Download as pdf or txt
Download as pdf or txt
You are on page 1of 71

KTH ROYAL INSTITUTE

OF TECHNOLOGY

Doctoral Thesis in Solid Mechanics

Impact of paperboard deformation


and damage mechanisms on
packaging performance
GUSTAV MARIN

Stockholm, Sweden 2023


Impact of paperboard deformation
and damage mechanisms on
packaging performance
GUSTAV MARIN

Academic Dissertation which, with due permission of the KTH Royal Institute of Technology,
is submitted for public defence for the Degree of Doctor of Philosophy on Monday the 20th
March 2023, at 13:00 in F3, Lindstedtsvägen 26, Stockholm.

Doctoral Thesis in Solid Mechanics


KTH Royal Institute of Technology
Stockholm, Sweden 2023
© Gustav Marin
© Mikael Nygårds: Paper A, B, C, D, E
© Sören Östlund: Paper A, B, C, D, E
© Prashanth Srinivasa: Paper C
© Anton Hagman: Paper E

ISBN 978-91-8040-488-4
TRITA-SCI-FOU 2023:05

Printed by: Universitetsservice US-AB, Sweden 2023


Det är en massa som jobbar med massa
De kör flis för att tjäna flis
Papper som jobbar med papper
Allt under skorstenens dis

Kalle Baah – Billie Rude Boy

Perfect is the enemy of good

– Italian proverb
This page intentionally left blank
Abstract

Paper-based materials, such as paperboard, are commonly used as packaging


materials. In addition to the fact that paper is renewable, there are also many other
benefits of paperboard. From a mechanical point of view, paperboard has a high
bending stiffness compared to its relatively low weight and high foldability, which
are properties of significance in the design of packages. However, a distinct
drawback of paperboard is its significant sensitivity to moisture. The moisture
reduces the mechanical properties of the paperboard and consequently reduces
the performance of the package. This thesis addresses the impact of paperboard
deformation and damage mechanisms on packaging performance, with the
characterization of the material properties as a starting point. Initially, the
relations between moisture and different mechanical properties on a continuum
material level were investigated. Then, experimental testing and finite element
(FE) simulations were applied to evaluate these relations at the packaging design
level.

In Paper A, a material characterization was performed on five commercial


paperboards with different basis weights, from the same producer. Five types of
mechanical tests to characterize the paperboards material properties were
performed:

• In-plane tensile test,


• Out-of-plane tensile test,
• Short-span Compression Test (SCT),
• Two-point folding,
• Double-notch shear test.

All tests were performed at several levels of relative humidity (RH). Linear
relations between the mechanical properties normalized with their respective
value at 50 % RH and moisture ratio were found.

Paper B examined whether the linear relationships discovered in Paper A are


also valid for other paperboard series. Therefore, this study investigated 15
paperboards from four producers at the same RH levels as in Paper A. The
paperboards were chosen to be different in furnish and construction, where four
recycled boards were included. Here, the in-plane stiffnesses, strengths and SCT
values were evaluated as a function of moisture. When the investigated
paperboards’ moisture ratios were also normalized, all paperboards followed a
linear master curve between normalized mechanical property and normalized
moisture ratio. Additionally, a bilinear elastic-plastic in‑plane model was
developed to predict the stress-strain relation of an arbitrary paperboard at an
arbitrary moisture level without requiring mechanical testing except at standard
conditions (50% RH, 23 °C).

In Paper C, the master curve developed in Paper B was used to estimate material
input parameters for simulating a Box Compression Test (BCT) at different
moisture levels by using an orthotropic material model with a stress-based failure
criterion, i.e., a relatively simple material model with few input parameters. The
result showed that it was possible to accurately predict the load-compression
curve of a BCT when accounting for moisture. Furthermore, it was concluded
that modeling the creases’ mechanical properties is vital for capturing the stiffness
response of the package. Here, a measurable approach for reducing the creases’
mechanical properties was suggested, based on a folding test to obtain the relative
creasing strength (RCS) and a short-span tensile test to obtain the relative tensile
strength (RTS). It should be emphasized that the model does not include any
fitting parameters. All input data is based on measured values. Due to the
importance of creases, the RCS and RTS ratios were investigated further in
Paper D. When evaluated against normative shear strength during creasing, the
RCS and RTS values together formed a creasing window, where the RTS values
corresponded to in-plane cracks (upper limit) and the RCS values corresponded
to delamination damage (lower limit). It was observed that both the lower and
upper limits exhibit linear relations as functions of shear strain.

Since creases have an evident effect on the packaging performance from a


stacking point of view, it was interesting to investigate a load case exposing the
package to shear. Therefore, an additional load case was investigated in Paper E:
torsion of paperboard packages, where the experimental data was accurately
predicted. Additionally, the effect of bending stiffness was investigated by
developing two FE models. Model 1 (used in Paper C) treated the paperboard as
a homogeneous material, and Model 2 considered the paperboard a three-ply
laminate structure. No significant effect was noted, and it was concluded that the
strength has a more significant effect on the BCT than the bending stiffness. It
should also be mentioned that there were no problems with cracks when the
paperboards were creased and mounted to packages used in Papers C and E.
This correlates to the creasing window developed in Paper D since the creasing
depth used for the packages is located within the creasing window.

To conclude, the primary procedure in this thesis is developing an easy-to-use


model with few material parameters that demonstrably can predict the load-
deformation curves for two different load cases. The purpose of the model is not
to be used for the precise prediction of failure loads but to gain knowledge about
damage mechanisms during the testing procedures. A clear advantage of this
approach is that the model can be used to either change the package’s geometry
or perform a parametric study on the ingoing material parameters. This can also
be varied for each ply separately, which helps converters and paperboard
producers. It has also been shown that the model can account for different
moisture levels if the master curve developed within this thesis is applied. Finally,
it should be emphasized again that the model does not include any fitting
parameters. All input data is based on measured values.
Sammanfattning

Kartong är ett exempel på ett vanligt, pappersbaserat, förpackningsmaterial.


Utöver från att det är återvinningsbart så finns det även andra fördelar med att
använda kartong i förpackningar. Ur ett mekaniskt perspektiv har exempelvis
kartong hög böjstyvhet i förhållande till sin vikt samt är enkelt att vika, vilket är
två betydelsefulla egenskaper vid förpackningsdesign. En uppenbar nackdel med
kartong däremot, är dess känslighet för fukt, som försämrar de mekaniska
egenskaperna hos kartongmaterialet och således även förpackningsprestandan.
Med materialkaraktärisering som utgångspunkt, tar den här avhandlingen upp
deformationer och skademekanismers påverkan på kartongförpackningars
prestanda. Initialt avhandlas relationen mellan fukt och olika mekaniska
egenskaper hos kartongmaterialet på kontinuumnivå. Därefter lyfts dessa
relationer upp på förpackningsnivå, genom experimentell förpackningsprovning
och datorsimuleringar med finita elementmetoden (FEM).

I Artikel A genomfördes en materialkaraktärisering på fem kommersiella


kartonger från samma producent, men med olika ytvikt. Följande provmetoder
användes vid materialkaraktäriseringen:

• Dragprov (i planet),
• Dragprov (ut ur planet),
• Korta kompressionsprov (SCT),
• Böjstyvhetsprov,
• Skjuvprofilsprov.

Samtliga prov utfördes vid flera olika nivåer av relativ fuktighet (RH). Linjära
relationer mellan mekanisk egenskap normerad med motsvarande värde vid
50 % RH och fukt noterades.

Artikel B utvärderade huruvida de linjära samband som noterades i Artikel A


gäller för andra kartongserier. Således användes totalt 15 kartongkvalitéer från
fyra olika tillverkare. Provningen skedde under samma förhållanden som i
Artikel A. Kartongerna valdes noga ut för att säkerställa att de hade olika
konstruktion och innehåll. Exempelvis innehöll fyra av kartongerna endast
återvunna fibrer. Artikel B avgränsades till att endast utvärdera dragprov i planet,
samt SCT som funktion av fukt. Till skillnad från Artikel A så normerades även
fuktkvoterna i diagrammet, vilket resulterade i att alla kartonger, oberoende av
tillverkare, sammanföll kring samma linjära masterkurva som beskriver
sambandet mellan normerad mekanisk egenskap och normerad fuktkvot. Utöver
detta utvecklades en bilinjär elastisk-plastisk i-planet-modell som kan förutsäga
spännings-töjningskurvor för en godtycklig kartong vid valfri fukthalt utan att
genomföra några mekaniska test utöver dragprov vid standardklimat (50% RH,
23 °C).

I Artikel C utnyttjades det linjära sambandet mellan mekanisk egenskap och


fuktkvot till att prediktera materialegenskaperna som användes till
ingångsparametrar vid simuleringar av boxkompressionsprovning (BCT) vid olika
fuktnivåer. Simuleringarna baserades på en ortotropisk materialmodell med ett
kollapskriterium, således en relativt enkel materialmodell med få
materialparametrar. Simuleringarna jämfördes med experimentell provning och
visade sig kunna prediktera experimentella resultat vid olika fukthalter bra. Utöver
detta drogs slutsatsen att bigarnas materialegenskaper är vitala för att simulera
förpackningens styvhetsrespons. För att modellera detta användes ett mätbart
tillvägagångssätt baserat på vikprover för att bestämma relativ bigstyrka (RCS)
och korta dragprover för att utvärdera bigens relativa dragstyrka (RTS). Alla
materialparametrar baserades på mätbara data, och ingen kurvanpassning gjordes.
Till följd av bigarnas stora påverkan undersöktes reduktionskvoterna, RCS och
RTS, ytterligare i Artikel D. När kvoterna utvärderas mot normativ skjuvtöjning
bildar de tillsammans ett bigfönster där RTS utgör övre gränsen och RCS undre
gränsen i fönstret. Det noterades att båda dessa gränser betedde sig linjärt.

På grund av bigarnas betydande påverkan på förpackningsprestandan var det av


yttersta intresse att utvärdera ett lastfall där förpackningen utsattes för skjuvning.
Ett sådant introducerades i Artikel E: vridning av lådor, i vilken simuleringar
överensstämde väl med fysiska experiment. Dessutom utvärderades effekten av
böjstyvhet genom att två olika FE-modeller, där Modell 1 (tillämpad i Artikel C)
simulerade kartongen som ett homogent material. I Modell 2 modellerades
kartongen som en treskiktsstruktur. Trots olikheterna i uppbyggnad märktes
ingen signifikant effekt på slutresultatet. Istället drogs slutsatsen att styrkan har
större effekt på förpackningens BCT-värde än böjstyvheten. Inga sprickor
noterades heller under bigning och montering av förpackningarna som användes
i Artiklarna C och E. Det här stämmer väl överens med resultaten i Artikel D
eftersom bigdjupet som användes för förpackningarna förhöll sig inom ramarna
för det utvecklade bigfönstret.

Sammanfattningsvis var det primära syftet med den här avhandlingen att utveckla
en lättanvänd modell med få materialparametrar som kan prediktera
last-deformationskurvor för olika lastfall. Syftet var inte att modellen skulle
användas för att simulera fysiska experiment så noggrant som möjligt, utan istället
att öka kunskapen om skademekanismer under provningsförfaranden. En tydlig
fördel med det här tillvägagångssättet är att modellen kan användas för att
antingen variera förpackningsgeometrin, eller göra en parameterstudie av de
ingående materialparametrarna. Den kan också användas för varje skikt separat,
vilket skulle hjälpa både konverterare och producenter. Utöver detta har det visats
att modellen kan ta hänsyn till fukt, om masterkurvan från Artikel B används.
Slutligen bör det förtydligas att modellen inte innehåller några
parameteranpassningar och att all ingångsdata bygger på mätbara värden.
Preface

The work presented in this thesis was carried out at the Research Institutes of
Sweden (RISE) and the Department of Engineering Mechanics, KTH Royal
Institute of Technology, Stockholm, Sweden, between June 2018 and January
2023. The work would not have been possible without the generous financial
support of STFI’s Association of Interested Parties (STFIs Intressentförening).

I want to express my sincere gratitude to my supervisor, Docent Mikael Nygårds,


for giving me this opportunity, convincing me to return to the field of paper
mechanics, and constantly pushing me forward. I appreciate our discussions; your
guidance in this thesis has been tremendously valuable.

Furthermore, I would like to show my appreciation to my supervisor,


Professor Sören Östlund, for introducing me to paper mechanics and providing
invaluable guidance and feedback throughout this project.

I want to thank all my colleagues at RISE and KTH. In particular,


Dr. Prashanth Srinivasa and Dr. Anton Hagman, without whom, I would still be
struggling with the thesis. I furthermore want to thank Cecilia Rydefalk for
sharing the struggles of the position as an industry-employed Ph.D. student. I
appreciate that I have had the opportunity to disturb all your work with my daily
questions.

Finally, I would like to thank my family and friends for all their support. Especially
my wife Matilda, for supporting me through these years of strange working hours.
A final thanks to my sons Harald and Folke, who ensure I do not oversleep in the
mornings.

Stockholm, January 2023

Gustav Marin
List of appended papers

Paper A: Stiffness and strength properties of five paperboards and their moisture
dependency

Gustav Marin, Mikael Nygårds and Sören Östlund

Tappi Journal, Vol. 19, No. 2:71–85, 2020

Paper B: Elastic-plastic model for the mechanical properties of paperboard as a


function of moisture

Gustav Marin, Mikael Nygårds and Sören Östlund

Nordic Pulp and Paper Research Journal, May 2020

Paper C: Experimental and finite element simulated box compression tests on paperboard
packages at different moisture levels

Gustav Marin, Prashanth Srinivasa, Mikael Nygårds and Sören Östlund

Packaging Technology and Science, Vol. 34, No. 4: 229-243, 2021

Paper D: Experimental quantification of differences in damage due to in-plane tensile test


and bending of paperboard

Gustav Marin, Mikael Nygårds and Sören Östlund

Packaging Technology and Science, Vol. 35, No. 1: 69-80, 2021

Paper E: Torsional and compression loading of paperboard packages: Experimental and


FE analysis
Gustav Marin, Anton Hagman, Sören Östlund and Mikael Nygårds

Packaging Technology and Science, Vol. 36, No. 1: 31-44, 2022


Contribution to the papers

Paper A: Principal author and performed all experimental work and analysis.
Marin, Nygårds and Östlund jointly evaluated and interpreted the results.

Paper B: Principal author and performed all experimental work and analysis.
Marin, Nygårds and Östlund jointly evaluated and interpreted the results.

Paper C: Principal author and performed all experimental work, simulations and
analysis. Srinivasa wrote the initial code that Marin developed for this study.
Marin, Nygårds, Srinivasa and Östlund jointly evaluated and interpreted the
results.

Paper D: Principal author and performed the experimental work (together with
Nygårds) and analysis. Marin, Nygårds and Östlund jointly evaluated and
interpreted the results.

Paper E: Principal author and performed experimental work (together with


Hagman), simulations and analysis. Marin, Hagman, Östlund and Nygårds jointly
evaluated and interpreted the results.

