitated a gradual change in tangential
stiffness from the cylinder to the boss neck. In contrast to the steep dome shape of
the subscale geometry, the fullscale geometry required multiple mid- to high-angle
helical layers to ensure a smooth change in tangential stiffness once the circumferen-
tial plies are dropped off. For the Layup E, which was composed of only low-angle
helical and circumferential layers, a large tangential expansion was recognized in
the transition from cylinder to dome. This further led to a preliminary failure in this
region. For the Layup F, which featured a variety of mid- and high-angle helical lay-
ers, a much lower tangential expansion in this area was noticed. This not only shifted
the failure location to the cylindrical section, but also yielded an increase in burst
pressure by 66% compared to Layup E while maintaining a similar CFRP mass.
The design prototype with the Layup F-C1 further provided a mass saving
potential of 17% compared to the state-of-the-art design, while fullfilling the cer-
tification requirements in regard to burst pressure and cyclic loading. Besides, the
prototype also showed a noticeably lower wall thickness and higher laminate qual-
ity, which was reasoned in the used towpreg material, higher fiber tensions and the
stacking sequence definition.
Design considerations to composite
pressure vessels 8

Gathered from the previous chapters related to the manufacturing, the modeling
and the testing of CPVs, this chapter is dedicated to the considerations related to
the design of CPVs. There is neither a claim on completeness of this chapter nor is
it intented to provide an optimum vessel design. Instead, it is aimed to shed light
on relevant phenomena that were encountered in this work, which shall provide
insights for future research. To this purpose, the last research question is answered
step by step by highlighting the most relevant aspects.

Research question 5:
What general design considerations can be drawn from the conducted structural
analysis?

Normal stress gradient through-the-thickness


The thick-wall condition of CPVs manifests itself in a normal stress gradient
through-the-thickness, where radial and tangential stress components are highest
on the inside and decrease towards the outside. This is of particular importance
for the layers whose alignment is closest to the tangential direction, such as cir-
cumferential layers, that exhibit the highest longitudinal stresses in the cylinder.
Therefore, the positioning and grouping of these layers has a profound impact on
the through-the-thickness stress gradient, as Fig. 8.1 shows.
Grouping circumferential layers together decreases the longitudinal stress gra-
dient between the innermost and outermost circumferential layer (e.g. Sequence
A and A-S2). Contrarily, separating circumferential layers into multiple groups,
which are distributed through-the-thickness, increases the stress gradient between
innermost and outermost layer (e.g. Sequence A-S1 and A-S3). The difference be-

Note: Parts of the content presented in this chapter have been previously published [82, 83]
or are derived from supervised student theses [86, 87, 97].

© The Editor(s) (if applicable) and The Author(s), under exclusive license to 149
Springer Fachmedien Wiesbaden GmbH, part of Springer Nature 2022
M. Nebe, In Situ Characterization Methodology for the Design and Analysis of
Composite Pressure Vessels, Werkstofftechnische Berichte j Reports of Materials
Science and Engineering, https://doi.org/10.1007/978-3-658-35797-9_8
150 8 Design considerations to composite pressure vessels

Longitudinal Stress σ11 [MPa]


2400

2200

2000

1800

1600
CFRP
A-S1 A-S3 A-S2 A
1400 0 1
0 0.2 0.4 0.6 0.8 1
Normalized Radial Coordinate r/ra [-]

Fig. 8.1 Distribution of longitudinal stress through-the-thickness at an internal pressure of


160 MPa for the Sequences A, A-S1, A-S2 and A-S3

tween the position of the circumferential stack is found marginal as the comparison
of Sequence A and A-S2 indicates.
Generally, decreasing the longitudinal stress gradient between innermost and
outermost plies will yield more performant vessel designs if failure is preserved
in the cylinder. Nonetheless, attention should be paid towards the load redistribu-
tion process that takes place whenever IFF is developing. Here, the existence of
multiple angle plies requires a further detailed analysis of the non-linear behavior,
whenever damage is present.

Laminate consolidation
The laminate consolidation level is determined by the stacking sequence of layers.
Once applied onto the mandrel, the fiber tension translates into radial compression
towards the subsequently wound layers. Here, the layer angle of incidence with
respect to the mandrel’s longitudinal axis determines the magnitude of applied
radial compression. For the cylindrical section this means that high-angle layers are
particularly effective at compacting the layers below, while low-angle layers do not
contribute noticeably to the laminate consolidation within the cylindrical section.
As a consequence, changes in the stacking sequence have a large repercussion on
the laminate quality through-the-thickness, as the comparison of CT cross-sections
in Fig. 8.2 shows. Locating high-angle layers towards the outside achieves high
FVF and low porosity (e.g. Sequence A), whilst the opposite is true whenever the
majority of high-angle layers are located on the laminate inside (e.g. Sequence
A-S2). Grouped sequences of circumferential layers were found to promote an
8 Design considerations to composite pressure vessels 151

Fig. 8.2 Comparison of CT


cross-section (longitudinal
view) of the Sequences A,
A-S1 and A-S2

5 mm

a) A b) A-S1 c) A-S2

increased compacting effect on the layers below. Yet, at the expense of a substantial
resin outflow, which accumulates at the layer boundary after the circumferential
stack.

Development of damage & crack propagation


When loaded with internal pressure, interfiber failure develops inevitably at already
mild pressure stages from 20 to 30 MPa throughout the composite. Between helical
and circumferential layers an intrinsic difference in crack propagation exists. While
the continous tow overlap within helical layers hampers crack growth in-plane
and through-the-thickness, the close unidirectional alignment of circumferential
layers allows cracks to grow in-plane and in case of multiple stacked circumfer-
ential layers, also through-the-thickness. The stacking of multiple circumferential
layers causes large fracture surfaces to be generated, whereas a frequent change
in layer orientation stops crack growth through-the-thickness. A comparison of
through-the-thickness crack growth is seen in Fig. 8.3, where the constant alter-
nation between circumferential and helical layer groups leads to much smaller
fracture surfaces for Sequence A-S1 as compared to the fractures seen in the cir-
cumferential stack of Sequence A. The implications of these differences are hard
to quantify at the time writing. Likely, continous crack propagation through-the-
thickness is not envisioned with respect to long term fatigue strength, which would
advocate for a continous alternation of winding angles as seen in Sequence A-
S1. On the other hand, the Layup F-C1 for the fullscale vessel geometry, which
exhibited a similar circumferential layer stack as seen for the Sequence A, sur-
passed 22,000 loading cycles and provided a relative decrease in burst pressure of
17% compared to the non-cycled average, which appears as a comparably minor
reduction of maximum strength. Nonetheless, to make a sound statement about
these implications, the long term behavior and the damage evolution process under
cyclic loading need to be further evaluated.
152 8 Design considerations to composite pressure vessels

