Sie sind auf Seite 1von 232

Zhuoping Shao

Fuli Wang

The Fracture
Mechanics of Plant
Materials
Wood and Bamboo
The Fracture Mechanics of Plant Materials
Zhuoping Shao Fuli Wang

The Fracture Mechanics


of Plant Materials
Wood and Bamboo

123
Zhuoping Shao Fuli Wang
School of Forestry and Landscape School of Forestry and Landscape
Anhui Agricultural University Anhui Agricultural University
Hefei, Anhui Hefei, Anhui
China China

ISBN 978-981-10-9016-5 ISBN 978-981-10-9017-2 (eBook)


https://doi.org/10.1007/978-981-10-9017-2
Library of Congress Control Number: 2018937331

© Springer Nature Singapore Pte Ltd. 2018


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

Printed on acid-free paper

This Springer imprint is published by the registered company Springer Nature Singapore Pte Ltd.
part of Springer Nature
The registered company address is: 152 Beach Road, #21-01/04 Gateway East, Singapore 189721,
Singapore
Preface

Wood and bamboo are the plant materials that can be directly used as structural
materials, meanwhile, they are also the oldest and still the most widely used natural
structure materials. Many of them are used as structural materials, such as beam,
frame, floor, and support. And bamboo is also a fine engineering structure material
with high strength, good stiffness, and high wear resistance, which is used to build
bamboo house, as construction scaffolding, and as bamboo ladder. Thus, to study
the strength, toughness, and the failure behaviors of wood and bamboo is very
important for the safety assessment and structural design of wood and bamboo.
Wood and bamboo are natural composites that possess obvious meso-structure
and can be studied in multiscale. Because of inhomogeneous and anisotropic
structure and microscopic or macroscopic natural defects or damages, when loaded,
the macroscopic mechanical behavior of wood and bamboo would be determined
by the irregular evolution behaviors of the defects or damages. Although wood and
bamboo both are cell body plant materials, the differences in macroscopic and
microscopic structure bring wood and bamboo different failure mechanisms cor-
responding to different study methods. Thus, it is significant for the design and
safety analysis of wood and bamboo components to understand how to use fracture
mechanics and meso-mechanics to analyze the fracture behaviors of wood and
bamboo, what the changes of inner microstructure are when loaded, and what the
relation between the changes and macro-mechanical response is. Meanwhile, it will
provide guiding function for the development of new biocomposites that possess
special strength and toughness properties and overcome the defects of biomaterials.
This is a book on the fracture behaviors and toughness mechanism of bamboo
and wood, which reflects the research work of authors in the past decade. In the
sections on wood, varieties of trees are selected, for example, softwoods: China fir
(Cunninghamia Lanceolata), Mongolian pine (Pinus sylvestris var. mongolica
Litv.), Picea jezoensis (Picea asperata), Larch (Larix gmelinii), and so on are
chosen, which possess growth rings in different clarity because of the changes of
early wood and late wood; hardwoods: Populus spp I-69, Castanopsis hystrix,
Koompassia spp, Melia azedarach, and so on are chosen considering their differ-
ence in construction such as diffuse-porous wood, ring-porous wood, wood ray,

v
vi Preface

grain, etc. And moso bamboo (Phyllostachys pubescens) is the study object in the
sections on bamboo. Theoretical analysis is combined with experiments assisted
with various mathematical tools and experiment means.
There are nine chapters in total. The content involves the mechanical charac-
teristics and stress–strain relationship of wood structure, the fracture of wood along
grain, the transverse fracture of wood, the finite element analysis of wood crack tip
stress field and prediction of the crack propagation direction, acoustic emission
characteristics and Felicity effect of wood fracture perpendicular to the grains, the
mechanical characteristics of bamboo structure and its components, the interlaminar
fracture properties of bamboo, and the toughness fracture model and energy
absorbing mechanism of bamboo. Many of the researches are studied for the first
time. Most of the chapters are written by Shao Zhuoping, and Sects. 8.4 and 8.5 are
written by Wang Fuli. In addition, Chap. 9 is written by the corporation of Shao
Zhuoping and Wang Fuli. There might be defects and mistakes inevitably for the
limited specialized knowledge of authors, so please put out the mistakes so that they
can be corrected.

Hefei, China Zhuoping Shao


November 2017
Acknowledgements

Here, we want to express heartfelt thanks to the National Natural Science


Foundation of China [grant number: 30571452, 11072001, 11008250, 31570715]
for the foundation support. And we also appreciate the help of Wu Yijun, Fang
Changhua, Tian Genlin, Zhou Liang, Liu Yamei, Huang Tianlai, Li Qizhi, and Wu
Dong.

vii
Contents

1 Introduction to the Application of the Fracture Mechanics


in Wood and Bamboo . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1 Brief History of Fracture Mechanics . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Mechanics of Materials and Fracture Mechanics . . . . . . . . . . . . . . 2
1.3 Brief Review of the Fracture Mechanics of Wood . . . . . . . . . . . . 3
1.3.1 The Strength Prediction of Wood Materials . . . . . . . . . . . . 3
1.3.2 The Application of Fracture Mechanics Combined with
Acoustic Emission (AE) in the Propagation Mechanism
of Wood Crack . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.3.3 Researches on the Fracture Property of Bamboo . . . . . . . . 6
1.4 The Main Contents of This Book . . . . . . . . . . . . . . . . . . . . . . . . 7
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2 Mechanical Characteristics and Stress–Strain Relationship
of Wood Structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.1 The Mechanical Characteristics of Wood Structure . . . . . . . . . . . . 11
2.2 The Stress–Strain Relation of Solid Material . . . . . . . . . . . . . . . . 13
2.3 Engineering Elastic Constants . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.4 Engineering Elastic Constants of Wood . . . . . . . . . . . . . . . . . . . . 16
2.5 The Concept of Plane Stress and Plane Strain . . . . . . . . . . . . . . . 16
2.5.1 Uniform Thickness Plate and Plane Stress . . . . . . . . . . . . . 17
2.5.2 Infinite Cylinder and Plane Strain . . . . . . . . . . . . . . . . . . . 17
2.5.3 The Stress–Strain Relationship in Plane Problem . . . . . . . . 18
2.6 Tests of Wood Elastic Coefficients . . . . . . . . . . . . . . . . . . . . . . . 19
2.6.1 The Application of Electrometric Method on Wood
Elastic Coefficients . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
2.6.2 The Application of DSCM on Wood Elastic Coefficients . . . 22
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26

ix
x Contents

3 Fracture of Wood Along Grain . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27


3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
3.2 Theory of LEFM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.2.1 Crack . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.2.2 Stress Intensity Factor K and K Criterion . . . . . . . . . . . . . 28
3.2.3 Energy Release Rate G and G Criterion . . . . . . . . . . . . . . 30
3.2.4 Relationship Between K and G . . . . . . . . . . . . . . . . . . . . . 32
3.3 Fracture Mechanics of Anisotropic Material . . . . . . . . . . . . . . . . . 36
3.4 The Special Application of LEFM on Wood . . . . . . . . . . . . . . . . 39
3.5 The Stress Intensity Factor KIC of Wood Fracture Along Grain . . . 41
3.5.1 The Methods to Test Stress Intensity Factor . . . . . . . . . . . 41
3.5.2 The K TL IC of CT Samples with Different Thickness . . . . . . . 44
3.5.3 The K TL IC of WOL Samples with Different Crack Length . . . 46
3.6 The Fracture Toughness GTL IC Along Grain of Wood by Energy
Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
3.6.1 Materials and Samples . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
3.6.2 Test and Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
3.6.3 The Relationship Between Stress Intensity Factor and
Energy Release Rate . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
3.7 Mode III Fracture Property of Wood Along Grain . . . . . . . . . . . . 58
3.7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
3.7.2 Material and Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
3.7.3 Test and Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
3.8 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
4 Transverse Fracture of Wood . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
4.2 Analysis on Stress Field at Crack Tip . . . . . . . . . . . . . . . . . . . . . 65
4.3 The Cracking Direction of Transverse Crack . . . . . . . . . . . . . . . . 68
4.4 Test of Critical Stress Intensity Factor . . . . . . . . . . . . . . . . . . . . . 70
4.4.1 Material and Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
4.4.2 Test and Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
4.5 The Influence of Transverse Crack on the Normal Strength
of Wood . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .... 74
4.5.1 Influence of Crack Perpendicular to Grain on MOR
of Wood . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .... 74
4.5.2 Influence of Crack Perpendicular to Grain on Impact
Toughness of Wood . . . . . . . . . . . . . . . . . . . . . . . . . .... 76
4.5.3 Influence of Crack Perpendicular to Grain on Tensile
Strength of Wood . . . . . . . . . . . . . . . . . . . . . . . . . . .... 77
4.5.4 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .... 78
Contents xi

4.6 Energy Release Rate of the Mode I Interlaminar Fracture


of Wood Beam and the Bending Delamination Damage
of Plywood . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
5 Finite Element Analysis of Wood Crack Tip Stress Field and
Prediction of the Crack Propagation Direction . . . . . . . . . . . . . . . . . 87
5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
5.2 Materials and Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
5.2.1 Materials and Fundamental Data . . . . . . . . . . . . . . . . . . . . 89
5.2.2 Fracture Specimens . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
5.3 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
5.3.1 Stress Field of the Crack Tip . . . . . . . . . . . . . . . . . . . . . . 91
5.3.2 Prediction of Crack Propagation . . . . . . . . . . . . . . . . . . . . 91
5.4 The Relationship Between Interfacial Strength and Toughness
of Wood . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ... 96
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
6 Fractal Features and Acoustic Emission Characteristics
of Wood Fracture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
6.1 The Fractal Features of Wood Fracture . . . . . . . . . . . . . . . . . . . . 103
6.1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
6.1.2 Theories . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
6.1.3 Tests and Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
6.1.4 The Relationship Between Fracture Toughness and
Fractal Dimension . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
6.2 Acoustic Emission Characteristics and Felicity Effect
of Wood Fracture Perpendicular to the Grains . . . . . . . . . . . . . . . 112
6.2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
6.2.2 Materials and Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
6.2.3 Results and Discussions . . . . . . . . . . . . . . . . . . . . . . . . . . 114
6.2.4 Felicity Effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
7 Mechanical Characteristics of Bamboo Structure and Its
Components . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
7.2 Mechanical Characteristics of Bamboo . . . . . . . . . . . . . . . . . . . . . 127
7.3 The Mechanical Characteristics of the Components Bamboo . . . . . 129
7.3.1 “Mixture Law” Method . . . . . . . . . . . . . . . . . . . . . . . . . . 129
7.3.2 Test on Single Fiber Bundle . . . . . . . . . . . . . . . . . . . . . . . 132
7.3.3 Analysis on Fracture Surface . . . . . . . . . . . . . . . . . . . . . . 136
7.4 Difference of Structure and Strength Between Internodes
Part and Node Part of Moso Bamboo . . . . . . . . . . . . . . . . . . . . . 138
xii Contents

7.4.1 Material and Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138


7.4.2 Structure Comparison Between Internodes Part
and Node Part Culm . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139
7.4.3 Comparison of Mechanical Properties Between
Internodes Part and Node Part of Bamboo . . . . . . . . . . . . 139
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145
8 Interlaminar Fracture Properties of Bamboo . . . . . . . . . . . . . . . . . . 147
8.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147
8.2 Mode I Interlaminar Fracture of Bamboo . . . . . . . . . . . . . . . . . . . 148
8.2.1 Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148
8.2.2 Materials and Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . 149
8.2.3 Results and Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152
8.2.4 Analysis of Fracture Surface . . . . . . . . . . . . . . . . . . . . . . . 154
8.3 Mode II Interlaminar Fracture of Bamboo . . . . . . . . . . . . . . . . . . 156
8.3.1 Test and Specimens . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156
8.3.2 The Data Processing Methods . . . . . . . . . . . . . . . . . . . . . 159
8.3.3 Results and Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162
8.3.4 The Fracture Surface Analysis . . . . . . . . . . . . . . . . . . . . . 166
8.4 Mode III Interlaminar Fracture of Bamboo . . . . . . . . . . . . . . . . . . 167
8.4.1 Material and Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . 168
8.4.2 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . 171
8.4.3 Compared with Artificial Fiber-Reinforced Polymer
(FRP) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173
8.5 The Toughness Contribution of Bamboo Node to the Interlaminar
Fracture Toughness of Bamboo . . . . . . . . . . . . . . . . . . . . . . . . . . 177
8.5.1 Mode I . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177
8.5.2 Mode II . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183
8.5.3 Mode III . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187
8.5.4 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 191
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 195
9 Modeling on the Toughness Fracture and Energy-Absorbing
Mechanism of Biomaterial—Bamboo (Phyllostachys pubescens) . . . . 199
9.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 199
9.2 Development of Theoretical Model . . . . . . . . . . . . . . . . . . . . . . . 201
9.2.1 The Damage Patterns of Bamboo Transverse Fracture . . . . 201
9.2.2 The Simplification of Bamboo Mechanical Model . . . . . . . 201
9.3 Theoretical Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 202
9.3.1 Ground Tissue Cracking . . . . . . . . . . . . . . . . . . . . . . . . . . 202
9.3.2 Interface Debonding . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 203
9.3.3 The Fracture of Fiber Bundle . . . . . . . . . . . . . . . . . . . . . . 205
9.3.4 The Pulling-Out of Fiber Bundle . . . . . . . . . . . . . . . . . . . 207
9.3.5 On the Calculation of Lp . . . . . . . . . . . . . . . . . . . . . . . . . 207
Contents xiii

9.4 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 208


9.4.1 Result of Experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . 208
9.4.2 Theoretical Result . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 209
9.4.3 Analysis and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . 211
9.5 Verification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 212
9.5.1 Theoretical Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . 213
9.5.2 Experiments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 215
9.5.3 Analysis and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . 216
9.6 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219
Chapter 1
Introduction to the Application
of the Fracture Mechanics in Wood
and Bamboo

Abstract In this chapter, the history of fracture mechanics is introduced, and the
difference between fracture mechanics and mechanics of materials is discussed.
Then, the application history and research progress of fracture mechanics applied to
wood are reviewed. And it is described briefly that although wood and bamboo are
both cell body plant materials, the differences in macroscopic and microscopic
structure bring wood and bamboo different failure mechanisms corresponding to
different study methods. When studied the fracture properties, wood can be treated
as orthotropic macro-homogeneous body, while bamboo is composite reinforced by
fiber distributed non-uniformly, so meso-mechanics should be used to build
mesoscale model to study the fracture of bamboo.

1.1 Brief History of Fracture Mechanics

In 1920, Griffith studied the fracture problem of ideal brittle materials such as glass
and ceramics and obtained some new concepts on material strength [1]. However,
Griffith’s theory has not attracted extensive attention, because the brittle materials
above were not used as engineering structure material in that time, and other
structure materials that showed brittle fracture were few.
Since World War II, many large welding elements were manufactured with
high-strength steel; meanwhile, major accidents increased obviously. Although the
welding fabrications all met the traditional design requirements, brittle fracture that
always gave rise to catastrophic devastation often happened under low stress level
without any forewarning. For example, 40 welded steel bridges collapsed contin-
uously with no omen from 1938 to 1940 [2]. During World War II, for about five
thousand welding ships, more than one thousand fracture faults happened and 238
ships had to be scrapped, some of which fractured into two suddenly in a calm sea
[3]. In 1965, Britain, the offshore drilling platform sank because of the brittle
fracture of the pull rod [4]. In 1969, America, because of the brittle fracture of wing
spindles, warplane F-111 and C-56 crashed [5], and so on. The frequent occur-
rences of the series major accidents shocked the engineering community, because

© Springer Nature Singapore Pte Ltd. 2018 1


Z. Shao and F. Wang, The Fracture Mechanics of Plant Materials,
https://doi.org/10.1007/978-981-10-9017-2_1
2 1 Introduction to the Application of the Fracture Mechanics …

the structures all met the traditional design requirements when failed. After realized
that there must be something neglected in traditional design idea, people found that
crack in structure was the immediate cause of failure for most accidents by
extensive investigation and research.
In the past 50 years, a lot of theoretical and experimental researches have been
performed on the failure of structure with crack, thus fracture mechanics comes into
being. Fracture mechanics provides the safety design of structure new idea and
method, which is extensively applied to metal, inorganic nonmetal, polymer, tim-
bers, and composites. Hence, fracture mechanics is a subject to study the strength of
crack body, and its generation and development has a close relationship with the
occurrence of engineering major accidents.

1.2 Mechanics of Materials and Fracture Mechanics

Studying the fracture of material is one of the primary missions of mechanics of


materials. Due to the design idea of mechanics of materials [5, 6], four mechanical
performance indexes of material should be tested, as follows:

Yield limited rs
Intensity index
Ultimate strength rb

Elongation rate d
Toughness index
Im pacttoughness Ka

Traditional design idea:

rStructure  ½r

where for plastic materials, ½r ¼ rns ; for brittle materials, ½r ¼ rnb . n is safety factor
and n > 1. And for the structure bearing impact load and pressure vessels, toughness
must be checked.
However, the traditional design idea above is based on the hypothesis of
material’s continuity and uniformity with no crack and defect. In fact, crack and
defect of engineering material are inevitable, and the fracture of material with crack
is not only determined by crack length and external force but also by the sensitivity
of material to crack, namely, fracture toughness [6].
Taken elasticity and plastic mechanics as theoretical approach, fracture
mechanics abandon the hypothesis of material’s continuity and uniformity, by
which the stress field and displacement field of material with crack can be deter-
mined. And the physical quantity that determines the propagation ability of crack
can be obtained; thus, a new strength design idea is proposed. The new strength
criterion could be developed by measuring the ability of material to resist crack
1.2 Mechanics of Materials and Fracture Mechanics 3

propagation, namely, fracture toughness KC and GC, for example, the fracture
criterion of brittle material is shown as follows:

K criterion: K ¼ KC

G criterion: G ¼ GC

In summary, mechanics of materials is a direct and empirical method and con-


venient for application in simple form but neglects the inherent defect of material,
while fracture mechanics provides compensation for the deficiency and irrationality
of traditional design idea. Consequently, significant influence is brought to the
whole material manufacturing industry, for example, the toughened mechanism
with weakened interface is a good example of the application of fracture mechanics
in artificial composites [7].

1.3 Brief Review of the Fracture Mechanics of Wood

1.3.1 The Strength Prediction of Wood Materials

Fracture mechanics was first used in wood by Porter [8]. Based on energy balance
principle, linear elastic fracture mechanics (LEFM) was successfully applied by
Porter to the opening fracture in longitudinal and tangential (LT) and longitudinal
and radial (LR) plan of Western white pine (Pinus monticola). The results showed
that the strain energy release rate GIC, which represented the ability of materials to
resist the propagation of crack along grain, had no relation with the geometric
dimensions of sample and crack length. And the comprehensive value of stress and
crack length could be predicted by the knowledge on GIC, which would cause the
rapid propagation of crack and fracture of sample. After that, researchers from
various countries have done tremendous work on wood fracture, and various wood
fracture modes, fracture toughness test methods, and fracture criterions have been
proposed. Preliminary results have been obtained when fracture mechanics was
applied to solve some practical problems of wood and wood structure.
So far, there are two fracture criterions for the macro fracture of wood: one has
thermodynamics meaning, based on the energy criterion in crack tip process zone,
such as energy release rate criterion (G criterion) [9, 10]; the other mechanics
meaning, based on the strength criterion in crack tip process zone, such as stress
intensity factor criterion (K criterion) [11, 12]. However, wood possesses many
characteristics which are obviously different from those of other orthotropic
materials. And the major characteristic of wood is the high anisotropy among three
principal directions from the perspective of the composition and structure of wood,
which makes it difficult to apply LEFM to wood. The complexity of orthotropic
case is that crack may not propagate along its initial direction in plan. Until now,
regarding the problem of wood fracture, most discussions are on the case of
4 1 Introduction to the Application of the Fracture Mechanics …

self-similar propagation of crack, because it is too difficult to mathematically deal


with cracks in different angles to principal direction of wood.
As most cracks and defects formed during tree growth and wood processing are
along fiber direction, early researches on fracture of wood mainly focus on the
cracks parallel to the grain (TL crack) and the cracks perpendicular to the grain (LT
crack), but the ability of wood to resist the propagation of crack along fiber
direction is low. The application of LEFM in the fracture of material with crack
along grain is successful, and a lot of achievements have been obtained. In recent
years, the propagation models of TR crack (crack along radial direction) and RT
crack (crack along tangential direction) and the interface mechanism of wood are
the focus of research works. As the strong anisotropy of wood stiffness and
strength, the tensile force caused by external load or the sudden change in sur-
rounding condition is negative to crack perpendicular to grain. In experiments, TR
crack and RT crack are similar to radial shake and ring shake of wood, respectively,
so the formation and propagation of crack in wood caused by tensile force along
grain is always an active subject in the research field of wood fracture mechanics.
Orthotropic model was first applied in wood by Wu, and tensile strength across
grain and shear strength along grain of wood samples with cracks were predicted by
LEFM; meanwhile, an empirical model was proposed predict the failure of wood
under mixture stress pattern [13, 14]. Later, a great quantity of research work on the
prediction of wood strength has been done by Mindness et al. [15], Barrett and
Foschi [12], Schnewind [16], Smith and Penny [17], and White and Green [18] with
the means of LEFM. Triboulot and Pluvinage [19] compared the results obtained by
finite element method (FEM) and experiment and demonstrated that it was feasible
to treat wood as orthotropic and elastic body and to apply the concept of fracture
mechanics in wood.
Fracture mechanics could be used to predict the initiation of failure in wood
caused by stress concentration due to node, notch, crack, and so on. For example, for
panel with node, Boatright and Garrentt [20] replaced “equivalent crack length Le”
as node, where Le was crack length generated by the fracture of clear wood sample
under the stress level equaled to the stress that caused the initial cracking of wood
sample with node, but the method was just appropriate for tensile load along grain.
Murphy [21] used fracture mechanics to estimate bending strength of panel with
node, notch, edge crack, or end crack, and showed that even under parallel-to-grain
stress, when crack propagated along the direction perpendicular to the notch, fracture
mechanics could still predict the strength of wood sample. When studied the
influence of end crack length on the bending strength of wood beam, Foschi and
Barrett [22] obtained the maximum end crack length that did not affect the strength
of wood beam, which was adopted in Structure Design of Timber Engineering by
Canadian Standards Association [23]. Sun and Lu [24] studied the stress field and
stress strength factor of mode I fracture at crack tip of Fraxinus mandshurica Rupr.
by finite element method and gave the distribution diagram of stress strength factor
at crack tip due to data processing.
At present, further studies have been carried out on the application of fracture
mechanics in the strength design and safety assessment of wood and wood
1.3 Brief Review of the Fracture Mechanics of Wood 5

structures. Fracture mechanics has been used by French Institute of Wood Science
to predict the crack direction and crack depth in wood caused by the accumulation
and release of growth stress, which is a breakthrough and innovation research idea.
It is also a new development tendency that fracture of wood is combined with its
microscopic structure to explore the toughness mechanism. For example, Ren and
Jiang [25] tested the fracture toughness of Chinese Fir and Masson Pine by LEFM,
and microscopic structure of wood had an influence on wood fracture due to the
fracture surface observed by SEM. Stefanie and Stanzl [26] studied the evolution of
wood microscopic structure under load and the fracture mechanical response in the
chamber of environment scanning electron microscope (ESEM).

1.3.2 The Application of Fracture Mechanics Combined


with Acoustic Emission (AE) in the Propagation
Mechanism of Wood Crack

Acoustic emission (AE) was first applied in the fracture of wood by Prof. Porter [8],
and then Knuffel proposed that the fracture of wood included three stages: crack
initiation, crack propagation, and failure, and due to the viewpoints, the failure of
wood was treated as a continuous process with different AE characteristics in
different stages but not a single event [27].
Later, AE was used to measure the elastic modulus of wood due to the sound
transmission feature in wood. Ansell [28] found that the shape of AE strain curve
was affected by earlywood or latewood percent, when studied the tensile properties
of three softwoods. Sato et al. [29] found a negative correlation between the number
of AE events and wood strength in tensile test, and differentiated between slow AE
response when microcrack propagated cross growth rings and fast AE response
when macrocrack propagated cross growth rings.
The research of Ogino et al. [30] showed that AE energy would increase rapidly
soon after the first microcrack occurred. Spectral analysis indicated that
high-frequency component would occur before cracking began, and low frequency
only occurred after cracking (once crack began, the frequency would decrease).
Ogino also found that the AE signal from wood drying could be divided into four
patterns and by observation, the appearance of any two patterns could be used as an
early warning signal of cracking.
Suzuki and Schniewind [31] found that there was a linear relation between
fracture toughness and the number of AE event in unit crack zone, when studied the
AE characteristic of the fracture of timber composites with different adhesive
contents. Rice and Skarr [32] studied the AE characteristic of steam beech (Zelkova
schneideriana) panel under transverse bending force and found that the AE char-
acteristic of dry wood was different with that of green wood. Ando et al. [33]
studied the relation between fracture toughness and AE signal of sample with single
edge crack and the influence of different grain angles.
6 1 Introduction to the Application of the Fracture Mechanics …

Schniewind et al. [34] collected AE signals during mode I and mixed mode tests
at different moisture contents and temperatures and found that the AE activity in
mixed mode tests was much higher than that for mode I. Dill-Langer et al. [35] used
AE technique to monitor the fracture of clear spruce wood under tensile loading and
found that there was an onset of AE prior to the first visible crack growth
step. Aicher et al. [36] used AE to localize crack nucleation in glulam loaded in
tension perpendicular to the grains. Reiterer et al. [37, 38] used AE to monitor mode
I fracture of softwoods (spruce and pine) and hardwoods (alder, oak, and ash) and
stated that the AE counts up to maximum force are much higher for the softwoods.
Chen et al. [39] used AE to monitor the failure process of hardwood and softwood
test-pieces under static and fatigue torsion loading and found it was possible to
monitor and analyze the failure process in wood by AE techniques. Choi et al. [40]
studied the fracture processes of typical fiber-reinforced plastic composites lami-
nates with continuous fiber reinforcement and the results showed that the AE
characteristics might represent the process of fiber breakages according to the
various loading stages, which expressed characteristic fracture processes for indi-
vidual fiber-reinforced composite laminates. The feature of the AE hit-event rate, in
combination with AE amplitude classifications, could be utilized for nondestructive
identification of different fracture mechanisms.