In addition to the appended papers, the project has resulted in the following
presentations:

Stiffness and strength properties of five paperboards and their moisture


dependency
Gustav Marin, Mikael Nygårds and Sören Östlund
Presented at TAPPI PaperCon, Indianapolis, USA, 2019

Box compression strength of packages in different climates


Gustav Marin and Mikael Nygårds
Presented at IAPRI symposium, Enschede, Netherlands, 2019

Modelling the moisture dependent elastic-plastic properties of


paperboard
Gustav Marin, Mikael Nygårds and Sören Östlund
Presented at Progress in Paper Physics Seminars, Jyväskylä, Finland, 2020
Contents

1 Introduction ............................................................................................................ 1
2 State-of-the-art........................................................................................................ 3
2.1 Packaging performance ................................................................................ 3
2.1.1 How packages break ................................................................................ 3
2.2 Corrugated board .......................................................................................... 7
2.3 Paperboard ..................................................................................................... 8
2.3.1 Converting a paperboard to a package ...............................................10
2.3.2 Structural properties of paperboard ....................................................12
2.3.3 Mechanical properties of paperboard .................................................13
2.4 Influence of moisture .................................................................................18
2.5 Finite element simulations .........................................................................24
3 Materials and Methods ........................................................................................27
3.1 Material characterization ............................................................................27
3.1.1 In-plane tensile test ................................................................................27
3.1.2 In-plane compression test .....................................................................28
3.1.3 Out-of-plane tension .............................................................................28
3.1.4 Out-of-plane shear test ..........................................................................28
3.1.5 Creasing and folding ..............................................................................29
3.2 Packaging testing .........................................................................................29
3.2.1 Box compression test ............................................................................30
3.2.2 Torsion test .............................................................................................30
3.3 Finite element model ..................................................................................31
4 Results and discussion .........................................................................................35
5 Conclusions ...........................................................................................................43
6 Future work...........................................................................................................45
7 Bibliography ..........................................................................................................47
1 Introduction
The world is facing innumerable environmentally-related challenges, and plastic
waste is one of them. More than 25 billion metric tons of plastic waste are
generated yearly in Europe only. Approximately 1.5 – 4 % of global plastics
production ends up in the ocean annually. These numbers are relevant for the
packaging industry since 59 % of plastic waste is due to packaging [1]. Partly
because of this, the European Union has been acting by banning single-use
plastics [2]. The required reduction of plastic use in future environmentally
friendly packages enables the huge potential for the paper industry since paper-
based material is relatively cheap but also a recyclable material. Regarding the
recyclability of paper, a recent study indicates that fibers may be recycled 25 times
without any reduction in the tensile index [3]. This shift in the usage of packaging
materials can be observed in a regular grocery store, where packages previously
made of plastics are now made of paper-based materials; an example is illustrated
in Figure 1.

Figure 1. Two applications (straw and cover) that were made of plastic previously.

The function and shape of paper-based packaging material may vary. For
example, it can be used as wrapping or converted into a box or container. The
latter function is typically based on a corrugated board or paperboard. A
corrugated board combines thin flat papers (liners) with a wavy-shaped paper
(fluting) in between. A paperboard is typically thick paper with high basis weight,
or grammage. In this thesis, the focus will primarily be on paperboard.

From a mechanical point of view, paper-based materials are appropriate for


packaging since they have low weight compared to their stiffness. However,
replacing plastic with paper-based materials entails new requirements and
challenges, which means that more extensive knowledge and new developing
tools are required. On the material level, paper-based materials have many
drawbacks as well. One of the most obvious drawbacks is the sensitivity to
moisture, which makes applications such as the one in Figure 1 quite challenging
to make successful. Furthermore, paper is a fiber-based material and has, due to

1
production, different material properties in different directions, often referred to
as machine direction (MD), cross direction (CD), and thickness direction (ZD).

With these drawbacks as a starting point, this thesis aims to evaluate how a
package’s performance is affected by its material properties, geometry, or
construction on the material level. Physical testing and finite element (FE)
simulations have been performed to succeed with the project’s goal. The work
performed for this thesis complements the literature by focusing on the packaging
level of paperboard packages. The approach uses a basic material model based on
measurable input parameters and FEM for varying design, geometry, and load
case. From this, it is possible to gain knowledge about the material parameters
affecting the behavior of the package, and deformation mechanisms in general,
instead of focusing on simulating a specific load case as accurately as possible.
Specific issues on the packaging performance that the thesis addresses are:

• how moisture affects the mechanical properties and the packaging


performance,
• the impact of creases on a package’s strength and stiffness response,
and
• the effect of bending stiffness of the paperboard.

2
2 State-of-the-art

2.1 Packaging performance

In 1856, Edward G Healy and Edward E. Allen, in England, patented a process


resulting in a wavy paper. The purpose of the patent was to increase the durability
and comfort of top hats. Several years later, in 1871, a wavy paper was used as
packaging material to protect glass bottles by Albert L. Jones. In 1874, Healy and
Allen’s patent was developed further by Oliver Long, and the wavy-shaped paper
was placed between two other papers (liners). From this discovery, it took 20
years until the first corrugated box was produced. During the same century,
another paper-based material, paperboard, was developed and used as packaging
material, of which Sir Malcolm Thornhill developed the first package in 1817 [4].

Today, paper-based packages made of materials such as corrugated board and


paperboard frequently occur in our modern society. In fact, between 1992 and
2018, the number of tons of produced packages worldwide made of paper
materials doubled [5]; today, approximately one-third of the produced packages
are made of paper-based material [6]. With new regulations on plastic packages
[1] and increased e-commerce, this number is expected to increase further [6].

Three different categories of packages can be specified through the distribution


chain:

• Primary package
• Secondary package
• Tertiary package

where the primary package refers to the package in contact with the product, for
instance, a milk package. Several primary packages are grouped into a secondary
package to facilitate logistics. Secondary packages are after that grouped into a
larger tertiary package. The tertiary package is dimensioned to be placed on a
pallet and is typically used for transportation and warehouse storage. These
different categories naturally have different purposes, which is why the design of
the different packages differs. However, there are different requirements for
different types of packages within the same category as well. Therefore, even if a
milk package and a shoe box categorize as primary packages, they have different
requirements and are designed differently.

The purpose of the package is, therefore, essential for the developing process,
and several aspects need to be considered. From a mechanical point of view,
external loads and grip stiffness, for example, are essential to consider in the
design process. Furthermore, visual aspects and printability are vital as well. For
packages that contain more luxurious products, 3D effects, such as deep drawing,
in the package are a common method to visualize the exclusivity of the product.
Different and appropriate properties must be evaluated with different methods
to analyze whether a package fulfills its purpose.

2.1.1 How packages break


From a mechanical point of view, the most crucial design criterion when
constructing a package is that it should not break. A damaged package is not

3
desirable for several reasons. For instance, cracks in a milk package may lead to
leakage, and the function criterion is consequently not fulfilled. However, even if
the function is not affected, the customer easily deprioritizes a damaged primary
package such as, e.g., the one in Figure 2.

Figure 2. Damaged package.

Packages are exposed to different situations that may damage them. They should,
for instance, endure a long-time stacking and transportation and ideally also
survive dropping to the ground. It is, therefore, essential to ensure that one
designs a package and mechanically optimizes it for its purpose, i.e., it is necessary
to analyze the packaging performance. To succeed with this, well‑developed
methods, such as physical testing, analytical methods, or finite element (FE)
simulations, are required.

One of the most analyzed load cases for paper-based packages is the stacking
performance, which can be evaluated by performing a Box Compression Test
(BCT), where a testing apparatus compresses a package between two plates and
records a load-compression curve. The test is established and well described in
the literature, where Frank [7] summarizes the background of the test method and
presents challenges with a focus on corrugated boxes. The box compression
strength, defined by the BCT value, is obtained from the test procedure, i.e., the
maximum force a box can carry before failure. Here, it is essential to emphasize
that this critical force is evaluated at the first local maximum from the load-
compression curve, not necessarily the ultimate peak of the curve. If the global
maximum occurs after the first peak, the testing apparatus might compress the
package so that any potential content would be damaged. Figure 3 shows a typical
load-compression curve for a BCT. As shown, an initial slack behavior is present
due to the presence of horizontal creases.

4
Figure 3. Typical load-compression curve from a box compression test (BCT) [8].

As mentioned, one can use analytical methods to evaluate packaging performance.


This is preferable since testing is quite expensive and time-consuming. For BCT,
the work by McKee et al. [9] establishes a relation between the box compression
strength and elastic buckling of panels, described by the local in-plane compression
strength and bending stiffness, for boxes made of corrugated board. A similar relation
to predicting the box compression strength for paperboard boxes is described by
Grangård [10] in Eqs (1) and (2). Here, the compression strength is denoted 𝐹𝑐
and the bending stiffness, S.

BCT = 𝑘√𝐹𝑐 𝑆, (1)


where, S, due to the anisotropy of the material, is expressed as

𝑆 = √𝑆𝑀𝐷 𝑆𝐶𝐷 . (2)

Analytical models for predicting the BCT value, such as McKee or Grangård,
typically depend on physical testing. For example, if the package design is
changed, the parameter 𝑘 in Eq. (1) might change, and additional testing must be
performed before these models are applicable.

A critical drawback with invariably evaluating a package on its box compression


strength is the limited amount of information available for design: it does not
provide any information about deformation, compression stiffness response or
failure mechanisms. However, the BCT strength of a package has been predicted
in an alternative approach for packages of paperboard by Ristinmaa et al. [11] by
analyzing the deformation behaviors of panels loaded in compression. The
approach predicts the strength by introducing yield curves representing folds
created during testing. A panel with a low height/width aspect ratio will have
different yield lines compared to a high aspect ratio. Ristinmaa et al. establish that
the corners of a panel carry a large part of the load during a box compression test
and that the yield lines commonly propagate from the corners into the panel’s
interior. The corner region fails at peak load due to a collapse mechanism rather
than a stability issue. Consequently, a local failure criterion is activated at peak
load.

The drawbacks of physical experiments and analytical methods requires the usage
of new development tools when the packaging industry faces a demanding future.

5
One possible way forward is to use finite element analysis (FEA) in the
development process. One can use FEA in the package design process to account
for other variables besides strength. For BCT, several studies investigate this on
packages of corrugated board. Nordstrand et al. [12] developed a FE method for
predicting strength for corrugated board containers, including a failure criterion
based on the Tsai-Wu tensor failure theory [13], and compared the predicted
failure load to experimental data. In the same year, Biancolini and Brutti [14]
accurately predicted initial buckling during box compression and estimated the
critical load by an eigenvalue buckling analysis. A clear advantage of
FE simulations is the possibility to evaluate different designs, which is utilized by,
for instance, Han and Park [15], who simulate corrugated boxes with ventilation
and hand holes. Fadiji et al. [16] have similar motivations and report the
differences in results when changing the vent’s number, orientation, and shape.
Jiménez-Caballero et al. [17] also illustrate the advantages of FEM by simulating
BCT for different geometries. Garbowski et al. have recently developed analytical-
numerical approaches for predicting BCT of corrugated boxes with various
openings [18], similar to the vents, and for corrugated boxes with various
perforations [19].

For paperboard packages, Beldie and Sandberg [20] were early in the field,
analyzing different parts of a package during a BCT, and concluded that the
horizontal creases affect the stiffness of a paperboard package. Later, similar
studies on BCT of paperboard packages were performed with a more accurate
shape of the load-displacement curve. Luong et al. [21] accurately predicted a
paperboard package’s strength and stiffness response during a BCT by explicitly
implementing an orthotropic elastoplastic model for in-plane stresses in Abaqus
developed by Mäkelä and Östlund [22]. However, this method requires several
parameters determined by fitting them to experimental data. Luong [23] also notes
that the creases are crucial for the stiffness response of the package, which is why
he reduces the stiffness of the creases. Zaheer et al. [24] investigate critical
buckling loads and consider different geometries and influences of creases.

Using FEA allows the studying of stress and strain fields that arise during loading,
contributing to a better understanding of the crucial mechanisms activated in the
loaded package. In addition, it is possible to isolate and modify different design
parameters, such as the package’s perimeter and the packaging material’s
mechanical properties and see how they affect the behavior of the package. In the
end, this may lead to better knowledge and the development of improved
sustainable packages and paperboard materials. FEA also enables the possibility
of investigating varying load cases. Analyses, where FE simulations have been
performed and compared with experimental observations, have also been done
on a macroscopic level for 3D forming [25], brim forming of cups [26], loading
of cigarette packages [27], and drop testing [28]. Common for all these analyses
was that the authors used relatively simple material models to make accurate
predictions. When less advanced material models are used in the analysis, one can
reduce time and focus on activated mechanisms during testing.

Regardless of which method is used to analyze the packaging performance, one


should note that external loads are not always why a package breaks. Cracks may
occur during the converting process, i.e., when the paper-based material is

6
converted to a package. It is vital to design the material appropriately to avoid
these cracks; therefore, detailed knowledge of the material’s construction is
required.

2.2 Corrugated board

One main advantage that makes paper-based materials suitable for packaging is
the stiffness-weight ratio; paper generally has high stiffness relative to its low
weight. In paper physics, the literature commonly refers to bending stiffness. It is the
board’s resistance against the bending deformation, expressed as a function of the
elastic modulus, 𝐸, and the area moment of inertia, 𝐼. By modifying the board’s
construction, e.g., the ZD density profile, its area moment of inertia can be
altered. This allows for optimization for increased bending stiffness.

A corrugated board visualizes this theory well since it follows the design of a
sandwich construction and the concept of an I-beam: outer facings (liner) with
high stiffnesses and strengths, separated by a wave-shaped core (fluting),
according to Figure 4. The ingoing materials of a corrugated board, i.e., the liner
and the fluting, are called container boards and typically glued together with a
starch-based adhesive.

Figure 4. Schematic view of a single wall corrugated board.

As noted, Figure 4 illustrates the shape of a single-wall corrugated board,


explaining the purpose of sandwich construction. It is also possible to increase
the number of layers by adding several flutings and liners. Single-faced corrugated
board is also frequently occurring in the packaging industry, i.e., a liner glued onto
a fluting.

There are different profiles of corrugated boards. A profile specifies the geometry
of the flute (i.e., the wavelength and height). A letter categorizes the profile, where
profile A is the highest flute (4.0-4.8 mm) with the longest wavelength (8.3-10
mm). Since a larger flute increases the distance between the liners, the
manufacturers use this when high bending stiffness is required. On the other
hand, small flutes are commonly used when a high-quality print is preferred.

As indicated in Figure 4, a corrugated board is described as an anisotropic


(commonly orthotropic) material at the continuum level, where a machine direction
(MD) is defined along the production line. The other in-plane direction is called
the cross direction (CD), and the out-of-plane direction is called ZD. The anisotropy
comes from both the shape of the fluting and the manufacturing process of the
container boards.