Fig. 8.3 Comparison of


crack growth through-
the-thickness between
Sequence A and A-S1

p = 105 MPa 5 mm p = 105 MPa 5 mm

a) A b) A-S1

Concerning the development of IFF, a multitude of variables is influencing


the likelihood and sequence of events, which hinders the establishment of clear
guidelines with respect to CPV design. Yet, certain trends were observed and are
highlighted in the following. The positioning of directional stiffness through-the-
thickness also determines the load transfer in between the different layer types.
Locating a stack of low-angle helical layers close to the liner surface, as in Se-
quence A, requires the compliant helical layers to transfer the high tangential loads
from the inside to the circumferential layers on the laminate outside. This leads
to a large cylinder expansion and in turn, increases the likelihood of developing
damage within the layers that transfer the load to the circumferential stack. The
circumferential layer stack located above the low-angle helical layers leads to a
high degree of consolidation, which, as seen in material characterization campaign
in Sect. 5.1.1, has a substantial impact on the composite’s transverse and shear
mechanical properties. Contrary to this, as in the case of Sequence A-S2, the low-
angle helical layers, are not required to transfer the high tangential loads as they
are located above the circumferential layer stack which carries the loads closely to
the liner surface. This however, comes at the expense of a noticeably lower con-
solidation level within the low-angle helical layers due to the missing compaction,
which lowers their transverse and shear strength.
Finally, the radial stress component of the internal pressure appears to have
a noticeable impact on the development of IFF. Highest for the inner layers, the
radial compression lowers the transverse tensile stress state and therefore delays
the development of IFF. As the radial stress decreases to zero towards the outside,
the transverse stress state for the outer layer increases. Combined with the fact that
outer layers generally present a lower consolidation level than the inner ones leads
to the fact that IFF is likely to initiate from the outside towards the inside, which
was seen for the investigated sequences experimentally and numerically.
8 Design considerations to composite pressure vessels 153

Besides the development of intralaminar damage, interlaminar damage in the


form of a growing delamination was found within layer interfaces at the cylinder-
dome transition of Sequence A-S2. The subsequent drop-off of circumferential
layers leads to sudden changes in stiffness and the so-called free edge condition,
which is likely to further promote interlaminar stresses. Resulting from this, pre-
mature failure can be further amplified such as in the case of Sequence A-S2.
However, more efforts need to be made to gain a comprehensive understanding of
the collective effects that are present at these interfaces.

Behavior at cylinder-dome transition


The behavior at the cylinder-dome transition is impacted by the dome shape, the
individual expansion of dome and cylinder, the local distribution of stiffness at the
transition and the laminate’s response concerning bending and extensional-bending
coupling. As seen in Sect. 6.4.3, the strain distribution is sensitively affected by the
ply drop locations of circumferential layers, that represent sudden changes in stiff-
ness along the meridional coordinate. The severity of these ply drops however is
determined by the layup composition. The fewer circumferential layers are used,
the smaller the impact of the ply drop locations on the tangential stiffness distribu-
tion is. Contrarily, if circumferential layers represent a vast majority of the utilized
layers, the sensitivity of the tangential stiffness distribution on the ply drop loca-
tions increases correspondingly, which was seen for the Sequence A-S2.
The strain and displacement continuity at the transition triggers the existence
of meridional bending, which is influenced by the geometrical definition (e.g.
dome shape), the individual expansion of cylinder and dome, and the laminate’s
out-of-plane response. The choice of directional stiffness distribution through-the-
thickness has a profound impact on the cylinder expansion and therefore deter-
mines the magnitude of bending experienced at the transition. A prevailing cylinder
expansion, which is for example achieved by locating the majority of circumfer-
ential layers towards the laminate outside, is likely to contribute to an increased
bending at the transition. Whereas positioning the circumferential layers on the
laminate inside allows to contain the cylinder expansion, which leads to less bend-
ing at the transition. Additionally, the stacking sequence asymmetry introduces
meridional extensional-bending coupling, that either promotes or counteracts the
geometrical induced bending. This effect is difficult to quantify without further
detailed numerical analysis. A comparison of the distinct responses at the cylinder-
dome transition is seen in Fig. 8.4, where Layup B-2 exhibits the inverted sequence
of Layup B whilst the circumferential ply drop locations are virtually the same.
The larger tangential expansion of the Layup B yields an increase in merid-
ional strain prior to the geometrical cylinder-dome transition, which indicates an
154 8 Design considerations to composite pressure vessels

Layup B Meridional Strain Contour


Layup B-S2 Tangential Strain SD
1.5
p = 105 MPa Nominal liner geometry B B-S2
100
Measured contours

Radial Coordinate r [mm]


1
Strain ε [%]

0.5
50

0
100 150 200
Axial Coordinate z [mm]

Fig. 8.4 Strain components along the meridional surface path at an internal pressure of
105 MPa for the Layups B and B-S2

inward bending cylinder end. Contrarily, the noticeably lower tangential expan-
sion of Layup B-S2 leads to an almost constant meridional strain from cylinder
to dome. At the time writing, the implications of this difference are not fully un-
derstood. Likely, an increased bending at the transition is not envisioned, which
would argue for equalizing the expansion of dome and cylinder and thus reduce
the meridional bending.

Liner deformation
Besides the different phenomena related to the mechanical response under internal
pressure loading, a consideration that needs to be taken into account with respect
to the manufacturing of CPVs, is the plastical deformation of the liner. As seen in
Sect. 4.1.2, local denting arises whenever the applied compression resulting from
the tensioned fibers is too excessive and the liner deforms plastically. Besides fiber
tension, internal pressure and liner rigidity, the sequence of layers also impacts the
mandrel’s deformation pattern. In Sect. 6.3.2, differences in the circumferential
ply drop locations were noticed between the Sequences A, A-S1 and A-S2 which
resulted from the fact, that the liner geometry shrank differently in the axial direc-
tion depending on the stacking sequence. Particularly for the first few layers, the
fiber tension directly translates into out-of-plane compression that is applied on
8 Design considerations to composite pressure vessels 155

Axial Liner Length l [mm]


465
Low-angle High-angle Circumferential layers
463 helical layers helical layers

461

459
Layup B
457
0 0.25 0.5 0.75 1
Normalized No. of Layers [-]

Fig. 8.5 Variation of axial liner length over normalized number of layers for the Layup B

the highly compliant liner mandrel. In this regard, grouping layers of similar an-
gles leads to a directional deformation. Figure 8.5 depicts the liner length variation
of Layup B during filament winding, where an axial shrinkage due to the applied
low-angle helical layers is noticed.
The force equilibrium during filament winding is dependent on the internal
mandrel pressure and the applied fiber tension. Because of its high compliance,
the liner can only be subjected to low internal pressures prior to the winding. To
maintain an equilibrium at every position of the mandrel, an early alternation of
the layer types (circumferential/low-angle helical layer) appears as more process
stable, because of the reinforcement it would provide in radial and axial direction,
respectively. This would then allow to increase the internal pressure step-by-step
accordingly to the used fiber tension. To further analyze the resultant stresses
within the liner wall, a detailed FE analysis can provide additional insights, as
recently shown by Zhang et al. [50].

Dome contour & layup composition


The most influential factor for the choice of winding angles is the contour of the
dome. The dome slope determines the rate at which the loading component of
the internal pressure changes from radial to axial dominated. Protruding shapes
appear, whenever the directional stiffness of the laminate does not match the re-
sultant forces, which was seen in the case of Layup E on the fullscale geometry.
A gradual decrease in slope requires high-angle helical layers to contribute to the
tangential stiffness distribution once the circumferential layers are dropped off.
Contrarily, for steep dome slopes the loading component changes rapidly, which
requires the directional stiffness of the laminate to change accordingly. As such,
the layup composition varies depending on the dome shape.
156 8 Design considerations to composite pressure vessels

Whilst certain phenomena related to the dome layup were highlighted in this
work, much remains open to future research. The double curved surface, where
layer angle, wall thickness and material properties are constantly varying along
the meridional coordinate places a high sophistication to the task. Aside from the
experimental accessibility of the local strain distribution, which was established
in this work, more emphasis needs to be put towards a sound numerical depic-
tion of the resultant thickness build-up, layer trajectories, material properties and
damage evolution process in this region. This in turn, will permit to derive further
insights to make clear arguments about the structural efficiency of different layup
compositions.
Summary and outlook
9