1.3.3 Researches on the Fracture Property of Bamboo

Bamboo, “The Second Forest”, is an important part of forest resource in China and
has the advantages of short growth cycle, easy renewal, high production, strong
regeneration capacity, extensive use, and high economic value. There are a wide
variety of bamboo species, among which moso bamboo (Phyllostachys pubescens)
is the most widely distributed one, mostly in Yangtze River Basin. It has a straight
stem, high stiffness, strong strength, and some other fine mechanical properties,
such as high static bending strength and tensile strength, high elastic modulus, and
hardness.
Through millions of years’ evolution, bamboo has formed a particular structure
in order to bear the bending load mainly caused by snow or wind. This structure
contributes to its high transverse bending strength and toughness. In contrast, the
anti-cleaving and anti-shearing strength along the grain of bamboo are relatively
low. Therefore, bamboo often cracks along the direction of grain while drying.
Many studies have been reported on the general mechanical properties of bamboo
[41–43]. But researches on fracture characteristics of bamboo are few [44, 45], and
the transverse fracture toughness KIC of bamboo measured by fracture toughness
test method of metal materials was just a nominal toughness that had limited
physical significance, because stress intensity factor method is not feasible to deal
with the fracture problem of fiber composites.
In view of macro-mechanical behavior, bamboo is a typical unidirectional
long-fiber-reinforced bio-composite. It exhibits significant anisotropy in strength
1.3 Brief Review of the Fracture Mechanics of Wood 7

and stiffness corresponding to longitudinal, radial, and transverse directions. The


tensile strength along longitudinal direction can be as large as 150–300 Mpa, but
the transverse tensile strength and shear strength parallel to grain are fairly low [41].
Therefore, the lateral tensiling and interlaminar shearing force exerted exteriorly or
caused by surrounding conditions’ changes are prone to bring interlaminar fracture
parallel to grain of bamboo. Then, the interlaminar cracking is controlled by the
interlaminar fracture toughness rather than by the transverse strength of bamboo.
Even the crack or lacuna perpendicular to grain of bamboo may easily be trans-
formed to develop along the direction of grains under loads, and influences the
mechanical behavior thereby. However, because of the short industrialized uti-
lization history of bamboo, the understanding of bamboo properties is restricted to
general properties, lacking more extensive and further study on the basic properties
such as fracture and its mechanism.
In sum, although the study on wood fracture has already been 40 years of history
and much research work has been done by researchers at home and abroad, there is
no similar book that comprehensively and systematically introduces the fracture
mechanics theory on plant materials including wood and bamboo. Although wood
and bamboo both are cell body plant materials, the differences in macroscopic and
microscopic structure bring wood and bamboo different failure mechanisms cor-
responding to different study methods. When studied the fracture properties, wood
is treated as orthotropic macro-homogeneous body, while bamboo is composite
reinforced by fiber distributed non-uniformly, so meso-mechanics should be used to
build mesoscale model to study the fracture of bamboo.

1.4 The Main Contents of This Book

To study the fracture behavior and fracture mechanism of plant materials com-
prehensively, varieties of trees are selected as experimental subjects, for example,
softwood: China fir (Cunninghamia lanceolata), Mongolian pine (Pinus sylvestnis
var. mongolica Litv.), Picea jezoensis (Picea asperata), Larch (Larix gmelinii), and
so on, which possess growth rings in different clarity because of the changes of
early wood and late wood; hardwood: I-69 poplar (Populus spp.), Castanopsis
hystrix, Koompassia spp., Melia azedarach, and so on are chosen considering their
difference in construction such as diffuse-porus wood, ring-porus wood, wood ray,
and grain. And moso bamboo (Phyllostachys pubescens) is the study object in the
sections on bamboo.
Theoretical analysis is combined with experiments assisted with various math-
ematical tools and experiment means. This book includes six principal aspects,
many of which are studied for the first time, involving:
(1) The fracture properties of wood along grain and test method of wood fracture
toughness; the influence of sample size on test result; the relationship between
critical stress intensity factor KIC and critical energy release rate GIC and the
8 1 Introduction to the Application of the Fracture Mechanics …

influence factors; fractal feature of the fracture surface of wood along grain and
the internal relationship between the fracture toughness and meso-structure of
wood due to the relation between fractal dimension of fracture surface and
fracture toughness; Mode III interlaminar fracture property of wood and the test
method.
(2) The transverse fracture properties of wood and the analysis on stress filed at
crack tip by finite element method, prediction of the crack propagation direc-
tion, the transverse fracture mechanism; the influence of transverse crack on the
tensile strength, bending strength, and impact strength; the suggestion of a net
stress criterion for the safety assessment and strength design of wood compo-
nent with crack perpendicular to grain; analytic solution of energy release rate
when crack propagates laterally based on energy theory.
(3) The raise and preliminary definition of four basic damage structure elements;
the evolution character of microstructure during the fracture of wood; the
application of AE technique to the identification of wood fracture pattern;
research on the felicity effect of wood structure bearing load.
(4) The mechanical characteristics of bamboo structure; analysis on the strength
and elastic modulus of bamboo fiber and ground tissue by Mixture law and
Shear lag theory of meso-mechanics; the difference between bamboo internodes
and node in construction and strength.
(5) Mode I, II, III interlaminar fracture properties of bamboo and the fracture
mechanism due to fracture morphology observed by SEM.
(6) The transverse fracture model of bamboo and the toughness contribution of
each damage pattern; the energy absorbing mechanism of bamboo transverse
fracture.

References

1. Griffith AA (1920) The phenomenon of rupture and flow in solid. Philos Trans R Soc Lond A
221:163–198
2. Xiang H (2001) Advanced theory of bridge structure. China Communication Press, Beijing
(in Chinese)
3. Gao Q (1986) Engineering fracture mechanics. Chongqing University Press, Chongqing
(in Chinese)
4. Yang G (1995) The disasters of offshore engineering and the environment load. China
Offshore Platform 10(5):202–203
5. Li H, Zhou C (1990) Engineering fracture mechanics. Dalian University of Technology Press,
Dalian (in Chinese)
6. Kuang Z, Ma F (2002) Crack tip fields. Xi’an Jiaotong University Press, Xi’an (in Chinese)
7. Gordon JE (1968) The new science of strong materials or why you don’t fall through the
floor. Penguin Books Limited, Haromondsworth
8. Porter AW (1964) On the mechanics of fracture in wood. Forest Prod J 8:325–331
9. Larsen HJ, Gustafsson PJ (1990) The fracture energy of wood in tension perpendicular to the
grain. In: 23th CIB-W18 meeting, Lisbon Portugal, p 23-19-2
References 9

10. Stanzl-Tschegg SE, Tschegg EK, Teischinger A (1994) Fracture energy of spruce wood after
different drying procedures. Wood Fiber Sci 26:467–478
11. Ewing PD, Williams JG (1979) Thickness and moisture content effect in the fracture
toughness of Scots Pine. J Mater Sci 14(12):2959–2966
12. Barrett JD, Foschi RO (1977) Model II stress-intensity factors for cracked wood beams. Eng
Fract Mech 9(2):371–378
13. Wu EM (1967) Application of fracture mechanics to anisotropic plates. J Appl Mech 34:967–
974
14. Wu EM (1968) Fracture mechanics of anisotropic plates. In: Tsai SW (ed) Composite material
workshop, Technomic Publishing Company, Lancaster, p 23
15. Mindness S, Nadeau JS, Barrett JD (1976) Stow crack growth in Douglas-fir. Wood Sci
1:389–396
16. Schnewind AP (1977) Fracture toughness and duration of load factor. Duration factor for
cracks propagating perpendicular to grain. Wood Fiber 9(3):216–226
17. Smith TW, Penny DT (1980) Fracture mechanics of butt joints in laminated wood beams.
Wood Sci 12(4):227–235
18. White MS, Green DW (1980) Effect of substrate on the fracture toughness of wood-adhesive
bonds. Wood Sci 12(3):149–153
19. Triboulot P, Pluvinage G (1984) Validity of fracture mechanics concepts applied to wood by
finite element calculation. Wood Sci Technol 18(1):51–58
20. Boatright SWJ, Garrentt GG (1983) The effect of microstructure and stress state on the
fracture behaviour of wood. J Mater Sci 18:2181–2199
21. Murphy JF (1979) Strength of wood beams with end splits. Research Paper. FPL 347. USDA
Forest Service Products Laboratory. Madison, WI, p 12
22. Foschi RO, Barrett JD (1976) Stress intensity factors in anisotropic plates using singular
isoparametric elements. Int J Numer Meth Eng 10(6):1281–1287
23. Canadaian Stanards Association (1984) Engineering design in wood (working stress design).
CAN3-086-M84
24. Sun Y, Lu Z (1999) Calculation of stress strength factor at crack tip of Fraxinus mandshurica
Rupr. by finite element. J Beijing Forest Univ 21(3):53–57 (in Chinese)
25. Ren H, Jiang Z (2001) Morphology of wood fracture of Chinese Fir and Masson Pine.
Scientia Silvae Sinicae 37(3):118–121(in Chinese)
26. Stefanie E, Stanzl T (2006) Microstructure and fracture mechanical response of wood. Int J
Fract 139:495–508
27. Knuffel WE (1988) Acoustic emission as Strength predictor in structural timber.
Holzforschung 42:336–348
28. Ansell MP (1982) Acoustic Emission from softwood in tension. Wood Sci Technol 16(1):35–
58
29. Sato K, Okano T, Asano I, Fushitani M (1985) Application of AE to mechanical testing of
wood. In: 2nd internatioal conference on acoustic emission, Lake Tahoe, pp 240–243
30. Ogino S, Kaino K, Suzuki M (1986) Prediction of lumber checking during drying by means
of acoustic emission technique. J Acoustic Emission 5(2):61–65
31. Suzuki M, Schniewind AP (1987) Relationship between fracture toughness and acoustic
emission during cleavage failure in adhesive joints. Wood Sci Technol 21:121–130
32. Rice RW, Skarr C (1990) Acoustic emission patterns from the surfaces of red oak wafers
under transverse bending stress. Wood Sci Technol 24:123–129
33. Ando K, Sato K, Fushitani M (1992) Fracturetoughness and acoustic emission Characteristics
of Wood II. J Japan Wood Res Soc 38:342–349
34. Schniewind AP, Quarles SL, Lee H (1996) Wood fracture, acoustic emission, and the drying
process Part 1. Acoustic emission associated with fracture. Wood Sci Technol 30:273–281
35. Dill-Langer G, Aicher S (2000) Monitoring of microfracture by microscopy and acoustic
emission. In: International conference on wood and wood fiber composites, Stuttgart, pp 93–
104
10 1 Introduction to the Application of the Fracture Mechanics …

36. Aicher S, Höfflin L, Dill-Langer G (2001) Damage evolution and acoustic emission of wood
at tension perpendicular to fiber. Holz als Roh Werkst 59:104–116
37. Reiter A, Stanzl-Tschegg SE, Tschegg EK (2000) Mode I fracture and acoustic emission of
softwood and hardwood. Wood Sci Technol 34(5):417–430
38. Reiter A, Stanzl-Tschegg SE, Tschegg EK (2002) Fracture characteristics of different wood
species under Mode I loading perpendicular to the grain. Mater Sci Eng A 332:29–36
39. Chen Z, Gabbitas B, Hunt D (2006) Monitoring the fracture of wood in torsion using acoustic
emission. J Mater Sci 41:3645–3655
40. Choi N-S, Woo S-C, Rhee K-Y (2007) Effects of fiber orientation on the acoustic emission
and fracture characteristics of composite laminates. J Mater Sci 42:1162–1168
41. Zeng QY, Li SH, Bao XR (1992) Effect of bamboo nodal on mechanical properties of
bamboo wood. Sci Silvae Sinica 28(3):247–252 (in Chinese)
42. Ahmad M, Kamke FA (2005) Analysis of calculate bamboo for structural composite
materials: physical and mechanical properties. Wood Sci Technol 39:448–459
43. Obataya E, Kitin P, Yamauchi H (2007) Bending characteristics of bamboo (Phyllostachys
pubescens) with respect to its fiber-foam composite structure. Wood Sci Technol 41:385–400
44. Xi X, Xi D (1991) Fracture behaviour of bamboo. Mater Sci Progr 5(4):336–341(in Chinese)
45. Amada S, Untao S (2001) Fracture properties of bamboo. Compos B 32:451–459
Chapter 2
Mechanical Characteristics
and Stress–Strain Relationship
of Wood Structure

Abstract In this chapter, the mechanical characteristics of wood structure are


introduced that wood stem possesses cylindrical symmetry, and when cut a cube
with three orthogonal principal axes: longitudinal (L), radial (R), and tangential
(T) axes at a certain distance from the pith, it could be regarded as orthogonal
anisotropic body. Then, the stress–strain relationship equation of orthogonal ani-
sotropic body, the engineering elastic constants of wood, and the concept and
mechanical characteristics of plane stress and plane strain are introduced. At last,
electrometric method and digital speckle correlation method (DSCM) used to
measure the elastic coefficients of wood are described.

2.1 The Mechanical Characteristics of Wood Structure

Wood is a kind of porous layered biological composite. The physical and


mechanical properties of wood depend both on its chemical compositions and its
physical construction. The cellulose chain molecules of wood are composed of
carbon, hydrogen, and oxygen, which are used to form unit lattice according to
Meyer–Misch structure model [1]; then, cellulose basic fibril is constituted of unit
lattices according to Roelofsen structure model [2]. And it is easy for cellulose basic
fibrils to gather together to form microfibrils with larger diameter, which will be
embedded in the matrix composed of hemicellulose and lignin to construct lamella
structure; then, multiple lamella structures are arranged concentrically to structure
wood cell wall in polygon or cylinder [3], as seen in Fig. 2.1. Wood is constructed
of various cells with different functions, while the ultrastructure of wood cell wall
can be treated as a multilayered composite with microfibrils as reinforcement and
hemicellulose and lignin as matrix. What’s more, the microfibrils in each layer are
arranged in different angles, which can heavily influence the mechanical properties
of wood.
In macroscale, wood is a composite composed of bark, sapwood, and heartwood.
Wood stem consists of thin layers (growth ring layers) in the shape of cylinder
concentrically that endow wood with cylindrical symmetry, which is reflected in

© Springer Nature Singapore Pte Ltd. 2018 11


Z. Shao and F. Wang, The Fracture Mechanics of Plant Materials,
https://doi.org/10.1007/978-981-10-9017-2_2
12 2 Mechanical Characteristics and Stress–Strain Relationship …

Fibril

Amorphous
substance
Microfibril

Crystal-like
arrangement
Hydrogen bond
Cellulose chain
molecules
OH H CH2OH
H H OH O
H OH H H
OH
OH H H H OH OH
O
CH2OH OH H

Fig. 2.1 The sketch of wood structure

many physical properties of wood, such as elasticity, strength, thermal conductivity,


and electrical conductivity. In 1928, orthogonality principle was first used in wood
to explain the anisotropy of wood property by Price [4]. Let us cut a cube with three
orthogonal principal axes: longitudinal (L), radial (R), and tangential (T) axes at a
certain distance from the pith, as seen in Fig. 2.2. The three axes are called elastic
axes, and the coordinate directions are called principal elasticity directions. Wood
can be regarded as orthogonal anisotropic body, when studied in LRT rectangular
coordinate system [5].

Fig. 2.2 The principal axes L(x)


and plane of wood T(z)

R(y)
2.2 The Stress–Strain Relation of Solid Material 13

2.2 The Stress–Strain Relation of Solid Material

According to elasticity theory [6], the stress state or strain state of an arbitrary point
in solid, as seen in Fig. 2.3, could be expressed by six stress components or strain
components, namely, normal stress: rx ; ry ; rz , shear stress: syz ; szx ; sxy ; normal
strain: ex ; ey ; ez , shear strain: cyz ; czx ; cxy .
In the elastic range, stress component is a linear function of strain component
due to generalization Hooke law, as follows:
2 3 2 32 3
rx C11 C12 C13 C14 C15 C16 ex
6 ry 7 6 C21 C22 C23 C24 C25 C26 76 ey 7
7 6
6 7 6 7
6 rz 7 6 C31 C36 7 6 7
6 7¼6 C32 C33 C34 C35 7 6 ez 7 ð2:1Þ
6 syz 7 6 C41 C46 7 6 7
6 7 6 C42 C43 C44 C45 76 cyz 7
4 szx 5 4 C51 C52 C53 C54 C55 C56 4 czx 5
5
sxy C61 C62 C63 C64 C65 C66 cxy

In contrary, strain component is a linear function of stress component as well, as


follows:
2 3 2 32 3
ex S11 S12 S13 S14 S15 S16 rx
6 ey 7 6 S21 S22 S23 S24 S25 S26 7 6 ry 7
6 7 6 76 7
6 ez 7 6 S31 S36 7 6 7
6 7¼6 S32 S33 S34 S35 76 rz 7 ð2:2Þ
6 cyz 7 6 S41 7
S46 76 7
6 7 6 S42 S43 S44 S45 6 syz 7
4 czx 5 4 S51 S52 S53 S54 S55 S56 5 4 szx 5
cxy S61 S62 S63 S64 S65 S66 sxy

There are 36 constants in stiffness matrix C or flexibility matrix S, respectively.


And strain energy or strain energy density is a function of strain component ei ,
which is only determined by the final state of strain and has no relation with process
due to strain energy principle of elastic body. Hence, when stress ri is applied to

Fig. 2.3 The stress state of


an element
14 2 Mechanical Characteristics and Stress–Strain Relationship …

elastic body, the elementary work increment per unit volume dW has no relation
with order, namely:

dW ¼ ri dei ¼ Cij ej dei ¼ rj dej ¼ Cji ei dej ð2:3Þ

@W @W
¼ Cji ej ) ¼ Cji ð2:4Þ
@ej @ej @ei

@W @W
¼ Cij ej ) ¼ Cij ð2:5Þ
@ej @ei @ej

So: cij ¼ cji . Similarly, sij ¼ sji .


Stiffness matrix or flexibility matrix has a property of symmetry, so only 21
constants are independent. Therefore, in the linear elastic range, the relationship
expression between stress and strain is
2 3 2 32 3
rx C11 C12 C13 C14 C15 C16 ex
6 ry 7 6 C12 C22 C23 C24 C25 C26 7 6 ey 7
6 7 6 76 7
6 rz 7 6 C13 C36 7 6 7
6 7¼6 C23 C33 C34 C35 7 6 ez 7 ð2:6Þ
6 syz 7 6 C14 C46 7 6 7
6 7 6 C24 C34 C44 C45 76 cyz 7
4 szx 5 4 C15 C25 C35 C45 C55 C56 4 czx 5
5
sxy C16 C26 C36 C46 C56 C66 cxy

For anisotropic material, there is no interaction between normal stress and shear
strain, as well as shear stress and normal strain, and shear stress and shear strain, so
there are only nine independent constants in stiffness matrix. The stress–strain
relation in coordinate system along principal direction of anisotropic material can
be expressed as
2 3 2 32 3
rx C11 C12 C13 0 0 0 ex
6 ry 7 6 C12 C22 C23 0 0 0 7 6 ey 7
7 6
6 7 6 7
6 rz 7 6 C13 0 7 6 7
6 7¼6 C23 C33 0 0 7 6 ez 7 ð2:7Þ
6 syz 7 6 0 0 7 6 7
6 7 6 0 0 C44 0 76 cyz 7
4 szx 5 4 0 0 0 0 C55 0 4 czx 5
5
sxy 0 0 0 0 0 C66 cxy

If there are infinite symmetry planes in material, the relationship expression


could be simplified down to the situation in isotropic material that there are only
two independent constants. The stress–strain relation of isotropic material is
2 3 2 32 3
rx C11 C12 C13 0 0 0 ex
6 ry 7 6 C12 C22 C23 0 0 0 76 e y 7
6 7 6 76 7
6 rz 7 6 C13 C23 C33 0 0 0 76 e z 7
6 7¼6 76 7 ð2:8Þ
6 syz 7 6 0 0 0 ðC11  C12 Þ=2 0 0 76 cyz 7
6 7 6 76 7
4 szx 5 4 0 0 0 0 ðC11  C12 Þ=2 0 54 czx 5
sxy 0 0 0 0 0 ðC11  C12 Þ=2 cxy
2.3 Engineering Elastic Constants 15

2.3 Engineering Elastic Constants

Engineering elastic constants are the elastic constants of material including gen-
eralized elastic modulus, Poisson’s ratio, and shear modulus, which can be mea-
sured by simple tests, such as axial tensile test and pure shear test. Engineering
constants are more intuitive compared with stiffness coefficients and flexibility
coefficients in stiffness matrix and flexibility matrix, so engineers are used to
express generalized Hooke’s law in the form of engineering constants Ei, Gi, and
lij. The flexibility matrix of anisotropic material can be expressed with engineering
constants as follows:
2 l l
3
1
Ex  Eyxy  Ezxz 0 0 0
6 lxy l 7
6  Ey 1
 Ezyz 0 0 0 7
6 l Ey 7
  6  xz l
 Eyzy 1
0 7
6 Ex 0 0 7
Sij ¼ 6 Ez
7 ð2:9Þ
6 0 0 0 1
0 0 7
6 Gyz 7
6 0 1
0 7
4 0 0 0 Gzx 5
1
0 0 0 0 Gxy

where Ex, Ey, and Ez are the elastic modulus along three elastic axes x, y, and z,
respectively. lij is transverse sensitivity ratio of the strain in direction i caused by
the stress in direction j, called Poisson’s ratio.
ej
lij ¼  ð2:10Þ
ei

Gyz, Gzx, and Gxy are the shear modulus in x–y, z–x, and x–y planes, respectively.
For anisotropic material, the flexibility matrix has a property of symmetry, and it is
lij lij
¼ i; j ¼ x; y; z ð2:11Þ
Ei Ej

So, there are nine independent engineering constants.


Metal material applied in engineering is polycrystalline material, which can be
treated as isotropic body, so there are only two independent engineering constants,
namely, normal elastic modulus (Young’s modulus) E and shear elastic modulus
G. And the relationship expression among Poisson’s ratio (l), E, and G is

E
G¼ ð2:12Þ
2ð1 þ lÞ
16 2 Mechanical Characteristics and Stress–Strain Relationship …

2.4 Engineering Elastic Constants of Wood

Wood, as an anisotropic material, possesses cylindrical symmetry because of the


concentrically growth ring layers. When cut a cube with a plane tangent to growth
ring layer at a certain distance from the pith, it could be regarded as orthogonal
anisotropic body. So, the generalized Hooke’s law can be expressed as follows by
the engineering elastic constants:
2 3 2 32 3
eL 1=EL lLR =ER lLT =ET 0 0 0 sL
6 eR 7 6 lRL =E lRT =ET 0 7 6 7
6 7 6 L 1=ER 0 0 7 6 sR 7
6 eT 7 6 lTL =E lTR =ER 1=ET 0 0 0 7 6 sT 7
7 6
6 7¼6 L 7
6 cRT 7 6 0 7 6 7
6 7 6 0 0 0 1=GRT 0 76 sRT 7
4 cTL 5 4 0 0 0 0 1=GTL 0 54 sTL 5
cLR 0 0 0 0 0 1=GLR sLR
ð2:13Þ

Due to symmetry principle, the flexibility matrix above has the property of
symmetry [6], so the following relationship expressions can be obtained:
lLR lRL lLT lTL lRT lRT
¼ ; ¼ ¼ ð2:14Þ
ER EL ET EL ET ER

However, wood is porous, anisotropic, and inhomogeneous body in fact, so the


elastic coefficients measured in tests cannot satisfy the symmetry of above
expressions. For this reason, 12 engineering elastic constants need to be measured
in tests, and they are EL, ER, ET, GRT, GTL, GLR, lRT, lTR, lTL, lLT, lLR, and lRL.
The 12 engineering elastic constants can be obtained by electrometric method or
digital speckle method as shown in Sect. 2.6.

2.5 The Concept of Plane Stress and Plane Strain

The main study subject of mechanics of materials is bar member with


cross-sectional dimension much less than axis length. However, there may be load
carriers in other shapes, among which the most common two are plane and infinite
cylinder corresponding to plane stress state and plane strain state. It is necessary to
understand the concepts of the two stress states and the corresponding stress–strain
relationship, because the fracture property for the same material is different in
different stress states.
2.5 The Concept of Plane Stress and Plane Strain 17

2.5.1 Uniform Thickness Plate and Plane Stress

1. Geometric feature and mechanical characteristics


① Thickness is much less than length and width; ② external forces are loaded in
the peripheral area of plate and parallel to it middle plane; and ③ the upper surface
and lower surface are free, as seen in Fig. 2.4.

2. Stress characteristics
Due to description ③, for the upper surface and lower surface, there is
rz ¼ szx ¼ szy ¼ 0;
Due to description ①, at any point, there is rz ¼ szx ¼ szy ¼ sxz ¼ syz ¼ 0;
Due to description ②, there are only three stress components at any point, rx , ry
sxy ¼ syx , which are parallel to xy plane, and they are only the function of x and
y for the uniform thickness, as seen in Fig. 2.5.
The situation above is called as plane stress problem.

2.5.2 Infinite Cylinder and Plane Strain

1. Geometric feature and mechanical characteristics


① The size of the axis is great; ② external forces are perpendicular to axes and
keep uniform along axes, as seen in Fig. 2.6.

2. Stress characteristics
Cut a thin cross section at where both far from the two end, then:
➀ the stress on both sides are symmetric, szx = szy = sxz= syz = 0;

Fig. 2.4 An uniform thick y


flat
z

x
18 2 Mechanical Characteristics and Stress–Strain Relationship …

Fig. 2.5 Stress component of


plane stress state

Fig. 2.6 Long cylinder and


slice of cross section

➁ the cross section is restrained by both sides, and e = 0, so deformation just


occurs in xy plane;
➂ the stress in cross section is rx, ry, rz, and sxy, which are only the function of
x and y, as seen in Fig. 2.7.
The situation above is called as plane strain problem. Plane stress problem and
plane strain problem are called collectively as plane problem.

2.5.3 The Stress–Strain Relationship in Plane Problem

The stress–strain relationship of isotropic body in general stress state can be


described by generalized Hooke’s law, as follows:

Fig. 2.7 Stress component of


plane strain state
2.5 The Concept of Plane Stress and Plane Strain 19

8   8 s
< ex ¼ E1 rx  lðry þ rz Þ < cyz ¼ Gyz
e ¼ 1 r  lðrz þ rx Þ : c ¼ szx : ð2:15Þ
: y E1  y : zx sGxy
ez ¼ E rz  lðrx þ ry Þ cxy ¼ G

In plane stress state, as rz ¼ 0, szx ¼ 0 and szy ¼ 0, Hooke’s law can be sim-
plified as follows:
8
< ex ¼ E ðrx  lry Þ
1

ey ¼ E1 ðry  lrx Þ : ð2:16Þ


: 2ð1 þ lÞ
cxy ¼ G sxy ¼ E sxy
1

And ez ¼  El ðrx þ ry Þ, so it can be seen that ez 6¼ 0 in plane stress state.