7
2.3 Paperboard

The main principle of a paperboard’s construction is generally the same as for a


corrugated board. The design follows the same goal: stiff and strong outer plies
with a bulk contribution in the middle. However, the geometry on the continuum
level looks different than for a corrugated board. Figure 5 shows a close-up image
of a multi-ply commercially produced coated paperboard and illustrates three
essential things of papermaking: orientation of fibers, density, and fiber length.

Figure 5. Close-up image of a multi-ply paperboard.

When manufacturing a paper-based material, a liquid suspension with a solid fiber


content of approximately 1 % uniformly spreads out on a moving wire. A
consequence of reducing nonuniformities in the mass distribution of the solid
content on the wire is that the fibers rotate parallel to the paper machine’s
direction due to a necessary speed difference between the furnish jet and the wire.
Water is later removed through suction units and pressing cylinders forming a
web with a solid content of approximately 50 % [29, pp. 22-24]. In this section,
the density and thickness of the paper are determined. Finally, the web dries
through a series of hot cylinders. When water disappears from the web, joints or
bonds form between the fibers. Higher density generally means more bonds in
the web and, therefore, stronger and stiffer paper [29, p. 198]. Tension is applied
to the web to prevent fluttering during drying. This tension prevents the web
from shrinking in MD and contributes further to the anisotropy of the paper.
Hence, the manufacturing process of the paper results in an anisotropy of the
paper that, to a good approximation, can be described by three principal
directions: machine direction (MD), cross-machine direction (CD), and the
through-thickness direction (ZD).

Achieving a sandwich construction requires different properties for the different


plies of the paperboard. It might be challenging to see, but the paperboard in

8
Figure 5 consists of outer plies with higher density, stiffness, and strength than
the middle ply, which generally consists of fibers with low density. Changing each
ply’s furnish enables modification of the mechanical properties. For example,
fibers from different wood sources have different tensile indexes (corresponding
to specific strength, i.e., tensile strength divided by density), which Figure 6 (a)
illustrates.

(a) (b)
Figure 6. Tensile indexes for different fibers, redrawn from (a) [30] and (b) [31, p. 34].

The reader should also note that fibers from different sources have different
lengths. The dimensions of typical wood fibers are 20-30 µm in width and 1 mm
in length for hardwood and up to 3 mm for softwood (commonly pine and
spruce), respectively. The fiber length may affect the mechanical properties as well
as the printability of a paper [32, 33]. Nygårds [34] recently introduced an
approach based on material characterization of split plies, where he expresses the
in-plane stiffness and strength as a linear equation of density and fiber length.
Hence, it is possible to adjust the mechanical properties of each ply by modifying
the orientation, the density, and the fiber length. Depending on whether
mechanical or chemically based methods are used for separating the fibers in the
pulp manufacturing process, they will also achieve different properties. The main
difference from a mechanical point of view is that mechanical pulp generally has
lower strength and stiffness since a significant fraction of the fibers is
damaged during the mechanical process [35, 29, p. 21]. This difference is
illustrated in Figure 6 (b), where groundwood (GWD) and mechanical pulp (MP)
have the lowest tensile indexes. Termo-mechanical pulp (TMP) and chemi-
thermomechanical pulp are also mechanical pulps treated with heat and
chemicals. In a more recent review paper, Petterson et al. [36] publish similar
results as in Figure 6 (b), but the tensile index for CTMP reaches 50 kN/kg.
However, when manufacturing the CTMP after preheating to a temperature
above the softening temperature of lignin (HTCTMP), the tensile index reaches
the same levels as in Figure 6 (b). SBK is an acronym for semi-bleached kraft
pulp, and UBS is unbleached kraft pulp. For example, a typical standard folding

9
box board (FBB) comprises two outer plies with chemical pulp (kraft pulp) and a
middle ply consisting of mechanical pulp, such as CTMP.

2.3.1 Converting a paperboard to a package


Everyone who ever made a paper airplane has utilized the foldability of paper.
Paperboard is also suitable for folding but requires some preparation first since it
is thicker than copy paper. The preparation procedure is called creasing and is
well-defined in the literature on paperboard. The procedure was recently
thoroughly reviewed by Coffin and Nygårds [37], where a patent on creases from
1876 is mentioned [38]. The primary purpose of creases is to enable folding along
well-define lines, i.e., create hinge-like performance zones, as Carlsson et al. [39]
discuss. The creasing process relies on purpose-made damage initiated in the
paperboard structure; ideally, the paperboard should delaminate internally.
Several studies investigate the importance of delamination during creasing. For
instance, Halladay and Ulm [40] performed experiments in 1939, marking vertical
lines on the paperboard’s edges and capturing photographs during the creasing
procedure. The experiments were repeated by Coffin and Nygårds [37], and by
studying the shapes of the vertical lines, they concluded that shear deformation
during the creasing procedure is significant.

During the creasing procedure, a ruler is punched into a die with the paperboard
in between. By doing this, shear strains develop in the material and cause
delamination, as illustrated in Figure 7. Hence, the delamination procedure is
associated with the breaking of fiber-fiber joints. Furthermore, the structure of
the paperboard is crucial for increasing the extent of delamination in the creased
zone, which Carlsson et al. [39] conclude. A study by Nygårds et al. [41] points out
that an ideal structure for enabling delamination and facilitating the plies to
deform plastically in shear internally is a paperboard with well-defined plies and
weak interfaces.

(a) (b)
Figure 7. Before (a) and during (b) a creasing procedure.

Delamination is needed to form a bulge on the inner part of the paperboards


during the subsequent folding. As the initiation of delamination cracks is essential,
the procedure must balance the initiation and propagation of in-plane cracks.
In-plane cracks are associated with fibers torn out of the fiber network and will
contribute to the cracking of the outer plies, which can jeopardize the integrity of
the package and its content.

10
To illustrate the importance of the creasing procedure, Figure 8 shows
cross-section pictures of folded specimens at different creasing depths, d.
In Figure 8 (a), where the specimen was uncreased, one can observe how the
inner surface, or the bottom ply, has been compressed. Cavlin [42] noted this in
1988 and concluded that compression failure occurs at the peak value of the
curve.

(a) (b) (c)


Figure 8. Folded MD sampled that are: uncreased (a), creased to d=0.0 mm (b) and d=0.2 mm (c)
[43].

In the literature, several authors present different approaches to define the shear
strains caused by the creasing procedure, for instance, Halladay and Ulm [40],
Donaldson [44], and Nagasawa et al. [45, 46]. All approaches interpret geometrical
relations between parameters such as the creasing depth, the channel width, and
the ruler’s width. The definition of a shear measure is further developed by Coffin
and Panek [47] and Panek et al. [48], where the paperboard draw is introduced to
define a creasing window. The draw defines how much paperboard draws into the
die during the creasing operation. With this approach, creasing windows ensure
that the relative crease strength (RCS), also referred to as the bending resistance
ratio, is sufficiently low, i.e., avoiding cracking and securing good crease quality.
The relative crease strength (RCS) is the ratio between the maximum bending
creased
force achieved during the folding of a creased board, Ff , and an uncreased
uncreased
board, Ff , hence

𝐹𝑓creased
𝑅𝐶𝑆 = . (3)
𝐹𝑓uncreased
The folding operation is evaluated by different methods, where one common way
of evaluating the procedure is two-point bending, which provides a force-
deflection curve. From this curve, the bending stiffness can be evaluated
according to the Euler-Bernoulli beam theory, and the peak value characterizes
the maximum bending force discussed in Eq. (3). The bending stiffness at two-
point bending is given by [49]:

19.1𝐹𝐿2 (4)
𝑆=
Φ𝑏
where Φ is the angular deflection, 𝑏, is the width and 𝐿 is the length of the
specimen. The force, 𝐹, refers to the value measured at 5°, where paperboards
generally remain in the folding curve’s elastic region. Figure 9 shows typical force-
deflection curves in MD from two-point bending and visualizes the effect of
creasing on the maximum bending force and the bending stiffness as a
complement to Figure 8.

11
Figure 9. Two-point bending curves of uncreased and creased specimens [50].

Published studies on creases generally use apparatus performing creases at low


speed, where the deformation and damage are quasi-static compared to the high-
speed machines typically used for industrial creasing.

2.3.2 Structural properties of paperboard


In the literature on paper mechanics, it is common to report the in-plane
mechanical properties for paper-based materials in kN/m, i.e., the results do not
account for the thickness of the board. Moreover, one excludes the thickness
when referring to the basis weight or grammage, i.e., the weight per square meter
[g/m2]. The reason for excluding thickness is generally the difficulty of measuring
the thickness of the paperboard. However, when implementing material
properties in FE software, it is necessary to also account for the thickness. An
established method [51] for this is measuring the structural thickness, also referred
to as the STFI method, developed by Fellers et al. [52]. This method records the
thickness along a line as the distance between two spherical measurement probes.
It provides the user with a thickness profile, a mean value, and a standard
deviation. An apparatus developed for further developing the STFI method by
measuring the thickness in two dimensions is the STFI local thickness equipment [53],
constructed by TJT-teknik [54]. Figure 10 illustrates the variation in the measured
thickness of a commercially produced liquid packaging board (LPB) with a basis
weight of 450 g/m2 [55]. The basis weight and thickness vary for different
paperboards. However, as an example, a typical folding box board (FBB) is
produced with the basis weight range of 170-400 g/m2 [56], with the
corresponding thicknesses of 205-605 µm.

12
Figure 10. Typical surface profile for a paperboard.

2.3.3 Mechanical properties of paperboard


Paperboard is often considered orthotropic, i.e., the stiffness properties are
symmetric with respect to the three material directions. In the elastic region, the
strains and stresses are connected by the symmetric compliance matrix in Voigt
notation:
1 𝑣21 𝑣31
− − 0 0 0
𝐸1 𝐸2 𝐸3
𝑣12 1 𝑣32
− − 0 0 0
𝜀𝑥 𝐸1 𝐸2 𝐸3 𝜎𝑥
𝜀𝑦 𝑣13 𝑣23 1 𝜎𝑦
− − 0 0 0
𝜀𝑧 𝐸1 𝐸2 𝐸3 𝜎𝑧
𝛾𝑥𝑦 = 𝜏𝑥𝑦 (5)
1
𝛾𝑥𝑧 0 0 0 0 0 𝜏𝑥𝑧
𝐺12
[ 𝛾𝑦𝑧 ] 1 [𝜏𝑦𝑧 ]
0 0 0 0 0
𝐺13
1
0 0 0 0 0
[ 𝐺23 ]

As seen in Eq. (5), knowing the in- and out-of-plane stiffnesses is essential for
characterizing the material. Moreover, it is vital to know the strength of the
material, in tension, compression, and shear, for the in-plane and out-of-plane
directions, respectively. Noteworthy is that these properties are not always
published in data sheets for commercial use. For paperboard, other properties
such as bending stiffness, are commonly presented. Some applications or material
models require additional parameters as well. However, the stiffness properties
from Eq. (5), in combination with the strength for the different material
directions, are a good starting point for material characterization.

According to Niskanen [29, p. 8], most published data on the mechanical


properties of paperboard are based on laboratory sheets and not commercial
boards. Therefore, the purpose of this section is to review what has previously
been done in the field and provide the reader with an approximate value for each
property. The section will focus on commercially produced paperboard.

13
2.3.3.1 In-plane properties of paperboard
The in-plane properties of paperboard, especially the strength and stiffness in
tension, have been thoroughly investigated in the literature since they are vital for
analyzing the material’s performance, whereas Niskanen [57] reviewed the field
until 1993. Niskanen concludes that two factors govern the strength of paper: the
strength of fiber-to-fiber bonds and the formation of paper, i.e., a uniform paper
tolerates higher stresses than non-uniform paper. Fiber breaks are not considered
the critical factor in tension unless for specific scenarios when the bonds are very
strong. The explanation behind bond failure is not discussed further in this thesis.
However, it is well analyzed in the literature, where Cox [58] in 1952 discusses a
shear-lag mechanism where the stress is transmitted from one fiber to another
through the bonds. However, on a board level, the pure in-plane tensile properties
are obtained from uniaxial tensile tests, where different standard procedures are
established (such as ISO [59] and ASTM [60]), and different types of equipment
following the standards have been developed.

Even if this thesis primarily discusses the stiffness and strength properties of
paperboard, it is essential to mention the size dependency of in-plane tensile tests.
The characterization of tensile fracture and critical length is discussed by Cavlin
[61] in 1974, where the brittleness of a material is related to the critical length of
a specimen, i.e., the shortest length for which the load-elongation curve rapidly
drops from peak load to zero [62]. This is explained by the ratio of incrementally
released elastic energy and fracture energy within the specimen. In large
specimens, the rate of released elastic energy is sufficiently large to cause an
unstable fracture of the specimen, i.e., a brittle and rapid fracture. Figure 11 shows
the length dependency for a specimen, where specimens of the same paperboard
with the same widths but different lengths were exposed to tensile load. For
shorter lengths, a stable reduction of the load-elongation curve is observed since
the released rate of elastic energy is lower than the rate of required fracture energy
needed for progressing the fracture. Tryding [63] analyzes this concept further,
and Tryding et al. [64] characterize the crack growth in paperboard by introducing
stress-widening curves for analyzing the post-peak cohesive behavior. Hagman
and Nygårds [65] further investigate the sample size effects with an alternative
approach, using tensile test and speckle analysis. They conclude that there is a
difference in strain behavior depending on the length-to-width ratio and that this
was dependent on the board composition.

14
Figure 11. Tensile test with different specimen lengths (width = 15 mm) [66].

The magnitude of the properties in paperboard varies with density [34]. However,
to introduce the reader to representative numbers of in-plane tensile properties,
an early work by Persson [67, p. 26] is emphasized. The study was performed on
a three-plied commercial duplex board, where Persson reports the in-plane
stiffnesses to be 5.4 GPa (MD) and 1.9 GPa (CD). The basis weight of the board
was 240 g/m2 and a thickness of 0.374 mm, i.e., a density of around 640 kg/m3.
The tensile strengths were 46 MPa in MD and 22 MPa in CD. These are
representative paperboard numbers confirmed in other studies [68, 69].