Reminiscing on the state-of-the-art in terms of structural analysis related to com-


posite pressure vessels, apparent challenges are faced whenever the correlation
between experimental and numerical work is limited. The complexity of multi-
layered filament wound pressure vessels requires a synergy of both to derive an
in-depth knowledge of the vessel’s mechanical response under internal pressure
loading. Yet, existing experimental capabilities inevitably narrow the correlation
scales, where agreement between simulation and reality can be sought. Mostly con-
fined to the determination of final burst pressure and discrete strain information, the
extent to which the vessel’s mechanical response can be captured is limited.
This leads to the importance of the in situ characterization methodology estab-
lished in this work. The substantive framework, composed of different experimen-
tal measurement methods, permits to derive insights about the vessel’s meridional
thickness profile, its deformation behavior and the damage evolution process dur-
ing internal pressure loading. Scrutinized through the correlation to alternative
experimental methods and its application onto a multitude of vessels, the estab-
lished framework provided detailed insights into phenomena related to the testing
of CPVs. Amongst these, the accessibility of the cylinder-dome transition through
the use of DIC as well as the monitoring and localization of failure by means of
AAE were deemed key elements to gain a comprehensive understanding.
As a result, the variety of acquired experimental data permitted to establish
and validate a high fidelity FE model with implemented CDM approach. The cor-
relation scales, for which an alignment of experimental and numerical data was
achieved, included the vessel’s meridional thickness profile, its deformation be-
havior, the onset of interfiber failure and the final burst pressure. In this regard,
the accurate description of the vessel’s geometry as well as the precise determi-
nation of material properties stood out as crucial factors to reproduce the vessel’s
mechanical response reasonably well. The model further expanded the structural

© The Editor(s) (if applicable) and The Author(s), under exclusive license to 157
Springer Fachmedien Wiesbaden GmbH, part of Springer Nature 2022
M. Nebe, In Situ Characterization Methodology for the Design and Analysis of
Composite Pressure Vessels, Werkstofftechnische Berichte j Reports of Materials
Science and Engineering, https://doi.org/10.1007/978-3-658-35797-9_9
158 9 Summary and outlook

understanding by providing access to stress-strain data, which the experimental


methods could not provide.
Subsequently, the application of in situ characterization methodology and FE
model for the investigation of stacking sequence further allowed to unveil critical
aspects in regard to the design of CPVs. The stacking sequence strongly impacts
the consolidation level, the stress-strain distribution as well as the crack propa-
gation through-the-thickness. Yet, in the investigated experimental set, the most
significant effect was encountered in the cylinder-dome transition region. Here,
the variation of circumferential ply drop locations had a profound impact on the
deformation behavior and on the final strength of the selected sequences. In this
matter, the developed FE model demonstrated its capability of reproducing the
distinct mechanical responses of various stacking sequences and predicted their
final failure reasonably well. The consideration of damage was found of particular
importance for the prediction of failure located outside of the vessel’s cylindrical
section. Contrarily, failure preserved in the cylinder was already reasonably well
predicted by considering a constitutively elastic solution.
Ultimately, the transfer of the derived insights onto the fullscale vessel geometry
permitted to further expand the knowledge on CPV laminate design and exposed
the severe dependency of the layup composition to the chosen dome geometry. The
resulting design provided a noticeable material saving potential of 17% CFRP mass
in comparison to the state-of-the-art, while maintaining the certification require-
ments with regard to burst pressure and cyclic loading. The investigation moreover
highlighted the applicability of the established framework to capture these struc-
tural differences regardless of vessel geometry or size.
In essence, this work provides a sound foundation for future developments re-
lated to the design and analysis of composite pressure vessels as it unveils critical
aspects in regard to the modeling, the filament winding and the experimental char-
acterization of CPVs. Figure 9.1 outlines potential applications in future research.
First, the in situ characterization methodology can be further expanded with re-
gard to the analysis of airborne acoustic emissions. The unambigous identification
and interpretation of emissions at higher pressure stages will yield a deeper un-
derstanding of the progression of damage leading to final failure. Yet, it requires
a thorough understanding of the arising acoustic phenomena (e.g. dispersion, at-
tenuation, propagation). Once achieved, the combination of optic and acoustic
characterization methods could be further used for an end-of-line test for CPVs.
Second, an accurate numerical representation of the vessel’s mechanical re-
sponse was found to be strongly impacted by a sound definition of the geometry.
Thus, more emphasis needs to be put towards the realistic description of thickness
9 Summary and outlook 159

In situ characterization methodology


for the design and analysis of composite pressure vessels

Applications in future research

In situ characterization Computationally efficient and Design optimization of


methodology as end-of-line test accurate FE modeling composite pressure vessels

Principal aim

Reliable, safe and cost-efficient commercialization of composite pressure vessels

Fig. 9.1 Overview of potential applications of this work in future research

build-up, that in turn, will lead to more computationally-efficient and accurate FE


modeling strategies, where less experimental testing will be required.
Third, by evaluating other layup compositions and vessel geometries, the pre-
sented design considerations can be further extended. This will allow for more
generic conclusions about critical aspects and will lead to more optimized CPV
designs.
Finally, the perhaps simplest conclusion that can be drawn from this work is
that CPVs remain a highly complex issue and require tremendous effort to ensure a
thorough structural understanding. This however should always be the first priority
to ensure a reliable, safe and cost-efficient commercialization. Whilst this work
provided answers to existing questions, it more importantly initiates a small infinity
of new questions that eventually will contribute to further breakthroughs in CPV
design and analysis.
List of publications

Patent applications

 Nebe, M.; Braun, C.; Kasses, A.; Johman, A.


Segmentierte Ablegerolle zur Separation der Fadenzugkraft beim Faserwickel-
Prozess
 Steffens, D.; Nebe, M.; Obergfell, C.
Leichtbaubatteriegehäuse aus Faserverstärktem Kunststoff

Journal papers and proceedings

 Nebe, M.; Soriano, A.; Braun, C.; Middendorf, P.; Walther, F.


Analysis on the internal pressure loading of composite pressure vessels:
FE modeling and experimental correlation
Composites Part B: Engineering 212, 108550 (2021).
https://doi.org/10.1016/j.compositesb.2020.108550
 Nebe, M.; Asijee, T.; Braun, C.; van Campen, J.M.J.F.; Walther, F.
Experimental and analytical analysis on the stacking sequence of composite
pressure vessels
Composite Structures 247, 112429 (2020).
https://doi.org/10.1016/j.compstruct.2020.112429
 Nebe, M.; Maraite, D.; Braun, C.; Hülsbusch, D.; Walther, F.:
Experimental characterization of the structural deformation of type IV pressure
vessels subjected to internal pressure
Key Engineering Materials 809, ISSN 1662-9795 (2019) 47–52.
https://doi.org/10.4028/www.scientific.net/KEM.809.47

© The Editor(s) (if applicable) and The Author(s), under exclusive license to 161
Springer Fachmedien Wiesbaden GmbH, part of Springer Nature 2022
M. Nebe, In Situ Characterization Methodology for the Design and Analysis of
Composite Pressure Vessels, Werkstofftechnische Berichte j Reports of Materials
Science and Engineering, https://doi.org/10.1007/978-3-658-35797-9
162 List of publications

 Nebe, M.; Braun, C.; Gebhardt, T.; Hülsbusch, D.; Walther, F.:
Optische und akustische Analyse der Schadens- und Versagensmechanismen in
CFK-Druckbehältern
Werkstoffprüfung 2018 – Werkstoffe und Bauteile auf dem Prüfstand, Hrsg.:
G. Moninger, Stahleisen, ISBN 978-3-941269-99-6 (2018) 113–118.
 Nebe, M.; Braun, C.; Gebhardt, T.; Hülsbusch, D.; Walther, F.:
Combining optical and acoustical characterization methods to investigate dam-
age and failure mechanisms in composite pressure vessels
ICFC7, Proceedings of the 7th International Conference on Fatigue of Compos-
ites (2018) 1–8.