In plane strain state, as ez ¼ 0, rz ¼ lðrx þ ry Þ, Hooke’s law can be simplified
as follows:
8
> 1l2 l
< ex ¼ E ðrx  1l ry Þ
>
2
l
ey ¼ 1l
E ðry  1l rx Þ
: ð2:17Þ
>
>
: c ¼ 1 s ¼ 2ð1 þ l2 Þ s
xy G xy E xy

l
E 0 and l0 are used to substitute 1l
E
2 and 1l, and then, the relations above can be

simplified as follows: 8 0
< ex ¼ E= ðrx  l ry Þ
1
0
e ¼ ðr  l rx Þ :
1 ð2:18Þ
: y E10 y
cxy ¼ G sxy

It can be seen that the form of Hook’s law in plane stress state is same with that
in plane strain state. Thus, if the geometrical shape of xy plane and the load
condition in plane stress state are same with that in plane strain state, once the
solutions of plane stress problem are obtained, the solutions of plane strain problem
can be obtained by substituting E 0 and l0 for E and l.

2.6 Tests of Wood Elastic Coefficients

There are many methods to measure the elastic coefficients of wood, such as
extensometer method, laser speckle interferometry method, ultrasonic method, and
electrometric method. Electrometric method was used to measure the elastic
coefficients of wood since 1960s. Digital speckle correlation method (DSCM) was
applied on wood by Xu et al. [7], Viotti et al. [8], and Wang et al. [9]. The elastic
coefficients of wood are important parameters for the design of wood structures and
20 2 Mechanical Characteristics and Stress–Strain Relationship …

Fig. 2.8 Six kinds of speckle F F F


sample for testing elastic
1# 2# 3#
modulus

EL µTL µRL ER µTR µLR ET µRT µLT

F F F
4# 5# 6#

GTL GLR GRT

wood composite, while the experimental data of wood elastic coefficients are not
sufficient to be used for design.
Among the 12 elastic coefficients, three elastic modulus (EL, ER, and ET) and six
Poisson’s ratio (lRT, lTR, lTL, lLT, lLR, and lRL) can be tested by electrometric
method and DSCM with the rectangle samples as shown in Fig. 2.8. The three
shear elastic modulus (GRT, GTL, and GLR) can be tested by three rectangle samples
with 45° off-axis due to the relationship between elastic coefficients of normal axis
sample and those of off-axis sample, as seen in Eq. (2.19).

EX45
Gij ¼ ð2:19Þ
2ð1 þ lXY Þ

where EX45 is the elastic modulus of sample with 45° off-axis.

2.6.1 The Application of Electrometric Method on Wood


Elastic Coefficients

Resistance strain gage is the sensor in electrometric method. It is that the strain of
sample can be converted into the resistance variation of resistance strain gage, and
the resistance variation can be detected and converted into the strain of sample by
resistance strain gage with an accuracy of 1 le. The testing bridge is shown in
Fig. 2.9, where R1 is working gage, R2 is temperature compensating gage, and R3
and R4 are the fixed internal resistances.
As wood is inhomogeneous body, the resistance strain gage should have large
gage length to extend test range. In this book, BQ12O-IOAA foil paper-based strain
gages are used, and they are attached on sample by No.502 glue, room temperature
2.6 Tests of Wood Elastic Coefficients 21

Fig. 2.9 Testing bridge B

R1
R2

A C U
R3

R4

curing. Before this, as wood is a porous material, the surfaces of sample should be
sanded to ensure the pores are filled with wood flour. The arrangement of resistance
strain gages has been shown in Fig. 2.8. And the coordinate lines on resistance
strain gage should overlap with the longitudinal line and transverse line drawn on
the surfaces of sample in advance.
The tensile, compressive, and bending elastic modulus of wood are about equal
[5], so uniaxial compression test can be performed to measure the elastic coeffi-
cients of wood. Before that, failure test should be performed on study subject to
determine the lower limit load and upper limit load, and they are 0.3 and 0.7 time of
proportional limit. Stepwise loading method is often adopted, and it is that test is
performed on a computer-controlled testing machine, and after setting up step
number k and step length DF, loading and unloading will be performed by test
machine automatically. The stepwise loading curve of spruce sample 1# (as seen in
Fig. 2.8) is shown in Fig. 2.10 with a load speed of 20 N/s. The initial load is
0.5 kN keeping 5 s, DF = 500 N, until the fifth step with a load of 2.5 kN keeping
5 s, then unloaded with a speed of 100 N/s, and the process should be repeated six
times. However, the maximum load and step length should be decreased corre-
spondingly for samples 2# to 6#.
The strain data can be read on test machine directly or automatically collected by
the assistance of strain indicator. Creep is inevitable for wood during loading, but if
test time is not long and increment formula is adopted, the influence of viscous

Fig. 2.10 Stepwise loading


curve of spruce sample
22 2 Mechanical Characteristics and Stress–Strain Relationship …

deformation on elastic strain rate can be ignored. The increment formula of wood
elastic coefficients is shown as follows:

DFi REi
Ei ¼ ; E¼
A  Dei n
 0 ð2:20Þ
Dei  Rl
li ¼  ; l ¼ i
De i n

2.6.2 The Application of DSCM on Wood Elastic


Coefficients

DSCM is a non-contacting optical experimental mechanical method proposed by


Yamaguchi [10] and Ranson and Peters [11], respectively. The fundamental of
DSCM is that two images of surface before and after deformation are collected, and
then the displacement and strain of the surface can be determined by analysis on the
probability statistics correlation of randomly distributed speckles before and after
deformation; consequently, the deformation field can be obtained. The measuring
tools for DSCM vary in a large range, such as optical microscope, electron
microscope, and atomic force microscope that can measure tiny displacement with a
sensitivity of 0.01–0.05 pixel. What’s more, dynamic measurement can be achieved
by DSCM with the help of high-speed video recording system or camera system.
The strain of wood sample surface is often measured by electrometric method
which is a mature experimental mechanical method. However, in electrometric
method, data are collected by a limited number of strain gages on the base of point
measurement, so the whole deformation field of sample cannot be obtained. In
addition, strain gages are attached on sample by glue in that method, and when the
stiffness of the base material of strain gage is higher than that of wood, some errors
will be caused. Xu et al. [7] used DSCM to measure the compression elastic
modulus, EL, ET, and ER, of loblolly pine (Pinus taeda), and the difference between
the results obtained by DSCM and electrometric method was in the range of 1–7%,
which indicates that it is feasible for DSCM to be applied to measure the elastic
modulus of wood. What’s more, DSCM has the advantages of simple optical path
and easy implementation. The experimental device and experimental process to
measure the 12 elastic coefficients of wood are introduced as follows.

(1) Speckle
After the surfaces of the six samples were polished, black matted paint and white
matted paint were sprayed on the surfaces of sample by high-pressure spray method
in sequence to form artificial speckle. The size of speckle would affect the accuracy
of associated processing, which should satisfy the requirement of CCD’s resolution
and sampling theorem. Too small speckle would cause a great error during the
2.6 Tests of Wood Elastic Coefficients 23

Fig. 2.11 Schematic diagram Loading


of hardware equipment force F Speckle area
installation of DSCM
Light

PC
CCD Image card

processing of different values, while too large speckle would affect the sensitivity of
the variation of correlation coefficient to the deformation of sample surface and
cause an error in related calculation process.

(2) Image acquisition system


As shown in Fig. 2.11, image acquisition system was composed of CCD camera,
image collection card, white light source, and computer with image processing
software, among which cold light source was adopted to avoid the influence of heat
on the strain of sample.

(3) Test
Stepwise loading method was applied in DSCM just like that of electrometric
method. Step number k, step length DF, and load maintaining time were set up in
advance. Loading and unloading will be performed by test machine automatically,
and speckle image would be collected during each load maintaining stage, as seen
in Fig. 2.12.

Fig. 2.12 Two speckle images of 1 and 1.5 kN load


24 2 Mechanical Characteristics and Stress–Strain Relationship …

(4) Image processing


Input the two speckle images collected before and after deformation of sample
surface corresponding to the load increment DFi into image processing software,
the elastic coefficients were calculated by computer automatically, as seen in
Fig. 2.13.

(5) Calculation of elastic modulus


The strain increment De corresponding to load increment DF could be calculated
according to the following method: taken a square area as calculation area in the
middle of speckle zone, calculation points were selected at intervals of m pixels,
and 2n  2n points in total, then the vertical strain increment and horizontal strain
increment corresponding to load increment DFi could be calculated as follows
(Fig. 2.14):

1X n
vi;j þ n  vi;j 1 X2n
Dei ¼ ; De ¼ Dei
n j¼1 nm 2n i¼1
ð2:21Þ
0 1X n
ui þ n;j  ui;j 0 1 X2n
Dej ¼ ; De ¼ Dej
n j¼1 nm 2n j¼1

So, the elastic coefficients of wood could be calculated:

Fig. 2.13 The document of


data
2.6 Tests of Wood Elastic Coefficients 25

Fig. 2.14 Calculation area of F


the strain increment

(1,1) (2n,1)

(1,2n) (2n,2n)

PK
DFk Ek
Ek ¼ ; E ¼ k¼1
A  Dek K
 0 PK ð2:22Þ
Dek  l
lk ¼  ; l ¼ k¼1 k
De k K

where DFk was the load increment, and A was the cross-sectional area of sample.

2.6.2.1 The Limitation of Elastic Coefficients

The elastic coefficients of orthotropic material should meet the following require-
ments: the product of stress component and corresponding strain component should
be the work done by stress; the sum of work done by all stress components should
be a positive value; thus, flexibility or stiffness matrix must be positive definite [12].
The requirements above present a thermodynamic limitation on the value of elastic
coefficients, and if described with engineering coefficients, the elastic coefficients of
orthotropic material should satisfy the following equations:

Ei [ 0; Gij [ 0 ð1  lij lji Þ [ 0 l2ij \Ei =Ej ði; j ¼ L; R; TÞ ð2:23Þ

Then, flexibility or stiffness matrix of elastic material is symmetric theoretically,


as shown in the following equation:

lij =Ei ¼ lji =Ej ði; j ¼ L; R; TÞ ð2:24Þ

However, researches indicated that there were always some differences between
the measured value and theoretical value of wood elastic coefficients, and the
26 2 Mechanical Characteristics and Stress–Strain Relationship …

relative error was in the range of 15–25% [12–14]. The main reasons are as follows:
wood is approximately cylindrically symmetric, but rectangular coordinate system
is adopted in the orthotropic hypothesis of wood, which leads to a certain model
error especially to wood with small diameter and growth ring with large curvature;
wood properties vary greatly; the errors come from sample processing and the
attachment of strain gage, such as deviation longitudinal axis of sample and fiber
orientation, error on the angle of strain gage and the attachment quality of strain
gage, and so on.

References

1. Meyer KH, Misch L (1937) Posiondes atemes dans le nouveru modele spatial do la cellulose.
Helv Chim 20:232–244
2. Roelofsen PA (1959) The plant cell-wall. Gegruder Borntraeger, Berline-Nikolassee, pp 126–
189
3. Junqing Cheng (1985) Wood science. China Forestry Publishing House, Beijing, p 1379
4. Price AT (1928) A mathematical discussion on the structure of wood in relation to its elastic
properties. Philos Trans R Soc Lond 228:1–62
5. Kollmann FFP (1991) Wood science and principle of wood technology (translation). China
Forestry Publishing House, Beijing, pp 278–281
6. Jones RM (1972) Mechanics of composite materials. Scripta Book Company, New York
7. Xu M, Jin G, Lu Z (2003) Digital spackle correlation method (DSCM) for measurement of
wood compression elastic modulus. Scientia Silvae Sinicae 39(2):174–176
8. Viotti MR, Kaufmann GH, Galizzi GE (2006) Measurement of elastic moduli using spherical
indentation and digital speckle pattern interferometry with automated data processing. Opt
Laser Eng 44(6):495–508
9. Wang QH, Xie HM, Tang PF et al (2009) A study on the mechanical properties of beagle
femoral head using the digital speckle correlation method. Med Eng Phy 31(10):1228–1234
10. Yamaguchi IA (1981) Laser-speckle strain gage. J Phys E Sci Instrum 14:1270–1273
11. Ranson WF, Peters WH (1982) Digital image techniques in experimental stress analysis. Opt
Eng 21(3):427–431
12. Lempriere BM (1968) Poisson’s ratio in orthotropic materials. AIAA 6:2226–2227
13. Liu Z, Liu Y, Yu H et al (2004) Research progress of automatic measuring technology of
wood modulus of elasticity. Forest Sci Technol 29(1):45–48
14. Wang X, Ross RJ et al (2001) Several nondestructive evaluation techniques for assessing
stiffness and MOE of small-diameter logs. Research Paper, FPL-RP-600, USDA, Forest
Service, FPL, Madison, WI, p 12
Chapter 3
Fracture of Wood Along Grain

Abstract In this chapter, the theory of LEFM and the special application of LEFM
on wood are introduced. Then, different samples and methods are applied to
measure the Mode I fracture toughness of wood along grain, and the results show
that the fracture toughness of wood along grain is the basic attribute of wood and
has nothing to do with test method, geometrical shape, and size of crack. The
research indicates that LEFM based on isotropic body is applicable to the crack
propagation of wood along grain. And based on energy theory, double cantilever
inversion symmetry bending load method is applied to measure the Mode III
fracture toughness of spruce. The results show that the average Mode III inter-
laminar fracture toughness 1.05 kJ/m2 can be seen as a basic attribute of spruce that
represents the capacity of spruce to resist the propagation of Mode III crack.

3.1 Introduction

In recent years, with the further application of fracture mechanics on wood, the
application fields have developed from safety assessment to processing and uti-
lization of wood. Currently, there are many test methods to determine the fracture
parameters of wood, among which fracture toughness is one of the most important
indexes. Fracture toughness represents the capacity of wood to resist the unstable
propagation of opening crack, which has significant meaning for the quality
assessment, safety design, and optimization of processing method of wood.
However, there may be great difference in the results of different test methods,
because there is no universal test standard. What’s more, comparative studies
among different test methods are not sufficient as well. In this chapter, the partic-
ularity of the application of LEFM on wood is discussed, then the fracture property
of wood along grain and test method are studied, at last, fractal theory is applied to
study the fractal feature of the fracture surface of wood, and the relationship
between fractal dimension of fracture surface and fracture toughness is developed.

© Springer Nature Singapore Pte Ltd. 2018 27


Z. Shao and F. Wang, The Fracture Mechanics of Plant Materials,
https://doi.org/10.1007/978-981-10-9017-2_3
28 3 Fracture of Wood Along Grain

3.2 Theory of LEFM

3.2.1 Crack

Crack and defects are inevitable in engineering materials, which might be generated
during manufacture, processing, or using, such as fatigue crack caused by alter-
nating loads, radial or ring shake caused by compression damage or drying. As seen
in Fig. 3.1, crack can be classified into three types due to the characters of load and
fracture.
Mode I (Opening mode) crack—external normal stress is perpendicular to crack;
Mode II (Sloping mode) crack—shearing stress is parallel to crack;
Mode III (Tearing mode) crack—stress staggers crack surfaces.
Among the three types of crack, Mode I and Mode II cracks are more common
and dangerous than Mode II crack. If there are normal stress and shearing stress or
tensile stress is not perpendicular to crack line on crack body, Mode I and Mode II
(Or Mode III) cracks will occur in material at the same time, which is called mixture
crack. And in fact, mixture crack is the most common situation, and it is treated as
Mode I crack sometimes from the angle of safety and convenience. Thus, Mode I
crack is the research focus of engineering fracture mechanics.

3.2.2 Stress Intensity Factor K and K Criterion

Fracture mechanics is a subject on the fracture strength of structure with crack or


defect. Before the rise of fracture mechanics, Sneddon [1] demonstrated that stress
field at crack tip had singularity of r−1/2 step by elasticity mechanics, namely, when
r ! 0, all the stress components would increase infinitely (rij ! 1). Obviously, it
is meaningless to apply traditional strength condition to assess the safety of
structure with crack, as shown in the following equation:

rmember  ½r ð3:1Þ

Fig. 3.1 Three typical y


interlaminar fracture modes

r x

Mode I Mode II Mode III


3.2 Theory of LEFM 29

To study the stress distribution at crack tip, Irwin illuminated the general form of
stress field at crack tip by Westergaard stress function [2, 3] as follows:
pffiffiffiffiffiffiffiffiffi
r1 p  a
rij ¼ p ffiffiffiffiffiffiffiffiffiffiffi  fij ðhÞ ð3:2Þ
2p  r

where a is the size of crack; r1 is the working stress far away from crack tip; r and
h are polar coordinates with crack tip as origin; fij ðhÞ is azimuth function neglecting
the higher order term of r.
As seen in Fig. 3.2, there is a central through crack with a length of 2a in a plate.
It is assumed that the thickness of plate is one unit, and length and width are much
greater than 2a, so the crack can be treated as Mode I crack in an infinite plate under
tensile stress. The stress and strain fields in the small area at crack tip are
pffiffiffiffiffiffi  
r pa h h 3h KI
rx ¼ pffiffiffiffiffiffiffiffi  cos 1  sin sin ¼ pffiffiffiffiffiffiffiffi f ðhÞ
2pr 2 2 2 2pr
pffiffiffiffiffiffi  
r pa h h 3h KI
ry ¼ pffiffiffiffiffiffiffiffi  cos 1 þ sin sin ¼ pffiffiffiffiffiffiffiffi f ðhÞ
2pr 2 2 2 2pr
pffiffiffiffiffiffi
r pa h h 3h KI
sxy ¼ pffiffiffiffiffiffiffiffi  sin cos cos ¼ pffiffiffiffiffiffiffiffi f ðhÞ ð3:3Þ
2pr 2 2 2 2pr
pffiffiffiffiffiffi rffiffiffiffiffi   rffiffiffiffiffi
r pa 2r 2h h KI 2r
u¼  ð1  lÞ þ ð1 þ lÞ sin cos ¼ f ðhÞ
E p 2 2 E p
pffiffiffiffiffiffi r ffiffiffiffi
ffi   r ffiffiffiffi

r pa 2r h h KI 2r
v¼  ð1  lÞ þ ð1 þ lÞ sin2 sin ¼ f ðhÞ
E p 2 2 E p
E
where for plane stress: E ¼ E, l ¼ l; for plane strain: E ¼ 1l 2 , l ¼ 1l.
l

Fig. 3.2 Schematic diagram


of a central through crack
within a plate and the feature
of stress element at crack tip y
y
xy
r
x σx σx
2a

y
30 3 Fracture of Wood Along Grain

It can be seen that for an arbitrary point in front of crack, due to Eq. 3.3, the
pffiffiffiffiffiffiffiffiffi
stress at the point is determined by r p  a called as stress intensity factor K, which
is a new physical quantity. Generally, stress intensity factor can be written as
pffiffiffiffiffiffiffiffiffi
K ¼r paY ð3:4Þ

where r is the working stress assuming there was no crack in the center, and Y is
shape factor related to the shape and location of crack.
Thus, stress intensity factor K is the function of working stress, crack size, and
shape. And according to the comprehensive factors, when K reaches to a critical
value KC, crack will propagate unstably, so a basic criterion in LEFM is established:

K ¼ KC ð3:5Þ

The international unit of K is MN m3=2 or MPa m1=2 .


For the three types of crack body, the corresponding stress intensity factors are
KI, KII, and KIII. If the sample has enough thickness that can be seen in plane strain
state, when crack propagated unstably, the lower value of stress intensity factors KC
is usually a constant, namely, the basic attribute of material. Critical stress intensity
factor represents the capacity of material with crack to resist fracture in the range of
linear elasticity, which is called as plane strain fracture toughness of material or
fracture toughness. Thus, it is an important task in LEFM to solve the stress
intensity factor or measure the fracture toughness of material with crack.

3.2.3 Energy Release Rate G and G Criterion

As energy will be consumed in the propagation process of crack, the relationship


expression when fracture happens can be obtained according to the balance of
energy transformation before and after the moment crack propagates unstably.
For a crack body with a thickness of B and a crack area of A, if the area of crack
increased is dA, the external work is dW, the change of elastic strain energy is dU in
the propagation process of crack, the increment of surface energy for new crack
surface is dC, and the consumed plastic work is dP. It is assumed that the propa-
gation of crack is a thermal isolation process under static load with no change of
thermal or inertia force, so due to the law of the conservation of energy that external
work is equal to the internal energy of the system, there is

dW ¼ dU þ dC þ dP ð3:6Þ

Or
3.2 Theory of LEFM 31

dW  dU ¼ dC þ dP ð3:7Þ

In Eq. (3.7), dC þ dP is the energy consumed when crack area increases dA,
namely the energy needed to resist the propagation of crack, which is determined by
the toughness of material. For ideal brittle material, dP ¼ 0, while for metal
material, dP  dC, where dC can be neglected.
dW  dU in Eq. (3.7) represents the energy released by the system when crack
area increases dA, which is used to drive the propagation of crack. If П is total
potential energy, P ¼ U  W, then dP ¼ dW  dU represents the energy
released when crack increases da, which is the active force to drive the propagation
of crack. So there is:

dP ¼ dW  dU ¼ dC þ dP ð3:8Þ

The energy release rate (G) of crack is defined as the energy released by elastic
system when crack propagates per unit area, so there is

@P @W @U
G¼ ¼  ð3:9Þ
@A @A @A

If the thickness of crack body is B, the length of crack is a, when crack increases
da, the increased area is dA ¼ Bda; so Eq. (3.9) can be written as

1 @P
G¼ ð3:10Þ
B @a

The energy consumed when crack propagates per unit area is defined as ratio of
crack propagation resistance (R or GC), so

@C @P
R ¼ GC ¼ þ ð3:11Þ
@A @A

To a certain material, the crack surface work and plastic work of crack propa-
gation are material constants that have nothing to do with external load and the
geometry of crack, so R or GC reflects the ability of material to resist fracture
damage, called as fracture toughness of material and is determined by test.
When energy release rate G reaches to GC, the crack system would loss balance
and crack begin to propagate unstably. So G criterion is

G ¼ GC ð3:12Þ

The international unit of G and GC is N m1 or J m2 .


32 3 Fracture of Wood Along Grain

3.2.4 Relationship Between K and G

There are two views on the study of crack propagation law by linear elastic fracture
mechanics: K criterion and G criterion. Though the starting points of these two
criteria are different, they are actually the same, because there is a certain rela-
tionship between G and K under the linear elastic conditions. Taking the mode-I
crack of isotropic material as an example, due to Westergaard [3] stress function,
the calculation formulas of stress and displacement perpendicular to crack surface at
crack tip are
 
KI h h 3h
ry ¼ pffiffiffiffiffiffiffiffi cos 1 þ sin  sin ð3:13Þ
2pr 2 2 2
rffiffiffiffiffi 
ð1 þ lÞKI 2r h 3h
v¼ ð2v þ 1Þ sin  sin ð3:14Þ
4E p 2 2

ðPlanestressÞ
3l
where E is elastic modulus, l is Poisson’s ratio, v ¼ 1þl .
3  4l ðPlanestrainÞ
When crack length increases Da, if the energy released equals to the work done
to close the opening crack to its original state, as seen in Fig. 3.3, the energy
released can be calculated as follows:

Fig. 3.3 Schematic diagram


of stress field at crack tip and
the propagation of crack
3.2 Theory of LEFM 33

ZDa
1
U¼ ry  2v  Bdn ð3:15Þ
2
0

Notice that the coordinates origin of ry is O, and the coordinates origin of


opening displacement v is O′. When crack propagates per unit area, the energy
release rate is

ZDa
U 1  
GI ¼ ¼ ry h¼0;r¼n  vh¼p;r¼Dan  dn
B  Da Da
0
ðv þ 1Þð1 þ lÞ 2 ð3:16Þ
¼ KI
4E
1
¼ 0 KI2 ¼ S0 KI2
E

0
E ðPlane stressÞ 0
where E ¼ E
ðPlane strainÞ , S ¼ E0 is compliance.
1
ð1lÞ2
For Mode II and Mode III crack, there are similar relationships (Assuming crack
propagates along the original crack line):

1 2
GII ¼ K ð3:17Þ
E0 II
1l 2
GIII ¼ KIII ð3:18Þ
E0

The relationship between K and G indicates that K is the strength of elastic stress
field around crack tip, while the square of K could determine the energy release rate
of crack propagation, therefore, K criterion equals to G criterion when solving linear
elastic fracture problems. G criterion has a simple form and clear concept, but the
analytical expression of G should be given when a crack is introduced to a structure,
which will encounter great difficulty in mathematical and mechanical analysis.
Irwin pointed out that for a linear elastic crack body, G could be expressed by the
compliance of the system [4]. For a plate with a central through crack, G is dis-
cussed in constant displacement situation and constant load situation, respectively.
1. Constant displacement situation
As seen in Fig. 3.4, for an elastic crack body, the deformation energy is U ¼ 12 FC,
and the displacement is D ¼ CF, where C is compliance. And U is only in related
34 3 Fracture of Wood Along Grain

(a) (b)
F
A
FA
U
G=−
2a 2a +2da A

FB B

C
0

Fig. 3.4 The energy release rate of crack propagation in the situation of constant displacement

with the final state and has nothing to do with process, so when crack propagates
from a to a + da, DU = Ua+da − Ua.
Case A: Crack length 2a remained unchanged, when plate is stretched over D, the
deformation energy UA =FAD/2;
Case B: Crack length 2a +2da remained unchanged, when plate is stretched over D,
the deformation energy UB = FBD/2.
Thus, after elastic plate is stretched over D under load F, the up and down ends
of plate are fixed to form a energy closed system with constant displacement. When
crack propagates from 2a to 2a +2da, deformation energy of plate will be released
and decrease, namely @U @A ¼ ðSOBC  SOAC Þ\0.
As the displacement of load point is unchanged, namely, dD = 0, the increment
of external work is dW = 0, thus Eq. (3.9) can written as:
   
@ðW  UÞ @U 1 @U
GI ¼ ¼ ¼ ð3:19Þ
@A @A D B @a D

The equation above shows that in the constant displacement situation, the strain
energy released by the system is used to drive the propagation of crack, so the
energy rate of crack propagation is the energy release rate of elastic body. And the
compliance C of elastic body is the function of crack length, namely, C = C(a).
As

dD ¼ FdC þ CdF ¼ 0
1 1 1 1
dU ¼ FdD þ DdF ¼ FCdF ¼  F 2 dC
2 2 2 2
3.2 Theory of LEFM 35

So substituting dU into Eq. (3.19), the following can be obtained:


   
@ðW  UÞ @U 1 @U F 2 @C
GI ¼ ¼ ¼ ¼  ð3:20Þ
@A @A D B @a D 2B @a

2. Constant load situation


As seen in Fig. 3.5, for an elastic crack body under constant load F, when crack
increases da, the increment of displacement is dD ¼ FdC þ CdF, thus the variation
of strain energy is

1 1 1
dU ¼ FdD ¼ FðFdC þ CdFÞ ¼ F 2 dC ð3:21Þ
2 2 2

The variation of external work is

dW ¼ FdD ¼ F 2 dC ¼ 2dU ð3:22Þ

Thus, the energy release rate of elastic body is


   
@ðW  UÞ @U 1 @U F 2 @C
GI ¼ ¼ ¼ ¼  ð3:23Þ
@A @A F B @a F 2B @a

It can be seen from the equation above that the energy consumed by the prop-
agation of crack is the energy of external work minus the increment of elastic strain
energy in constant load condition.
The comparison between Eqs. (3.20) and (3.23) indicates that no matter the
situation is constant displacement or constant load, energy release rate GI has a
uniform expression:

(a) (b)
F
A B
F

2a 2a +2da

F
F 0 A B

Fig. 3.5 The energy release rate of crack propagation in the situation of constant load
36 3 Fracture of Wood Along Grain

F 2 @C
GI ¼  ð3:24Þ
2B @a

Or dimensionless crack length a/W can replace crack length a, then Eq. (3.24)
can be written as

P2 @C
GI ¼  ð3:25Þ
2BW @ða=WÞ

Equations (3.24) or (3.25) are called Irwin–Kies relationship expression and it is


the base of calibration test of energy release rate, which does not depend on the
compliance of test machine but only depend on the change rate of compliance
caused by the propagation of crack. Thus, in experiment, it only need to measure
the change rate of compliance C with the variation of crack length a, then energy
release rate G can be calculated due to Eq. (3.24) or Eq. (3.25), at last according to
K–G relationship expression, stress intensity factor K can be obtained.