When describing an anisotropic material, it is crucial to observe the difference


between the material in-plane properties in compression and tension. As noted in
Eq. (1), the in-plane compression strength (also referred to as the edgewise
compression strength) is vital for analytical models on box compression strength.
There are obvious difficulties in evaluating compression properties for a thin
sheet material such as paperboard since buckling must be avoided. In 1975, Cavlin
and Fellers [70] suggested two methods to achieve edgewise compressive
strength: either by performing tests on a long strip, with lateral support along its
free span (later in the literature referred to as long span compression test, LCT),
or compression on a specimen with a significantly small span. The same year,
Fellers and Jonsson [71] performed short-span tests on paper-based materials,
where the free span decreases until the strength-span curve reaches a plateau. In
their study, this plateau occurs slightly below a span length of 1 mm. Today,
measuring the edgewise compression strength by compressing a significantly
small span is the most common method and is referred to as the short-span
compression test (SCT) [72]. Clear disadvantages of this method are the lack of a
stress-strain curve of the procedure and the influence of the clamping conditions.
The SCT only reports the maximum strength. Bugiel et al. [73] discuss the
difficulties of measuring the in-plane compression properties of thin materials
and developed a new single-curved test device in 2017, which they used to
characterize two different paper-like materials. Brandberg and Kulachenko [74]
use a network model to analyze failure during SCT. They conclude that the failure
mode is elasto-plastic buckling and that the most significant improvement to

15
sheet strength is increasing the elastic properties of the board. They also highlight
the effect of the clamping procedure and the effect it has on the delamination.
Regarding numbers, representative values of the compressive strength for a multi-
ply board with a thickness of 0.390 mm and a density of 740 kg/m3 are 15.2 MPa
in MD and 12.5 MPa in CD, respectively [75].

According to Fellers and Donner [76], the causes of compression failure are
elastic buckling of fiber segments (low-density sheets) or shear dislocations in the
fiber walls (high-density sheets). In 1983, Habeger and Whitsitt [77] developed a
mathematical theory describing compressive failure related to in-plane and out-
of-plane shear stiffnesses. Lately, Hagman et al. [75] have investigated the cause of
compression failure on multi-plied commercial boards. Hagman et al. state that
shearing the interfaces, combined with the onset of plasticity in the loading
direction, initiates failure during SCT. They further concluded that the main
parameters determining compressive strength were the elastic modulus, local
shear strength, and out-of-plane strength. Hagman and Nygårds [78] also show
that the short-span compression test is governed by in-plane stiffness and
through-thickness delamination. Hence, the in-plane compression is also related
to the out-of-plane properties.

The final in-plane property discussed here is the in-plane shear modulus, which is
important when describing paperboard as an orthotropic material. Persson [67]
estimated it by performing in-plane tensile loading at 45° to the MD and CD,
page 16. Baum et al. [79] derived the expression

(𝐸𝑥 𝐸𝑦 )2
1
(6)
𝐺𝑥𝑦 = 1
2(1+(𝑣𝑥 𝑣𝑦 )2 )

for the in-plane shear modulus in 1981, where the in-plane Poisson’s ratio was
assumed to follow:

√𝑣𝑥𝑦 𝑣𝑦𝑥 ≈ 0.293 (7)

𝑣𝑥𝑦 𝑣𝑦𝑥 (8)


𝐸𝑥
= 𝐸𝑦

𝐸 (9)
𝑣𝑥𝑦 = 0.293√𝐸𝑥
𝑦

and Poisson’s ratios in the out-of-plane directions were set to zero [80]. These
assumptions are used in several studies in the literature [69, 75].

2.3.3.2 Out-of-plane properties of paperboard


Compared to the in-plane strength, determining the out-of-plane strength, also
referred to as ZD strength, is a more complex procedure. The out-of-plane
strength of hand-made sheets was analyzed in 1973 in a doctoral thesis by
Van Liew [81], who fastened circular specimens onto cylinders with an adhesive,
which were mounted in uniaxial testing equipment and then separated during a
displacement-controlled test procedure in the thickness direction. Girlanda and
Fellers [82] on handsheets and Nygårds [83] on multi-plied paperboard for each
ply individually described a similar approach for determining the ZD properties.
Furthermore, Stenberg et al. [84, 85] thoroughly analyzed the out-of-plane
properties in paperboard and developed an elastic-plastic constitutive model for

16
combined normal and shear through-thickness loadings. The tensile strength in
the thickness direction varies in the literature, depending on the paper-based
material investigated. For handsheets, Van Liew reports 0.55 and 0.95 MPa for
two different handsheets, and Girlanda and Fellers report 0.2-1.75 MPa,
depending on the furnish. Persson measured the out-of-plane strength of the
paperboard to 0.33 MPa, which is similar to the values reported for the middle
ply by Nygårds: estimated to be 1.3 MPa for the top ply, 0.25 MPa for the middle
ply and 0.95 MPa for the bottom ply, respectively.

In the literature, the out-of-plane stiffness is measured in different ways, where


the spread of results by different techniques is noted and summarized by
Girlanda and Fellers [82] in 2007. The elastic modulus may be evaluated through
a stress-strain curve. However, several studies use the ultrasonic technique, i.e.,
measuring the velocity of a sound wave within a material [86]. For paperboard,
Nygårds [83] determines the elastic modulus from the stress-strain curve and
reports 2520 MPa in the top, 16 MPa in the middle, and 220 MPa in the bottom
plies, respectively. Again, the reported result from the middle ply is in the same
range as Persson’s [67] reported values for the entire board: 17 MPa.

A paperboard may also be compressed in the out-of-plane direction. This load


case is not thoroughly investigated in this thesis but is essential in converting
operations such as printing, calendaring, and die-cutting. The cylinder distance
controls the contact pressure during such operations, where the paperboard runs
through rotating rolls. Instead of discussing the pressure expressed in MPa, the
pressure is occasionally referred to as the distance between cylinders [87]. For
flexographic printing of paperboard, Johnson et al. [88] measured the pressure at
the print nip to 1-2 MPa. In the literature, when measuring the compression in
ZD, it is common to use a universal testing machine (UTM) and expose the
paperboard to a displacement-controlled testing procedure. The force is typically
recorded by a load cell, resulting in a load-displacement curve. The deformation
is, however, difficult to estimate since the measured deformation also includes the
deformation of the surface roughness [89]. Nygårds [83] investigates the
compressive response in ZD of different plies of paperboard by compressing
them with a rate of 200 N/s up to 20 MPa and concludes that the middle ply is
more compressible than the outer plies due to its lower density. Nygårds [90]
further describes the compression behavior as an exponential function dependent
on elastic and plastic strains. ZD compression has also been investigated with the
same method by Li et al. [91] when modeling the out-of-plane behavior of paper.
However, the properties of a paperboard during ZD compression can also be
evaluated by other techniques. Rättö and Rigdahl [92] studied the mechanical
properties in 1998, with a UTM, but with short pulses. In a recent study,
Rydefalk et al. [87] used a rapid ZD-tester to subject the paperboard to a rapid pulse
of the same magnitude as in a printing press. The device is equipped with a probe
with a diameter representing the length of a printing nip, aiming to resemble the
printing conditions better.

For the characterization of paperboard from an orthotropic perspective, the


out-of-plane shear properties need to be evaluated. This is a complex procedure
where the specimen is prepared with an adhesive in a similar way to out-of-plane
tension tests. Here, the specimen is commonly fastened between two metal blocks

17
separated with a loading direction oriented in MD or CD, with a prescribed
deformation. The shear testing procedure results in a load-deformation curve,
from which the strength and stiffness properties can be obtained. Fellers [93]
published data (not commercially produced board) on these rigid block tests in 1977
with an out-of-plane shear modulus of 70 MPa and a shear strength of 1.1 MPa.
A more recent study on commercial multiply boards reports that the middle ply
has an out-of-plane modulus of 30 MPa [25]. As an alternative approach,
Nygårds et al. [68, 94] developed a new way of evaluating shear strength: The
notched shear test (NTS). A tensile specimen is notched on the top and the
bottom (with an in-plane distance between the notches) and is then exposed to a
tensile load. Hence, a shear zone is created at the depth where the notches meet.
This method does not evaluate the out-of-plane shear modulus but is useful when
the shear strength profile is of interest [95, 96, 97, 98]. An alternative method for
evaluating the shear properties is the Arcan device, designed for combined
loadings in normal and shear. This method was used by Stenberg [99] to measure
the out-of-plane shear properties during high compressive loads.

2.4 Influence of moisture

The most distinct drawback of paperboard as a packaging material is its significant


sensitivity to moisture. Since moisture is the weakest point of paper-based
materials, it is also thoroughly investigated in the literature. Since this thesis
focuses on the mechanical properties of paperboard, the influence of moisture
must be discussed. The sensitivity is explained by the fact that water acts as a
softener on the polymers in the material. Briefly, the fibers contain several
microfibrils, which are made of cellulose chains (crystalline structure). The
microfibrils are considered to have disordered regions along their lengths and are
surrounded by a matrix of hemicellulose, lignin, and loose cellulose chain ends,
modeled by Salmén [100] in 1982. The disordered regions represent restrained
cellulose chains which display a softening at higher moisture content than
hemicellulose. Figure 12 shows the difference in water uptake for cellulose, lignin,
and hemicellulose.

Figure 12. Differences in water uptake for hemicellulose, cellulose, and lignin. Redrawn by
Salmén from data by [101].

18
Hence, complete softening may not be achieved until the fibers are immersed in
water. This is illustrated in the model by Salmén in Figure 13. In Figure 13, the
humid state corresponds to a relative humidity level above 80 %.

ellulose
microfibril
crystalline

isordered
one glass glass soft

atri
emicellulose, glass soft soft
lignin and
cellulose chain ends

Figure 13. Illustration of microfibril with surrounding matrix during different environmental
conditions, redrawn from [100].

There are, however, several studies on how different pulps interact with moisture,
but this thesis focuses on research on primarily commercial paperboard. The
paper-water interaction could either be through contact between paperboard and
liquid, i.e., hydrosorption, or by changing the relative humidity (RH): hygrosorption. In
this thesis, the effect of moisture will be analyzed through the perspective of
changes in RH, and hydrosorption will not be discussed further, but the reader
interested in hydrosorption is referred to the work of Larsson [102].

The relation between the amount of moisture in the material and the surrounding
environment is non-linear and irreversible. However, when the relative humidity
changes, the paper-based material will either gain water (adsorption) or lose water
(desorption). This history dependence is referred to as a hysteresis effect and is
not unique to paper. It is, however, well discussed in the literature of paper
materials. Even though the moisture sorption in paper is well investigated, there
is still a lack of consensus regarding the origin of the hysteresis phenomenon
[103]. A conceptual view of the effect is shown in Figure 14.

19
oisture measure

Figure 14. Conceptual view of the hysteresis effect in paper materials.

From a mechanical point of view, this phenomenon is fundamental since it means


that the result when measuring a mechanical property at 23 °C, 50 % RH
(standard climate [104]) depends on if the material state is in adsorption or
desorption. The testing standard, therefore, says that specimens must be pre-
conditioned at lower levels of RH before conditioning for mechanical testing.
When the paperboard is exposed to higher levels of RH, the fibers will swell. Since
the fibers are bonded together, the swelling affects neighboring fibers and
separates them further apart. The hygroexpansion is a measure of how much the
paperboard material expands during hygrosorption.

The effect on the in-plane mechanical properties from testing at different climates
is well documented, where the stiffness and strength decrease with increased
RH [103]. Benson [105] noted in 1971 that the in-plane strength and stiffness are
reduced by 50 % when the relative humidity changes from ca 20 % to 90 % RH.
This is valid for both MD and CD. Typical in-plane stress-strain curves are shown
in Figure 15, where the observations from Benson are verified. Similar curves as
those in Figure 15 were evaluated by Mäkelä [106] in 2010, although he also
investigated the impact of moisture on in-plane compression. Tensile tests at
different moisture levels have also been performed and simulated by
van der Sman et al. [107].

(a) (b)
Figure 15. In-plane stress-strain curves at different levels of RH for MD (a) and CD (b)
[108].

20
Out-of-plane properties are also affected by moisture but are not reported in the
literature to the same extent as the in-plane properties. Tryding and Ristinmaa
[109] published curves on tensile test in ZD at two different moisture levels in
2016 where the cohesive post-peak behavior was present, just like the curves in
Figure 16. More recently, Spiewak et al. [110] performed out-of-plane tensile tests
at different moisture contents when modeling the crack growth behavior of filter
paper.

Figure 16. Tensile test in ZD at different levels of RH [111].

Regarding out-of-plane compression, Land et al. [112] evaluated the relationship


between plastic ZD compression and moisture. However, the compression in ZD
was caused by calendaring, not a UTM. Compression-strain curves measured with
a UTM in ZD at different levels of RH are shown in Figure 17.

Figure 17. Compression in ZD at different levels of RH [66].

The folding procedure is also affected by the influence of moisture, as shown


in Figure 18. It also illustrates the different shapes of the folding curves
depending on the material direction.

21
(a) (b)
Figure 18. Folding at different levels of RH [111].

In the literature, mechanical properties such as stress and stiffness are evaluated
against RH, as in Figure 15 -Figure 18. This is relevant for some applications, but
the relative humidity primarily gives information regarding the testing
environment, not the amount of moisture within the material. Therefore,
methods for describing this have been developed in the literature. Perhaps the
most common way of evaluating the moisture in paper is the moisture content
[113], which is the ratio between the water and the total mass, i.e.:
𝑚𝑤𝑎𝑡𝑒𝑟 (10)
𝑚𝑐 =
𝑚𝑡𝑜𝑡𝑎𝑙
An alternative method is using the moisture ratio:
𝑚𝑤𝑎𝑡𝑒𝑟 (11)
𝑚𝑟 =
𝑚𝑑𝑟𝑦
This method does not have an ISO standard but has the clear advantage that it is
strictly linear to the mass of water. It has also been shown in the literature that
mechanical properties are linear to moisture ratio (within a limited range). Fellers
shows the linearity for strength and stiffness [31, p. 41 and 141]. Further,
Mäkelä [106] reports that the in-plane tensile strain at break and the compression
stiffness and strength are also linear to moisture ratio. The linear interval for in-
plane stiffness was studied by Mäkelä [114], which is shown in the initial regions
of the curves in Figure 19. It should be noted that the study was a drying study
performed on wet paper specimens. Hence, the material was not exposed to
moisture through hygrosorption, i.e., changing the relative humidity, which
explains the huge values of the moisture ratio.

22
Figure 19. Tensile stiffness index as a function of moisture ratio, redrawn from [114].

Linear relations to the moisture ratio are also valid for tensile strength in ZD,
according to measurements by Fellers [115]. In the same study, Fellers notes that
if the relative strength is plotted against the relative moisture ratio (i.e., strength and
moisture ratio normalized with the corresponding values at 50 % RH), a linear
relation is established independent of pulp. Other authors also report relative
values of the mechanical properties. Popil [116] studies relative values of SCT
when describing the relation between compression strength and moisture
content, and Back [117], studies the difference in reduction of in-plane tensile and
compression strength.

Besides the impact of moisture on the mechanical properties of paper-based


materials, there are other moisture-related issues that will not be discussed further
in this thesis. One example of such issues is curl or warp of board, i.e., when the
board reacts to moisture by bending. This is a common problem, for instance,
when producing corrugated boards. One further moisture-related problem is
creep. Most people who store things in corrugated boxes have experienced
this phenomenon — the strain over time increases under a constant state of
stress. The increase in strain is susceptible to moisture, and it is well established
that cyclic humidity is more severe for creep of paper than a constant high level
of RH. This effect is referred to as accelerated creep or the mechanosorptive
creep [118].