Presentations

 Nebe, M. (V.); Braun, C.; Hülsbusch, D.; Walther, F.


Filament winding and advanced experimental characterization of composite
pressure vessels
World of Filament Winding, Paris, France, 26.–27. November (2019)
 Nebe, M. (V.); Torres, A.; Braun, C.; Hülsbusch, D.; Walther, F.:
Experimental and numerical analysis of interfiber fracture in composite pres-
sure vessels subjected to internal pressure
MECHCOMP 2019, 5th International Conference on Mechanics of Compos-
ites, Lisboa, Portugal, 01.–04. July (2019)
 Nebe, M. (V.); Maraite, D.; Braun, C.; Hülsbusch, D.; Walther, F.:
Experimental characterization on the structural deformation of type IV pressure
vessels subjected to internal pressure
22. Symposium Verbundwerkstoffe und Werkstoffverbunde, Kaiserslautern,
Germany, 26.–28. Juni (2019)
 Nebe, M. (V.); Braun, C.; Gebhardt, T.; Hülsbusch, D.; Walther, F.:
Optische und akustische Analyse der Schadens- und Versagensmechanismen in
CFK-Druckbehältern
Tagung Werkstoffprüfung 2018, Bad Neuenahr, Germany, 06.–07. December
(2018)
 Nebe, M. (V.); Braun, C.; Hülsbusch, D.; Walther, F.:
Charakterisierung des Verformungs- und Schädigungsverhaltens von Typ IV-
Druckbehältern unter Zyklischer Belastung
DGM/DVM-AG Materialermüdung, Technische Universität Dortmund, Dort-
mund, Germany, 25.–26. October. (2018)
List of publications 163

 Nebe, M. (V.); Braun, C.; Gebhardt, T.; Hülsbusch, D.; Walther, F.:
Combining optical and acoustical characterization methods to investigate dam-
age and failure mechanisms in composite pressure vessels
ICFC7, 7th International Conference on Fatigue of Composites, Vicenza, Italy,
04.–06. July (2018)
Supervised student theses

In the subject area of the present work, the author supervised the following student
theses:

 Maraite, D.: Dreidimensionale Verformungsanalyse von Hochdruckbehältern


aus kohlenstofffaserverstärktem Kunststoff mittels optischer Messverfahren,
Diploma Thesis, TU Dresden University, Institute of Textile Machinery and
High Performance Material Technology, Supervisors: Dr.-Ing. Andreas Nocke,
Dipl.-Ing. Reimar Unger, 2019.
 Torres Guijarro, A.I.: Experimental and analytical determination of interfiber
fracture mechanisms and patterns in type IV composite pressure vessels, Mas-
ter Thesis, Delft University of Technology, Faculty of Aerospace Engineering,
Supervisor: Dr. ir. Julien Van Campen, 2019.
 Asijee, T.: Experimental and numerical investigation into the influence of layup
sequence on the mechanical performance of composite pressure vessels, Mas-
ter Thesis, Delft University of Technology, Faculty of Aerospace Engineering,
Supervisor: Dr. ir. Julien Van Campen, 2020.
 Cesari, E.: Combined acoustic and optic characterization of damage mecha-
nisms in internally pressurized composite pressure vessels, Master Thesis, Delft
University of Technology, Faculty of Aerospace Engineering, Supervisor: Dr. ir.
Julien Van Campen, 2020.
 Soriano Sutil, A.: Analysis strategies for as-manufactured composite pressure
vessels, Master Thesis, Delft University of Technology, Faculty of Aerospace
Engineering, Supervisor: Dr. ir. Julien Van Campen, 2020.
 Johman, A.: Development of a computationally efficient modelling framework
for type IV pressure vessels, Master Thesis, Delft University of Technology,
Faculty of Aerospace Engineering, Supervisor: Dr. ir. Julien Van Campen, 2021.

165
Curriculum vitae

Personal

Name: Martin Nebe


Date of birth: 26.11.1992
Place of birth: Quedlinburg
Nationality: German

Education

11.2017– Dr.-Ing. Mechanical Engineering


TU Dortmund University
03.2015–08.2015 Exchange, Mechanical Engineering
University of Califoria Santa Barbara
10.2011–04.2017 Dipl.-Ing. Mechanical Engineering
TU Dresden University

Professional career

11.2020– Process development engineer, cellcentric GmbH & Co. KG


11.2017–10.2020 PhD student, Mercedes-Benz AG
10.2016–04.2017 Graduate, BMW AG
09.2015–09.2016 Research assistant, TU Dresden University
09.2014–03.2015 Intern, Audi AG
11.2013–08.2014 Research assistant, Fraunhofer IWU

167
Published volumes

Volume 6 Hülsbusch, D.: Charakterisierung des temperaturabhängigen Ermü-


dungs- und Schädigungsverhaltens von glasfaserverstärktem Polyur-
ethan und Epoxid im LCF- bis VHCF-Bereich. Dissertation, Technische
Universität Dortmund, Springer Vieweg Verlag, Wiesbaden (2021).
https://doi.org/10.1007/978-3-658-34643-0
Volume 5 Schmiedt-Kalenborn, A.: Mikrostrukturbasierte Charakterisierung des
Ermüdungs- und Korrosionsermüdungsverhaltens von Lötverbindungen
des Austenits X2CrNi18-9 mit Nickel- und Goldbasislot. Dissertation,
Technische Universität Dortmund, Springer Vieweg Verlag, Wiesbaden
(2020). https://doi.org/10.1007/978-3-658-30105-7
Volume 4 Wittke, P.: Charakterisierung spanlos gefertigter Innengewinde in Alu-
minium- und Magnesium-Leichtbauwerkstoffen. Dissertation, Techni-
sche Universität Dortmund, Springer Vieweg Verlag, Wiesbaden (2020).
https://doi.org/10.1007/978-3-658-27943-1
Volume 3 Schmack, T.: Entwicklung einer ganzheitlichen Methode zur Bestim-
mung des dehnratenabhängigen Verhaltens faserverstärkter Kunststoffe.
Dissertation, Technische Universität Dortmund, Springer Vieweg Ver-
lag, Wiesbaden, ISBN 978-3-658-26931-9 (2019). https://doi.org/10.
1007/978-3-658-26931-9
Volume 2 Klein, M.: Mikrostrukturbasierte Bewertung des Korrosionsermü-
dungsverhaltens der Magnesiumlegierungen DieMag422 und AE42.
Dissertation, Technische Universität Dortmund, Springer Vieweg Ver-
lag, Wiesbaden, ISBN 978-3-658-25310-3 (2019). https://doi.org/10.
1007/978-3-658-25310-3
Volume 1 Siddique, S.: Reliability of Selective Laser Melted AlSi12 Alloy
for Quasistatic and Fatigue Applications. Dissertation, Technische
Universität Dortmund, Springer Vieweg Verlag, Wiesbaden, ISBN 978-
3-65823425-6 (2019). https://doi.org/10.1007/978-3-658-23425-6
169
References