3.3 Fracture Mechanics of Anisotropic Material

For the fracture problem of anisotropic material, the propagation of crack is com-
plex, and even for Mode I crack in anisotropic material, it may not propagate along
the extension of the original crack. The fracture toughness of anisotropic material is
in close relation to the direction of material property.
The analytical solution of stress, strain, and displacement field at crack tip has
been given in the early research work of Sih et al. [4] and Wu [5, 6]. The stress and
strain around the Mode I crack tip in the center of infinite anisotropic plate
(Fig. 3.2) can be expressed as
  
KI n1 n2 n2 n1
rx ¼ pffiffiffiffiffiffiffiffi  Re 
2pr n  n2 u2 u1
 1  
KI 1 n1 n2
ry ¼ pffiffiffiffiffiffiffiffi  Re 
2pr n1  n2 u2 u1
  
KI n1 n2 1 1
sxy ¼ pffiffiffiffiffiffiffi
ffi  Re  ð3:26Þ
2pr n1  n2 u1 u2
rffiffiffiffiffi  
2r 1
u ¼ KI  Re ðn p2 u2  n2 p1 u1 Þ
p n1  n2 1
rffiffiffiffiffi  
2r 1
v ¼ KI  Re ðn1 q2 u2  n2 q1 u1 Þ
p n1  n2

where
3.3 Fracture Mechanics of Anisotropic Material 37

u1 ¼ ðcos h þ n1 sin hÞ1=2


u2 ¼ ðcos h þ n2 sin hÞ1=2
pffiffiffiffiffiffi
KI ¼ r pa
p1 ¼ S11 n21 þ S12  S16 n1
ð3:27Þ
p2 ¼ S11 n22 þ S12  S16 n2
S22
q1 ¼ S11 n1 þ  S26
n1
S22
q2 ¼ S12 n2 þ  S26
n2
n1 and n2 are the unequal complex roots of the following equation:

S11 n4  2S16 n3 þ ðS12 þ S66 Þn2  S26 n þ S22 ¼ 0, where, Sij is the coefficient of
flexibility matrix.
For Mode I crack problem, as load is acted on crack body, the energy released by
crack body still equals to the work done to close the opening crack to its original
state:
ZDa
1
U¼ ry  2vBdn
2
0
ZDa
  ð3:28Þ
¼ ry h¼0;r¼n  vh¼p;r¼Dan  Bdn
0
 
1 n þ n2
¼ BKI2 S22 Re i 1
2 n1 n2
pffiffiffiffiffiffi
where K ¼ r pa.
So the energy release rate is
 
U K2 n þ n2
GI ¼ ¼ I S22 Re i 1 ð3:29Þ
B  Da 2 n1 n2

Similarly,
KII2
GII ¼ S11 Re½iðn1 n2 Þ ð3:30Þ
2

For orthotropic material, when Mode I crack is parallel to one of the symmetry
planes, as S16 ¼ S26 ¼ 0, the relationship expression of K and G with four inde-
pendent elastic constants can be obtained. n1 and n2 are the roots of the following
equation:
38 3 Fracture of Wood Along Grain

S11 n4 þ ðS12 þ S66 Þn2 þ S22 ¼ 0 ð3:31Þ

And n1 and n2 satisfy the following equations:


"  #1=2
pffiffiffi S22 1=2 2S12 þ S66
n 1 þ n2 ¼ i 2 þ
S11 2S11
ð3:32Þ
 1=2
S11
n1 n2 ¼ 
S12

So, energy release rate is


  1   
S11 S12 2 S22 1 2S12 þ S66 1
GI ¼ KI2  S ¼ KI2  þ ð3:33Þ
2 S11 2 2S11 2
  1
 S11 S22 1 2S12 þ S66 2
GII ¼ KII2 S ¼ KII2  pffiffiffi þ ð3:34Þ
2 S11 2 2S11

S11S12
12 h S22
12 i12
where S* is equivalent flexibility, and S ¼ 2 S11 þ 2S12 þ S66
2S11 .
Sij can be appropriately replaced by engineering elastic parameters, thus the
equivalent flexibility of Mode I crack body can be expressed as
2 !1=2 31=2
 1=2  1=2 lTL
1 E 2 þ 1
S ¼ 4 ET GLT 5
L
þ ð3:35Þ
2EL ET ET 2
EL

The small value term with Poisson’s ratio can be neglected, so the equivalent
flexibility can be written as
 1=2 " 1=2   #1=2
 1 EL EL 1=2
S ¼  þ ð3:36Þ
2EL ET ET 2GLT

Sih et al. [4] proved that the concept and relationship between elastic stress and
input energy of isotropic crack body can be generalized to the similar situation of
anisotropic material. And if self-balancing forces are acted on crack body, the form
of 2D stress intensity factor of anisotropic material is same with that of isotropic
material. Therefore, the related concepts in classical fracture theory are appropriate
for the crack problem of anisotropic material, for example,
P2 @C
GI ¼  ð3:37Þ
2BW @ða=WÞ

So
3.3 Fracture Mechanics of Anisotropic Material 39

P2 @C
KI2 ¼  ð3:38Þ
2BWS @ða=WÞ

Conclusion can be obtained that if the crack of anisotropic material is very


similar to that of isotropic material, LEFM can be applied to general anisotropic
material [7], namely, if the KC is known, the analysis method of anisotropic material
in most actual cases is same with that of isotropic material. However, the conclu-
sion is valid only when there is no unbalanced force on crack surface and the
propagation of crack is self-similar.

3.4 The Special Application of LEFM on Wood

As known wood is anisotropic, inhomogeneous material, and the stress–strain curve


of air-dried wood under a certain load has linear characteristic conforming with
linear elastic behavior, so it can be approximately seen as an orthotropic material
with three elastic symmetry planes perpendicular to the longitudinal direction (L),
radial direction (R) and transverse direction (T) of wood, respectively. Thus, if the
normal direction of crack surface is represented by one symbol, and the propagation
direction of crack is represented by another symbol, there are six kinds of cracks
basically, as seen in Fig. 3.6, and they are TL, RL, LT, LR, TR, and RT.
The fracture of anisotropic material is much more complex than that of an
isotropic material. Sih et al. [4] derived the equations of stress and displacement
field at crack tip in isotropic material by complex function, as follows:

RL RT LR

TL TR LT

Fig. 3.6 Six planes of crack propagation


40 3 Fracture of Wood Along Grain

Fig. 3.7 Schematic diagram


of crack and fiber direction y
x

2a

K

rij ¼ pffiffiffiffiffiffiffiffi  Re fij h; a; aij ; u1 ; u2 ; u3
2pr
rffiffiffiffiffi ð3:39Þ
K 2r

vij ¼   Re fij h; a; aij ; u1 ; u2 ; u3
G p

where aij is elastic constant of material, Re is the real part of complex function fij.
u1, u2, and u3 are complex parameters depending on the anisotropic degree of
material and the angle a between crack and longitudinal fiber (Fig. 3.7).
From the equations above, it can be found that the fracture of wood is much
more complex than that of isotropic material. The difference between anisotropic
material and isotropic material has been discussed by Shahrokh Parhiagar et al. [8],
when LEFM was used in composite materials. However, for wood, the great dif-
ferences lie in the following aspects:
1. In general, crack in wood always propagates along the direction of fibers instead
of its original direction, while LEFM presupposes that crack propagates along
its original direction all the time.
2. Even under simple load, compound form displacement may occur at crack tip in
wood, which is different with Mode I, Mode II, and Mode III crack in LEFM.
3. The stress field at crack tip in wood is the function of complex parameters that
are the function of wood property and the angle a between crack and longitu-
dinal fiber, which is different with LEFM developed from the stress field at
crack tip and in no relation with material property and crack direction.
Generally, LEFM is not applicable to approximately orthotropic wood consid-
ering the three facts above. And there are no material constants KIC, KIIC, and KIIIC
for wood to represent its fracture toughness, unless all the cases that cracks are at
different angles to fiber direction are studied and the applicability of LEFM to each
case is verified, which is very inconvenient even impossible in practice. However, if
crack is along fiber direction and the orthotropic principal axes are superposed with
the direction of crack surface and the propagation direction of crack, respectively,
the three differences can be eliminated, and this particular situation has been proved
by many experiments that:
3.4 The Special Application of LEFM on Wood 41

1. Crack propagates along its original direction in the same direction with fibers;
2. Displacement is not in compound form;
3. To a fixed angle between crack and fiber (a = 0), the complex parameters of
material are constants, so the stress at crack tip is only the function of r and h.
The situation above indicates that LEFM is applicable to material when original
crack is along fiber direction. For wood, most of the crack and defects formed
during growth and processing are along fiber direction, while the resistance of wood
to the propagation of crack along fiber direction is minimum. In experiments, TR
crack and RT crack are similar to radial shake and ring shake of wood, respectively,
so LEFM is applicable to wood with a crack propagating along fiber direction,
namely grain. There will be significant meaning for the quality assessment, safety
design, and optimization of processing method of wood to study and measure the
fracture toughness along grain of wood.

3.5 The Stress Intensity Factor KIC of Wood Fracture


Along Grain

3.5.1 The Methods to Test Stress Intensity Factor

Now there are test standards at home and abroad to measure KIC of materials, such
as ASTM-E399 [9] in America, BS5447 [10] in Britain and GB4161-84 [11] in
China. And there are various samples, such as compact tension (CT) sample, single
edge notched bending (SEB) sample, tensile (TS) sample and double cantilever
beam (DCB) sample, and so on (Fig. 3.8), and the improved forms of above
samples.
In theory, KIC is the natural property of material and has nothing to do with load
mode, sample form, and size of sample. From the view of measurement, the CT
method is simple and easy to perform on specimen with small size, which is
particularly suitable for wood with small diameter class; SEB sample often used to
measure transverse fracture toughness needs large diameter timber to make crack
along grain; TS sample has high requirements for tensile grips; DCB sample is
adopted by compliance method by which the energy release rate can be measured,
while nine elastic coefficients need to be measured to calculate KIC, so the workload
is heavy. Therefore, CT sample is always applied, and the validity of result can be
improved by increasing sample number.
There are three types of load–displacement (F-d) curves in the test of KIC, as
seen in Fig. 3.9. To a certain sample, it needs to measure crack size and geometrical
dimensions of sample before test and determine the critical load Fcr after test to
calculate KIC. It should follow the requirements below to determine Fcr. For brittle
sample or sample with large size, it will fracture unstably once crack initiates with
no obvious subcritical propagation stage, then the maximum load Fmax is the critical
load Fcr. In general, crack would propagate slowly before sample fractures, and
42 3 Fracture of Wood Along Grain

Fig. 3.8 The basic samples F


of testing for fracture F
toughness

H
a

F B
a
e W

CT sample TS sample
F

e a W
F
DCB sample

B
L/2 L/2

SEB sample

there is no visible sign unstable of propagation, thus the maximum load is not the
critical load. For metal material without visible yield stage, the rule is that the stress
corresponding to 0.2% of strain is regarded as nominal yield limit. And in many
literature and national standard, the critical load is the corresponding load when the
propagated length Da equals to 2% of the initial crack length.
3.5 The Stress Intensity Factor … 43

Fig. 3.9 Basic forms of F Fmax Fmax


load–displacement (F-d)
curve
Fcr =F 5 Fcr Fcr = Fmax
5%

Type 1 Type 2 Type 3

However, in practical test, the curve obtained is F-d curve but not F-Da curve, so
the corresponding load (Da/a = 2%) should be determined on F-d curve. In theo-
retical analysis and approximation [12], it is found that the slope of secant line of
the corresponding load on the F-d curve decreases 5% compared with the slope of
the line segment of the F-d curve before the propagation begins, so graphing
method can be used to determine the critical load on the F-d curve.
The third type of curve will be obtained if sample has enough thickness. When
loaded, sample is in plane strain state except the surfaces of sample, and there is no
propagation in crack front zone. When load reaches its maximum, brittle fracture
will happen suddenly, and the maximum load is Fcr. For sample with smaller
thickness, the second type of curve can be obtained. There is an obvious “burst”
platform on the curve during loaded, the reason is that the central layer of sample is
in plane strain state, where crack propagates first, while there is no crack propa-
gation on surfaces of sample that are in plane stress state, so the propagation of
crack in central layer will be dragged by surfaces. During test, when the “burst”
platform appears, clear sound of crack can be heard, so the corresponding load is
Fcr.
In the two cases above, there is no need to record the F-d curve during test, but
to obtain the maximum load or load at the “burst” platform, while sample with large
dimensions is required, and correspondingly test machine with wide range.
To overcome the difficulties above, sample with minimal thickness is always
adopted, and the first type of curve can be obtained. In this case, the maximum load
cannot be used to calculate fracture toughness, because crack propagates slowly
before load reaches its maximum value, which is imperceptible. Thus, particular
engineering assumption is applied to determine Fcr, as seen in Fig. 3.9, and it is that
the load at the intersection point of the F-d curve and a secant line with a slope 5%
less than that of F-d curve is the Fcr.
To verify that the fracture toughness along grain is a basic attribute of wood and
has nothing to do with shape and dimensions of sample, experiments are performed
on Pinus sylvestris (Pinus sylvestris var. mongolica Litv.) with TL crack. CT sample
and its improved sample (WOL sample) are adopted considering wood diameter,
the manufacture of sample and load way, and the influence of sample size on KIC is
44 3 Fracture of Wood Along Grain

also studied. Then KIC values of CT samples with TL crack and GIC values of DCB
samples of four kinds of woods (China fir, spruce, poplar, and Castanopsis hystrix)
are tested, and the tested KIC values are compared with the critical stress intensity
factor calculated from GIC.

3.5.2 The K TL
IC of CT Samples with Different Thickness

The CT sample of wood is made according to national standard “Metallic materials


—Determination of plane-strain fracture toughness” (GB4161-84) [11]. The size of
sample is W = 50 mm, e = 12.5 mm, a = 25 mm, H = 60 mm, as seen in
Fig. 3.10.
To explore whether the thickness of sample has influence on fracture toughness
of wood, Barrent [13] measured the fracture toughness of Douglas-fir by four-point
bending sample, and concluded that for sample with thickness greater than 5 mm,
the KIC was a constant. And Boatright and Garrentt [14] believed that when the
thickness of sample was greater than 10 mm, it had no obvious influence on
fracture toughness of wood. To confirm the situation above, two groups of CT
sample are made considering annual ring width of P. sylvestris. There are 12
samples with a thickness (B1) of 17 mm in group CT1, and 12 samples with a
thickness (B2) of 30 mm in group CT2. To make a pre-crack, a straight slot is
sawed by a band saw, then the slot is cut forward 1–2 mm by a sharp blade, at last a
sharp crack is obtained with a length of a, as seen in Fig. 3.10.
The CT specimen is connected with a steel U hook by a steel pin loaded by the
computer-controlled testing machine with a constant speed (For wood, the cross-
head speed is 2 mm/min). A computer recorded the load–displacement (F-d) curve
automatically, where d is the displacement between the two load points.
Due to GB4161-84 [11], the fracture toughness of CT sample can be calculated
by the following equation:

Fig. 3.10 Sample of CT F


specimen with TL crack

F B
a
e W
3.5 The Stress Intensity Factor … 45

Fcr a
KIC ¼ pffiffiffiffiffi  f ð3:40Þ
B W W

where Fcr is the critical load, a/W is the dimensionless crack length, and f(a/W) is
the function of sample geometry given by the formula below:

h a

2
3
4 i
a 2þ a
W 0:886 þ 4:64 W  13:32 Wa þ 14:72 Wa 5:6 Wa
f ¼
3=2 :
W 1  Wa

It can be seen that the F-d curve of air-dried CT sample is kept straight basically
from the initiation of crack to the failure of sample under tensile force. And all the
F-d curves are the third type according to GB4161-84 [11], so Fmax was the critical
load Fcr. Substituting sample dimensions and Fmax into Eq. 3.40, the fracture
toughness of sample can be obtained.
The sizes and KIC values of CT samples are shown in Table 3.1, and statistical
description results and variance analysis results are shown in Tables 3.2 and 3.3. It

Table 3.1 Size of CT sample and test value of Pinus sylvestris


No. W (mm) a (mm) H (mm) B (mm) Fmax (N) f(a/W) KIC (Nmm−3/2)
CT1-1 50 25 60 17 122.5 9.659 9.84
CT1-2 50 25 60 17 117.8 9.659 9.46
CT1-3 50 25 60 17 123.6 9.659 9.93
CT1-4 50 25 60 17 134.3 9.659 10.79
CT1-5 50 25 60 17 128.6 9.659 10.33
CT1-6 50 25 60 17 123.6 9.659 9.93
CT1-7 50 25 60 17 122.4 9.659 9.84
CT1-8 50 25 60 17 118.5 9.659 9.52
CT1-9 50 25 60 17 139.1 9.659 11.8
CT1-10 50 25 60 17 105.8 9.659 8.51
CT2-1 50 25 60 30 236.3 9.659 10.76
CT2-2 50 25 60 30 219.5 9.659 10.00
CT2-3 50 25 60 30 203.0 9.659 9.25
CT2-4 50 25 60 30 213.7 9.659 9.73
CT2-5 50 25 60 30 203.0 9.659 9.25
CT2-6 50 25 60 30 203.4 9.659 9.26
CT2-7 50 25 60 30 211.7 9.659 9.64
CT2-8 50 25 60 30 196.4 9.659 8.94
CT2-9 50 25 60 30 201.6 9.659 9.18
CT2-10 50 25 60 30 220.4 9.659 10.04
CT2-11 50 25 60 30 238.4 9.659 10.85
CT2-12 50 25 60 30 236.9 9.659 10.79
46 3 Fracture of Wood Along Grain

Table 3.2 Statistical analysis of KIC


TL
of CT samples of Pinus sylvestris
Group Number Sum Mean value SD CV Accuracy
(Nmm−3/2) (Nmm−3/ (%) index (%)
2
)
CT1 12 119.03 9.92 0.73 7.39 4.27
(B = 17 mm)
CT2 12 117.68 9.81 0.68 6.99 4.03
(B = 30 mm)
Total 24 236.71 9.86 0.70 7.06 2.88

Table 3.3 Variance analysis of KIC


TL
for different thicknesses of CT samples of Pinus sylvestris
Source of variance SS df MS F F crit Sig.
Between groups 0.0768 1 0.076792 0.152639 4.300949 No
Within groups 11.068 22 0.503095
Total 11.145 23

can be seen that the coefficients of variation (CV) of the KIC values are less than
20%, and the accuracy indexes are all less than 5%, which indicates that the average
KIC value of CT samples is effective, and it is 9.86 N mm−3/2. The results
demonstrate that the thickness of sample nearly has no influence on the fracture
toughness of wood, which is in accordance with the conclusions of Barrent [13] and
Boatright and Garrentt [14].

3.5.3 The K TL
IC of WOL Samples with Different Crack
Length

WOL sample is an improved one of CT sample, which is similar with CT sample in


shape but longer, so the fracture toughness corresponding to different crack length
can be measured. The sizes of WOL sample of P. sylvestris are B = 20 mm,
W = 2.55 B, H = 2.48 B, and e = 0.13 mm. To explore the influence of crack
length on wood fracture toughness, a/W is 0.3, 0.4, 0.5, 0.6, 0.7, 0.8, respectively,
and there are two samples for each crack length, 12 samples in total. The prepa-
rations of sample and test process are the same as that of CT sample.
Fracture toughness can be calculated as follows:

Fcr a
KIC ¼ pffiffiffiffiffi  f ð3:41Þ
B W W
a
a
1=2 a
3=2
5=2
7=2
9=2
where f W ¼ 30:96 W 195:8 W þ 730:6 Wa 1186:3 Wa þ 754:6 Wa .
3.5 The Stress Intensity Factor … 47

The sizes and KIC values of WOL samples are shown in Table 3.4, and the
average value of fracture toughness is 10.09 Nmm−3/2 that is close to the average
fracture toughness of CT samples. The statistical description results and variance
analysis results are shown in Tables 3.5 and 3.6. It can be seen that the CV of the
results of WOL sample is larger than that of CT sample, while the variance analysis
results in Table 3.6 show that the sizes and crack length (0.3 < a/W < 0.8) of CT
sample have no obvious influence on fracture toughness.

Table 3.4 Sizes of WOL samples and test value of Pinus sylvestris
No. W (mm) a (mm) a/W B (mm) H (mm) Fmax (N) f(a/W) KIC (Nmm−3/2)
W1-1 56 16.8 0.3 21.93 54.56 291.44 6.603 11.73
W1-2 56 16.8 0.3 21.96 54.56 290.48 6.603 11.67
W2-1 56 22.4 0.4 21.93 54.56 251.28 8.178 12.52
W2-2 56 22.4 0.4 21.94 54.56 266.36 8.178 13.27
W3-1 56 28 0.5 21.91 54.56 158.80 10.313 9.99
W3-2 56 28 0.5 22.01 54.56 174.88 10.313 10.95
W4-1 56 33.6 0.6 21.91 54.56 122.60 13.983 10.46
W4-2 56 33.6 0.6 21.94 54.56 97.48 13.983 8.30
W5-1 56 39.2 0.7 21.96 54.56 68.36 21.898 9.11
W5-2 56 39.2 0.7 21.91 54.56 48.24 21.898 6.44
W6-1 56 44.8 0.8 21.98 54.56 40.20 38.999 9.53
W6-2 56 44.8 0.8 21.93 54.56 30.16 38.999 7.17

Table 3.5 Statistical analysis of KIC


TL
of CT samples and WOL samples of Pinus sylvestris
Group Number Sum Mean value (N/mm3/2) SD (N/mm3/2) CV (%)
CT1(B = 17 mm) 12 119.03 9.92 0.73 7.39
CT2(B = 30 mm) 12 117.68 9.81 0.68 6.99
WOL 12 121.14 10.09 2.10 20.78
Total 36 357.85 9.94 1.17 13.17

Table 3.6 Variance analysis of KIC


TL
for different thicknesses of CT samples and WOL samples of
Pinus sylvestris
Source of variance SS df MS F F crit Sig.
Between groups 0.505842 2 0.252921 0.140364 3.284918 No
Within groups 59.46253 33 1.801895
Total 59.96837 35
48 3 Fracture of Wood Along Grain

3.6 The Fracture Toughness GTL


IC Along Grain of Wood
by Energy Method

DCB sample is always used to test pure Mode I critical strain energy release rate
GIC of wood. GIC can be calculated by the compliance of crack measured by the
symmetrical bending test method. Triboulot et al. [15] measured the fracture
toughness of wood with TL crack by DCB method, and the results are compared
with the solution of finite element analysis (FEA). The results tally well with each
other.

3.6.1 Materials and Samples

DCB samples of China fir, spruce, poplar, and C. hystrix are prepared, and a is the
length of initial crack, as seen in Fig. 3.11. There are 6–9 samples for each wood.
China fir samples are divided into two groups: one with TL crack and the other with
RL crack, and only TL crack is prepared on the samples of spruce, poplar, and C.
hystrix. The sizes of sample and crack are shown in Table 3.7. As thickness of

Fig. 3.11 Schematic diagram


F
of DCB specimen
B

e a W
F

Table 3.7 Sizes of DCB sample and test value of Cunninghamia lanceolata
No. W (mm) H (mm) B (mm) a (mm) a/W Fmax (N) C (mm/N) GIC (J/m2)
1 260 84 21.84 49.5 0.190 366.284 1.692 87.37
2 260 84 21.84 59.5 0.229 341.268 2.053 89.27
3 260 84 21.84 71 0.273 313.612 2.598 90.93
4 260 84 21.84 82.5 0.317 302.066 3.014 101.75
5 260 84 21.84 95.5 0.367 273.310 3.641 102.96
6 260 84 21.84 107.5 0.413 247.798 4.516 102.93
7 260 84 21.84 120 0.462 221.682 5.611 100.99
8 260 84 21.84 137 0.527 188.144 7.225 95.97
9 260 84 21.84 155 0.596 156.750 9.727 89.33
3.6 The Fracture Toughness … 49

samples is larger than 20 mm, the fracture can be treated as plane strain problem.
The moisture content (MC) of specimens is about 12%, the temperature is about
20 °C and relative humidity is about 60% in laboratory.