As a final remark, this chapter has examined moisture’s impact on paperboard


properties. It should also be noted that temperature affects the mechanical
properties. An increased temperature reduces paper’s in-plane strength and
stiffness, which is investigated by, i.a., Salmén and Back [119] in 1978, and
Linvill and Östlund 2014 [120]. These studies both cover a wide range of high
temperatures. For paperboard, Figure 20 illustrates the effect of temperature
between 10 - 40 °C, at 50 % RH, where the reduced stiffnesses and strengths are
shown.

23
(a) (b)
Figure 20. Impact of temperature on in-plane tension for MD (a) and CD (b) [66].

2.5 Finite element simulations

In a perfect world, simulations of paperboard packages would be easy. What is


needed is typically a simple geometry and a relevant load case. Additionally,
element type and mesh settings need to be considered. However, due to its
complexity, several approaches to describing paper-based material exist. First, the
analysis can be performed on a macro- or micromechanical level, i.e., one needs to
consider if the material could be treated as a continuum or if a network model is
required. This is, of course, an application‑driven issue. When studying, for
instance, the effect of fibers and fiber-fiber bonds on paper as a material, it is
helpful to use network modeling on the microscopic level. Such analyses are
performed by Borodulina et al. [121]. When investigating the stress-strain curves
for networks with weak and strong bonds, Kulachenko and Uesaka [122] showed
in 2012 through a network model with a disordered structure, that the strength
and stiffness do not vary significantly from a statistical point of view. The average
number of fibers in the simulated specimen and the degree of fibers were kept
constant. The strain was, however, sensitive to disordered structures.
Network modeling has successfully been applied to simulate the effects of
moisture [123, 124]. It has recently been used for describing phenomena such as
curl caused by moisture application during the printing process [125]. However,
another way of describing paperboard is through a continuum approach. Instead
of treating the material as a complex network with explicit fiber-fiber bonds, it is
considered an anisotropic continuum with three principal directions (MD, CD,
and ZD). This is typically a more relevant approach for analysis of packaging
performance and therefore used in this thesis.

Due to its three principal material directions on continuum level, it is common to


describe it as an orthotropic material in the elastic region. For this approach, ill’s

24
criterion [126] may be used, which originally was developed in 1948 to describe
anisotropic metals. ill’s model suggests a linear, but not anisotropic, hardening.
For ductile materials, a power-law relation was introduced by Ramberg-Osgood
[127] in 1943 and was then applied to paper through an orthotropic elastic-plastic
model by Mäkelä and Östlund [22] in 2003. The model accurately predicts the in-
plane stress-strain response of paperboard. The most simplified way of describing
the elastic-plastic behavior of paperboard is a bilinear model, i.e., a model with
two straight lines connected in the yield point. This approach has been
investigated in the literature and is accurate for several cellulose-based materials
[128]. Its simplicity makes it, however, sensitive to the determination of the yield
point, which is discussed by Erkkilä et al. [129], where different empirical plasticity
models are reviewed. Commonly, the offset strain is set to be 0.1 or 0.2 % for
metals but may differ for paper materials. Several approaches to describing a
paperboard from a continuum perspective have been performed in the literature.
It gets more complicated if the model should also account for damage
mechanisms, such as delamination. Since delamination is crucial for converting
operations, damage mechanisms must be considered when modeling paperboard;
consequently, some damage criterion might be needed within the model. This can
be introduced in different ways; one method is to introduce cohesive interfaces
between the plies of the paperboard, which for instance, is done by Xia et al. [130]
in 2002, Nygårds et al. [41] in 2009. Another method is to introduce continuum
damage. This approach was used by Isaksson et al. [131] in 2004, who included a
gradient-enhanced damage model in the constitutive relations when describing
the material behavior of test liner and kraft liner. Borgqvist et al. [132] investigate
paperboard material through a continuum approach, accounting for non-linear
kinematics and large strains. To account for anisotropy, they introduce three
structural vectors corresponding to the principal directions of a paperboard. By
letting the in-plane vectors deform as line segments and the out-of-plane direction
deform as a normal vector, they enable the decoupling of the in-plane and the
out-of-plane responses in shearing. Stenberg presents an elastic-plastic material
model for the out-of-plane mechanical behavior, which enables simulations under
high compressive loads in ZD. Stenberg [133] uses the peak load to identify
critical stress due to difficulties determining the load level where the deformation
changes from elastic to elastic-plastic. It should be noted that these models are
quite complex and require a large number of material input parameters.
Additionally, the complexity and the number of required parameters increase
further if moisture and temperature are included [120, 134].

Section 2.1 mentions some FE simulations on packages reported in the literature.


However, the literature also shows several studies where specific test procedures
on paperboard are simulated. Creasing, folding, and short-span compression tests
are such experiments that have thoughtfully been investigated by using FE
simulations. Borqvist et al. [132] simulate creasing using the aforementioned
approach, and later investigate the relation between folding and SCT [135].
Robertsson [136] recently developed a rate-dependent paperboard model with
Borgqvist’s model as a starting point when investigating creasing and folding.
Folding and SCT have been investigated by other authors as well. Hagman and
Nygårds [78] used a layered continuum model, described by a Hill criterion with
cohesive interfaces when evaluating SCT on multi-plied paperboard. A similar

25
approach was used by Huang and Nygårds when simulating creasing [69] and
folding [137], or the combined scenario, which was evaluated by Huang et al. [138].
Beex and Peerlings [139] analyzed creasing and folding by an orthotropic elasto-
plastic continuum model and a delamination model between the plies in order to
enable an opening behavior between the plies, and Giampieri et al. [140]
investigated the mechanical response of creasing and folding. In their study, they
developed a constitutive model where a damage variable is introduced, related to
the development of delamination inside the creased region. Another scenario that
has been investigated is deep drawing of paperboard. Wallmeier et al. [141] use an
anisotropic material model that accounts for temperature. Furthermore, they
proposed a plasticity model with a rectangular yield surface as a starting point.
However, since the in-plane shear stress reduces the size of the yield surface, they
introduce four equations that reduce the rectangle when increasing the yield stress
in shear.

Simon [142] summarizes the complexity of paperboard well in a paper in 2020,


where he reviews recent trends and challenges in the computational modeling of
paperboard: the material behavior of paperboard is nonlinear, anisotropic, and
dependent on time, loading history, and strain rate. Additionally, it is sensitive to
moisture and temperature. All these aspects have been accounted for in different
models. However, Simon concludes that no model exists (yet) capable of
accounting for all these effects simultaneously.

One of the problems with advanced FE models describing the material behavior
in detail is that the model might be too complex to manage for other users. During
Simon’s review of elasto-plasticity models, he notes in-plane models with up to 28
non-zero parameters [142, p. 2414]. Hence, pursuing good accuracy for the
simulations compared to experimental data might be counterproductive. An
alternative approach is to use as straightforward material models as possible,
based on measurable input parameters, and use FEM for varying design,
geometry, and load case. From this, it is possible to gain knowledge about the
material parameters affecting the behavior of the package, and deformation
mechanisms in general, instead of focusing on simulating a specific load case as
accurately as possible.

26
3 Materials and Methods
To investigate the relations between material properties and their impact on the
performance of packages, the procedure of this thesis is categorized into three
different main areas:

• Material characterization of paperboard.


• Package testing (BCT and torsion test).
• FE simulations of BCT and torsion test.

3.1 Material characterization

As mentioned in Chapter 2.5, material data is crucial for performing accurate


FE simulations. For some materials, this can be found in table values. However,
there are only a few published studies with material data on paperboard, which
motivated the purpose of Paper A. In Paper A, in which the testing procedures
are described more thoughtfully, a material characterization was performed on
five commercial multi-plied paperboards with different basis weight but from the
same paperboard machine, to achieve material data. The experiments were
generally performed at moisture levels 20, 50, 70, and 90 % RH at 23 °C, and the
test pieces were, prior to testing, conditioned for 24 hours to ensure equilibrium.
The in-plane tension and compression properties were further investigated for
different levels of RH in Paper B, in which several paperboards from four separate
manufacturers were evaluated.

3.1.1 In-plane tensile test


The material properties, tensile strength, tensile stiffness, and strain at break were
obtained from an in-plane tensile test. Following ISO standard [59] with an
Alwetron TH1 apparatus, developed by L&W [143], a specimen with dimensions
length, L = 100 mm, and width, b = 15 mm, was subjected to a displacement-
controlled test at a rate of 100 mm/mm. The force was measured with a load cell
to obtain a force-elongation curve, from which stress, , and strain, , were
evaluated according to

𝜎=
𝐹
= ,
𝐹 (12)
𝐴 𝑏𝑡

and
𝛿 (13)
𝜀=𝐿,
where t is the thickness of the board and δ is the elongation of the specimen. The
elastic modulus was calculated following ISO standard [59].

In Paper D, it was of interest to evaluate the impact of creasing depth on tensile


strength. Therefore, short-span tensile tests were performed on the creases. The
length of the specimens was 10 mm, and the width was 15 mm. A universal
testing machine (UTM) from MTS [144] was used, where the clamps, shown in
Figure 21, were equipped with a cylindrical-shaped grip to achieve a single contact
line between the clamps and the paperboard. Furthermore, these clamps were
operated with pneumatics to get an equal clamping pressure (6 bar) for each
specimen.

27
Figure 21. The clamps used for short-span in-plane tensile tests.

3.1.2 In-plane compression test


Short-span compression tests (SCT) were performed to evaluate the in-plane
compression strength. The tests were performed using an L&W [143] SCT
equipment, in MD and CD, for each level of RH. The width of the specimens
was 15 mm, and the length between the clamps was 0.7 mm. The length between
the clamps was reduced during the procedure, and the maximum force was
recorded.

3.1.3 Out-of-plane tension


A testing procedure, well-described by Nygårds [83], evaluated the out-of-plane
tensile properties. In short, a 40 mm × 40 mm sandwich construction was created.
The core consisted of a paperboard surrounded by tissue paper, with outer layers
of plastic laminate sheets. This was done at each RH level. The sandwich ran
through a laminator at 142 °C. In the lamination process, the conditioned
paperboard was therefore encapsulated. Here, it should be mentioned that the
temperature of the laminator presumably reduced the moisture within the
paperboard. The specimen was then glued between two rigid cylinders with a
diameter of 38 mm, which, after the glue set, were mounted in a UTM from
MTS [144] and then separated with a displacement-controlled procedure. Before
the procedure, all leftover material outside the cylinder was removed. This was
performed just before the test to minimi e the paperboard’s contact with the
surrounding environment.

3.1.4 Out-of-plane shear test


To determine the out-of-plane shear stiffness, a rigid block shear test was
performed at 50 % RH at 23 °C. This test was thus not performed at different
levels of RH. The experimental setup consisted of two steel plates and a
21 × 70 mm2 specimen glued in between them, according to Figure 22, mounted
in a uniaxial tensile tester [144]. The blocks were exposed to a tensile force and
allowed failure in any layer of the paperboard.

28
Specimen
Figure 22. Conceptual illustration of rigid block shear test.

A double-notched shear test is a more efficient way of determining a paperboard’s


out-of-plane shear strength, as mentioned in section 2.3. By varying the notch
depth through the paperboard, it is possible to obtain shear profiles of the
paperboard. Due to the unevenness of the surfaces, the minimum depth from the
surfaces was 30 µm. Each test consisted of 13 specimens with a width of 15 mm,
different notch depth positions, z, of the notches, and L = 5 mm between the
notches, which gives the shear zone an area of 5×15 mm2. For each board and
level of relative humidity, five samples were tested. The specimens were then
subjected to a standard testing procedure for in-plane tension tests in the
equipment: Alwetron Th1 tensile strength tester, developed by L&W [143].

3.1.5 Creasing and folding


In Paper A, folding was performed on an uncreased board to evaluate the
bending stiffness and the maximum bending force. The specimens (width 25 mm,
length 70 mm, and bending length 20 mm) were subjected to two-point bending
in a folding apparatus from Marbach [145]. The equipment measured the bending
force as a function of the deflection angle from 0° to 90°. In Paper D, the
creasing depth was evaluated, and a Marbach [145] crease press equipment was
used to achieve the creases. The apparatus performs creasing at low speed, and
the deformation and damage were considered quasi-static compared to the high-
speed machines used in industrial creasing.

3.2 Packaging testing

As previously mentioned, packages are exposed to several external loads, for


instance, during handling and transportation. With FEA, it is possible to simulate
situations such as these and understand what damage mechanisms are activated.
However, to validate the simulations and evaluate if the simulations are accurate,
physical packaging testing is required. In this thesis, a package with the geometry
according to Figure 23 was exposed to two different load cases. In Paper C, the
stacking performance is evaluated by performing box compression tests. As
discussed in chapter 2.1.1, this load case is probably the most investigated on the
packaging level. Additionally, due to the well-documented importance of creases
during BCT, it was essential to investigate which damage mechanisms were
activated when the package was exposed to shear. Therefore, a less investigated
load case was introduced in Paper E: torsion loading of packages.

29
(a) (b)
Figure 23. Drawing of the blank (a) and mounted package (b).

3.2.1 Box compression test


The paperboard packages were compressed between two parallel steel plates in a
uniaxial MTS system, see Figure 24. The tests were performed in a controlled
climate room at 50, 70, and 90 % RH at 23 °C. From the tests, force-compression
curves were obtained using an MTS load cell.

Figure 24. Test equipment for BCT.

3.2.2 Torsion test


The packages were subjected to torsional loading at standard conditions (50 %
RH at 23 °C) in a combined axial and torsional tester (ZwickRoell Machine type
Z005 [146]), capable of combined axial and torsional loading. The packages were
clamped with 3-point-chuck clamps, shown in Figure 25 (c). Two wooden plates
with a radius of 30 mm were attached to the package: one at the top and one at
the bottom. They were attached with melted glue to ensure that the clamps did
not affect the panels or the corners of the package. The plates were engraved with
a laser cutter to facilitate the alignment of the wooden plates, following the pattern

30
in Figure 25 (b). The package was then placed within the square. After the glue
set, the specimen was fixed into the torsion tester, as shown in Figure 25 (a).

(a) (b) (c)


Figure 25. (a) Test setup for torsion test with a detailed view of the: (b) pattern of a wooden
plate attached to the package (c) clamps used for torsion test.

The mounted package was exposed to a rotation with a loading rate of 1°/s. The
torque was measured by the load cell of the equipment. To maintain zero axial
force, avoid axial deformation, and achieve a state of loading approaching simple
shear; axial load control was used. Thus, as the package was clamped at both ends
and became susceptible to compressive buckling under torsional loading, a tensile
load was applied to negate this such that the net axial force on the box was zero.

3.3 Finite element model

A geometry was created using LS-PrePost [147]. The same geometry was used for
BCT in Paper C and the torsion test in Paper E and followed the measures and
material directions from Figure 23 with a sheet thickness, t. A transparent view of
the FE model is shown in Figure 26, where the modeled creases are green, the
panels yellow, and the flaps orange, analogous to Figure 23. At the top and
bottom, the three small flaps are visible under one large flap. The manufacturer’s
joint is seen at the back of the right front panel.