1. Global technical regulation No. 13 (GTR13) of United Nations: Hydrogen and fuel cell
vehicle safety (2013)
2. Economic Commission for Europe of the United Nations (UN/ECE): Regulation No
134 – Uniform provisions concerning the approval of motor vehicles and their com-
ponents with regard to the safety-related performance of hydrogen fuelled vehicles
(HFCV) (2015)
3. Daimler AG: https://media.daimler.com/marsmediasite/ko/de/29181457 (03.10.2018)
4. Fuel Cell Technologies Office, U.S. Department of Energy: Multi-year research, devel-
opment, and demonstration plan – section 3.3 hydrogen storage (2015)
5. Peters, S.T.: Composite filament winding. ASM International (2011)
6. Munro, M.: Review of manufacturing of fiber composite components by filament wind-
ing. Polymer Composites 9(5):352–359 (1988) https://doi.org/10.1002/pc.750090508
7. Shen, F.C.: A filament-wound structure technology overview. Materials Chemistry and
Physics 42(2):96–100 (1995)
8. Cohen, D.: Influence of filament winding parameters on composite vessel quality and
strength. Composites Part A: Applied Science and Manufacturing 28(12):1035–1047
(1997) https://doi.org/10.1016/S1359-835X(97)00073-0
9. Cai, Z.; Gutowski, T.; Allen, S.: Winding and consolidation analysis for cylindrical
composite structures. Journal of Composite Materials 26(9):1374–1399 (1992) https://
doi.org/10.1177/002199839202600908
10. Cohen, D.; Mantell, S.; Zhao, L.: The effect of fiber volume fraction on filament wound
composite pressure vessel strength. Composites Part B: Engineering 32(5):413–429
(2001) https://doi.org/10.1016/S1359-8368(01)00009-9
11. Christ, T.K.: Rechnerische und experimentelle Untersuchungen zum Versagensver-
halten CFK-umwickelter Kryo-Druckbehälter. PhD Thesis: Technical University of
Munich (2017)
12. Fukunaga, H.; Uemura, M.: Optimum design of helically wound composite pres-
sure vessels. Composite Structures 1(1):31–49 (1983) https://doi.org/10.1016/0263-
8223(83)90015-6
13. Liang, C.C.; Chen, H.W.; Wang, C.H.: Optimum design of dome contour for filament-
wound composite pressure vessels based on a shape factor. Composite Structures
58(4):469–482 (2002) https://doi.org/10.1016/S0263-8223(02)00136-8

171
172 References

14. Koussios, S.; Bergsma, O.K.; Beukers, A.: Filament winding. Part 1: determination of
the wound body related parameters. Composites Part A: Applied Science and Manu-
facturing 35(2):181–195 (2004) https://doi.org/10.1016/j.compositesa.2003.10.003
15. Zu, L.; Koussios, S.; Beukers, A.: Design of filament–wound domes based on
continuum theory and non-geodesic roving trajectories. Composites Part A: Ap-
plied Science and Manufacturing 41(9):1312–1320 (2010) https://doi.org/10.1016/j.
compositesa.2010.05.015
16. Koussios, S.; Bergsma, O.K.; Mitchell, G.: Non-geodesic filament winding on generic
shells of revolution. Proceedings of the Institution of Mechanical Engineers, Part L:
Journal of Materials: Design and Applications 219(1):25–35 (2005) https://doi.org/10.
1243/146442005X10184
17. Zu, L.; Xu, H.; Wang, H.; Zhang, B.; Zi, B.: Design and analysis of filament-wound
composite pressure vessels based on non-geodesic winding. Composite Structures
207:41–52 (2019) https://doi.org/10.1016/j.compstruct.2018.09.007
18. Leh, D.; Saffré, P.; Francescato, P.; Arrieux, R.: Multi-sequence dome lay-up simula-
tions for hydrogen hyper-bar composite pressure vessels. Composites Part A: Applied
Science and Manufacturing 52:106–117 (2013) https://doi.org/10.1016/j.compositesa.
2013.05.007
19. De Carvalho, J.; Lossie, M.; Vandepitte, D.; Van Brussel, H.: Optimization of filament-
wound parts based on non-geodesic winding. Composites Manufacturing 6(2):79–84
(1995) https://doi.org/10.1016/0956-7143(95)99647-B
20. Trench, W.F.: Elementary differential equations with boundary value problems. Math-
ematics Series: Brooks/Cole-Thomson Learning (2001)
21. Wang, R.; Jiao, W.; Liu, W.; Yang, F.; He, X.: Slippage coefficient measurement for
non-geodesic filament-winding process. Composites Part A: Applied Science and Man-
ufacturing 42(3):303–309 (2011) https://doi.org/10.1016/j.compositesa.2010.12.002
22. Park, J.S.; Hong, C.S.; Kim, C.G.; Kim, C.U.: Analysis of filament wound com-
posite structures considering the change of winding angles through the thickness
direction. Composite Structures 55(1):63–71 (2002) https://doi.org/10.1016/S0263-
8223(01)00137-4
23. Krikanov, A.A.: Refined thickness of filament wound shells. Science and Engineering
of Composite Materials 10(4):241–248 (2002)
24. Vasiliev, V.V.; Krikanov, A.A.; Razin, A.F.: New generation of filament-wound
composite pressure vessels for commercial applications. Composite Structures 62(3-
4):449–459 (2003) https://doi.org/10.1016/j.compstruct.2003.09.019
25. Wang, R.; Jiao, W.; Liu, W.; Yang, F.: A new method for predicting dome thick-
ness of composite pressure vessels. Journal of Reinforced Plastics and Composites
29(22):3345–3352 (2010) https://doi.org/10.1177/0731684410376330
26. Wang, R.; Jiao, W.; Liu, W.; Yang, F.: Dome thickness prediction of composite pressure
vessels by a cubic spline function and finite element analysis. Polymers and Polymer
Composites 19(2–3):227–234 (2011)
27. Zu, L.; Xu, H.; Wang, H.; Zhang, B.; Zi, B.: Design and analysis of filament-wound
composite pressure vessels based on non-geodesic winding. Composite Structures
207:41–52 (2019) https://doi.org/10.1016/j.compstruct.2018.09.007
References 173