3.6.2 Test and Results

Tests were performed on the computer-controlled testing machine with a crosshead


speed between 1 and 5 mm/min. At the beginning of the test, a low crosshead speed
was used because the cantilever beams were short. Then the crosshead speed was
increased when the cantilever beams were relatively long [16]. Computer recorded
the curve of the applied load versus opening displacement (F-d) automatically, as
seen in Fig. 3.12. The fracture of air-dried wood along the grain presented brit-
tleness approximately. Except the curve in the initial loading stage caused by the
space between the specimen and U-shaped hook, almost all the F-d curves
remained straight. Once crack along grain initiated, the crack propagation parallel to
grain was unstable. The bearing capacity of the specimen decreased sharply, so the
top point of F-d curve represented the critical point of rapid cracking. After the load
went down, the test machine was stopped immediately and the recorded data was
stored. The crack tips were marked by means of optical microscope. Then the
specimen was unloaded and reloaded. The same procedure was repeated until the
specimen was fractured completely. Took the sample off test machine, measured
the crack length after each increment of crack propagation. The F-d curves belong
to the third type due to GB4161-84, so Fcr = Fmax. With the increase of crack
length, the slope of the linear part of F-d curves decreased. The reciprocal of the
slope is the corresponding compliance Ci of the F-d curve of DCB specimen with
the certain crack length ai or (a/W)i, thus C = f(a) or C = f(a/W).
The sizes and calculated energy release rate GIC are shown in Table 3.7. The F-d
curves corresponding to different crack length of one DCB specimen and the

Fig. 3.12 Typical F-d curve a1


0.4 a2
of DCB specimen of a3
0.35
Cunninghamia lanceolata a4
a5
0.3 a6
a7
F(kN)

0.25
a8
0.2 a9
0.15
0.1
0.05
0
0 0.5 1 1.5 2
(mm)
50 3 Fracture of Wood Along Grain

relationship between the corresponding compliance and a/W are shown in


Figs. 3.12 and 3.13.
Exponential curve is chosen to characterize the relationship between compliance
and crack length, and it is given by

C ¼ qemða=WÞ ð3:42Þ

The fracture toughness can be calculated by the following equation:

Fmax
2
@C
GIC ¼  ð3:43Þ
2BW @ða=WÞ

where q and m are the fitting coefficients of the compliance curve. And the coef-
ficients of determination (R2) of compliance and crack length are all above 0.98.
Figure 3.13 Relationship between corresponding compliance and a/W of
Cunninghamia lanceolata DCB sample
Mode I fracture toughness GIC can be calculated by Eq. (3.43). The arithmetic
mean values of toughness of one specimen x and one group of specimens X are
calculated, respectively, as

1X k
x¼ ðGIC Þi ð3:44Þ
k i¼1

1X n
X¼ xj ð3:45Þ
n j¼1

where k is the amount of measured points of one specimen and n is the amount of
the specimens in one group.
Figure 3.14 shows the distribution relationship between the fracture toughness
and crack length of DCB samples of China fir. The average energy release rate of

Fig. 3.13 Relationship


between corresponding
compliance and a/W of
Compliance (mm/N)

Cunninghamia lanceolata
DCB sample

a/W
3.6 The Fracture Toughness … 51

Fig. 3.14 Relationship


TL crack
between GIC and a of
Cunninghamia lanceolata

GIC (J/m 2)
104 J/m 2
DCB sampler

TL
a ( mm)

RL crack
GIC (J/m2)
101 J/m 2
RL

a ( mm )

China fir samples with TL crack and RL crack are 104.17 J/m2 (SD = 13.35 J/m2)
and 101.49 J/m2 (SD = 22.07 J/m2) respectively, and there is no significant
difference.
The fracture toughness of China fir samples with TL crack and RL crack are
about the same because of the fine wood ray of China fir. However, for many
woods, there may be difference between the results of samples with TL crack and
RL crack caused by the size and content of wood ray and the construction of wood.

3.6.3 The Relationship Between Stress Intensity Factor


and Energy Release Rate

For the opening crack of homogeneous linear anisotropic material, the relationship
expression between strain energy release rate and fracture toughness was given by
Sih, Paris, and Irwin [17], called S.P.I. relationship, as follows:

 12 " 12 #12


S11 S12 S22 2S12 þ S66
GI ¼ KI2  S ¼ KI2  þ ð3:46Þ
2 S11 2S11

Substituting Sij by engineering elastic parameters, and neglecting the small value
term with Poisson’s ratio, the equation above can be written as
52 3 Fracture of Wood Along Grain

 1=2 " 1=2   #1=2


 1 EL EL 1=2
S ¼ p  þ ð3:47Þ
2EL ET ET 2GLT

Equation (3.47) is similar to the fracture mechanics formula of isotropic mate-


rial, as follows:

KI2
GI ¼ ðIsotropic materialÞ ð3:48Þ
E

KI2
GI ¼ ðAnisotropic materialÞ ð3:49Þ
E

where E* is equivalent modulus, and

1 ð2EL ET Þ1=2
E ¼ ¼  1=2 1=2 1=2 ð3:50Þ
S EL EL
ET þ 2GLT

Wood is anisotropic porous biomaterial, but when cutting a small rectangle piece
at a certain distance from the pith and making one of its symmetry plane perpen-
dicular to the growth rings, then it can be seen as an orthotropic body in macro-
scopic scale [18]. Triboulot et al. [15] had converted the measured energy release
rate to stress intensity factor according to Eq. (3.46) or Eq. (3.49) to demonstrate
that it was feasible to treat wood as orthotropic and elastic body, and fracture
mechanics was applicable to wood. However, although the converted values of
stress intensity factor of two type samples were in accordance with each other, no
experiment was performed to verify the S.P.I. relationship, so it is not stated that
whether the KIC values of one wood obtained by different test methods are in
agreement.
Ashby et al. [19] considered that when crack peeled or layered in opening mode
along grain, the fracture process was the same as the peeling of bonded points. As
the composition and structure of cell wall had almost no difference in different kind
of woods, the energy absorbed per unit area was approximately a constant when
peeled (under given moisture content) for all woods. In fracture process, the fracture
energy was provided by elastic energy released by the surrounding tissue and the
work done by the applied load. Therefore, by the principle of energy balance, the
energy release rate of Mode I delamination could be sufficiently approximated as
follows (Called Ashby relationship):

KIC
2
GIC ¼ ð3:51Þ
ER
3.6 The Fracture Toughness … 53

where ER was Young’s modulus in radial direction, and KIC was the stress intensity
factor of delaminating fracture.
In the same experimental environment, there are two main factors that affect the
equivalent relation between KIC and GIC, and they are the construction of wood and
test method, which both have relation with the occurrence of bridge. KIC is a
physical quantity in instantaneous state when wood fractures along grain unstably,
and the measurement is based on pre-crack, so there is nearly no fiber bridge.
However, GIC is an average physical quantity during the multiple continuous
propagation process of crack, thus fiber bridge in different levels would occur in
different woods. Generally, for softwood with straight grain, there is nearly no fiber
bridge during the propagation process of crack in DCB sample, while for hardwood
with interlocked grain, lots of fiber bridges will occur during crack propagation
process in DCB sample, which would increase the energy consumption, and con-
sequently, the equivalent relation between KIC and GIC is affected.
To illustrate this point, the KIC of the CT samples of China fir, spruce, poplar,
and C. hystrix are measured according to GB4161-84 [11]. And the GIC of DCB
samples of the four kinds of woods are measured by compliance method in two
ways, namely multi-sample with single-point method and single sample with
multipoint method. For multi-sample with single-point method, the pre-crack of
DCB sample satisfies a/W = 0.2–0.8, 6–7 samples in each group, and only one
point will be tested; for single sample with multipoint method, that is, for one single
DCB sample, multiple points will be tested during the multiple continuous prop-
agation process of crack. The process of multi-sample with single-point method is
the same as that of CT sample, during which there is nearly no fiber bridge, while,
fiber bridges may occur during the multiple continuous propagation process of
crack by single sample with multipoint method.
China fir possesses the characters of straight grain, uniform structure, little and
very thin wood ray, so no bridge occurred in DCB sample during the continuing
propagation of crack. Thus, there is no significant difference between the results
obtained by multi-sample with single-point method and single sample with multi-
point method. The relationship between the compliance of F-d curves obtained by
multi-sample with single-point method and the corresponding a/W is shown in
Fig. 3.15. Then, the energy release rate GIC can be calculated according to
Eq. (3.43), and the results are shown in Table 3.8. The average energy release rate
of China fir DCB samples is 99.03 J/m2 obtained by multi-sample with single-point
method, which is very close to the average value 101.49 J/m2 obtained by single
sample with multipoint method.
The statistical description results of energy release rate of the four kinds of
woods by multi-sample with single-point method and single sample with multipoint
method are shown in Table 3.9. It can be seen that there are significant differences
between the results obtained by the two methods for spruce, poplar, and C. hystrix.
As seen in Fig. 3.16, for poplar, fiber bridges occurred during the continuing
propagation of crack leading to the high fracture toughness.
The elastic coefficients of the four kinds of woods measured by electrometric
method are shown in Table 3.10, then the equivalent modulus E* can be calculated.
54 3 Fracture of Wood Along Grain

Compliance (mm/N)

a/W

Fig. 3.15 Relationship between corresponding compliance and a/W of Cunninghamia lanceolata
DCB sample (single sample with multipoint method)

Table 3.8 Sizes of DCB samples and test value of Cunninghamia lanceolata (single sample with
multipoint method)
No. W (mm) H (mm) B (mm) a (mm) a/W Fmax (N) C (mm/N) GIC (J/m2)
1 200 80 20.01 40 0.2 329.76 2.352 114.73
2 200 80 20.12 60 0.3 270.30 3.216 113.39
3 200 80 19.98 80 0.4 213.00 4.513 103.57
4 200 80 20.07 100 0.5 179.60 6.359 108.32
5 200 80 20.11 120 0.6 140.54 10.225 97.56
6 200 80 20.09 140 0.7 110.40 15.544 88.56
7 200 80 20.15 160 0.8 90.30 22.967 87.15

Table 3.9 Statistical analysis of GIC between many measure and single measure for DCB samples
Species Single sample with multipoint method Multi-sample with single-point
method
Crack Sample Test GIC SD Crack Sample GIC SD
point (N/m or (N/m) (N/m or (N/m)
J/m2) J/m2)
Cunninghamia TL 7 50 104.17 13.35 TL
lanceolata
Cunninghamia RL 6 48 101.49 22.7 RL 6 99.03 26.7
lanceolata
Picea asperata TL 9 69 306.07 63.33 TL 6 163.38 57.9
Populus sp TL 7 42 582.92 151.03 TL 6 300.3 51.13
Castanopsis TL 6 51 1222.09 333.39 TL 6 392.7 134.8
hystrix
3.6 The Fracture Toughness … 55

Populus sp.

Fig. 3.16 Bridge phenomena of wood DCB sample

The measured stress intensity factor by CT samples, energy release rate by DCB
samples, and the corresponding stress intensity factor converted by S.P.I. rela-
tionship or Ashby relationship are shown in Table 3.11. Conclusions can be drawn
that:
(1) The stress intensity factor of China fir CT samples is 7.98 Nmm−3/2. The stress
intensity factor converted by S.P.I. relationship from the energy release rate of
China fir DCB samples obtained by two methods are 8.09 and 8.19 Nmm−3/2,
respectively. The stress intensity factors converted by Ashby relationship from
the energy release rate of China fir DCB samples obtained by the two methods
are 7.18 and 7.26 Nmm−3/2, respectively. It can be seen that as there is no fiber
bridge during the propagation process of crack in China fir, there is fine
equivalent relation between stress intensity factor converted by S.P.I. rela-
tionship or Ashby relationship and energy release rate of China fir.
(2) The stress intensity factors of CT samples of spruce and poplar have good
equivalent relation with the energy release rate of DCB samples obtained by
multi-sample with single-point method, but no equivalent relation with the
energy release rate is obtained by single sample with multipoint method. And
the stress intensity factor converted from the energy release rate obtained by
single sample with multipoint method is larger than the measured value and the
value converted from the energy release rate is obtained by multi-sample with
single-point method, because fiber bridges occurred during the continuing
propagation of crack leading to the high fracture toughness.
(3) For C. hystrix, there is no equivalent relation between stress intensity factor and
energy release rate obtained by the two methods, because the orthotropic
property is affected by the interlocked grain and wide ray of C. hystrix. Fiber
bridges occurred, and the fracture surface is rough and uneven that much
rougher than that of other woods. And the relationship between the feature of
fracture surface and fracture toughness along grain will be discussed in Chap. 6.

In addition, it must be pointed out that the elastic coefficients measured at a


certain point by electrometric method some limitation because of the heterogeneity
and variability of wood, and there is no quite effective method to measure the 12
56

Table 3.10 Elastic coefficients of woods


Species EL (N/ lLR lLT ER (N/ lRT lRL ET (N/ lTR lTL GRL (N/ GTL (N/ GRT (N/ S* (mm2/N) E* (mm2/
mm2) mm2) mm2) mm2) mm2) mm2) N)
Cunninghamia 8500 0.49 0.6 520 300 0.419 0.035 106 0.00151171 661.50103
lanceolata
Picea asperata 11,733 0.34 0.442 1204 0.455 0.037 620.7 0.407 0.07 495 209 84.7 0.00081378 1228.8288
Populus sp 5352 0.64 0.66 1204.7 0.452 0.037 1039 0.399 0.033 621 466.3 94.3 0.00064767 1543.9971
Castanopsis 19,637 0.41 0.51 3837 0.713 0.079 1514 0.329 0.049 1949 645.3 161 0.0003552 2815.2816
hystrix
3 Fracture of Wood Along Grain
Table 3.11 The relation between KIC and GIC of woods
3.6 The Fracture Toughness …

Species CT GIC (N/m or J/m2) S.P.I. KIC = (GIC  E*)1/2 (N/mm3/2) Ashby KIC = (GIC  ER)1/2 (N/mm3/2)
sample relation relation
Crack KIC (N/ DCB DCB E* = 1/ DCB DCB E* = ER DCB DCB
mm3/2) (Single-point (Multipoint S*(N/ (Single-point (Multipoint (N/mm2) (Single-point (Multipoint
method) method) mm2) method) method) method) method)
Cunninghamia RL 7.98 99.03 101.49 662 8.09 8.19 520 7.18 7.26
lanceolata
Picea asperata TL 13.83 163.38 306.07 1229 14.17 19.39 1204 14.03 19.20
Populus sp TL 20.59 300.30 582.92 1544 21.53 30.00 1205 19.02 26.50
Castanopsis TL 21.64 392.70 1222.09 2815 33.25 59.66 3873 39.00 68.80
hystrix
For Cunninghamia lanceolata, GIC(TL)  GIC(RL) of DCB sample by multipoint method
57
58 3 Fracture of Wood Along Grain

elastic coefficients, which will bring errors to the description and analysis on wood
mechanical properties.

3.7 Mode III Fracture Property of Wood Along Grain

3.7.1 Introduction

Mode I fracture along grain is the most common and dangerous failure way of
wood, so it is the focus of wood fracture mechanics. And there are various samples,
such as compact tension (CT) sample, single edge notched bending (SEB) sample,
tensile (TS) sample and double cantilever beam (DCB) sample, and the improved
forms of above samples. There are also many researches on the Mode II fracture
toughness of wood based on end notched flexure (ENF) sample. However, the
research on the Mode III fracture of wood is rarely reported. Ehart et al. [20]
performed the Mode III fracture test of wood sample by a twist machine, but the
complex experimental facility hindered the spread of the method.
Donaldson [21, 22] characterized the Mode III interlaminar fracture of com-
posites successively by using split cantilever beam (SCB) specimens and proposed
that torsion issue could be dealt well by applying the split cantilever beam with
initial crack bonded between two parallel aluminum bars. Gary Becht [23] tested
the Mode III interlaminar fracture toughness of continuous-fiber-reinforced com-
posite materials by using crack rail shear (CRS) specimens. Shaw Ming Lee [24]
measured the Mode III interlaminar fracture toughness of several carbon fiber/
epoxy composites by edge crack torsion (ECT) method. Xiao Jun [25] modified the
Donaldson SCB fixture device by using guide limited post to simplify the Mode III
interlaminar fracture toughness test method of laminates. Mohammad Reza
Khoshravan and Mohsen Moslemi [26] calculated and analyzed the Mode III
interlaminar fracture behavior of woven fabric-reinforced glass/epoxy composite
laminates by compliance method and virtual crack closure technique. Here, the
Mode III interlaminar fracture toughness of spruce (Picea jezoensis) is tested by
improved designed SCB fixture device with the compliance method.

3.7.2 Material and Method

Air-dried spruce (P. jezoensis) is selected to make sample with straight grain, and
the size of sample is 240 (L) 20 (T) 20 (R) (mm). To simulate a naturally
sharp crack, the pre-crack is cleaved along middle level parallel to grain by knife,
and the length of initial crack is controlled about 40 mm from the center of loading
points.
3.7 Mode III Fracture Property of Wood Along Grain 59

3.7.3 Test and Results

Double cantilever inversion symmetry bending load method was used to test
Mode III interlaminar fracture (Fig. 3.17). Sample was placed in the fixture device,
then loaded by a computer-controlled testing machine, and the load–load point
deflection (F-d) curve was recorded by computer automatically during the test
(Fig. 3.18). Single specimen with multipoint method was applied, that is, on one
specimen, repetitive loading-unloading-loading was performed and the loading
speed was 1 mm/min. Temperature was 25 °C and the humidity was 65–70% in
laboratory.
At the initial of loading, the F-d curve kept straight until the crack began, then
the slope of F-d curve increased suddenly, meanwhile, the crack propagated slowly
along the grain. After the F-d curve rose nonlinearly to the maximum load, it began
to decline. At this moment, stopped loading to save the F-d curve and mark the
crack tip. The critical load Fcr was the load when the crack began, namely the point

Fig. 3.17 The test device of


improved split cantilever
Sample
beam method F
Crack

Rollers
Crack

Fig. 3.18 Typical F-d curves a1 a2


of Mode III specimen. a3 a4
a5 a6 a7
a1 = 49 mm, a2 = 60 mm, a8 a9
a10
a3 = 69 mm, a4 = 77 mm,
Force (kN)

a5 = 87 mm, a6 = 99 mm,
a7 = 105 mm, a8 = 117 mm,
a9 = 123 mm,
a10 = 129 mm

(mm)
60 3 Fracture of Wood Along Grain

the slope of the F-d curve increased. After the test is over, length between the crack
tip marked each time and the center of the load point was measured. The average
value of the crack length on to sides of sample was taken as the final crack length.
As seen in Fig. 3.18, it was the test curve of one spruce sample. It can be seen
that the slope of the F-d curve straight portion is inverse to crack length and the
reciprocal of the slope is the compliance Ci corresponding to different crack lengths
ai. And by curve fitting, the exponential relationship between the compliance and
crack length is obtained

C ¼ qema ð3:52Þ

where q and m are the fitting coefficients of the compliance curves and the deter-
mination coefficients (R2) were all above 0.95. As shown in Fig. 3.19, the fitting
curve of compliance C and crack length a of one spruce sample was shown.
So the propagation resistance of Mode III interlaminar crack along grain, namely
the interlaminar fracture toughness can be calculated by the following formula:

@U @ðFcr
2
C=2Þ Fcr
2
@C
GTL
IIIC ¼ ¼ ¼  ð3:53Þ
@A @ðBaÞ 2B @a

The distribution relationships of crack length a and GTL


IIIC are shown in Fig. 3.20,
and the statistical description of results is shown in. Although GTL IIIC increases
slightly with the increase of crack length, by statistical description, there are 6
samples, 57 test points, and the average value is GIIIC-average = 1051.42 Jm−2
(SD = 207.02 Jm−2) with a CV of 19.69 and accuracy index of 5.21, which indi-
cates that the difference is not significant, so the average Mode III interlaminar
fracture toughness 1.05 kJ/m2 can be seen as a basic characteristic of spruce that
represents the capacity of spruce to resist the propagation of Mode III crack. As
seen in Fig. 3.21, it is the failed spruce sample.

Fig. 3.19 Relationship 0.0164 a


C = 0.003e
between compliance (C) and
crack length (a) R = 0.9975
2
C (mm/N)

a (mm)
3.8 Conclusions 61

Fig. 3.20 Relationship


between GIIIC and a

GIIIC (kJ/m2)
1.05

a (mm)

20mm

Fig. 3.21 Mode III fracture specimen of Picea jezoensis var.

3.8 Conclusions

In this chapter, the theory of LEFM and the special application of LEFM on wood
are introduced. Then different samples and methods are applied to measure the
Mode I fracture toughness of wood along grain, and the results show that the
fracture toughness of wood along grain is the basic attribute of wood and has
nothing to do with test method and the geometrical shape and size of crack. The
research indicates that LEFM based on isotropic body is applicable to the crack
propagation of wood along grain.
Fractal theory is applied to study the fractal feature of the fracture surfaces of five
woods along grain. The relationship between fractal dimension of fracture surface
and fracture toughness of wood is established. The results show that fracture
toughness parallel to the grains of various woods is different because of their
textural diversity and such differences are also shown on the morphology of fracture
surfaces. Furthermore, there is evident and direct proportional relation between the
fractal dimension and fracture toughness along grains.
Double cantilever inversion symmetry bending load method is applied to mea-
sure the Mode III fracture toughness of spruce based on energy theory. The results
have shown that the average Mode III interlaminar fracture toughness 1.05 kJ/m2
can be seen as a basic characteristic of spruce that represents the capacity of spruce
to resist the propagation of Mode III crack.
62 3 Fracture of Wood Along Grain

References

1. Snedden IN (1946) The distribution of stress in the neighborhood of a crack in an elastic solid.
Proc R Soc Lond A 187:229–260
2. Irwin GR (1957) Analysis of stresses and strain near the end of a crack traversing a plate.
J Appl Mech 24:361–364
3. Westergaard HM (1939) Bearing pressures and cracks. J Appl Mech 6:49–53
4. Sih GC, Prais PC, Irwin GR (1965) On cracks in rectilinearly anisotropic bodies. J Fract Mech
1(3):189–203
5. Wu EM (1967) Application of fracture mechanics to anisotropic plates. J Appl Mech 34:967–
974
6. Wu EM (1968) Fracture Mechanics of anisotropic plate. In: Tsai SW (ed) Composite material
workshop. Technomic Publishing Company, p 23
7. Kanninen MF, Popelar CH (1985) Advanced fracture mechanics. Oxford University
8. Parhizgar Shahrokh, Zachary Loren W, Sun CT (1982) Application of the principles of linear
fracture mechanics to the composite materials. Int J Fract 20:3–15
9. American Society of Testing Materials (1981) Standard methods for plain-strain fracture
toughness of metallic materials. ASTM E399-74
10. British Standards Institution (1997) Methods of test for plane strain fracture toughness.
BS5447
11. National Standard (1984) Metallic materials-determination of plane-strain fracture toughness.
GB4161-84
12. Cui Z (1981) Theory and test method of fracture toughness. Shanghai Scientific and Technical
Publishers, Shanghai, pp 15–34
13. Barrent JD (1976) Effect of crack-front width on fracture toughness of Doouglas-fir. Eng Fract
Mech 8(4):711–717
14. Boatright SWJ, Garrentt GG (1983) The effect of microstructure and stress state on the
fracture behaviour of wood. J Mater Sci 18:2181–2199
15. Triboulot P, Jodin P, Pluvinage G (1984) Validity of fracture mechanics concept applied to
wood by finit element calculation. Wood Sci Technol 18(6):448–459
16. Hodgkinson JM (2000) Mechanical testing of advanced fiber composites. Woodhead
Publishing Limited, Niklewicz
17. Sih GC, Prais PC, Irwin GR (1965) On cracks in rectilinearly anisotropic bodies. J Fract Mech
1(3):189–203
18. Coleman (1991) Wood and wood process principle. China Forestry Publishing House,
Beijing, pp 227–235
19. Ashby MF, Easterling KE, Harrysson R et al (1985) The fracture and toughness of woods.
Proc R Soc Lond A 398:261–280
20. Ehart RJA, Stanzl-Tschegg SE, Tschegg EK (1999) Mode III fracture energy of wood
composites in comparison to solid wood. Wood Sci Technol 33:391–405
21. Donaldson SL (1988) Mode III interlaminar fracture characterization of composite material.
Compos Sci Technol 32:225–249
22. Donaldson SL, Mall Shankar, Lingg Cynthia (1991) The split cantilever beam test for
characterizing Mode III interlaminar fracture toughness. J Compos Tech Res 13(1):41–47
23. Becht Gary, Gillespie John W (1988) Design and analysis of the crack rail shear specimen for
Mode III interlaminar fracture. Compos Sci Technol 31:143–157
24. Lee SM (1993) An edge crack torsion method for Mode III delamination fracture testing.
J Compos Tech Res 15(3):193–201
25. Jun X (1996) Experimental study on Mode III fracture toughness of multi-directional interface
of composite materials. J Nanjing Univ Aeronaut Astronaut 28(2):267–270
26. Khoshravan MR, Moslemi M (2014) Investigation on Mode III interlaminar fracture of glass/
epoxy laminates using a modified split cantilever beam test. Eng Fract Mech 127:267-279
Chapter 4
Transverse Fracture of Wood

Abstract In this chapter, first, the stress field at crack tip of wood component with
crack perpendicular to grain is analyzed by finite element method to explore the
reason why the crack always initiates along grain. The critical stress intensity factor
of four kinds of woods is measured with acoustic emission (AE) technique, and the
results show that Mode I crack in wood component initiates along fiber, and then
ductile fracture happens along the cross section of the component. The whole
propagation process of crack involves three stages: linear stage, stable stage, and
unstable stage. Then the influence of crack perpendicular to grain on the strength of
wood component is studied. Bending test, impact test, and tensile test as experi-
mental verification are performed on standard samples and samples with crack
perpendicular to grain of two softwoods and two hardwoods. The results show that
wood possessed fine capacity to resist transverse fracture because of its specific
organization structure, so it would not fail under low work stress in spite of the
stress singularity at crack tip. So for the safety assessment and strength design of
wood component with crack perpendicular to grain, net stress criterion should be
adopted, namely the strength of the clear part of component apart from the size of
crack. At last, the analytical formula of energy release rate of the Mode I inter-
laminar fracture of wood and the bending delamination damage of plywood are
derived.