31
Figure 26. Transparent view of the FE model.

In Paper C, the paperboard was modeled as a homogeneous material, neglecting


the layered structure of the material. Due to this, the model could not
simultaneously represent the bending stiffness and the in-plane stiffness
accurately. Here, the in-plane stiffness was prioritized to reduce the errors at the
initial stage of compressive loading. The advantage of this approach was that the
material characterization was performed at the full board level. Hence, the input
parameters, such as the in-plane stiffnesses and stresses, were quite easily
characterized. In Paper E, however, the effect of the bending stiffness was
investigated, and it was, therefore, necessary to introduce a laminate structure
considering the properties of the different plies. In short, this was achieved by
utilizing a relationship between material property and density, defined by Nygårds
2022 [34], shown in Eq. (14) for the strength of the top ply as an example. Hence,
no material characterization was performed on the separate plies. Instead, the
paperboards were split with a Fortuna splitter [148], and the density was
measured.
𝜌top (14)
𝜎top = 𝜎
𝜌full full

The material was modeled as an elastic orthotropic material with an additional


stress-based failure criterion, using Mat_054 in LS Dyna [149] and is further
described in Paper C and Paper E. The elements were four-noded, fully
integrated shell elements with four in-plane integration points and four
integration points through the shell thickness. The problems were solved
implicitly using the Massively Parallel Processing (MPP) solver. The input
parameters for one of the paperboards described as a homogeneous material (i.e.,
characterized as a full board) are shown in Table 1.

32
Table 1. Material parameters for Paperboard A.

Property Paperboard A
Density [kg/m3] 764
EMD [MPa] 5810
ECD [MPa] 2980
σfMD [MPa] 59
σfCD [MPa] 35
σf,cMD [MPa] 59
σf,cCD [MPa] 35
GMD/ZD [MPa] 16.50
GCD/ZD [MPa] 16.50
GMD/CD [MPa] 1440

As mentioned, the creases are essential for simulating packages. Here, the
properties of the creases were reduced by multiplying all stiffness parameters with
the relative crease strength (RCS), previously described in section 2.3.1.
Additionally, a similar measure was used to reduce the tensile properties of the
creases: RTS, i.e., the ratio between the tensile strength of the creased and
uncreased board. Hence, the properties of the creases were reduced with measured
values. No fitting parameters were used.

33
This page intentionally left blank

34
4 Results and discussion
In Paper A, a thorough material characterization was performed on five
commercial paperboards with different basis weights but from the same
paperboard series. The purpose was primarily to achieve material data for the FE
simulations, but the effect of moisture was also investigated. The material
properties were normalized with the value at standard climate (50 % RH, 23 °C),
and linear relations between normalized mechanical properties and moisture ratio
were observed. Since all the paperboards in the study were from the same
manufacturer, extending the study to several manufacturers was natural. This was
done in Paper B, where paperboards with significantly different furnish and
construction were consciously chosen to examine how their in-plane mechanical
properties were affected by moisture. The result for the normalized strength is
shown in Figure 27, where each color represents a manufacturer.

Figure 27. Normalized strength as a function of moisture ratio for paperboards from four
different paperboards.

However, when both the material properties and the moisture ratio were
normalized, all paperboards coincided with the same master curve, shown in
Figure 28, where data for both MD and CD are included.

35
Figure 28. Normalized tensile strength as a function of normalized moisture ratio for 15
paperboards in both MD and CD.

The physical interpretation from the results in Figure 28 is that if any paperboard’s
moisture is doubled, the strength would be reduced by approximately 35 %. Thus,
it was possible to express any of the investigated mechanical properties as a linear
equation:

𝑃𝑟𝑜𝑝𝑒𝑟𝑡𝑦 𝑚𝑟 (15)
= 𝑎𝑝𝑟𝑜𝑝𝑒𝑟𝑡𝑦 + 𝑏𝑝𝑟𝑜𝑝𝑒𝑟𝑡𝑦 .
𝑃𝑟𝑜𝑝𝑒𝑟𝑡𝑦50% RH 𝑚𝑟 50% RH
Analogously, this was also performed for the in-plane stiffness and the SCT
values, which were the mechanical properties investigated in Paper B. The
parameters a and b for the linear relations are summarized in Table 2.

Table 2. Parameters for the linear relation between normalized property and normalized
moisture ratio.

Normalized property a b R2
Strength -0.35 1.36 0.96
Stiffness -0.37 1.40 0.96
SCT -0.51 1.51 0.96

Combining Eq. (15) and the parameters a and b in Table 2, it is possible to


estimate the in-plane parameters at any moisture level by knowing the
moisture ratio for the specific moisture level and the mechanical property at
standard climate. This is valid for the investigated moisture range (i.e.,
corresponding to levels of 20 % to 90 % RH). Furthermore, the results in
Paper A and Paper B indicate that this is also valid for other mechanical
parameters, but the parameters a and b will have different values. Hence, it is
possible to estimate the mechanical properties at different moisture levels without
performing any mechanical tests, if the properties at standard condition are
known.

These results are significant since it means that if an orthotropic material model
based on the in-plane stiffness and strength is used in FEM, it is also possible to
predict the package response at different moisture levels. This was shown in

36
Paper C, where physical BCT experiments were compared to FE simulations at
different moisture levels. The simulations were based on measured properties at
standard climate, and Eq. (15) and Table 2 when simulating the different moisture
levels. The result is shown in Figure 29.

Figure 29. Comparison of experiments and simulations for different moisture levels.

There are several things interesting to note here. First, the BCT values are
accurately predicted using a simple material model with few input parameters. An
overprediction is expected since FEM does not account for structural and
material imperfections.

Secondly, it is interesting that the stiffness response of the simulated compression


of the package is similar to the experimental data, based on the approach of
reducing the properties of the creases and the fact that no curve fitting has been
applied. Apart from the initial parts of the curves, which are explained by the
presence of horizontal creases and perturbations of top and bottom flaps during
the physical experiments, the predictions of the curves are quite accurate.
The RCS and RTS values, mainly described in Paper D, reduced the creases’
properties, where their dependence on shear strains caused by creasing depth was
investigated. When evaluated against shear strength during creasing, the RCS and
RTS values formed a creasing window. The RTS values corresponded to in-plane
cracks (upper limit), and the RCS values corresponded to delamination damage
(lower limit). It was observed that both the lower limit and the upper limit exhibit
linear relations against shear strain. It should also be mentioned that there were
no problems with creasing cracks in the packages. This correlates to the creasing
window developed in Paper D since the creasing depth used is located within the
creasing window.

Thirdly, the impact of moisture on the packaging level can be implemented by a


linear master curve developed in Paper B. It should again be emphasized that the
input parameters are measured at 50 % RH and then estimated with Eq. (15) for
other moisture levels. Finally, trust in the model and the crease reduction method
was further investigated by simulating the same package geometry but using
another (thicker) paperboard material, and the results were adequately accurate.

37
Additionally, when BCT was analyzed further in Paper E, the material was
considered a three-ply laminated structure to investigate the effect of bending
stiffness. This effect was shown to be very small, illustrated by Figure 31, where
FE model 1 is the homogeneous model, calibrated for tensile stiffness and FE
model 2 is a three-ply laminated structure, calibrated for tensile and bending
stiffness.

Figure 30. Comparison of BCT curves for the experiments and the two FE models plotted up
to the first peak load for Paperboard A.

From this, two conclusions were drawn. First, the bending stiffness did not
significantly impact the critical force during BCT. Instead, it was shown that the
strength of the material had a more considerable impact. Secondly, determining
the mechanical properties of the separate plies, i.e., by calculating them from
Eq. (14), was representative. This is an essential observation since material
characterization on the ply level is quite complex.

All these observations are relevant if the purpose of the simulations is to predict
BCT for different moisture levels. However, the advantage of FEM is predicting
the compression strength of packages and observing damage mechanisms
during the procedure. Paper C results show that the maximum stresses coincide
with the occurrence of yield lines in the packages, see Figure 31. The results
support the theory that the package’s failure can be explained by local in-plane
failure in the elements close to the yield line, as analytically described by
Ristinmaa et al. [11].

38
(a) (b)
Figure 31. Comparison between FE simulation (a) and physical experiment (b).

A series of close-up photos of the permanent deformation at the corners of the


packages was captured at peak load. These are shown in Figure 32, where
surface cracks are observed in the corner region.

(a) (b)

(c) (d)
Figure 32. Close-up photos of the damaged package at peak load during BCT.

39
In Figure 32 (a), the parabolic yield line is seen. This yield line was investigated
further by cutting the package, as the red line indicates in Figure 33. A cross-
section photo showed that the yield line is caused by delamination in the interface
between the bottom and the middle ply. Similar delamination behavior has
previously been seen by Eriksson et al. [150].

Figure 33. Cross-section photo of the yield line at peak load during BCT.

Due to the well-documented importance of creases during BCT, it was interesting


to investigate which damage mechanisms activated when the package was
exposed to shear. Therefore, a less investigated load case was introduced in
Paper E: torsion loading of packages. For this load case, it was shown that the
model accurately predicted the experimental curve, with some over-prediction in
torque shown in Figure 34.

Figure 34. Experimental and FE simulations torque-twist curves for Paperboard A


packages.

The experimental deformation of a package was captured in a 3D scan at twists


of 0°, 1°, 2°, and 5°, where 5° was post-peak deformation. The scan was
performed with an EINSCAN SP [151] system at the corresponding steps, shown
in Figure 35, which illustrates the total deformation of two panels of a package
from 0.0 mm (blue) to 0.5 mm deformation (red).

40
Figure 35. 3D scanned images of a package at different stages of twist.

Figure 35 can be compared to the stress fields in MD and CD at peak torque,


expressed in MPa, shown in Figure 36 for Paperboard A. Here, "X-stress" and "Y-
stress" correspond to MD and CD stresses, respectively.

(a) (b)
Figure 36. Stresses in MD (a) and CD (b) for Paperboard A expressed in MPa.

The comparison between physical experiments and FE simulations concludes


that even if the simulations overpredicted the maximum torque and twist at
failure, the torque vs. twist curves were accurate for both paperboard qualities.
Hence, despite its lack of complexity, the model can also predict the package’s
response in compression (BCT) and the response when it is exposed to shear.

For both load cases, failure occurs in in-plane tension, not in in-plane
compression or in-plane shear. However, it should be emphasized that the model
does not account for out-of-plane properties in the failure criterion. When the
out-of-plane shear stresses are analyzed in the torsion test, the stress at maximum
torque reaches values around 1 MPa. This is in the same range as the measured
out-of-plane shear strength in Paper A. Thus, it would be interesting to further
investigate the effect of the transverse shear properties.

However, Johansson et al. [152] have recently observed that paperboard deforms
heavily in the thickness direction before in-plane failure, contributing to bond
failure in the paperboard, i.e., out-of-plane deformations are activated during
tensile loading.

41
This page intentionally left blank

42
5 Conclusions
This thesis investigated the impact of material parameters on packaging
performance. The analysis has been divided into material characterization,
packaging testing, and finite element simulations. From the initial investigation of
the material characterization, it was concluded that different mechanical
properties normalized with the properties at standard climate decrease linearly
when the moisture ratio of the paperboard is increased. The study was extended
to include paperboards from different manufacturers. It was concluded that when
the moisture ratio is normalized with the moisture ratio at standard climate, the
results from all investigated paperboards to a good approximation, followed the
same linear relation. This made it possible to estimate a mechanical property for
any board at different moisture levels by knowing only the following:

• the mechanical property for standard climate (50 % RH and 23 °C), and
• the parameters a and b that describe the linear relation.

This linear relation was used to estimate material input parameters for simulating
a Box Compression Test (BCT) at different moisture levels using an orthotropic
material model with a failure criterion, i.e., a quite simple material model with few
input parameters. The result showed that it was possible to accurately predict the
load-compression curve of a BCT when accounting for moisture. Furthermore,
it was concluded that modeling the creases’ mechanical properties is vital for
capturing the stiffness response of the package. Here, a measurable approach for
reducing the creases’ mechanical properties was suggested, based on a folding test
to obtain the relative creasing strength (RCS) and a short-span tensile test to
obtain the (RTS).

Since creases have an evident effect on the packaging performance from a


stacking point of view, it was interesting to investigate a load case where the
package was exposed to shear. An additional load case was therefore investigated:
torsion of paperboard packages, where the experimental data was accurately
predicted.

The effect of bending stiffness was investigated for both load cases by developing
two FE models. Model 1 treated the paperboard as a homogeneous material, and
Model 2 considered the paperboard to be a three-ply laminate structure. No
significant effect was noted, and it was concluded that the strength has a more
significant effect on the BCT than the bending stiffness.

As a final remark, the main procedure here is developing an easy-to-use model


with few material parameters that can predict the load-deformation curves for
two different load cases. The purpose of the model is not to be used for the
precise prediction of failure loads but to gain knowledge about damage
mechanisms during testing procedures. A clear advantage of this approach is that
the model can be used to either change the package’s geometry or perform a
parameter study on the relevant material parameters. This can also be varied for
each ply separately. It has also been shown that the model can account for
different moisture levels if applying the master curve developed within this thesis.
Finally, it should be emphasized that the model does not include any fitting
parameters. All input data is based on measured values.

43
This page intentionally left blank

44
6 Future work
The main advantage of the approach used in this thesis is the model’s simplicity.
However, for future studies, it would be of interest to explicitly include the out-
of-plane shear properties and investigate their impact on the package’s
performance. This will require modification in the material description since the
current material model does not account for out-of-plane shear or strength.
epending on the user’s needs, additional modifications might be interesting to
investigate, such as imperfections of the geometry and the introduction of
plasticity. From a paperboard manufacturer’s perspective, the laminate model,
developed in Paper E, is highly relevant since it will be possible to change the
properties of the separate plies individually. It is simultaneously easy to change
the package design, making this a valuable tool for evaluating the impact of
material properties and packaging performance. From a converting point of view,
it would be interesting to run complementing package simulations with different
reduction ratios of the creases based on the creasing window developed in
Paper D. A final suggestion is to investigate the impact of temperature. This has
not been included in this thesis, but it would be interesting to evaluate packaging
performance at different temperatures for some applications.

45
This page intentionally left blank

46
7 Bibliography
[1] European commission, "A European Strategy for Plastics in a Circular
Economy," 2018. [Online]. Available: https://eur-
lex.europa.eu/resource.html?uri=cellar:2df5d1d2-fac7-11e7-b8f5-
01aa75ed71a1.0001.02/DOC_1&format=PDF. [Accessed 20 September
2020].

[2] European parlament, 2019. [Online]. Available: https://eur-


lex.europa.eu/legal-
content/EN/TXT/PDF/?uri=CELEX:32019L0904&from=EN.
[Accessed 20 September 2020].