28. Roylance, D.K.: Netting analysis for filament-wound pressure vessels. Tech. Rep.
AMMRC TN 76-3: Composites Division Army Materials and Mechanics Research
Center (1976)
29. Koussios, S.; Beukers, A.; Bergsma, O.K.; Bersee, H.E.N.: Isotensoid related compos-
ite structures. In: Technical Conference of the American Society for Composites (2006)
30. De Jong, T.: A theory of filament wound pressure vessels. PhD Thesis. Faculty of
Aerospace Engineering, Delft University of Technology. Delft (1983)
31. Tew, B.W.: Preliminary design of tubular composite structures using netting theory and
composite degradation factors. Journal of Pressure Vessel Technology 117:390–394
(1995) https://doi.org/10.1115/1.2842141
32. Zu, L.: Design and optimization of filament wound composite pressure vessels. PhD
Thesis. Faculty of Aerospace Engineering, Delft University of Technology. Delft
(2012)
33. Jones, R.M.: Mechanics of Composite Materials. CRC Press, Virginia (1999)
34. Hojjati, M.; Safavi Ardebili, V.; Hoa, S.V.: Design of domes for polymeric composite
pressure vessels. Composites Engineering 5(1):51 – 59 (1995) https://doi.org/10.1016/
0961-9526(95)93979-6
35. Madhavi, M.; Rao, K.V.J.; Rao, K.N.: Design and analysis of filament wound compos-
ite pressure vessel with integrated-end domes. Defence Science Journal 59(1):73–81
(2009) https://doi.org/10.14429/dsj.59.1488
36. Musthak, M.; Madar Valli, P.; Narayanarao, S.; Madhavi, M.: Prediction of structural
behavior of FRP pressure vessel by using shear deformation theories. Materials Today:
Proceedings 4(2):872–882 (2017) https://doi.org/10.1016/j.matpr.2017.01.098
37. Alcántar, V.; Aceves, S.M.; Ledesma, E.; Ledesma, S.; Aguilera, E.: Optimization of
Type 4 composite pressure vessels using genetic algorithms and simulated annealing.
International Journal of Hydrogen Energy 42(24):15,770–15,781 (2017) https://doi.
org/10.1016/j.ijhydene.2017.03.032
38. Parnas, L.; Katirci, N.: Design of fiber-reinforced composite pressure vessels under
various loading conditions. Composite Structures 58(1):83–95 (2000) https://doi.org/
10.1016/S0263-8223(02)00037-5
39. Leissa, A.W.; Chang, J.D.: Elastic deformation of thick, laminated compos-
ite shells. Composite Structures 35:153–170 (1996) https://doi.org/10.1016/0263-
8223(96)00028-1
40. Qatu, M.S.: Accurate equations for laminated composite deep thick shells. International
Journal of Solids and Structures 36:2917–2941 (1999) https://doi.org/10.1016/S0020-
7683(98)00134-6
41. Asadi, E.; Wang, W.; Qatu, M.S.: Static and vibration analyses of thick deep laminated
cylindrical shells using 3D and various shear deformation theories. Composite Struc-
tures 94(2):494–500 (2012) https://doi.org/10.1016/j.compstruct.2011.08.011
42. Xia, M.; Takayanagi, H.; Kemmochi, K.: Analysis of multi-layered filament-wound
composite pipes under internal pressure. Composite Structures 53(4):483–491 (2001)
https://doi.org/10.1016/S0263-8223(01)00061-7
43. Baličević, P.; Kozak, D.; Mrčela, T.: Strength of pressure vessels with ellipsoidal heads.
Strojniski Vestnik/Journal of Mechanical Engineering 54(10):685–692 (2008)
44. Leh, D.; Magneville, B.; Saffré, P.; Francescato, P.; Arrieux, R.; Villalonga, S.: Opti-
misation of 700 bar type IV hydrogen pressure vessel considering composite damage
174 References

and dome multi-sequencing. International Journal of Hydrogen Energy 40(38):13,215–


13,230 (2015) https://doi.org/10.1016/j.ijhydene.2015.06.156
45. Berro Ramirez, J.P.; Halm, D.; Grandidier, J.C.; Villalonga, S.; Nony, F.: 700 bar type
IV high pressure hydrogen storage vessel burst – Simulation and experimental valida-
tion. International Journal of Hydrogen Energy 40(38):13,183–13,192 (2015) https://
doi.org/10.1016/j.ijhydene.2015.05.126
46. Leh, D.; Saffré, P.; Francescato, P.; Arrieux, R.; Villalonga, S.: A progressive failure
analysis of a 700-bar type IV hydrogen composite pressure vessel. International Journal
of Hydrogen Energy 40(38):13,206–13,214 (2015) https://doi.org/10.1016/j.ijhydene.
2015.05.061
47. Roh, H.S.; Hua, T.Q.; Ahluwalia, R.K.: Optimization of carbon fiber usage in Type 4
hydrogen storage tanks for fuel cell automobiles. International Journal of Hydrogen
Energy 38(29):12,795–12,802 (2013) https://doi.org/10.1016/j.ijhydene.2013.07.016
48. Wang, L.; Zheng, C.; Luo, H.; Wei, S.; Wei, Z.: Continuum damage modeling and pro-
gressive failure analysis of carbon fiber/epoxy composite pressure vessel. Composite
Structures 134:475–482 (2015) https://doi.org/10.1016/j.compstruct.2015.08.107
49. Son, D.; Chang, S.H.: Evaluation of modeling techniques for a type III hydrogen
pressure vessel (70 MPa) made of an aluminum liner and a thick carbon/epoxy compos-
ite for fuel cell vehicles. International Journal of Hydrogen Energy 37(3):2353–2369
(2012) https://doi.org/10.1016/j.ijhydene.2011.11.001
50. Zhang, Q.; Xu, H.; Jia, X.; Zu, L.; Cheng, S.; Wang, H.: Design of a 70 mpa type iv
hydrogen storage vessel using accurate modeling techniques for dome thickness predic-
tion. Composite Structures 236:111,915 (2020) https://doi.org/10.1016/j.compstruct.
2020.111915
51. Liu, P.F.; Chu, J.K.; Hou, S.J.; Zheng, J.Y.: Micromechanical damage modeling
and multiscale progressive failure analysis of composite pressure vessel. Computa-
tional Materials Science 60:137–148 (2012) https://doi.org/10.1016/j.commatsci.2012.
03.015
52. Wang, L.; Zheng, C.; Wei, S.; Wei, Z.: Micromechanics-based progressive failure anal-
ysis of carbon fiber/epoxy composite vessel under combined internal pressure and
thermomechanical loading. Composites Part B: Engineering 89:77–84 (2016) https://
doi.org/10.1016/j.compositesb.2015.11.018
53. Gentilleau, B.; Villalonga, S.; Nony, F.; Galiano, H.: A probabilistic damage behav-
ior law for composite material dedicated to composite pressure vessel. International
Journal of Hydrogen Energy 40(38):13,193–13,205 (2015) https://doi.org/10.1016/j.
ijhydene.2015.04.043
54. Harada, S.; Arai, Y.; Araki, W.; Iijima, T.; Kurosawa, A.; Ohbuchi, T.; Sasaki, N.: A
simplified method for predicting burst pressure of type III filament-wound CFRP com-
posite vessels considering the inhomogeneity of fiber packing. Composite Structures
190:79–90 (2018) https://doi.org/10.1016/j.compstruct.2018.02.011
55. Rafiee, R.; Torabi, M.A.: Stochastic prediction of burst pressure in composite pres-
sure vessels. Composite Structures 185:573–583 (2018) https://doi.org/10.1016/j.
compstruct.2017.11.068
56. Yao, X.F.; Meng, L.B.; Jin, J.C.; Yeh, H.Y.: Full-field deformation measurement of fiber
composite pressure vessel using digital speckle correlation method. Polymer Testing
24(2):245–251 (2005) https://doi.org/10.1016/j.polymertesting.2004.05.009
References 175