4.1 Introduction

Fracture Mechanics is extensively used in Material Science, as traditional view-


point of safety could not explain the phenomenon that material with crack fractures
under the stress much lower than the ultimate strength of material. The fracture of
material is not only determined by the length of crack and external load but also the
sensitivity of material to crack, namely fracture toughness. Mode I fracture along
grain of wood is brittle fracture and once crack initiates, material would fracture
because of the self-similar unstable propagation of crack. The application of LEFM
to the fracture along grain of wood is successful and there is a good deal of research

© Springer Nature Singapore Pte Ltd. 2018 63


Z. Shao and F. Wang, The Fracture Mechanics of Plant Materials,
https://doi.org/10.1007/978-981-10-9017-2_4
64 4 Transverse Fracture of Wood

on the crack body [1–4]. However, for the transverse fracture of wood, LEFM is not
applicable anymore because the crack does not propagate along its original direc-
tion. Jeronimidis [5] proposed that fracture energy or fracture work, namely the
ratio of energy absorbed during the failure process and double of cross-sectional
area, could be used to characterize the fracture property of wood component.
Although fracture energy could be used to assess the toughness of different
materials and could help to understand the fracture process, it could not be used as
design parameter directly.
Wood is anisotropic and inhomogeneous material, and for air-dried wood, the
strain–stress curve possessed linear characteristic, so the mechanical behavior of
wood fitted into the linear elastic behavior and wood could be approximately
assumed as an orthotropic material [6]. While wood possesses many characteristics
which are obviously different with those of other orthotropic materials, the major
characteristic of wood is the high anisotropy among the three principal directions
from the perspective of the composition and structure of wood. As known, most
wood cells arrange along axial direction and only a few ray cells arrange along
radial direction. What’s more, because of the great difference in chemical bond
energy between valence bonds of cellulose chain molecules in the axial directions
and in radial direction, the transverse tensile strength is only 1/50–1/24 of the
longitudinal tensile strength [7, 8]. Wood species, density, and load direction could
influence the fracture of wood. Meanwhile, the structure of wood was also an
important factor that determined the fracture of wood [9]. Bodner et al. [10, 11]
have studied the initiation and propagation of fracture behavior of wood clear
specimens by tensile along grain test and bending test. Bodner et al. [12] also
studied the bending and tensile fracture behavior of wood specimen with oblique
grain. Grekin [13] studied the perpendicular-to-grain tensile fracture behavior of
Scots pine and obtained that the TR crack propagation was in tortuous, stepwise
crack path. Galicki and Czech [14] studied the failure of pine with grain in different
angles to the longitudinal direction of specimen under uniaxial tensile stress. In
recent years, finite element technique is also extensively used to simulate the
fracture behaviors of composites as well as wood [15–18].
During the bending test of wood specimen with crack perpendicular to grain, we
found that crack propagated along grain to a certain length then stopped, at this
moment, specimen had the ability to bear load as well. It was found that the
maximum failure load was several times the critical load when crack initiated. It
was obvious that if the critical load when crack initiated was taken as the criterion
for the safety design of wood component, there must be a waste of strength. Thus, a
thought was inspired that for the safety design of wood component with crack
perpendicular to grain, only the strength of the clear part without crack needed to be
considered instead of considering the angle of fracture mechanics. To verify the
thought above, numerical simulation and experimental verification were performed.
ABAQUS finite element software was used to simulate the crack tip stress field of
Picea asperata SEB (Single edge notched bending) specimen with crack perpen-
dicular to grain to explore why the crack always initiated along grain. Then stan-
dard specimens and specimens with crack perpendicular to grain of two softwoods
4.1 Introduction 65

and two hardwoods were chosen as samples to perform bending test, impact test,
and tensile test. And a criterion applicable for the strength design and safety
assessment of wood component with crack perpendicular to grain was proposed.

4.2 Analysis on Stress Field at Crack Tip

Taking P. asperata with moisture content of 13% as example, digital speckle


correlation method (DSCM) [19, 20] was applied to test 12 elastic parameters as
seen in Table 4.1. According to Chinese national standard GB1938-91 [21] and
GB/T14017-92 [22], tensile strength parallel to grain (rL) and tensile strength
perpendicular to grain (rT) were tested, respectively. They were 95.54 and 3.8 MPa
respectively and rL/rT 25.
For the SEB specimen with a crack perpendicular to grain as seen in Fig. 4.1, the
stress field at crack tip was simulated numerically by ANSYS finite element soft-
ware. As the system is symmetrical, half of the structure is modeled to reduce
calculating amount. And degraded 1/4 singularity element was introduced around
the crack tip. The radius of the singularity element was 0.05 mm and the crack tip
was divided into 10 equal parts. Plane 2 element, namely 2-dimension, six nodes
and triangle elements are adopted, which is applicable to axisymmetric element. As
seen in Fig. 4.2, the mesh was dense near crack tip and sparse away from crack
tip. The applied load was 100 N. The nephograms and isograms of stress field are
shown in Figs. 4.3, 4.4, 4.5, and 4.6.

Table 4.1 12 elastic parameters of Picea asperata


EL ER ET GRL GTL GRT µLR µLT µRT µRL µTR µTL
GPa) (GPa) (GPa) (MPa) (MPa) (MPa)
12 1.2 0.6 495 209 85 0.4 0.4 0.45 0.04 0.23 0.02

Force sensor AE sensor

x
r
w
y a

Fig. 4.1 Schematic diagram of SEB sample and test device


66 4 Transverse Fracture of Wood

Fig. 4.2 The mesh generation around crack tip

Fig. 4.3 The nephogram of stress (ry) perpendicular to crack around crack tip
4.2 Analysis on Stress Field at Crack Tip 67

Fig. 4.4 The nephogram of stress (rx) parallel to crack around crack tip

Fig. 4.5 The isogram of stress (ry) perpendicular to crack


68 4 Transverse Fracture of Wood

Fig. 4.6 The isogram of stress (rx) parallel to crack

There are 113 nodes right above crack tip (the plane Y = 0) in total, and after the
nodes around load point and the nodes with negative stress in the bending com-
pression zone are neglected, stress analysis is performed on 70 nodes taken from the
bottom up, as seen in Table 4.2.
For the 70 nodes, the variation tendencies of tensile stress perpendicular to crack
(rY) and tensile stress parallel to crack (rX) are shown in Fig. 4.7.
It can be seen that there are both rY and rX within a large zone in front of the
crack tip, and rY/rX is almost a constant in the range of 4–5 within a large zone
around the crack tip apart from singular points around the crack tip (Table 4.2). The
stresses declined quickly with the increase of Distance from crack-tip and after
reached a certain value, stresses changed gently.

4.3 The Cracking Direction of Transverse Crack

The fracture of wood always begins at the micro-area in front of a crack tip. The
micro-area called the fracture process region is not only the high-stress area but also
where the microstructure has decisive influence. The stress analysis has shown that
there are both rY and rX in the fracture process region, and rY/rX is almost a
constant ranging from 4 to 5 (The values are different for different wood). Thus,
4.3 The Cracking Direction of Transverse Crack 69

Table 4.2 The stress of 70 nodes taken from the bottom up right ahead of crack tip
Node ry rx ry/rx Node ry rx ry/rx
(Mpa) (Mpa) (Mpa) (Mpa)
1 95.78 13.94 6.87 36 3.48 0.75 4.62
2 21.31 4.87 4.37 37 3.38 0.73 4.63
3 19.46 4.62 4.21 38 3.28 0.70 4.65
4 16.82 3.78 4.45 39 3.17 0.68 4.66
5 14.56 3.34 4.36 40 3.07 0.66 4.68
6 13.14 2.96 4.43 41 2.98 0.63 4.70
7 11.95 2.69 4.44 42 2.89 0.61 4.72
8 10.97 2.47 4.43 43 2.80 0.59 4.74
9 10.21 2.30 4.45 44 2.71 0.57 4.77
10 9.57 2.15 4.46 45 2.62 0.55 4.79
11 9.00 2.02 4.46 46 2.54 0.53 4.82
12 8.50 1.90 4.47 47 2.46 0.51 4.85
13 8.06 1.80 4.47 48 2.38 0.49 4.89
14 7.67 1.72 4.47 49 2.30 0.47 4.93
15 7.31 1.64 4.47 50 2.22 0.45 4.97
16 6.99 1.56 4.48 51 2.15 0.43 5.01
17 6.69 1.49 4.48 52 2.07 0.41 5.07
18 6.42 1.43 4.49 53 2.00 0.39 5.12
19 6.17 1.37 4.49 54 1.93 0.37 5.19
20 5.93 1.32 4.49 55 1.86 0.35 5.26
21 5.71 1.27 4.50 56 1.79 0.33 5.34
22 5.51 1.22 4.50 57 1.72 0.32 5.43
23 5.31 1.18 4.51 58 1.65 0.30 5.54
24 5.13 1.14 4.51 59 1.58 0.28 5.66
25 4.96 1.10 4.52 60 1.52 0.26 5.79
26 4.79 1.06 4.52 61 1.45 0.24 5.96
27 4.63 1.02 4.53 62 1.37 0.23 6.14
28 4.48 0.99 4.54 63 1.32 0.21 6.37
29 4.34 0.96 4.54 64 1.26 0.19 6.64
30 4.20 0.92 4.55 65 1.20 0.17 6.98
31 4.07 0.89 4.56 66 1.13 0.15 7.39
32 3.95 0.86 4.57 67 1.07 0.14 7.94
33 3.82 0.84 4.58 68 1.01 0.12 8.65
34 3.71 0.81 4.59 69 0.95 0.10 9.66
35 3.59 0.78 4.60 70 0.89 0.08 11.15
70 4 Transverse Fracture of Wood

160
20
120 15
80 10

40 σY 5 σX
σX
0 0
0 1 σY 2 3 4 5 6 0 1 2 3 4 5 6
Distance from crack-tip (mm) Distance from crack-tip (mm)

Fig. 4.7 The variation tendency of rY and rX along the crack initial direction within 5 mm from
the crack tip

when the tensile stress region in front of crack tip developed to the interface, if the
interfacial strength of the wood (namely tensile strength perpendicular to grain) was
higher than 1/5 of longitudinal strength, the interface would not crack. Then the
crack would propagate through the interface. At this moment, the mechanical
behavior of wood is just like that of common brittle materials, and brittle failure
would happen. On the contrary, if the transverse strength was lower than 1/5 of
longitudinal strength, the interface would be pulled open by the tensile stress rX
leading to the formation of a new crack perpendicular to the original crack. The
passivated crack could eliminate the stress concentration and the propagation of
original crack was prevented, then the unstable fracture was avoided. In fact, the
tensile strength perpendicular to grain is only about 1/40–1/20 of the tensile
strength parallel to grain (Here, the rT/rL of the P. asperata is only 1/25). As a
result, the crack in timber component always initiates along the grain.
Then for the timber beams with cracks in different angle to grain, when subjected
to transverse bending force, what are the directions that cracks propagate along?
The question will be discussed in Chap. 5. The next question is if the load when
crack initiates along the grain is treated as the critical load to calculate the fracture
toughness, whether the fracture toughness is still the basic attribution of wood. This
question will be discussed in next section.

4.4 Test of Critical Stress Intensity Factor

4.4.1 Material and Method

Air-dried samples of P. asperata, Castanopsis hystrix, Koompassia spp., and Melia


azedarach are taken with moisture content (MC) of 13%. And the densities of the
four air-dried woods are 0.427, 0.794, 0.958, and 0.487 g/cm3, respectively.
4.4 Test of Critical Stress Intensity Factor 71

The SEB sample (Fig. 4.1) is adopted to measure the critical stress intensity
factor (KIC) of wood transverse fracture according to national standard “Metallic
materials—Determination of plane-strain fracture toughness” (GB4161-84) [23].
As the accuracy of KIC depends on the load when crack initiates, acoustic emission
(AE) technique is applied to monitor the load when crack initiates along the grain.
According to GB4161-84 [23], there are three groups of samples for each wood.
In group A, samples are standard with a size of W/B = 2, S = 4 W; in group B,
samples are nonstandard with a size of S/W = 3–5, W/B = 2; in group C, samples
are nonstandard with a size of S/W = 4, 1 < W/B < 4. The crack is single edge
notched LT crack with a length of a = W/2, and the sizes of sample are the same for
different woods, as seen in Table 4.3.
Tests are performed on the computer-controlled mechanics of material testing
machine. Thin steel sheets are placed on the supports to minimize the friction
between sample and supports. AE sensor is attached on sample to monitor the
whole fracture process of the sample (Fig. 4.1). Loading speed is 2 mm/min, and
the load–deflection (F-d) curve is drawn automatically by the computer during
testing. Temperature was 15 °C and the humidity was 50–55% in laboratory.

4.4.2 Test and Results

From the F-d curve as seen in Fig. 4.8, it can be seen that the whole transverse
fracture process of wood includes three stages. (1) In the initial stage (OA), the F-d
curve is linear, and there is no crack along grain, while at point A, crack initiates
along grain, and the slope of F-d curve changes slightly. Also at that moment, AE
signals appear and slight sound of cracking can be heard by a medical stethoscope,

Table 4.3 The sizes of SEB samples for different woods


Species Groups Number S (mm) B (mm) W (mm) a (mm) a/
W
Picea asperata A 10 160 20.88 40 20 0.5
B 11 200 20.75 40 20 0.5
C 11 96 18.46 24 12 0.5
Castanopsis A 12 120 15 30 15 0.5
hystrix B 13 150 15 30 15 0.5
C 15 120 20 30 15 0.5
Koompassia A 16 144 19 36 18 0.5
spp. B 16 180 19 36 18 0.5
C 18 84 14 21 11 0.5
Melia A 16 144 18.5 36 18 0.5
azedarach B 15 180 18.8 36 18 0.5
C 15 120 16 30 15 0.5
72 4 Transverse Fracture of Wood

1.6 B Load

1.2

F (kN) 0.8
A
0.4
C

00 2 4 6 8 10 12
(mm)

Fig. 4.8 The F-d curve of wood transverse fracture

so the load at point A is the critical load Fcr. (2) In the stable propagation stage of
crack (AB), crack propagates slowly with the increase of load, then ceases grad-
ually, during which more cracks along grain occur accompanied with slight sound
of cracking corresponding to the curved section on the F-d curve. (3) In the unstable
fracture stage (BC), when load increases to its maximum Fmax, the tissues around
the load point collapse on the upside of sample by pressure, and the fibers on the
downside of sample fracture by tensile force, accompanied with high sound of
cracking. Meanwhile load decreases stepwise, and the crack opens quickly until the
sample fractures.
Figure 4.9 shows the curve of force versus time and AE cumulative counts
versus time compared with effective voltage of discrete events versus time. As the
effective voltage (mv) is low when crack initiates along grain, decibel (dB) is
adopted to identify the critical load easily, as seen in Fig. 4.10.

B
AE cumulative (counts)

Events RSM (mv)


Load (kN)

A C

Time (s)

Fig. 4.9 Load/time/cumulative events/relative energy (mV) curve for SEB specimen of wood
4.4 Test of Critical Stress Intensity Factor 73

AE cumulative (counts)

Events RSM (mv)


Load (kN)

A C

Time (s)

Fig. 4.10 Load/time/cumulative events/relative energy (dB) curve for SEB specimen of wood

Due to GB4161-84 [23] and substituting the critical load and the sizes of sample
into the following equation, the nominal fracture toughness of wood transverse
fracture can be obtained

PC S  a 
KIC ¼ f ð4:1Þ
BW 3=2 W

where f(a/W) is correction factor.


For SEB sample with single edge notched, the correction factor is
 a 1=2 h    i
a
2
W 3 1:99  Wa 1  Wa 2:15  3:93 Wa þ 2:7 Wa 2
f ¼    ð4:2Þ
W a 3=2
2 1 þ 2a
W 1W

According to ASTM E399, the correction factor is


a  a 1=2  a 3=2  a 5=2  a 7=2  a 9=2
f ¼ 2:9 4:6 þ 21:8 37:6 þ 38:7
W W W W W W
ð4:3Þ

The results of Eqs. (4.2) and (4.3) are close.


The statistics description of results is shown in Table 4.4, and it can be seen that
the coefficients of variation (CV) of several groups are greater than 20%, which is
caused by defects (cross grain or node) and differences in texture. However, the
analysis of variance (ANOVA) (Table 4.5) shows that there is no significant dif-
ference among groups, and the results are not affected by sample sizes on the whole.
The results above indicate that the critical stress intensity factor of wood transverse
74 4 Transverse Fracture of Wood

LT
Table 4.4 The statistics description of KIC for four kinds of woods
Species Group Number F0 Fmax Fmax/ KIC SD CV
(N) (N) F0 (Nmm−3/2) (Nmm−3/2) (%)
Picea A 10 715 2062 2.9 57.71 4.6 7.98
asperata B 11 518 1601 3.15 52.55 6.98 13.3
C 11 501 1345 2.72 59.47 8.28 13.93
Total 56.54 7.29 12.9
Castanopsis A 12 1136 2192 1.93 147.45 29.44 19.97
hystrix B 13 901 1758 1.95 146.09 27.18 18.61
C 15 1323 2825 2.21 128.75 28.58 22.2
Total 134 29.04 20.74
Koompassia A 16 1640 5748 3.55 152.11 24.94 16.4
spp. B 16 1271 5675 4.55 147.44 30.25 20.51
C 18 895 3388 3.87 147.55 53.84 36.49
Total 150.7 31.86 21.14
Melia A 16 936 2190 2.34 89.2 15.33 17.19
azedarach B 15 936 2190 2.34 78.17 25.1 32.11
C 15 703 1415 2.27 78.17 25.1 32.11
Total 83.06 20.92 25.19

fracture measured here, as the nominal fracture toughness of wood transverse


fracture, can be seen as the basic attribution of wood.

4.5 The Influence of Transverse Crack on the Normal


Strength of Wood

Two softwoods and two hardwoods were chosen as samples and they were spruce
(P. asperata), larch (Larix gmelinii), I-69 poplar (Populus spp.), and Castanopsis
hystrix. Standard specimens and specimens with crack perpendicular to grain were
used to perform bending test, impact test, and tensile test. The aim was to inves-
tigate the influence of crack perpendicular to grain on the bending strength, impact
toughness, and tensile strength of wood component.

4.5.1 Influence of Crack Perpendicular to Grain on MOR


of Wood

Bending strength of wood, also called modulus of rupture(MOR), representing the


capacity to bear transverse load, is one of the most important mechanical properties
4.5 The Influence of Transverse Crack on the Normal Strength of Wood 75

LT
Table 4.5 The analysis of variance (ANOVA) of KIC for four kinds of woods
Species Source of Sum of df Mean F F-crit Sig.
variance Squares Square
Picea Between 283.45 2 141.72 3 3.33 No
asperata groups
Within 1365.84 29 47.1
groups
Total 1649.29 31
Castanopsis Between 3045.22 2 1522.61 1.88 3.25 No
hystrix groups
Within 29,836.46 37 806.39
groups
Total 36,230.94 39
Koompassia Between 263.89 2 131.94 0.12 3.21 No
spp. groups
Within 45,420.5 43 1056.29
groups
Total 45,684.38 45
Melia Between 1003.71 2 501.86 1.15 3.21 No
azedarach groups
Within 18,687.98 43 434.6
groups
Total 19,691.69 45

of wood. To investigate the influence of crack perpendicular to grain on MOR of


wood, there were two groups of specimens for each wood species: one was standard
specimen with a size of 300 mm (L)  20 mm (T)  20 mm (R) according to
GB1927-1943-91 [24]; the other one was specimen with crack perpendicular to
grain with a size of 300 mm (L)  30 mm (T)  20 mm (R), and the crack was cut
at the middle of specimen with a depth of 10 mm as seen in Fig. 4.11. Tests were
performed on the WDW-100 computer-controlled mechanical testing machine.
Data were recorded automatically by a computer. And the MOR was calculated
according to GB1927-1943-91 [24]. It was necessary to illustrate specially that the
area parameter of the specimen with crack was the cross-sectional dimension of the
clear wood part and it was 20  20 (mm2) theoretically.

20
20

20
20
10

30

Fig. 4.11 Schematic diagrams of standard specimen and specimen with crack for bending test
76 4 Transverse Fracture of Wood

Fig. 4.12 Bending specimen with crack under load- crack initiated along grain

As seen in Fig. 4.12, for specimen with crack, crack initiated along grain. As
described in Sect. 4.2, the failure process of bending specimen with crack was that
the crack initiated along grain, and after propagated to a certain length, the speci-
men fractured with a fracture surface similar to that of standard specimen.
The statistical analysis results of the MOR of both groups of the four wood
species were shown in Table 4.6. It could be seen that the numerical values of the
two groups were similar.

4.5.2 Influence of Crack Perpendicular to Grain on Impact


Toughness of Wood

Impact toughness is the energy absorbed or consumed by per area when wood
fractures because of impact load and it is the inherent ability of wood to resist
impact damage. To investigate the influence of crack perpendicular to grain on the

Table 4.6 Statistical analysis of MOR of standard specimen and specimen with crack
Species Specimen Number Mean (MPa) SD (MPa) CV (%)
Picea asperata Standard 22 74.95 11.06 14.76
Crack 22 78.23 9.62 12.3
Larix gmelinii Standard 27 129.84 12.98 10
Crack 26 125.28 16.35 13.05
I-69 Populus spp. Standard 30 65.28 10.47 16.03
Crack 42 66.12 8.28 12.52
Castanopsis hystrix Standard 16 122.97 13.26 10.78
Crack 16 118.08 10.88 9.21
4.5 The Influence of Transverse Crack on the Normal Strength of Wood 77

Table 4.7 Statistical analysis of impact strength of standard specimen and specimen with crack
Species Specimen Number Mean (MPa) SD (MPa) CV (%)
Picea asperata Standard 30 55.4 10.51 18.97
Crack 30 55.87 5.38 9.62
Larix gmelinii Standard 29 68.01 14.63 21.51
Crack 31 68.33 11.23 16.43
I-69 Populus spp. Standard 30 139.16 46.46 33.38
Crack 30 159.11 31.07 19.53
Castanopsis hystrix Standard 31 83.51 18.38 22.01
Crack 30 88.87 15.3 17.22

impact toughness of wood, there were two groups of specimens for each wood
species: one was standard specimen with a size of 300 mm (L)  20 mm
(T)  20 mm (R) according to GB1927-1943-91 [24]; the other one was specimen
with crack perpendicular to grain with a size of 300 mm (L)  30 mm
(T)  20 mm (R), and the crack was cut at the middle of specimen with a depth of
10 mm.
Impact tests were performed in the mechanical testing machine. The failure
surface of specimen with crack was similar to that of standard specimen. While it
could be observed on the fracture surface that the crack did not initiate along its
original direction but to both sides along grain just like that of bending specimen
with crack. When calculated the impact toughness of the specimen with crack, the
area parameter was 20  20 (mm2) theoretically.
The statistical analysis results of the impact toughness of both groups of the four
kinds of wood species were shown in Table 4.7. It could be seen that the impact
toughness of specimen with crack was all equal to or higher than that of standard
specimen. The main reason was that during impact test, the specimen with crack
absorbed more energy than the standard specimen because of the big size and
quality of specimen with crack.

4.5.3 Influence of Crack Perpendicular to Grain on Tensile


Strength of Wood

Tensile strength parallel to grain represents the ability of wood to bear tensile load
along grain. To investigate the influence of crack perpendicular to grain on the
tensile strength parallel to grain of wood, there were three groups of specimens as
shown in Fig. 4.13. The specimens in first group were standard specimens
according to GB1927-1943-91 [24] (1991) and the cross-sectional dimensions in
the test section were 15 mm (R)  4 mm (T). The cross-sectional dimensions in
the test section of specimens in group 2 were 15 mm (R)  4 mm (T), while a
crack was cut at the middle of specimen with a depth of 3 mm, then the
78 4 Transverse Fracture of Wood

4
15

3 15
5

20
5

Fig. 4.13 Standard specimen and specimen with crack for tensile test

cross-sectional dimensions of the clear part were 12 mm (R)  4 mm (T). The


cross-sectional dimensions in the test section of specimens in group 3 were 20 mm
(R)  4 mm (T), while a crack was cut at the middle of specimen with a depth of
5 mm, then the cross-sectional dimensions of the clear part was 15 mm
(R)  4 mm (T). Tensile tests were performed on the WDW-100
computer-controlled mechanical testing machine and data were recorded automat-
ically by a computer. Tensile strength parallel to grain was calculated according
GB1927-1943-91 [24]. For specimens with crack, the area parameters were the
cross-sectional dimensions of the clear part.
The statistical analysis results of the tensile strength of both groups of the four
wood species were shown in Table 4.8. It could be seen that for each wood species,
the mean value varied while the difference was not great.

4.5.4 Discussion

The significant analysis results of the MOR, impact toughness, and tensile strength
of the four wood species were shown in Tables 4.9, 4.10, and 4.11. Variance

Table 4.8 Statistical analysis of tensile strength of standard specimen and specimen with crack
Species Specimen Number Mean (MPa) SD (MPa) CV (%)
Picea asperata Standard 20 95.54 9.77 10.23
Crack A 16 102.23 10.11 9.89
Crack B 22 107.32 10.36 9.65
Larix gmelinii Standard 20 109.79 26.93 24.53
Crack A 18 95.60 33.25 34.78
Crack B 18 87.20 23.13 26.53
I-69 Populus pp Standard 19 85.08 9.90 11.64
Crack A 20 80.85 17.47 21.60
Crack B 15 86.07 16.29 18.93
Castanopsis hystrix Standard 17 106.43 17.78 16.70
Crack A 15 92.58 17.46 18.86
Crack B 19 99.27 16.02 16.14
4.5 The Influence of Transverse Crack on the Normal Strength of Wood 79

Table 4.9 Significant analysis of bending test results of standard specimen and specimen with
crack
Species Source of Sum of df Mean F Fcrit Sig.
variance squares square
Picea Between 118.41 1 118.41 1.102 4.073 No
asperata groups
Within 4514.2 42 107.48
groups
Total 4632.6 43
Larix Between 275.63 1 275.63 1.27 4.03 No
gmelinii groups
Within 11,065.58 51 216.97
groups
Total 11,341.21 52
I-69 Populus Between 12.51 1 12.51 0.146 3.978 No
spp. groups
Within 5987.81 70 85.54
groups
Total 6000.32 71
Castanopsis Between 191.07 1 191.07 1.299 4.171 No
hystrix groups
Within 4411.43 30 147.05
groups
Total 4602.5 31

analysis showed that there was no significant difference between the results of
standard specimen and specimen with crack.
The tensile strength parallel to grain of wood is mainly determined by cellulose,
while the tensile failure of wood is not caused by the fracture of cellulose molecular
chain but the slide of cellulose molecular chains [25]. When subjected to tensile
stress, failure or slipping would first happen where the molecular chains at a dis-
advantage place born most of the stress. With the increase of tensile load, molecule
chains began to slide and flow, which caused the decrease of cross section and
resulted in the failure of wood. The fracture morphology of tensile specimen
embodied the fracture process above. It also could be seen that for standard
specimens, slide initiated randomly, so the initial point was the weakest place of
material (Fig. 4.14b); for specimen with crack, slide initiated from the pre-crack tip,
what’s more, crack propagated along the grain instead of along its original direction
during tensile process (Fig. 4.14a). And it could be obviously observed that the
distance between the cut pre-crack surfaces increased with the slide of material at
crack tip.
The results of bending test, impact test and tensile test demonstrated that wood
component with crack perpendicular to grain would not fail under low work stress
in spite of the stress singularity at crack tip. Wood possessed fine capacity to resist
transverse fracture because of its specific organization structure, so for the safety
80 4 Transverse Fracture of Wood

Table 4.10 Significant analysis of impact strength of standard specimen and specimen with crack
Species Source of Sum of df Mean F Fcrit Sig.
variance squares square
Picea Between 3.27 1 3.27 0.047 4.007 No
asperata groups
Within 4040.53 58 69.66
groups
Total 4043.8 59
Larix Between 1.5 1 1.5 0.01 4.01 No
gmelinii groups
Within 9773.9 58 168.52
groups
Total 9775.4 59
I-69 Populus Between 5968.28 1 5968.28 3.821 4.007 No
spp. groups
Within 90,584.6 58 1561.8
groups
Total 96,552.88 59
Castanopsis Between 436.89 1 436.89 1.523 4.004 No
hystrix groups
Within 16,924.03 59 286.85
groups
Total 17,360.92 60

Table 4.11 Variance analysis of tensile strength of standard specimen and specimen with crack
Species Source of Sum of df Mean F Fcrit Sig.
variance squares square
Picea Between 1457.98 2 728.99 2.552 3.165 No
asperata groups
Within 15,711.3 55 285.66
groups
Total 17,169.28 57
Larix Between 4981.37 2 2490.69 3.168 3.172 No
gmelinii groups
Within 41,671.2 53 786.25
groups
Total 46,652.57 55
I-69 Populus Between 282.49 2 141.24 0.639 3.179 No
spp. groups
Within 11,277.12 51 221.12
groups
Total 11,559.61 53
Castanopsis Between 1534.93 2 767.47 2.642 3.191 No
hystrix groups
Within 13,941.9 48 290.46
groups
Total 15,476.83 50
4.5 The Influence of Transverse Crack on the Normal Strength of Wood 81

Fig. 4.14 Tensile specimens: a for specimen with crack, crack initiated along grain; b failed
standard tensile specimen

assessment and strength design of wood component with crack perpendicular to


grain, net stress criterion should be adopted, namely the strength of the clear part of
component apart from the size of crack. However, the fracture mechanics criterion
such as K criterion or G criterion should be used for the fracture along grain of
wood.