[3] R. Eckhart, "Recyclability of Cartonboard and Carton," Wochenblatt für


Papierfabrikation, vol. 11, 2021.

[4] S Smith, ”The origin of corrugated cardboard,” 2018. Online .


Available:
https://www.dssmith.com/tecnicarton/about/newsroom/2018/11/the
-origin-of-corrugated-cardboard. [Använd 12 10 2022].

[5] McKinsey & Company, "Pulp, paper, and packaging in the next decade:
Transformational change," 7 August 2019. [Online]. Available:
https://www.mckinsey.com/industries/paper-forest-products-and-
packaging/our-insights/pulp-paper-and-packaging-in-the-next-decade-
transformational-change. [Accessed 11 January 2023].

[6] Metsä Group, "The global packaging market," 2022. [Online]. Available:
https://www.metsagroup.com/metsaboard/investors/operating-
environment/global-packaging-market/. [Accessed 11 January 2023].

[7] B. Frank, ” orrugated bo compression - a literature survey,” Packaging


technology and science, vol. 27, pp. 105-128, 2014.

[8] G. Marin, P. Srinivasa, M. Nygårds and S. Östlund, "Raw data from:


Experimental and FE-simulated box compression tests on paperboard
packages at different moisture levels," Packaging Technology and Science, vol.
34, no. 4, pp. 229-243, 2021.

[9] R. C. McKee, J. W. Gander and J. R. Wachuta, “ ompression strength


formula for corrugated bo es,” Paperboard Packaging, vol. 48, no. 8, pp.
149-159, 1963.

[10] . Grangård, ” ompression of board cartons. 1. orrelation between


actual tests and empirical equations,” Svensk Papperstidning - Nordisk
Cellulosa, vol. 73, nr 15, p. 462, 1970.

[11] M. Ristinmaa, M. Ottosen and C. Korin, "Analytical Prediction of Package


Collapse Loads - Basic Considerations," Nordic Pulp and Paper Research
Journal, vol. 27, no. 4, pp. 806-813, 2012.

[12] T. Nordstrand, M. Blackenfeldt and M. Renman, "A Strength Prediction


Method for Corrugated Board Containers," Report TVSM-3065, Div. of
Structural Mechanics. Lund University, Sweden, 2003.

[13] S. W. Tsai and E. M. Wu, "A general theory of strength for anisotropic
materials," Journal of Composite Materials, vol. 5, no. 1, pp. 58-80, 1971.

47
[14] M. E. Biancolini and C. Brutti, "Numerical and Experimental
Investigation of the Strength of Corrugated Board Packages," Packaging
Technology and Science, vol. 16, pp. 47-60, 2003.

[15] J. Han and J. M. Park, "Finite Element Analysis of Vent/Hand Hole


Designs for Corrugated Fibreboard Boxes," Packaging Technology and Science,
vol. 20, pp. 39-47, 2007.

[16] T. Fadiji, C. Coetzee and U. L. Opara, "Compression strength of


ventilated corrugated paperboard packages: Numerical modelling,
experimental validation and effects of vent geometric design," Biosystems
engeneering, vol. 151, pp. 231-247, 2016.

[17] M. A. Jiménez-Caballero, I. Conde, B. García and E. Liarte, "Design of


Different Types of Corrugated Board Packages Using Finite Element
Tools," SIMULIA customer conference, 2009.

[18] T. Garbowski, T. Gajewski and J. K. Grabski, "Estimation of the


Compressive Strength of Corrugated Cardboard Boxes with Various
Perforations," Energies, vol. 14, 2021.

[19] T. Garbowski, T. Gajewski and J. K. Grabski, "Estimation of the


Compressive Strength of Corrugated Cardboard Boxes with Various
Openings," Energies, vol. 14, pp. 155-175, 2021.

[20] L. Beldie and G. Sandberg, "Paperboard packages exposed to static


loads—finite element modeling and experiments," Packag Technol Sci., vol.
14, no. 4, pp. 171-178, 2001.

[21] V. D. Luong, F. Abbès, J. B. Nolot and D. Erre, "Experimental


Characterisation and Finite Element Modelling of Paperboard for the
Design of Paperboard Packaging," IOP Conf. Ser.: Mater. Sci. Eng., vol. 540,
2019.

[22] P. Mäkelä and S. Östlund, "Orthotropic elastic-plastic material model for


paper materials," International Journal of Solids and Structures, vol. 40, pp.
5599-5620, 2003.

[23] V. . Luong, ”Étude e périmentale et modélisation numérique des


phénomènes d’endommagement par fatigue des emballages IN
FREN ,” PhD thesis at Université de Reims Champagne-Ardenne, pp. 101-
130, 2020.

[24] M. Zaheer, M. Awais, L. Rautkari and J. Sorvari, "Finite element analysis


of paperboard package under compressional load," Procedia manufacturing,
vol. 17, pp. 1162-1170, 2018.

[25] A. Hagman, B. Timmerman, M. Nygårds and A. Lundin, "Experimental


and numerical verification of 3D forming," Advances in Paper Science and
Technology: Trans. 16th Fundamental Research Symposium, Oxford, UK, vol. 1,
pp. 3-26, 2017.

[26] M. Upadhyaya and M. Nygårds, "A finite element model to simulate brim
forming of paperboard," 28th IAPRI Symposium on packaging 9–12 May,
2017, Lausanne, Switzerland, pp. 395-408, 2017.

48
[27] J. E. Gustafsson and M. Nygårds, "Loading and deformation of cigarette
packages—experimental and numerical comparison of paperboards," 28th
IAPRI Symposium on packaging, 9–12 May, 2017, Lausanne, Switzerland, 2017.

[28] M. Nygårds, S. Sjöqvist, G. Marin and J. Sundström, "Simulation and


experimental verification of a drop test and compression test of a gable
top package," Packag Technol Sci., vol. 32, no. 7, pp. 325-333, 2019.

[29] K. Niskanen(ed), Mechanics of paper products, Walter de Gruyter GmbH


& CO, 2012.

[30] A. e. a. de Ruvo, ”The influence of raw material and design on mechanical


performance of bo board,” Svensk papperstidning, vol. 18, pp. 557-566,
1978.

[31] C. Fellers, Pulp and Paper Chemistry and Technology: Volume 4 Paper
Products Physics and Technology, M. Ek, G. Gellerstedt and G.
Henriksson, Eds., Göttingen, Germany: Walter de Gruyter GmbH & CO,
2009.

[32] P. T. Larsson, T. Lindström, L. A. arlsson and . Fellers, “Fiber length


and bonding effects on tensile strength and toughness of kraft paper,” J
Mater Sci, vol. 53, pp. 3006-3015, 2018.

[33] L. A. Carlsson and T. Lindström, "A shear-lag approach to the tensile


strength of paper," Composites Science and Technology, vol. 65, pp. 183-189,
2005.

[34] Nygårds, ”Relating papermaking process parameters to properties of


paperboard with special attention to through-thickness design,” MRS
Advances, 2022.

[35] S. Norgren, G. Pettersson and . öglund, “Strong paper from spruce


CTMP - Part II: Effect of pressing at nip press temperatures above the
lignin softening temperature,” Nordic Pulp & Paper Research Journal, vol. 33,
no. 1, pp. 142-149, 2018.

[36] G. Pettersson, S. Norgren, P. Engstrand, M. Rundlöf and H. Höglund,


"Aspects on bond strength in sheet structures from TMP and CTMP - a
review," Nordic Pulp & Paper Research Journal, vol. 36, no. 2, pp. 177-213,
2021.

[37] D. Coffin and M. Nygårds, "Creasing and folding," Advances in Paper Science
and Technology: Trans. 16th Fundamental Research Symposium Oxford 2017, pp.
69-136, 2018.

[38] F. E. Shelton e, ”Improvement in Paper Bo es, US Patent 183423,” Oct


17, 1876.

[39] L. Carlsson, A. De Ruvo and C. Fellers, "Bending propreties of creased


zones of paperboard related to interlaminar defects," Journal of materials
science, vol. 18, pp. 1365-1373, 1983.

[40] J. F. Halladay and R. W. K. Ulm, "Creasing and bending of folding


boxboard," Paper Trade Journal, Tappi Section, vol. 108, no. 5, pp. 36-40,
1939.

49
[41] M. Nygårds, M. Just and J. Tryding, "Experimental and numerical studies
of creasing of paperboard," International Journal of Solids and Structures, vol.
46, pp. 2493-2505, 2009.

[42] S. I. Cavlin, "The Unique Convertibility of Paperboard," Packaging


technology and science, vol. 1, pp. 77-92, 1988.

[43] G. Marin, M. Nygårds and S. Östlund, "Experimental quantification of


differences in damage due to in-plane tensile test and bending of
paperboard," Packaging technology and science, vol. 35, no. 1, pp. 69-80, 2021.

[44] G. R. onaldson, ”An instrument for numerical measurement of bending


quality of boards,” Proc. Appita, vol. 8, pp. 237-250, 1954.

[45] S. Nagasawa, Y. Fukuzawa, T. Yamaguchi, S. Tsukatani and I. Katayama,


"Effect of crease depth and crease deviation on folding deformation
characteristics of coated paperboard," Journal of Materials Processing
Technology, vol. 140, pp. 157-162, 2003.

[46] S. Nagasawa, R. Endo, Y. Fukuzawa, S. Uchino and I. Katayama,


"Creasing characteristic of aluminum foil coated paperboard," Journal of
Materials Processing Technology, vol. 201, pp. 401-407, 2008.

[47] . W. offin and J. . Panek, “An Operating Window for Acceptable


reasing of Paperboard,” Progress in Paper Physics Seminars Conference
Proceeding, Darmstadt, Germany, pp. 132-137, 2016.

[48] J. C. Panek, S. D. Smith and D. W. Coffin, "Creasing severity and reverse-


side cracking," Tappi Journal, vol. 19, no. 4, pp. 219-227, 2020.

[49] Paper and board - Determination of bending stiffness by stiffness by static methods -
General principles, ISO5628:1990(E).

[50] G. Marin, M. Nygårds and S. Östlund, "Raw data from: Experimental


quantification of differences in damage due to in-plane tensile test and
bending of paperboard," Packaging technology and science, vol. 35, no. 1, pp.
69-80, 2021.

[51] Structural thickness and structural density, SCAN-P 88:01.

[52] C. Fellers, H. Andersson and H. Hollmark, in Paper structure and


properties, J. Bristow and P. Kolseth, Eds., New York: Marcel Dekker inc,
1986, p. 151.

[53] O. Schultz-Eklund, C. Fellers and P.-Å. Johansson, "Method for the local
determination of the thickness and density of paper," Nordic Pulp & Paper
Research Journal, vol. 7, no. 3, pp. 133-139, 1992.

[54] TJT-Teknik, Järfälla, Stockholm, Sweden.

[55] . Rydefalk, ”Unpublished data from Rydefalk's Ph project”.

[56] Billerud, "Billerud CrownBoard Prestige Technical Data Sheet," Billerud


AB, 12 October 2022. [Online]. Available:
https://www.billerud.com/globalassets/billerud/our-offer/packaging-
materials/cartonboard/billerud_crownboard_prestige_tech_data_sheet_
2022.10.12.pdf. [Accessed 23 November 2022].

50
[57] K. Niskanen, ”Strength and fracture of paper,” Trans. of the Xth Fund. Res.
Symp. Oxford, pp. 641-725, 1993.

[58] . L. o , “The elasticity and strength of paper and other fibrous


materials,” Br. J. Appl. Phys., vol. 3, no. 72, 1952.

[59] Paper and board - Determination of tensile properties, ISO 194-3:2005.

[60] Standard Test Method for Tensile Properties of Paper and Paperboard Using
Constant-Rate-of-Elongation Apparatus, ASTM(1997): ASTM D 828.

[61] S. Cavlin , "Sprödhetsmodul och kritisk längd," Svenska träforskningsinstitutet


(In Swedish). PA B:78 Nr 290, 1974.

[62] J. Goldschmidt and D. Wahren, "On the Rupture Mechanism of Paper,"


Svensk Papperstidning, pp. 477-481, 1968.

[63] J. Tryding, ”In-plane fracture of paper Ph Thesis,” Lund Institute of


Technology, Division of Structural Mechanics, 1996.

[64] J. Tryding, G. Marin, M. Nygårds, P. Mäkelä and G. Ferrari,


"Experimental and theoretical analysis of in-plane cohesive testing of
paperboard," International Journal of Damage Mechanics, vol. 26, no. 6, pp.
895-918, 2017.

[65] A. Hagman and M. Nygårds, "Investigation of sample-size effects on in-


plane tensile testing of paperboard," Nordic Pulp and Paper Research Journal,
vol. 27, no. 2, pp. 295-304, 2012.

[66] G. arin , ”Unpublished data from a separate study within this Ph


project”.

[67] K. Persson, ” aterial model for paper - experimental and theoretical


aspects,” Diploma report, Solid mechanics, Lund Institute of Technology, 1991.

[68] M. Nygårds, C. Fellers and S. Östlund, "Measuring Out-of-Plane Shear


Properties of Paperboard," Journal of Pulp and Paper Science, vol. 33, no. 2,
pp. 105-109, 2007.

[69] H. Huang and M. Nygårds, "A simplified material model for finite element
analysis of paperboard creasing," Nordic Pulp & Paper Research Journal, 2010.

[70] C. Fellers and S. Cavlin, "A new method for measuring the edgewise
compression properties of paper," Svensk Papperstidning, vol. 78, no. 9, pp.
329-332, 1975.

[71] C. Fellers and P. Jonsson, "Edgewise Compression strength of liner and


corrugating medium - an analysis of testmethods," Svensk Papperstidning,
vol. 78, no. 5, pp. 172-175, 1975.

[72] Paper and board - Compressive strength - Short span test, ISO 9896:1989.

[73] A. Bugiel, F. Hähnel, K. Wolf, J. Strauß and T. Kuntzsch, "Enhanced test


devices for the development of novel paper-like materials for sandwich-
structures," Trans. of the XVIth Fund. Res. Symp. Oxford, pp. 27-41, 2017.

51
[74] A. Brandberg and A. Kulachenko, "Compression failure in dense non-
woven fiber networks," Cellulose, vol. 27, pp. 6065-6082, 2020.

[75] A. Hagman, H. Huang and M. Nygårds, "Investigation of shear


introduced failure during SCT loading of paperboards," Nordic Pulp &
Paper Research Journal, vol. 28, no. 3, pp. 415-429, 2013.

[76] C. Fellers and B. C. Donner, Edgewise Compression Strength of Paper


In:Handbook of Physical Testing of Paper, Second Ed ed., New York,
Basel: Marcek Dekker, Inc., 2002, p. 496.

[77] C. C. Habeger and W. J. Whitsitt, "A Mathematical Model of Compressive


Strength in Paperboard," Fibre Science and Technology, vol. 19, pp. 215-239,
1983.