57. Meng, L.B.; Jin, G.C.; Yao, X.F.; Yeh, H.Y.: 3D full-field deformation monitoring of
fiber composite pressure vessel using 3D digital speckle correlation method. Polymer
Testing (2006) https://doi.org/10.1016/j.polymertesting.2005.09.011
58. Hao, J.-C.; Leng, J.-S.; Wei, Z.: Non-destructive evaluation of composite pressure
vessel by using FBG sensors. Chinese Journal of Aeronautics 20(2):120–123 (2007)
https://doi.org/10.1016/S1000-9361(07)60017-X
59. Degrieck, J.; de Waele, W.; Verleysen, P.: Monitoring of fibre reinforced composites
with embedded optical fibre Bragg sensors, with application to filament wound pressure
vessels. NDT & E International 34(4):289–296 (2001) https://doi.org/10.1016/S0963-
8695(00)00069-4
60. Gasior, P.; Malesa, M.; Kaleta, J.; Kujawińska, M.; Malowany, K.; Rybczyński, R.: Ap-
plication of complementary optical methods for strain investigation in composite high
pressure vessel. Composite Structures 203:718–724 (2018) https://doi.org/10.1016/j.
compstruct.2018.07.060
61. Tapeinos, I.G.; Rajabzadeh, A.; Zarouchas, D.S.; Stief, M.; Groves, R.M.; Koussios,
S.; Benedictus, R.: Evaluation of the mechanical performance of a composite multi-cell
tank for cryogenic storage: Part II – Experimental assessment. International Journal of
Hydrogen Energy 44(7):3931–3943 (2019) https://doi.org/10.1016/j.ijhydene.2018.12.
063
62. Kang, D.H.; Kim, C.U.; Kim, C.G.: The embedment of fiber Bragg grating sensors
into filament wound pressure tanks considering multiplexing. NDT and E International
39(2):109–116 (2006) https://doi.org/10.1016/j.ndteint.2005.07.013
63. Sause, M.G.R.: In situ monitoring of fiber-reinforced composites: vol 242 (2016)
https://doi.org/10.1007/978-3-319-30954-5
64. Hamstad, M.A.: Thirty years of advances and some remaining challenges in the ap-
plication of acoustic emission to composite materials. Acoustic Emission Beyond the
Millennium, Elsevier Science, Amsterdam p 77–91 (2000)
65. Ono, K.; Gallego, A.: Research and applications of ae on advanced composites. Journal
of Acoustic Emission (30):180–229 (2012)
66. Saeedifar, M.; Zarouchas, D.: Damage characterization of laminated composites us-
ing acoustic emission: A review. Composites Part B: Engineering 195:108,039 (2020)
https://doi.org/10.1016/j.compositesb.2020.108039
67. ASTM E1067 / E1067M-18: Standard practice for acoustic emission examination of
fiberglass reinforced plastic resin (FRP) tanks/vessels. ASTM International, West Con-
shohocken, PA (2018)
68. Fowler, T.J.: Acoustic emission of fiber reinforced plastics. Journal of the Technical
Councils of ASCE 105(2):281–289 (1979)
69. Kaiser, J.: Untersuchungen über das Auftreten von Geräuschen beim Zugversuch. PhD
Thesis: Technische Hochschule München (1950)
70. Downs, K.S.; Hamstad, M.A.: Correlation of acoustic emission felicity ratios and hold-
based rate moments with burst strengths of spherical graphite/epoxy pressure vessels.
Journal of acoustic emission 13(3-4):45–55 (1995)
71. Downs, K.S.; Hamstad, M.A.: Acoustic emission from depressurization to detect/
evaluate significance of impact damage to graphite/epoxy pressure vessels. Journal of
Composite Materials (32):258–307 (1998)
176 References

72. Sause, M.G.R.; Schmitt, S.; Kalafat, S.: Failure load prediction for fiber-reinforced
composites based on acoustic emission. Composites Science and Technology 164:24–
33 (2018) https://doi.org/10.1016/j.compscitech.2018.04.033
73. Sause, M.G.R.; Schmitt, S.; Hoeck, B.; Monden, A.: Berstdruckvorhersage mittels
Schallemissionsanalyse. DACH-Jahrestagung 2019: Zerstörungsfreie Materialprüfung,
27.–29. Mai 2019, Friedrichshafen (2019)
74. Sause, M.G.R.; Schmitt, S.; Hoeck, B.; Monden, A.: Acoustic emission based predic-
tion of local stress exposure. Composites Science and Technology 173:90–98 (2019)
https://doi.org/10.1016/j.compscitech.2019.02.004
75. Kalafat, S.; Sause, M.G.R: Acoustic emission source localization by artificial neural
networks. Structural Health Monitoring 14(6):633–647 (2015) https://doi.org/10.1177/
1475921715607408
76. Chou, H.Y.; Mouritz, A.P.; Bannister, M.K.; Bunsell, A.R.: Acoustic emission analy-
sis of composite pressure vessels under constant and cyclic pressure. Composites Part
A: Applied Science and Manufacturing 70:111–120 (2015) https://doi.org/10.1016/j.
compositesa.2014.11.027
77. Gorman, M.R.: Modal AE analysis of fracture and failure in composite materials, and
the quality and life of high pressure vessels. J. Acoustic Emission 29:1–28, (2011)
78. Liao, B.B.; Wang, D.L.; Hamdi, M.; Zheng, J.Y.; Jiang, P.; Gu, C.H.; Hong, W.R.:
Acoustic emission-based damage characterization of 70 MPa type IV hydrogen com-
posite pressure vessels during hydraulic tests. International Journal of Hydrogen
Energy (2019) https://doi.org/10.1016/j.ijhydene.2019.02.217
79. Ellul, B.; Camilleri, D.: The influence of manufacturing variances on the progressive
failure of filament wound cylindrical pressure vessels. Composite Structures 133:853–
862 (2015) https://doi.org/10.1016/j.compstruct.2015.07.059
80. Kangal, S.; Kartav, O.; Tanoğlu, M.; Aktaş, E.; Artem, H.S.: Investigation of interlayer
hybridization effect on burst pressure performance of composite overwrapped pressure
vessels with load-sharing metallic liner. Journal of Composite Materials 54(7):961–980
(2020) https://doi.org/10.1177/0021998319870588
81. Nebe, M.; Braun, C.; Gebhardt, T.; Hülsbusch, D.; Walther, F.: Combining optical and
acoustical characterization methods to investigate damage and failure mechanisms in
composite pressure vessels. Proceedings of the 7th International Conference on Fatigue
of Composites (ICFC7) (2018)
82. Nebe, M.; Asijee, T.; Braun, C.; van Campen, J.M.J.F.; Walther, F.: Experimental and
analytical analysis on the stacking sequence of composite pressure vessels. Composite
Structures 247:112,429 (2020) https://doi.org/10.1016/j.compstruct.2020.112429
83. Nebe, M.; Soriano, A.; Braun, C.; Middendorf, P.; Walther, F.: Analysis on the internal
pressure loading of composite pressure vessels: FE modeling and experimental corre-
lation. Composites Part B: Engineering 212:108,550 (2021) https://doi.org/10.1016/j.
compositesb.2020.108550
84. Torres, A.I.: Experimental and analytical determination of interfiber fracture mecha-
nisms and patterns in type IV composite pressure vessels. MSc Thesis: Delft University
of Technology (2019)
85. Cesari, E.: Combined acoustic and optic characterization of damage mechanisms in
internally pressurized composite pressure vessels. MSc Thesis: Delft University of
Technology (2020)
References 177