4.6 Energy Release Rate of the Mode I Interlaminar


Fracture of Wood Beam and the Bending
Delamination Damage of Plywood

For wood, cracking along grain is another damage form that has non-negligible
influence on the stiffness and stability of wood beam, and it is common in plywood,
so it is necessary to study the Mode I delamination and debonding damage of wood
beam and plywood further. Plentiful research work has been done by Williams [26,
27], Wang and Williams [28], Reeder [29, 30], and Du and Wang [31] to analyze
82 4 Transverse Fracture of Wood

and calculate the strain energy release rate of Mode I delamination and debonding
damage of laminate composites.
As seen, the schematic diagram of delamination crack tip under load and corner
analysis of wood beam is shown in Fig. 4.15. It is supposed that the thickness of
wood beam is H, the width is B, and there is a penetrative lamination fracture zone
and load distributed uniformly in width of the beam. The thickness above the
delamination is h1 and the thickness below is h2, and the applied moments are M1
and M2. When crack tip spreads from A-A to B-B by da, the energy release rate of
the whole system is
 
1 dUe dUs
G¼  ð4:4Þ
B da da

where Ue is the work done by external load to the system; Us is the strain energy of
the system.
When there is no delamination crack, the corners of section A-A and B-B under
load are u0 and u0 þ ddua0  da respectively, as seen in Fig. 4.15a.
When there is delamination crack, the corner of section A-A at crack tip was still
u0 under load (Fig. 4.15a). When crack tip spreads from section A-A to section
B-B, the corners of the section above and below section A-A are (Fig. 4.15b)
 
du1 du0
  da ð4:5Þ
da da

du2 du0
ð  Þ  da ð4:6Þ
da da

Due to the flexure beam theory, the relationships between the corners and the
applied moments are

du0 M1 þ M2
¼ ð4:7Þ
da EI 0

(a) (b)

Fig. 4.15 Sketch of wood beam interlaminar fracture: a the corner of sectional A-A at
interlaminar crack tip. b While sectional A-A expanding to B-B with da, the corner of sectional
A-A and corner between superstratum and lower layer
4.6 Energy Release Rate of the Mode I Interlaminar Fracture … 83

du1 M1
¼ ð4:8Þ
da EI 1

du2 M2
¼ ð4:9Þ
da EI 2
3
where the inertia moment I0 ¼ BH
12 , and let n ¼ H , so:
h1

 
Bh31 BH 3 h1 3
I1 ¼ ¼ ¼ n3 I 0
12 12 H
 
Bh3 BH 3 H  h1 3
I2 ¼ 2 ¼ ¼ ð1  nÞ3 I0
12 12 H

So:

du1 M1 M1
¼ ¼
da EI 1 En3 I0
ð4:10Þ
du2 M2 M2
¼ ¼
da EI 2 Eð1  nÞ3 I0

When delamination crack existed, the corners of the beam above and below the
lamination crack are different from the corners of the beam with no delamination
crack. The variation of the corners leads to the decrease of system energy and it is
also the driving force of the appearance and spread of delamination. Thus, during
the propagation of crack from section A-A to section B-B, the variation of the work
done by external load is
   
du1 du0 du2 du0
@Ue ¼ M1   da þ M2   da ð4:11Þ
da da da da

Then, substituting Eq. (4.10) into Eq. (4.11), the variation rate of the work done
by external load in the propagation process of delamination crack is
   
dUe du1 du0 du2 du0
¼ M1  þ M2 
da da da da da
 " #
M1 M1 þ M2 M2 M1 þ M2
¼ M1  þ M2  ð4:12Þ
En3 I0 EI0 Eð1  nÞ3 I0 EI0
" #
1 M12 M22
¼ 3
þ  ðM1 þ M2 Þ2
EI0 n ð1  nÞ3

In the same way, when the delamination crack existed, the strain energy of the
beam above and below delamination crack were different from the strain energy of
84 4 Transverse Fracture of Wood

the system of those when there is no delamination crack. Due to the flexure beam
theory, the variation of system strain energy is

1 M12 1 M22 1 ðM1 þ M2 Þ


@Us ¼ da þ da  da ð4:13Þ
2 EI1 2 EI2 2 EI0

Thus, the variation rate of the strain energy of the system in the propagation
process of delamination crack is

dUe 1 M12 1 M22 1 ðM1 þ M2 Þ


¼ þ 
da 2 EI1 2 EI2 2 EI0
" # ð4:14Þ
2 2
1 M1 M2
¼ þ  ðM1 þ M2 Þ
2EI0 n3 ð1  nÞ3

So, the energy release rate of the system during the propagation of crack from
section A-A to section B-B is
  " #
1 dUe dUs 1 M12 M22 2
Gi ¼  ¼ þ  ðM1 þ M2 Þ ð4:15Þ
B da da 2BEI0 n3 ð1  nÞ3

If there is transverse crack perpendicular to the longitudinal axis direction of


beam, as seen in Fig. 4.16, e is the length of the transverse crack, C is the length of
the delamination crack, and S is the span length. Thus, when delamination crack
propagates under three-point bending load, the strain energy release rate can be
obtained by substituting M1 ¼ 0; M2 ¼ M ¼ FðS  CÞ=4 into Eq. (4.15):
" #
1 M2
G¼ M 2
2BEI0 ð1  e=HÞ3
" # ð4:16Þ
3F 2 ðS  CÞ2 1
¼ 1
8EB2 H 3 ð1  e=HÞ3

Fig. 4.16 Sketch of F


interlaminar fracture of wood
beam containing crack
perpendicular to grain

c H
e

F/2 S F/2
4.6 Energy Release Rate of the Mode I Interlaminar Fracture … 85

The strain energy release rate will decrease with the increase of C and when it is
less than the critical strain energy release rate, the delamination crack will cease.
Especially, when C = 0, and the transverse crack converted to side crack along the
grain of wood beam, the rate of critical strain energy released can be calculated by
the following equation:
" #
3F 2 S2 1
GC ¼ 1 ð4:17Þ
8EB2 H 3 ð1  e=HÞ3

where F0 was the load when the crack along the grain began.
The analysis above can also be applied to the debonding behavior of plywood.
The energy consumed by single crack in the process of transverse bending fracture
of wood can be calculated by the following equation:

ZC Z " #
C
3 1
Uii ¼ 2 Gi Bdd ¼  1 F 2 ðS  2dÞ2 dd ð4:18Þ
4EBH 3 0 ð1  a=HÞ3
0

The energy consumed by the cracking of a single layer is not high, thus no
significant change can be found on the F-d curve. However, multiple cracks will
happen in the whole process, so much energy will be consumed.

References

1. Schniewind AP, Pozniak RA (1971) On the fracture toughness of Douglas fir wood. Eng Fract
Mech 2:223–230
2. Prokopski G (1993) The application of fracture mechanics to the testing of wood. J Mater Sci
28:5995–5999
3. Reiterer A, Sinn G, Stanzl-Tschegg SE (2002) Fracture characteristics of different wood
species under mode I loading perpendicular to the grain. Mater Sci Eng A Struct 332:29–36
4. Keunecke D, Stanzl-Tschegg S, Niemz Peter (2007) Fracture characterisation of yew (Taxus
baccata L.) and spruce (Picea abies wL.x Karst.) in the radial-tangential and tangential-radial
crack propagation system by a micro wedge splitting test. Holzforschung 61:582–588
5. Jeronimidis G (1976) The fracture of wood in the relations to its structure. Leiden Bot Ser
3:253–265
6. Kollmann FFP, Côté WA (1968) Principles of wood science and technology I solid wood.
Springer-Verlag, New York
7. Xu YM (2006) Wood science. China Forestry Publishing, Beijing, p 181
8. Danielsson H, Gustafsson PJ (2013) A three dimensional plasticity model for perpendicular to
grain cohesive fracture in wood. Eng Fract Mech 98:137–152
9. Stanzl-Tschegg SE (2006) Microstructure and fracture mechanical response of wood. Int J
Fract 139:495–508
10. Bodner J, Schlag MG, Grüll G (1997) Fracture initiation and progress in wood specimens
stressed in tension. Part I. Clear wood specimens stressed parallel to the grain. Holzforschung
51:479–484
86 4 Transverse Fracture of Wood

11. Bodner J, Schlag MG, Grüll G (1997) Fracture Initiation and progress in wood specimens
stressed in tension part II compression wood specimens stressed parallel to the grain.
Holzforschung 51:571–576
12. Bodner J, Schlag MG, Grüll G (1998) Fracture initiation and progress in wood specimens
stressed in tension part III. Clear wood specimens with various slopes of grain. Holzforschung
52:95–101
13. Grekin M, Surini T (2008) Shear strength and perpendicular-to-grain tensile strength of
defect-free scots pine wood from mature stands in Finland and Sweden. Wood Sci Technol
42:75–91
14. Galicki J, Czech Michal (2015) Tensile strength of softwood in LR orthotropy plane. Mech
Mater 37:677–686
15. Vasic S, Stanzl-Tschegg S (2007) Experimental and numerical investigation of wood fracture
mechanisms at different humidity levels. Holzforschung 61:367–374
16. Schoenmakers JCM, Jorissen AJM, Leijten AJM (2010) Evaluation and modelling of
perpendicular to grain embedment strength. Wood Sci Technol 44:579–595
17. Qiu LP, Zhu EC, van de Kuilen JWG (2014) Modeling crack propagation in wood by
extended finite element method. Eur J Wood Prod 72:273–283
18. Cao J, Li FG, Wang QR et al (2016) Analysis of fracture criteria for 7050 aluminum alloy
with different geometries based on the elastic strain energy density. Theor Appl Fract Mect
81:50–66
19. Viotti MR, Kaufmann GH, Galizzi GE (2006) Measurement of elastic moduli using spherical
indentation and digital speckle pattern interferometry with automated data processing. Opt
Laser Eng 44:495–508
20. Wang QH, Xie HM, Tang P et al (2009) A study on the mechanical properties of beagle
femoral head using the digital speckle correlation method. Med Eng Phys 31:1228–1234
21. Method of testing in tensile strength parallel to grain of wood (GB 1938-91) (1991). China
Standards Press, Beijing
22. Method of testing in tensile strength perpendicular to grain of wood (GB/T 14017-92) (1991).
China Standards Press, Beijing
23. National standard Metallic materials-Determination of plane-strain fracture toughness
(GB4161-84) (1984)
24. Testing methods for physical and mechanical properties of wood (GB1927-1943-91) (1991).
China Standards Press, Beijing
25. Naoto W (1986) Basic of wood application. Shanghai Science and Technology Press,
Shanghai, pp 401–407
26. Williams JG (1988) On the calculation of energy release rates for crack laminates. Int J Fract
Mech 36:101
27. Williams JG (1989) End corrections for orthouropic DCB specimens. Compos Sci Technol
35:367
28. Wang Y, Williams JG (1992) Corrections for mode II feacture toughness specimens of
composite materials. Compos Sci Technol 43:251
29. Reeder JR, Crews JH (1990) Mixed-mode bending methed for delamination testing. AIAA 28
(7):1270
30. Reeder JR, Crews JH (1991) Redesigm of mixed-mode bending test for delamination
toughness. In: Proceedings of international conference on cognitive modeling, vol 8,
section 13-36
31. Du S, Biao W (1998) Micromechanics of composite materials. Science Press, Beijing,
pp 161–237
Chapter 5
Finite Element Analysis of Wood Crack
Tip Stress Field and Prediction
of the Crack Propagation Direction

Abstract The crack in timber components was always initiated along the grain no
matter what the original crack direction was. To investigate the fracture mechanism,
three-point bending (SEB) specimens of spruce (Picea jezoensis) are selected for
tests. And finite element method is applied to analyze the stress fields around crack
tips of 4 different SEB specimens with initial cracks orientated 0°, 30°, 60°, and 90°
to grain. The “tangential normal stress intensity factor ratio criterion” is adopted to
predict the propagation direction of crack. The results show that (1) the radial lines
of the maximum tensile stress of the 4 crack bodies are all along grain when a radial
plane is made at the crack tip; (2) The ratio of the tensile stress perpendicular to
crack surface (rY) and tensile stress parallel to the crack surface (rX) is almost a
constant ranging from 1 to 5 in a large fracture process region of the crack tip apart
from several singular points; (3) The “tangential normal stress intensity factor ratio”
gets maximum along the grain at the crack tip no matter how large the angle
between the crack and grain is. Both theoretical analysis and experimental results
prove that the propagation direction of crack will turn along the grain. Also, the
relationship between the interfacial strength and toughness of wood is discussed.

5.1 Introduction

Wood is one of the few plant materials that can be directly used as structural
materials, meanwhile, it is also the oldest and still the most widely used natural
structure material. The annual total production of wood is in the range of 500
million tons, among which about 220 million tons are used for material applications
[1]. What’s more, the demand quantity of wood increases day by day. And trans-
verse bending is the main load carrying way of timber components. So, it is very
important to predict the propagation direction of the cracks in different angles to the
longitudinal direction of wood under transverse force, which is also very important
for the design and safety evaluation of timber.
For air-dried wood, the strain–stress curve had a linear characteristic, so the
mechanical behavior of wood fitted into the linear elastic behavior and wood could

© Springer Nature Singapore Pte Ltd. 2018 87


Z. Shao and F. Wang, The Fracture Mechanics of Plant Materials,
https://doi.org/10.1007/978-981-10-9017-2_5
88 5 Finite Element Analysis of Wood Crack Tip Stress Field …

be approximately assumed as an orthotropic material [2]. As natural biocomposite,


wood possesses many characteristics which are obviously different from those of
other orthotropic materials. The major characteristic of wood is the high anisotropy
among three principal directions from the perspective of the composition and
structure of wood. When studied on the stress distribution around the crack tip of
anisotropic material, Wu [3] proposed that the stress field strength of crack tip was
not only determined by the stress intensity factor but also by the property of
anisotropy and the directional function for crack and the principal directions of
material. Fracture behaviors related to the grain of wood were always concerned by
researchers, and also there were many research methods. Boatright [4] and
Stanzl-Tschegg et al. [5, 6] have studied on the influence of grain direction on wood
fracture behaviors. Keunecke [7] evaluated the elasticity and fracture behaviors of
yew (Taxus baccata L.) and spruce (Picea abies wL.x Karst.) in the
radial-tangential plane. Nairn [8] simulated transverse fracture in solid wood on the
scale of growth rings by material point method (MPM) that was well suited for
modeling problems with complex geometries and with crack propagation in arbi-
trary directions. Grekin [9] studied on the tensile failure behavior of Scots pine
wood sample with the crack perpendicular to the grain. Danielsson [10] developed a
three-dimensional cohesive zone model for analysis perpendicular to grain fracture
of wood and the finite element method was implemented for numerical calculations.
Acoustic emission technique (AET) was also a favorable method for the research on
the damage and fracture of wood. In recent years, the application of ESEM or SEM
and the finite element technology enabled researchers to understand the fracture
behavior and mechanism of wood on different scales [11–13].
As known, most of the wood cells arrange along the axial direction and only a
few ray cells arrange along the radial direction. What’s more, the cellulose chain
molecules are connected by C–C and C–O in the axial direction and by C–H and
H–O in the radial direction. It leads to the result that the transverse tensile strength
is only 1/50 * 1/24 of the longitudinal tensile strength because of the great dif-
ference in chemical bond energy [14]. Therefore, a wood specimen with an initial
crack perpendicular to grain shows different fracture characteristics compared with
that of the wood specimen with an initial crack parallel to grain when subjected to
transverse bending force. That is, the initial crack perpendicular to grain would
propagate along grain instead of along the original initial crack direction at the
initial stage. Then, when timber beams with initial cracks in different angles to grain
subjected to transverse bending force, what are the directions that cracks propagate
along? In fact, this involves the problem to predict the crack direction of timber
beam with initial crack. van der Put [15] pointed out that it was necessary to know
the stresses at the crack boundary to determine the failure mode of the composite.
Thus, in this section, Picea asperata was chosen as a sample to perform three-point
bending tests. And for the timber beam with initial cracks in an arbitrary angle to
grain, stress field around the crack tip was analyzed by ABAQUS finite element
software. Then “tangential normal stress intensity factor ratio criterion” was applied
to predict the propagation direction of crack. At last, the relationship between the
interfacial strength and toughness of wood was discussed.
5.2 Materials and Methods 89

5.2 Materials and Methods

5.2.1 Materials and Fundamental Data

Picea asperata was selected as a sample with a moisture content of 13% by air
dried. Wood stem could be seen as cylindrical symmetry body. When a cube is cut
along three orthogonal principal axes: longitudinal (L), radial (R), and tangential
(T) axes at a certain distance from the pith, it could be regarded as the orthogonal
anisotropic body. So the stress–strain relationship could be expressed as follows by
the engineering elastic constants [16]:
8 9 2 38 9
>
> eL >
> 1=EL lLR =ER lLT =ET 0 0 0 >
> rL >>
>
> > >
> rR >
>
> eR >
>
>
6 lRL =EL
6 1=ER lRT =ET 0 0 0 7 7>>
>
>
>
< = 6 < =
eT lTL =EL lTR =ER 1=ET 0 0 0 7 r
¼6
6
7
7
T
> c
> RT >>
> 6 0 0 0 1=GRT 0 0 7> > s >
>
> > RT >>
>
> cTL >
>
>
4 0 0 0 0 1=GTL 0 5> >
> s > >
: ; : TL >;
cLR 0 0 0 0 0 1=GLR sLR

The elastic parameters and the tensile strength parallel to grain (rL) and per-
pendicular to grain (rT) should be measured before going to analyze the stress field
around crack tip and predict the crack direction. Three rectangle samples with the
positive axis and three with 45° off-axis (Fig. 5.1) were used to measure 12 elastic
parameters by the digital speckle correlation method (DSCM) [17, 18]. The 12
elastic parameters was EL = 12 GPa, ER = 1.2 GPa, ET = 0.6 GPa,
GRT = 85 Mpa, GTL = 209 Mpa, GLR = 495 Mpa, lLR = 0.4, lLT = 0.4,
lRL = 0.04, lRT = 0.45, lTL = 0.02, lTR = 0.23, respectively.
The tensile strength parallel to grain (rL) and tensile strength perpendicular to
grain (rT) were tested according to Chinese national standards GB1938-91 [19] and
GB/T14017-92 [20] respectively. The results were shown in Table 5.1. The ratio of
rL/rT was about 25. All the experiments were conducted at 23 ± 2 °C and relative
humidity of about 66 ± 3%.

F F F F F F

(d) GLR (e) GTL (f) GRT

Fig. 5.1 Six kinds of speckle specimens for testing elastic modulus
90 5 Finite Element Analysis of Wood Crack Tip Stress Field …

Table 5.1 Tensile strength of Picea asperata


Node Number Average Variance Standard Variant Nicety
(MPa) (MPa) deviation coefficient index
(MPa) (%) (%)
rL 20 95.54 11.52 2.51 12.05 5.26
rT 17 3.8 0.33 0.08 8.62 4.18

5.2.2 Fracture Specimens

The specimen was standard single-edge-notched bending (SEB) specimen. The size
of the specimen was S = 160 mm, W = 40 mm, B = 20 mm, a = 20 mm,
respectively, where S was the span length, W was the height of specimen, B was the
thickness of the specimen, and a was the length of the initial crack. The initial crack
was located in the middle of the specimen. The initial crack orientation (b) was 0°
(a crack parallel to the grain, i.e., TL crack), 30°, 60°, and 90° (a crack perpen-
dicular to the grain, i.e., LT crack) respectively (Fig. 5.2).
The numerical simulations of the three-point bending tests were performed by
ABAQUS finite element software. CPS8R (eight-node plane strain element,
reduced integration) was adopted to mesh the model. And degraded 1/4 singularity
element was introduced around the crack tip. The radius of the singularity element
was 0.05 mm and the crack tip was divided into 36 parts (10° per unit). The mesh
was dense near crack tip and sparse away from the crack tip (Fig. 5.3). The applied
load was 100 N.

F F
T
L X 60
X L
T 0
Y Y
a a
R R

F F
L
T X T 90
30 X
L W=40
Y Y
a a=20
R R

S=160 B=20

Fig. 5.2 Schematic diagram of cracks and fiber orientation of SEB samples
5.3 Results and Discussion 91

Fig. 5.3 Arrangement of the quarter point elements at the crack tip Mesh generation around crack
tip

5.3 Results and Discussion

5.3.1 Stress Field of the Crack Tip

The field cloud maps of the stress (rY) perpendicular to the initial crack surface and
field cloud maps of the stress (rX) parallel to the initial crack surface were shown in
Fig. 5.4. It could be seen that the maximum stress rY of the four crack bodies were
all along grain when a radial plane was made at the crack tip, while the maximum
stress rX tended to perpendicular to the grain. The value of rY, rX, and rY/rX around
the crack tips of the four crack bodies were shown in Table 5.2. Variation ten-
dencies of rY and rX within 5 mm from the crack tip were shown in Fig. 5.5. It
indicated that with the distance from the crack tip increased the stress declined
quickly, then tended to be stable and very low.

5.3.2 Prediction of Crack Propagation

The fracture of materials always began at a micro-area in front of a crack tip. The
micro-area was called the fracture process region, which was not only the area with
high stress but also the area where microstructure played a determined role. When
predicted the crack direction of composite, Gregory and Herakovich [21] found that
no matter on any periphery within a certain distance from the crack tip (region
where critical stress intensity factor K played a leading role), crack would propagate
along the direction where the “tangential normal stress intensity factor ratio” got its
maximum. According to the hypothesis of Gregory and Herakovich [21], the
strength of material on the direction tangent to the periphery of arbitrary angle h
was given by

Thh ¼ X sin2 ðb  hÞ þ Y cos2 ðb  hÞ ð5:1Þ

The tangential normal stress was given by


92 5 Finite Element Analysis of Wood Crack Tip Stress Field …

2.16
2.46
3.08

5.15 1.99
2.52
3.09
5.04

o
β =0
o
3mm β =0 1mm
σY σX

2.06
2.53 2.01
3.06
2.50
4.89
3.03

4.83

o o
β =30 β =30
σY 5mm σX 1mm

1.45
2.08 1.47
2.98
5.14 1.97

2.89

4.52

o o
β =60 5mm β =60
1mm
σY σX

0.47

4.79 2.84 1.99 1.48


1.01
1.47
o o
β =90 β =90
2.00
σY 5mm σX 1mm

Fig. 5.4 The stress field cloud maps of the 4 kinds of specimens (b was the angle between crack
and grain)
Table 5.2 The rY and rX around crack tip of the four crack bodies (F = 100 N)
D/mm 0° 30° 60° 90°
rY/MPa rX/MPa rY/rX rY/MPa rX/MPa rY/rX rY/MPa rX/MPa rY/rX rY/MPa rX/MPa rY/rX
0 227.5 83.47 2.726 256.5 71.64 3.58 217.09 42.45 5.114 142.1 14.48 9.814
0.0073 147.97 40.933 3.615 84.249 35.444 2.377 59.56 21.62 2.755 38.07 8.7734 4.339
0.05 63.346 14.431 4.39 31.362 12.898 2.432 21.625 8.1487 2.654 13.476 3.2714 4.119
0.1 45.428 10.09 4.502 21.23 8.9743 2.336 14.614 5.6156 2.602 9.3721 2.2142 4.233
5.3 Results and Discussion

0.15 37.415 8.3198 4.497 17.356 7.4037 2.344 11.978 4.6401 2.581 7.7176 1.8134 4.256
0.2 32.486 7.1922 4.517 14.969 6.4087 2.336 10.311 4.0237 2.563 6.6682 1.5646 4.262
0.25 29.048 6.4147 4.528 13.289 5.7222 2.322 9.1467 3.5994 2.541 5.9443 1.3933 4.266
0.3 26.492 5.833 4.542 12.046 5.21 2.312 8.2728 3.284 2.519 5.4034 1.2662 4.268
0.35 24.494 5.3775 4.555 11.069 4.8091 2.302 7.5904 3.0372 2.499 4.9798 1.1665 4.269
0.4 22.879 5.0077 4.568 10.272 4.4839 2.291 7.035 2.8369 2.48 4.6358 1.0856 4.27
0.45 21.531 4.6995 4.582 9.6054 4.2118 2.28 6.5703 2.6699 2.461 4.3491 1.0181 4.272
0.5 20.388 4.4372 4.595 9.0363 3.9821 2.269 6.1737 2.5278 2.442 4.1052 0.9606 4.274
0.55 19.4 4.2103 4.608 8.5421 3.7826 2.258 5.8295 2.4049 2.424 3.8942 0.9108 4.276
0.6 18.535 4.0113 4.621 8.1072 3.6076 2.247 5.5266 2.2971 2.406 3.7092 0.8671 4.278
0.65 17.769 3.8348 4.634 7.7201 3.4524 2.236 5.257 2.2015 2.388 3.5451 0.8283 4.28
0.7 17.083 3.6768 4.646 7.3725 3.3134 2.225 5.0149 2.1159 2.37 3.3982 0.7935 4.283
0.75 16.465 3.5341 4.659 7.0576 3.1879 2.214 4.7956 2.0386 2.352 3.2656 0.762 4.285
0.8 15.904 3.4044 4.672 6.7705 3.0738 2.203 4.5955 1.9683 2.335 3.145 0.7334 4.288
0.85 15.391 3.2857 4.684 6.5071 2.9694 2.191 4.4119 1.904 2.317 3.0347 0.7072 4.219
0.9 14.92 3.1764 4.697 6.2641 2.8733 2.18 4.2425 1.8449 2.3 2.9333 0.6831 4.294
(continued)
93
Table 5.2 (continued)
94