[78] A. Hagman and M. Nygårds, "Short compression testing of multi-ply


paperboard, influence from shear strength," Nordic Pulp & Paper Research
Journal, vol. 31, no. 1, pp. 123-134, 2016.

[79] G. A. Baum, D. C. Brennan and C. C. Habeger, "Orthotropic elastic


constants of paper," Tappi journal, vol. 64, no. 8, pp. 97-101, 1981.

[80] N. Stenberg and C. Fellers, "Out-of-plane Poisson’s ratios of paper and


paperboard," Nordic Pulp and Paper Research Journal, vol. 17, no. 4, pp. 387-
394, 2002.

[81] G. P. Van Liew, ”The deformation of paper in the Z-direction,” PhD thesis,
The institute of Paper Science and Technology, Atlanta, Georgia, 1973.

[82] O. Girlanda and C. Fellers, "Evaluation of the tensile stress-strain


properties in the thickness direction of paper materials," Nordic Pulp and
Paper Research Journal, vol. 22, no. 1, pp. 49-56, 2007.

[83] . Nygårds, ”E perimental techniques for characteri ation of


elasticplastic material properties in paperboard,” Nordic Pulp and Paper
Research Journal, vol. 23, nr 4, pp. 432-437, 2008.

[84] N. Stenberg, C. Fellers and S. Östlund, "Measuring the stress-strain


properties of paperboard in the thickness direction," Journal of Pulp and
Paper Science, vol. 27, no. 1, pp. 213-221, 2001.

[85] N. Stenberg, C. Fellers and S. Östlund, "Plasticity in the thickness


direction of paperboard under combined shear and normal loading,"
Journal of Engineering Materials and Technology, vol. 123, pp. 184-190, 2001.

[86] D. W. Coffin, J. O. Lif and C. Fellers, "Tensile and ultrasonic stiffness of


paper at different moistures - a clarification of the differences," Nordic Pulp
and Paper Research Journal, vol. 19, no. 2, pp. 257-263, 2004.

[87] C. Rydefalk, A. Hagman, L. Yang and A. Kulachenko, "Mechanical


response of paperboard in rapid compression - the rapid ZD-tester, a
measurement technique," 17th Fundamental Research Symposium, Cambridge,
August/September, pp. 311-332, 2022.

[88] J. Johnson, P. Rättö, M. Lestelius and E. Blohm, "Measuring the dynamic


pressure in a flexographic central impression printing press," Nordic Pulp
& Paper Research Journal, vol. 19, no. 1, pp. 84-88, 2004.

52
[89] H.-J. Schaffrath and L. Göttschling, "The behaviour of paper under
compression in z-direction," Proceedings of the 1991 International Paper Physics
Conference, Hawaii, 1991.

[90] . Nygårds, ” odelling the out-of-plane behaviour of paperboard,”


Nordic Pulp and Paper Research Journal, vol. 24, nr 1, pp. 72-76, 2009.

[91] Y. Li, S. E. Stapleton, S. Reese and J.-W. Simon, "Anisotropic elastic-


plastic deformation of paper: Out-of-plane model," International Journal of
Solids and Structures, pp. 172-182, 2017.

[92] P. Rättö and M. Rigdahl, "The deformation behaviour in the thickness


direction of paper subjected to a short pressure pulse," Nordic Pulp & Paper
Research Journal, vol. 13, no. 3, pp. 180-1851, 1998.

[93] . Fellers, ”Procedure for measuring the interlaminar shear properties of


paper,” Svensk papperstidning, vol. 3, pp. 89-93, 1977.

[94] M. Nygårds, C. Fellers and S. Östlund, "Development of the notched


shear test," Transactions of the 14th Fundamental Research Symposium, Oxford,
14-18 September, pp. 877-897, 2009.

[95] H. Huang and M. Nygårds, "The Dependency of Shear Zone Length on


the Shear Strength Profiles in Paperboard," Experimental Mechanics, vol. 52,
pp. 1047-1055, 2012.

[96] M. Nygårds and J. Malnory, "Measuring the out-of-plane shear strength


profiles in different paper qualities," Nordic Pulp & Paper Research Journal,
vol. 25, no. 3, pp. 366-371, 2010.

[97] M. Nygårds and J. Sundström, "Comparison and analysis of in-plane


compression and bending failure in paperboard," Nordic Pulp and Paper
Research Journal, vol. 31, no. 3, pp. 432-440, 2016.

[98] M. Nygårds, A. Bhattacharya and S. V. R. Krishnan, "Optimizing shear


strength profiles in paperboard for better crease formation," Nordic Pulp
and Paper Research Journal, vol. 29, no. 3, pp. 510-520, 2014.

[99] N. Stenberg, "Out-of-plane shear of paperboard under high compressive


loads," Journal of Pulp and Paper Science, vol. 30, no. 1, pp. 22-28, 2004.

[100] L. Salmén, ”Temperature and water induced softening behaviour of wood


fiber based materials, PhD thesis,” Department of Paper Technology, The Royal
Institute of Technology, 1982.

[101] J. Takamura, Japan Wood Res. Soc., vol. 14, pp. 75-79, 1968.

[102] P. A. Larsson, ” ygro- and hydroexpansion of paper: Influence of fibre-


joint formation and fibre sorptivity,” KTH Royal Institute of Technology,
Department of Fibre and Polymer Technology, Division of Fibre Technology, Doctoral
Thesis, 2010.

[103] L. Salmén and P. A. Larsson, "On the origin of sorption hysteresis in


cellulosic materials," Carbohydrate Polymers, vol. 182, pp. 15-20, 2018.

[104] Paper, board and pulps — Standard atmosphere for conditioning and testing and
procedure for monitoring the atmosphere and conditioning of samples, ISO 187:2022.

53
[105] R. E. Benson, ”Effect of relative humidity and temperature on tensile
stress-strain properties of kraft linerboards,” Tappi Journal, vol. 54, nr 4,
pp. 170-177, 1971.

[106] P. äkelä, ”In-plane compression properties of selected commercial


papers at different relative humidity,” Innventia Report No.: 117, November
2010.

[107] C. G. van der Sman, E. Bosco and R. H. Peerlings, "A model for moisture-
induced dimensional instability in printing paper," Nordic Pulp & Paper
Research Journal, vol. 31, no. 4, pp. 676-683, 2016.

[108] G. Marin, M. Nygårds and S. Östlund, "Raw data from: Elastic-plastic


model for the mechanical properties of paperboard as a function of
moisture," Nordic Pulp and Paper Research Journal, 2020.

[109] J. Tryding and M. Ristinmaa, "Normalization of cohesive laws for quasi-


brittle materials," Engineering Fracture Mechanics, vol. 178, pp. 333-345, 2017.

[110] R. Spiewak, G. S. Vankayalapati, J. M. Considine, K. T. Turner and P. K.


Purohit, "Humidity dependence of fracture toughness of cellulose
fibrous," Engineering Fracture Mechanics, vol. 164, 2022.

[111] G. Marin, M. Nygårds and S. Östlund, "Raw data from: Stiffness and
strength properties of five paperboards and their moisture dependency,"
Tappi Journal, vol. 19, no. 2, pp. 71-85, 2020.

[112] C. Land, L. Stolpe, T. Wahlström and L. Baghello, "MD strain and


bagginess due to calendering," Nordic Pulp and Paper Research Journal, vol.
26, no. 1, pp. 106-117, 2011.

[113] Paper and board – Determination of moisture content of a lot – Oven-drying method,
ISO 287:2017.

[114] P. äkelä, ”Analysis of the paper web during its passage through the
drying section,” International Paper Physics Conference, pp. 151-155, 2003.

[115] . Fellers, ”Z-properties for selected papers at different relative


humidity,” Innventia Report No. 2, 2009.

[116] R. E. Popil, Physical Testing of Paper, Shawbury, Shrewsbury, Shropshire,


United Kingdom: Smithers Group Company, 2017.

[117] E. L. Back, ”The relative moisture sensitivity of compression as compared


to tensile strength,” Trans. of the VIIIth Fund. Res. Symp. Oxford, pp. 497-
509, 1985.

[118] D. W. Coffin and C. Fellers, "Paper: Creep," Encyclopedia of Materials :


Science and Technology, p. 6656–6659, 2001.

[119] L. Salmén and E. L. Back, "Effect of temperature on stress-strain


properties of dry papers," Svensk papperstidning, vol. 10, pp. 341-346, 1978.

[120] E. Linvill and S. Östlund, "The Combined Effects of Moisture and


Temperature of the Mechanical Response of Paper," Experimental
Mechanics, vol. 54, pp. 1329-1341, 2014.

54
[121] S. Borodulina, A. Kulachenko, S. Galland and M. Nygårds, "Stress-strain
curve of paper revisited," Nordic Pulp and Paper Research Journal, vol. 27, no.
2, pp. 318-328, 2012.

[122] A. Kulachenko and T. Uesaka, "Direct simulations of fiber network


deformation and failure," Mechanics of Materials, vol. 51, pp. 1-14, 2012.

[123] E. Bosco, M. V. Bastawrous, R. H. Peerlings and J. Hoefnagels, "Bridging


network properties to the effective hygro-expansivity of paper:
Experiments and modelling," Philosophical Magazine, 2015.

[124] H. R. Motamedian and A. Kulachenko, "Simulating the hygroexpansion


of paper using a 3D beam network," International Journal of Solids and
Structures, vol. 161, pp. 23-41, 2019.

[125] A. Brandberg, H. R. Motamedian, A. Kulachenko and U. Hirn, "The Role


of the Fiber and the Bond in the Hygroexpansion and Curl of Thin
Freely," International Journal of Solids and Structures, Vols. 193-194, pp. 302-
313, 2020.

[126] R. ill, ”A theory of the yielding and plastic flow of anisotropic metals,”
The hydrodynamics of non-Newtonian fluids, pp. 281-297, 1948.

[127] W. Ramberg and W. R. Osgood, "Description of stress-strain curves by


three parameters," Technical Note No. 902, National Advisory Committee for
Aeronautics, Washington, 1943.

[128] P. Lipponen, T. Leppänen, J. Kouko and J. Hämäläinen, "Elasto-plastic


approach for paper cockling phenomenon: On the importance of
moisture gradient," International Journal of Solids and Structures, vol. 45, pp.
3596-3609, 2008.

[129] A.-L. Erkkilä, T. Leppänen and J. Hämäläinen, "Empirical plasticity


models applied for paper sheets having different anisotropy and dry solids
content levels," International Journal of Solids and Structures, vol. 50, pp. 2151-
2179, 2013.

[130] Q. S. Xia, M. C. Boyce and D. M. Parks, "A constitutive model for the
anisotropic elastic-plastic deformation of paper and paperboard,"
International Journal of Solids and Structures, vol. 39, pp. 4053-4071, 2002.

[131] P. Isaksson, R. Hägglund and P. Gradin, "Continuum damage mechanics


applied to paper," International Journal of Solids and Structures, vol. 41, pp.
4731-4755, 2004.

[132] E. Borgqvist, M. Wallin, M. Ristinmaa and J. Tryding, "An anisotropic in-


plane and out-of-plane elasto-plastic continuum model for paperboard,"
Composite Structures, vol. 126, pp. 184-195, 2015.

[133] N. Stenberg, "A model for the through-thickness elastic–plastic behaviour


of paper," International Journal of Solids and Structures, vol. 40, no. 26, pp.
7483-7498, 2003.

[134] E. Linvill and S. Östlund, "Parametric study of hydroforming of paper


materials using the explicit finite element method with a moisture-

55
dependent and temperature-dependent constitutive model," Packaging
technology and science, vol. 29, no. 3, pp. 145-160, 2016.

[135] E. Borgqvist, M. Wallin, J. Tryding, M. Ristinmaa and E. Tudisco,


"Localized Deformation in Compression and Folding of Paperboard,"
Packagin Technology and Science, vol. 29, pp. 397-414, 2016.

[136] K. Robertsson, M. Wallin, E. Borgqvist, M. Ristinmaa and J. Tryding, "A


rate-dependent continuum model for rapid converting of paperboard,"
Applied Mathematical Modelling, vol. 99, pp. 497-513, 2021.

[137] H. Huang and M. Nygårds, "Numerical and experimental investigation of


paperboard folding," Nordic Pulp and Paper Research Journal, vol. 26, no. 4,
pp. 452-467, 2011.

[138] H. Huang, A. Hagman and M. Nygårds, "Quasi static analysis of creasing


and folding for three paperboards," Mechanics of Materials, vol. 69, pp. 11-
34, 2014.

[139] L. A. A. Beex and R. H. J. Peerlings, "An experimental and computational


study of laminated paperboard creasing and folding," International Journal
of Solids and Structures, vol. 46, pp. 4192-4207, 2009.

[140] A. Giampieri, U. Perego and R. Borsari, "A constitutive model for the
mechanical response of the folding of creased paperboard," International
Journal of Solids and Structures, vol. 48, pp. 2275-2287, 2011.

[141] M. Wallmeier, E. Linvill, M. Hauptmann, J.-P. Majschak and S. Östlund,


"Explicit FEM analysis of the deep drawing of paperboard," Mechanics of
Materials, vol. 89, pp. 202-215, 2015.

[142] J.-W. Simon, ”A Review of Recent Trends and hallenges in


omputational odeling of Paper and Paperboard at ifferent Scale,”
Archives of Computational Methods in Engineering, vol. 28, pp. 2409-2428, 2021.

[143] Lorentzen & Wettre, Lorentzen & Wettre, ABB, Stockholm, Sweden.

[144] MTS System Corp, MTS System Corp.; Eden Praire, MN, US.

[145] M. Group, Marbach Group, Heilbronn, Germany.

[146] Zwick GmbH & Co. KG, August-Nagel-Straße 11, 89079 Ulm, Germany.

[147] Livermore Software Technology Corporation (LSTC), Livermore,


California, United States.

[148] F. GmbH, Splitting Machine AB 320 P, Fortuna GmbH, Eisenbahnstraße 15,


D-71263 Weil der Stadt (2004).

[149] Livermore Software Technology Corporation (LSTC), "LS-DYNA


Keyword user's manual, vol. II, material models, p. 463," LIVERMORE
SOFTWARE TECHNOLOGY (LST), AN ANSYS COMPANY, 27
September 2021. [Online]. Available:
https://www.dynasupport.com/manuals/ls-dyna-manuals/ls-
dyna_manual_volume_ii_r13.pdf. [Accessed 22 December 2022].

56
[150] . Eriksson, . Korin och F. Thuvander, ” amage to arton Board
Packages Subjected to oncentrated Loads,” Proceedings of the 19th IAPRI
World Conference on Packaging, pp. 172-182, 2014.

[151] S. 3D, SHINING 3D Technology GmbH. Stuttgart, Germany, Breitwiesenstraße


28, 70565 Stuttgart, Germany.

[152] S. Johansson, J. Engqvist, J. Tryding and S. A. Hall, "Microscale


deformation mechanisms in paperboard during continuous tensile loading
and 4D synchrotron X-ray," Strain, 2022.

57

You might also like