86. Soriano Sutil, A.: Analysis strategies for as-manufactured composite pressure vessels.
MSc Thesis: Delft University of Technology (2020)
87. Johman, A.: Analytical and experimental investigation of cylinder-dome transitions in
relation to laminate properties of composite pressure vessels. MSc Thesis: Delft Uni-
versity of Technology (2021)
88. Zanuttigh, P.; Minto, L.; Marin, G.; Dominio, F.; Cortelazzo, G.: Time-of-flight and
structured light depth cameras: Technology and applications (2016) https://doi.org/10.
1007/978-3-319-30973-6
89. Zhang, S.: High-speed 3d shape measurement with structured light methods: A re-
view. Optics and Lasers in Engineering 106:119–131 (2018) https://doi.org/10.1016/j.
optlaseng.2018.02.017
90. Kazhdan, M.; Bolitho, M.; Hoppe, H.: Poisson surface reconstruction. Proceedings of
Eurographics Symposium on Geometry Processing (2006)
91. Döbler, D.; Heilmann, G.: Perspectives of the acoustic camera. Internoise – The 2005
Congress and Exposition on Noise Control Engineering (2005)
92. DIN EN ISO 527-4:1997: Plastics – Determination of tensile properties – Part 4: Test
conditions for isotropic and orthotropic fibre-reinforced plastic composites. ISO/TC
61/SC 13 Composites and reinforcement fibres (1997)
93. DIN EN ISO 14129:1997: Fibre-reinforced plastic composites – Determination of the
in-plane shear stress/shear strain response, including the in-plane shear modulus and
strength, by the plus or minus 45 degree tension test method. ISO/TC 61/SC 13 Com-
posites and reinforcement fibres (1997)
94. ASTM D5528-13: Standard test method for Mode I interlaminar fracture toughness of
unidirectional fiber-reinforced polymer matrix composites. ASTM International, West
Conshohocken, PA (2013)
95. ASTM D7905 / D7905M-19e1: Standard test method for determination of the Mode II
interlaminar fracture toughness of unidirectional fiber-reinforced polymer matrix com-
posites. ASTM International, West Conshohocken, PA (2014)
96. DIN EN 2564:2019-08: Aerospace series – Carbon fibre laminates - Determination of
the fibre, resin and void contents. Deutsches Institut für Normung, Berlin, Germany
(2019) https://doi.org/10.1520/D6671_D6671M-19
97. Asijee, T.: Experimental and numerical investigation into the influence of layup se-
quence on the mechanical performance of composite pressure vessels. MSc Thesis:
Delft University of Technology (2020)
98. Knops, M.: Analysis of failure in fiber polymer laminates: The theory of Alfred Puck.
Springer, Berlin (2008) https://doi.org/10.1007/978-3-540-75765-8
99. Leone F; Bergan A; Dávila C: CompDam – Deformation Gradient Decomposition
(DGD), v2.5.0 (2019). Https://github.com/nasa/CompDam_DGD
100. Leone, F.A.: Deformation gradient tensor decomposition for representing matrix cracks
in fiber-reinforced materials. Composites Part A: Applied Science and Manufacturing
76:334–341 (2015) https://doi.org/10.1016/j.compositesa.2015.06.014
101. Benzeggagh, M.L.; Kenane, M.: Measurement of mixed-mode delamination fracture
toughness of unidirectional glass/epoxy composites with mixed-mode bending appa-
ratus. Composites Science and Technology 56(4):439–449 (1996) https://doi.org/10.
1016/0266-3538(96)00005-X
178 References

102. Turon, A.; Camanho, P.P.; Costa, J.; Dávila, C.G.: A damage model for the simulation
of delamination in advanced composites under variable-mode loading. Mechanics of
Materials 38(11):1072–1089 (2006)
103. Turon, A.; Camanho, P.P.; Costa, J.; Renart, J.: Accurate simulation of delamination
growth under mixed-mode loading using cohesive elements: Definition of interlaminar
strengths and elastic stiffness. Composite Structures 92(8):1857–1864 (2010)
104. Camanho, P.P.; Bessa, M.A.; Catalanotti, G.; Vogler, M.; Rolfes, R.: Modeling the in-
elastic deformation and fracture of polymer composites-Part II: Smeared crack model.
Mechanics of Materials 59:36–49 (2013) https://doi.org/10.1016/j.mechmat.2012.12.
001
105. Maimí, P.; Camanho, P.P.; Mayugo, J.A.; Dávila, C.G.: A continuum damage model for
composite laminates: Part I – Constitutive model. Mechanics of Materials 39(10):897–
908 (2007) https://doi.org/10.1016/j.mechmat.2007.03.005
106. Maimí, P.; Camanho, P.P.; Mayugo, J.A.; Dávila, C.G.: A continuum damage model
for composite laminates: Part II – Computational implementation and validation. Me-
chanics of Materials 39(10):909–919 (2007) https://doi.org/10.1016/j.mechmat.2007.
03.006
107. Dávila, C.G.; Rose, C.A.; Camanho, P.P.: A procedure for superposing linear cohesive
laws to represent multiple damage mechanisms in the fracture of composites. Interna-
tional Journal of Fracture 158(2):211–223 (2009) https://doi.org/10.1007/s10704-009-
9366-z
108. Bažant, Z.P.; Oh, B.H.: Crack band theory for fracture of concrete. Matériaux et Con-
structions 16:155–177 (1983) https://doi.org/10.1007/BF02486267
109. Nebe, M.; Braun, C.; Gebhardt, T.; Hülsbusch, D.; Walther, F.: Optische und akustische
Analyse der Schadens- und Versagensmechanismen in CFK-Druckbehältern. Werk-
stoffprüfung 2018 – Werkstoffe und Bauteile auf dem Prüfstand, Hrsg.: G. Moninger,
Stahleisen (2018)
110. Nebe, M.; Maraite, D.; Braun, C.; Hülsbusch, D.; Walther, F.: Experimental character-
ization of the structural deformation of type IV pressure vessels subjected to internal
pressure. Key Engineering Materials 809:47–52 (2019) https://doi.org/10.4028/www.
scientific.net/KEM.809.47
111. Maraite, D.: Dreidimensionale Verformungsanalyse von Hochdruckbehältern aus
kohlenstofffaserverstärktem Kunststoff mittels optischer Messverfahren. Diploma The-
sis: TU Dresden University (2019)
112. Wang, R.; Jiao, W.; Liu, W.; Yang, F.: A new method for predicting dome thick-
ness of composite pressure vessels. Journal of Reinforced Plastics and Composites
29(22):3345–3352 (2010) https://doi.org/10.1177/0731684410376330
113. Wang, R.; Jiao, W.; Liu, W.; Yang, F.: Dome thickness prediction of composite pressure
vessels by a cubic spline function and finite element analysis. Polymers and Polymer
Composites 19(2-3):227–234 (2018) https://doi.org/10.1177/0967391111019002-327
114. Hoffmann, K.: An introduction to measurements using strain gages. Hottinger Baldwin
Messtechnik GmbH, Darmstadt, Germany (1987)
115. Sutton, M.A.; Orteu, J.; Schreier, H.: Image correlation for shape, motion and de-
formation measurements: Basic concepts,theory and applications: 1st edn. Springer
Publishing Company, Incorporated (2009)
References 179

116. Prosser, W.H.: Advanced AE techniques in composite materials research. Journal of


Acoustic Emission 14:1–11 (1996)
117. Camanho, P.P.; Dávila, C.G.; Pinho, S.T.; Iannucci, L.; Robinson, P.: Prediction of in
situ strengths and matrix cracking in composites under transverse tension and in-plane
shear. Composites Part A: Applied Science and Manufacturing 37(2):165–176 (2006)
https://doi.org/10.1016/j.compositesa.2005.04.023
118. Talreja, R.; Veer Singh, C.: Damage and failure of composite materials. Cambridge
University Press (2012)
119. Kellas, S.; Morton, J.: Scaling effects in angle-ply laminates. Tech. Rep. NASA Con-
tractor Report 4423: NASA Langley Research Center (1992)
120. MIL-HDBK-5J: Metallic materials and elements for aerospace vehicle structures. Tech.
rep.: US Department of Defense (2003)
121. Catalanotti, G.; Camanho, P.P.; Xavier, J.; Dávila, C.G.; Marques, A.T.: Measure-
ment of resistance curves in the longitudinal failure of composites using digital
image correlation. Composites Science and Technology (2010) https://doi.org/10.1016/
j.compscitech.2010.07.022
122. Camanho, P.P.; Maimí, P.; Dávila, C.G.: Prediction of size effects in notched lam-
inates using continuum damage mechanics. Composites Science and Technology
67(13):2715–2727 (2007)
123. Shim, D.; Lagacé, P.A.: An analytical method for interlaminar stresses due to global
effects of ply drop-offs. Mechanics of Advanced Materials and Structures 12(1):21–32
(2005) https://doi.org/10.1080/15376490490491927

Das könnte Ihnen auch gefallen