D/mm 0° 30° 60° 90°


rY/MPa rX/MPa rY/rX rY/MPa rX/MPa rY/rX rY/MPa rX/MPa rY/rX rY/MPa rX/MPa rY/rX
0.95 14.484 3.0754 4.71 6.0389 2.7845 2.169 4.0854 1.7902 2.282 2.8395 0.6607 4.297
1 14.065 2.9782 4.723 5.8217 2.699 2.157 3.934 1.7376 2.264 2.7492 0.6392 4.301
1.2 12.692 2.6581 4.775 5.1026 2.4176 2.111 3.4319 1.5645 2.194 2.4524 0.5681 4.317
1.4 11.622 2.4073 4.828 4.535 2.1971 2.064 3.0347 1.429 2.124 2.2203 0.5112 4.335
1.6 10.745 2.2006 4.883 4.0651 2.0155 2.017 2.7052 1.3176 2.053 2.0295 0.4659 4.357
1.8 10.008 2.0257 4.941 3.6663 1.8619 1.969 2.4248 1.2234 1.982 1.8685 0.4265 4.381
2 9.3597 1.8713 5.002 3.3136 1.7264 1.919 2.1762 1.1404 1.908 1.7268 0.3916 4.41
2.5 8.087 1.5641 5.17 2.6143 1.4571 1.794 1.6797 0.976 1.721 1.4469 0.3215 4.5
3 7.1073 1.3238 5.369 2.0727 1.2468 1.662 1.291 0.8481 1.522 1.2303 0.266 4.625
4 5.6742 0.9626 5.895 1.2913 0.9318 1.386 0.7174 0.6578 1.091 0.9115 0.1809 5.04
5 4.6326 0.6899 6.715 0.7521 0.6953 1.082 0.3062 0.5164 0.593 0.6783 0.1144 5.929
D distance from crack tip
5 Finite Element Analysis of Wood Crack Tip Stress Field …
5.3 Results and Discussion 95

Fig. 5.5 The distribution of rY and rX in front of crack tips of the four crack bodies (b was the
angle between crack and grain)

rX þ rY r X  rY
rhh ¼ þ cos 2h  sXY sin 2h ð5:2Þ
2 2

where X and Y were the strength on the first and second principal directions of
material, respectively, sXY the shear strength, b the angle between the initial crack
and the first principal direction of material (Fig. 5.6). So the “tangential normal
stress intensity factor ratio” was defined as the ratio of tangential normal stress to
the strength of material in the same direction as follows:
96 5 Finite Element Analysis of Wood Crack Tip Stress Field …

Fig. 5.6 The schematic of


the normal stress ratio at the
crack tip

rhh
R¼ ð5:3Þ
Thh

Thus, the crack would propagate along the direction of h when R obtains the
maximum value.
The “tangential normal stress intensity factor ratio” hypothesis has been applied
successfully to predict the crack propagation direction of composite materials,
which was also called “tangential normal stress intensity factor ratio criterion”. It
was assumed that the criterion was appropriate for the orthotropic material––wood.
To verify the assumption, ABAQUS was applied to analyze the stress of each node
on the singularity element at the crack tip of P. asperata SEB specimens, as seen in
Fig. 5.3. The applied load was 100 N. The stress of each node and Thh, rhh and R
calculated by Eqs. (5.1)–(5.3) were shown in Table 5.3.
It was obvious from Table 5.3 that no matter how large the angle between crack
and grain was, the R (The load was 100 N) got maximum on the radial direction
along the grain. Therefore, according to the “tangential normal stress intensity
factor ratio criterion”, it could be determined that for timber beam with initial cracks
in arbitrary angle to grain, cracks would always initiate along the grain under
transverse load. As seen in Fig. 5.7, the crack propagation paths of the four crack
bodies were shown.

5.4 The Relationship Between Interfacial Strength


and Toughness of Wood

After millions of years’ evolution, as a response to the bending load caused by wind
and snow, tree has evolved into particular structure, which enables tree with high
resistance ability to transverse bending failure. As known, most wood cells are
oriented in the longitudinal axis. The first microstructure level is a tubular structure,
and the second microstructure level is cell wall with multilayer structure reinforced
by microfibrils which are bonded together with various non-cellulose components.
Thus, the transverse interfacial strength is much lower than the axial tensile strength
Table 5.3 Stress and tangential normal stress intensity factor ratio of singularity element around crack tip (F = 100 N)
b h 0° 10° 20° 30° 40° 50° 60° 70° 80° 90° 100° 110° 120° 130° 140° 150° 160° 170° 180°
0° rX / 63.35 43.91 30.54 24.67 22.07 20.51 18.9 16.82 14.78 13.45 12.65 12.04 12.1 13.39 15.85 19.11 22.31 23.15 3.6
MPa
r Y/ 14.43 14.88 15.44 16.02 16.63 17.21 17.66 17.75 17.13 15.51 13.1 10.54 8.34 6.61 5.16 3.81 2.45 1.05 0.86
MPa
sXY/ 0 −2.77 −2.4 −1.54 −0.62 0.32 1.39 2.69 4.21 5.7 6.77 7.28 7.38 7.26 6.99 6.48 5.57 3.85 1.25
MPa
Thh 3.8 6.57 14.53 26.74 41.7 57.64 72.61 84.81 92.77 95.54 92.77 84.81 72.61 57.64 41.7 26.74 14.53 6.57 3.8
rhh 63.35 43.99 30.32 23.84 20.44 18.26 16.77 15.92 15.62 15.51 15.4 15.39 15.67 16.57 18.32 20.9 23.57 23.8 3.6
R 16.67 6.7 2.09 0.89 0.49 0.32 0.23 0.19 0.17 0.16 0.17 0.18 0.22 0.29 0.44 0.78 1.62 3.62 0.95
30° rX / 31.36 38.05 49.61 57.57 29.93 17.17 13.29 12.75 12.69 11.74 9.38 6.45 4.04 1.94 −0.18 −1.5 −1.21 0.23 4.12
MPa
rY/ 12.9 12.42 11.99 11.8 12.42 12.96 13.46 13.93 14.35 14.66 14.68 14.11 12.71 10.68 8.57 6.83 5.45 4.24 4.05
MPa
sXY/ 0 0.22 −0.32 −3.9 −5.43 −4.25 −2.56 −1.86 −0.94 −0.03 1.02 2.25 3.46 4.33 4.75 4.87 4.83 4.61 3.76
MPa
Thh 26.77 14.56 6.58 3.8 6.55 14.51 26.7 41.67 57.6 72.57 84.78 92.76 95.54 92.79 84.83 72.64 57.67 41.74 26.77
rhh 31.36 37.2 45.41 49.51 28.04 18.89 15.63 14.98 14.62 14.66 14.87 14.66 13.54 11.34 8.11 4.8 2.68 1.93 4.12
R 1.17 2.56 6.69 13.03 4.28 1.3 0.59 0.36 0.25 0.2 0.18 0.16 0.14 0.12 0.1 0.07 0.05 0.05 0.15
60° rX / 21.63 23.29 25.66 29.13 34.64 43.82 47.55 20.96 10.47 8.37 9.11 9.84 9.31 7.07 4.05 1.36 −0.96 −2.72 0.74
MPa
r Y/ 8.15 7.86 7.52 7.18 6.87 6.58 6.49 7.05 7.43 7.73 7.96 8.14 8.25 8.2 7.8 6.97 5.87 4.88 4.78
5.4 The Relationship Between Interfacial Strength and Toughness …

MPa
sXY/ −1.16 −0.71 −0.41 −0.25 −0.32 −1.06 −4.44 −5.44 −4.18 −2.99 −2.12 −1.53 −1.05 −0.52 0.12 0.74 1.19 1.48 1.63
MPa
Thh 72.59 57.62 41.69 26.72 14.52 6.56 3.8 6.57 14.54 26.75 41.72 57.65 72.62 84.82 92.78 95.54 92.77 84.8 72.59
rhh 21.63 23.07 23.8 23.86 23.48 23.01 20.6 12.17 8.96 7.73 7.27 7.35 7.6 7.22 5.71 3.41 0.61 −1.98 0.74
R 0.3 0.4 0.57 0.89 1.62 3.51 5.42 1.85 0.62 0.29 0.17 0.13 0.1 0.09 0.06 0.04 0.01 −0.02 0.01
(continued)
97
Table 5.3 (continued)
98

b h 0° 10° 20° 30° 40° 50° 60° 70° 80° 90° 100° 110° 120° 130° 140° 150° 160° 170° 180°
90° rX / 13.48 13.64 14.11 14.89 16.01 17.59 19.83 23.27 28.8 29.62 11.41 5.18 4.64 5.72 6.49 6.16 4.53 2.32 −0.43
MPa
rY/ 3.27 3.18 2.97 2.7 2.44 2.19 1.95 1.72 1.45 1.38 1.54 1.6 1.57 1.49 1.36 1.2 0.96 0.64 0.46
MPa
sXY/ 0 0.27 0.49 0.62 0.7 0.71 0.66 0.47 −0.16 −2.48 −2.96 −2.07 −1.33 −0.89 −0.69 −0.6 −0.53 −0.45 −0.38
MPa
Thh 95.54 92.78 84.82 72.62 57.65 41.72 26.75 14.54 6.57 3.8 6.56 14.52 26.72 41.69 57.62 72.59 84.8 92.77 95.54
rhh 13.48 13.23 12.5 11.3 9.72 7.85 5.85 3.93 2.33 1.38 0.82 0.69 1.19 2.36 3.69 4.39 3.77 2.12 −0.43
R 0.14 0.14 0.15 0.16 0.17 0.19 0.22 0.27 0.35 0.36 0.13 0.05 0.04 0.06 0.06 0.06 0.04 0.02 0
5 Finite Element Analysis of Wood Crack Tip Stress Field …
5.4 The Relationship Between Interfacial Strength and Toughness … 99

Fig. 5.7 The crack propagation direction of the four crack bodies in the SEB tests

of wood. And it is just the weakened interfacial strength that enables tree stem to
possess excellent toughness.
The relationship curves of rY/rX and distance from crack tips of the four crack
bodies were shown in Fig. 5.8. It indicated that no matter how large the angle
between the crack and grain was, rY/rX was almost a constant ranged from 1 to 5
within 4 mm from the crack tip apart from singular points (Table 5.4). Although
with increase in the distance, the ratio increased, and the stress was much lower
there.
In order to explain the relationship between interfacial strength and toughness of
wood better, the specimen with the crack perpendicular to the grain was taken as an
example as shown in Fig. 5.8. The stresses declined quickly with the distance from
the crack tip increased and then changed gently after reaching a certain value.
While, the ratio of rY to rX was almost a constant ranged from 4 to 5 (Fig. 5.5)

Fig. 5.8 The relationship 12


curve of rY and rX around the 10
crack tips of the four crack
bodies 8
6
4
2
0
0 1 2 3 4 5 6
Distance from crack tip (mm)
100 5 Finite Element Analysis of Wood Crack Tip Stress Field …

Table 5.4 The values of rY/rX within 4 mm from the crack tip
Distance 0 0.0073 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4
from crack
tip (mm)
b (°) 90 9.81 4.34 4.12 4.23 4.26 4.26 4.27 4.27 4.27 4.27
60 5.11 2.75 2.65 2.6 2.58 2.56 0.54 2.52 2.5 2.48
30 3.58 2.38 2.43 2.37 2.34 2.34 2.32 2.31 2.3 2.29
0 2.73 3.61 4.39 4.5 4.5 4.52 4.53 4.54 4.55 4.57
Distance 0.45 0.5 0.55 0.6 0.65 0.7 0.75 0.8 0.85 0.9
from crack
tip (mm)
b (°) 90 4.27 4.27 4.28 4.28 4.28 4.28 4.29 4.29 4.29 4.29
60 2.46 2.44 2.42 2.41 2.39 2.37 2.35 2.33 2.32 2.3
30 2.28 2.27 2.26 2.25 2.24 2.23 2.21 2.2 2.19 2.18
0 4.58 4.59 4.61 4.62 4.63 4.65 4.66 4.67 4.68 4.7
Distance 0.95 1 1.2 1.4 1.6 1.8 2 2.5 3 4
from crack
tip (mm)
b (°) 90 4.3 4.3 4.32 4.34 4.36 4.38 4.41 4.5 4.62 5.04
60 2.28 2.26 2.19 2.12 2.05 1.98 1.91 1.72 1.52 1.09
30 2.17 2.16 2.11 2.06 2.02 1.97 1.92 1.79 1.66 1.39
0 4.71 4.72 4.77 4.83 4.88 4.94 5 5.17 5.37 5.89

apart from some singular points. Thus, when the tensile stress region in front of
crack tip developed to the interface, the interface would not crack if the interfacial
strength (namely tensile strength perpendicular to grain) was higher than 1/5 of
longitudinal strength of the wood. Then, the crack would propagate through the
interface. At this moment, wood would have the same failure behavior with
common brittle materials, and brittle failure would happen. On the contrary, if the
interfacial strength was lower than 1/5 of longitudinal strength, the interface would
be pulled open by the tensile stress rX leading to the formation of a new crack
perpendicular to the original crack, as seen in Fig. 5.9. The passivated crack could
eliminate the stress concentration and prevent the propagation of original crack to
avoid unstable fracture. In fact, the tensile strength perpendicular to grain was much
lower than the tensile strength parallel to grain, for example, the ratio of rT/rL of
the P. asperata was only 1/25 in this section, so cracks in timber beam always
propagate along the grain. The bonding force between interfaces of cells in a living
tree is lower than that in dried wood because of water permeation, thus the inter-
faces are easy to slip. This is why living trees have fine flexibility and the ability to
resist bending fracture even after bitten by beasts or chopped by knife or axe.
The theory to improve the toughness of material by means of interface control
has been successfully applied in the design of artificial composites. The structure
and property of interface would influence the overall performance of composite
directly. Studies [14, 22–25] indicated that the interface with low bonding force and
5.4 The Relationship Between Interfacial Strength and Toughness … 101

Fig. 5.9 Schematic diagram (a) (b)


of wood interface to resist
crack growth perpendicular to
the grain
x

good plasticity and easy to dissociate should be adopted to improve the strength and
toughness of composites. That was because low bonding force was beneficial for
the interfaces to slip, which would lead to the good plasticity of composites. And a
good plastic interface was helpful to eliminate the stress concentration. Therefore,
the key to produce high-performance composite lies in the study on the effect of the
formation, properties, and stress transfer behavior of interface on macroscopic
properties of the composite. Consequently, the study on interfacial debonding and
the numerical simulation of interface failure is the key research contents in
materials.

References

1. Mantau U et al (2010) Real potential for changes in growth and use of EU forests. Euwood
Final Report, Project: Call for tenders No. TREN/D2/491-2008. Hamburg
2. Kollmann FFP, Cate WA (1968) Principles of wood science and technology. I: solid wood.
Springer, New York
3. Wu EM (1967) Application of fracture mechanics to anisotropic plates. J Appl Mech 34
(4):967–974
4. Boatright SWJ, Ggrrette GG (1983) The effect of microstructure and stress state on the
fracture behaviour of wood. J Mater Sci 18:2181–2199
5. Stanzl-Tschegg SE, Tan DM, Tschegg EK (1996) Fracture resistance to the crack propagation
in wood. Int J Fract 75:347–356
6. Stanzl-Tschegg SE (2006) Microstructure and fracture mechanical response of wood. Int J
Fract 139:495–508
7. Keunecke D, Stanzl-Tschegg SE, Niemz P (2007) Fracture characterization of yew (Taxus
baccata L.) and spruce (Picea abies wL.x Karst.) in the radial-tangential and tangential-radial
crack propagation system by a microwedge splitting test. Holzforschung 61:582–588
8. Nairn JA (2007) Material point method simulations of transverse fracture in wood with
realistic morphologies. Holzforschung 61:375–381
102 5 Finite Element Analysis of Wood Crack Tip Stress Field …

9. Grekin M, Surini T (2008) Shear strength and perpendicular-to-grain tensile strength of


defect-free Scots pine wood from mature stands in Finland and Sweden. Wood Sci Technol
42:75–91
10. Danielsson H, Gustafsson PJ (2013) A three dimensional plasticity model for perpendicular to
grain cohesive fracture in wood. Eng Fract Mech 98:137–152
11. Frühmann K, Burgert I, Stanzl-Tschegg SE et al (2003) Mode I fracture behaviour on the
growth ring scale and cellular level of spruce (Picea abies [L.] Karst.) and Beech (Fagus
sylvatica L.) Loaded in the TR crack propagation system. Holzforschung 57:653–660
12. Stanzl-Tschegg SE, Keunecke D, Tschegg EK (2011) Fracture tolerance of reaction wood
(yew and spruce wood in the TR crack propagation system). J Mech Behav Biomed 4:688–
698
13. Qiu LP, Zhu EC, Kuilen JWG (2014) Modeling crack propagation in wood by extended finite
element method. Eur J Wood Wood Prod 72:273–283
14. Xu YM (2006) Wood science. China Forestry Publishing, Beijing, p 181
15. van der Put TACM (2015) Exact and complete fracture mechanics of wood-theory extension
and synthesis of all series C publications. Delft Wood Science Foundation Publication Series
No 3-ISSN1871-675X
16. Jones RM (1975) Mechanics of composite materials. Scripta Book Company, Washington
17. Viotti MR, Kaufmann GH, Galizzi GE (2006) Measurement of elastic moduli using spherical
indentation and digital speckle pattern interferometry with automated data processing. Opt
Laser Eng 44(6):495–508
18. Wang QH, Xie HM, Tang P et al (2009) A study on the mechanical properties of beagle
femoral head using the digital speckle correlation method. Med Eng Phys 31(10):1228–1234
19. Chinese national bureau of standards (1991) Method of testing in tensile strength parallel to
grain of wood (GB 1938-91). China Standards Press, Beijing
20. Chinese national bureau of standards (1991) Method of testing in tensile strength
perpendicular to grain of wood (GB/T 14017-92). China Standards Press, Beijing
21. Gregory MA, Herakovich CT (1986) Predicting crack growth direction in unidirectional
composites. J Compos Mater 20:67–85
22. Marshall DB, Cox BN, Evans AG (1985) The mechanics of matrix cracking in brittle-matrix
fiber composites. Acta Metall 33(11):2013–2021
23. Hughes JDH (1991) The carbon fibre/epoxy interface—a review. Compos Sci Tech 42(1):13–
45
24. Kim JK, Mai YW (1991) High strength, high fracture toughness fibre composites with
interface control—a review. Compos Sci Tech 41(4):333–378
25. Shen GL, Hu GK, Liu B (2006) Mechanics of composite materials. Tsinghua University Press
Chapter 6
Fractal Features and Acoustic Emission
Characteristics of Wood Fracture

Abstract In this chapter, the fractal theory is applied to study the fractal feature of
the fracture surfaces of five types of woods along the grain and the relationship
between the fractal dimension of fracture surface and fracture toughness of wood is
established. The results show that fracture toughness parallel to the grains of various
woods is different because of their textural diversity and such differences are also
shown in the morphology of fracture surfaces. Furthermore, there is an evident and
direct proportional relation between the fractal dimension and fracture toughness
along grains. Then the evolution characteristics of microstructure during the bending
failure of the clear samples and samples with the transverse crack of four types of
woods are studied by acoustic emission (AE) technique, and the initiation and
expansion of different types of damage of wood component are identified with AE
characteristic parameters. The results showed that: (1) AE event counts developed
slowly and most were the low-amplitude AE events at the low strains and a large
number of high-amplitude AE events appeared in peak load or fracture stage for the
standard sample. (2) The initiation and expansion of crack tip could be monitored
efficiently by AE technique in the whole process of wood three-point bending test for
the notched sample. (3) The AE signals were related to different damage patterns/
modes. The AE characteristics of cell wall fracture were high amplitude, high
energy, and long duration time AE events but the AE characteristics of cell wall
damage and spallation, cell wall buckling, and collapse were low-amplitude,
low-energy and short duration time AE events. (4) Kaiser effect appeared at low
loading and Felicity effect at high loading under repeated wood bending loading.
The Felicity ratio could better indicate the damage degree of wood structure.

6.1 The Fractal Features of Wood Fracture

6.1.1 Introduction

Observations have shown that fracture surfaces for different materials are statisti-
cally very well described by self-affine fractal [1], and fractal dimension has been
considered as a measure of fracture surface morphology of materials. It also reflects

© Springer Nature Singapore Pte Ltd. 2018 103


Z. Shao and F. Wang, The Fracture Mechanics of Plant Materials,
https://doi.org/10.1007/978-981-10-9017-2_6
104 6 Fractal Features and Acoustic Emission Characteristics of Wood Fracture

the materials inherent properties, such as the components of structure,


microstructure and so on. Therefore, the study of the morphology of fracture sur-
faces and mechanical properties is a very active field of research [2, 3].
Wood is a kind of porous-layered biological composite. Its damage fracture
process can be described as nuclear of mass micro-damages, expansion, connection,
and ultimately causing a fracture. Due to the nonhomogeneity and anisotropy in its
composition and structure, the fractography of wood is more complex than metal.
This makes it difficult to the quantitative morphology of wood fracture surface and
the fracture mechanisms. Thus, researchers are trying to look for a new index in the
quantitative morphology of wood fracture surfaces and expect to predict the wood
intensity and understand its fracture mechanism with it [4]. Therefore, fractal offers
a new theory for developing a better understanding of fracture mechanics of wood.
The fractal geometry was proposed by Mandelbrot [5] and first brought it into
materials science in 1984 [6]. Over the past nearly three decades, extensive studies
on the applications of fractal geometry in the field of materials science and society
have been conducted and becomes a key research in the world. However, the
applications of fractal geometry in wood began at 1990s, and focused on wood
surfaces from sorption isotherms [7], the process of water entering timber [8, 9],
and wood color variation [10]. In recent years, researchers apply fractal to study the
microstructure of wood fracture surfaces. Severa and Buchar [11] studied methods
of evaluation of fracture surface of the wood with fractal. Morel et al. [12] applied
fractal geometry of fracture surfaces in quasi-brittle materials wood. Ponson et al.
[13] found the fracture surface of the quasi-brittle material (wood) was shown to be
self-affine. Wang et al. [14] researched 60 different woods of surface roughness
based on fractal dimension. However, up to now, there is no report on the corre-
lation relationship between the fractal dimension of various woods and fracture
toughness of wood. In this article, the fractal dimension of various woods are
analyzed and the possible correlation with fracture toughness is studied.

6.1.2 Theories

The nature of fractals is reflected in the word itself, coined by Mandelbrot [5] from
the Latin verb frangere, ‘to break’, and the related adjective fractus, ‘irregular and
fragmented’. Fractal geometry indicates the property of self-similar or self-affine on
mathematics, and the fractal dimension is used to describe quantificationally the
degree of irregularity of geometric construction [15]. As a common phenomenon in
nature, observations have shown that material fracture surfaces are self-affine. The
application of fractal geometry in fracture studies is helpful to understand the
fracture itself and builds the possible relationship between fracture toughness and
fractal dimension.
The method to measure the fractal dimension of the material fracture surface is
that choose a kind of area element small enough to cover the fracture surface.
6.1 The Fractal Features of Wood Fracture 105

The size of the area element is e, and the number of the area elements needed to
cover the rupture surface is N(e), So

NðeÞ ¼ CesD ð6:1Þ

where C is a constant in the equation, while D is fractal dimension of the fracture


surface.
However, it is relatively difficult to find the area element in such a small size. So,
as a relatively mature method, measuring fractal dimension of contour line on
fracture surface is applied in practice. Besides fractal structure of contour line of
cross section is correlative to that of the fracture surface. The existing studies [3] of
other fields indicate that the measure of fractal dimension on the surface by means
of measuring fractal dimension of the contour line of the surface is a good indirect
method to measure and calculate the surface fractal dimension in the experiment.
Figure 6.1 shows the relation between fracture surface S and vertical section
contour line L, where S0 and L0 are the projections of S, L respectively. Assume the
linear element yardstick to measure contour line of cross section is eL, surface
element yardstick to measure surface S is e. Let C = S0, then

SðeÞ ¼ S0 e2D ð6:2Þ

By Richardson formula, we have

LðeL Þ ¼ L0 e1D
L
L
ð6:3Þ

where DL is the fractal dimension of contour line L of cross section. Actually, the
roughness of contour line is RL = L/L0. Similarly, we can see the surface roughness
as RS = S/S0. The relation between RS and RL is deduced by Underwood [16] as

4
RS ¼ þ1 ð6:4Þ
pðRL  1Þ

Then, substituting Eqs. (6.2) and (6.3) in Eq. (6.4), we have

4
e2DS ¼ þ1 ð6:5Þ
pðe1D
L
L
 1Þ

Fig. 6.1 Sketch of wood Fracture Surface


fracture surface and its Vertical Section
projection Contour Line S L
Vertical Section

Projection Face Projection Line


So Lo
106 6 Fractal Features and Acoustic Emission Characteristics of Wood Fracture

Since the fractal structure of S and contour line of section L exists in the same
scale.
Therefore, let e = eL, we get
 
pe4L
ln 1DL
4ðeL 1Þ þ p
DS ¼ ð6:6Þ
2 ln eL

The equation above is the analytical formula to calculate the fractal dimension
DS of the fracture surface. DL can be calculated by the method of box-counting
which is from the curve that is obtained by electronic scanning, and eL is a fixed
linear element yardstick.

6.1.3 Tests and Analysis

6.1.3.1 Material and Sample

Three kinds of air-dried softwood and two kinds of air-dried hardwood were chosen
for this experiment. They are Cunninghamia lanceolata, Picea jezoensis var.
microsperma, Larix gmelinii, Populus spp, and Castanopsis hystrix respectively.
The moisture content of the samples is about 13%.
According to the standard established by ASTM E399-74 [17], the size of the
CT sample containing TL crack is W = 50 mm, e = 12.5 mm, a = 25 mm,
H = 30 mm (Fig. 6.2). The number of samples of each kind of wood is 30, and the
total number is 150. In order to get such a CT sample, first, we cut a straight flute
about 30 mm with a band saw along the grain, then cut the straight flute forward 1–
2 mm with a sharp blade and press a knife with relatively thick back (or wedge)
into the straight flute to expand the crack port forward about 2–5 mm. After the
process above, the natural incisive crack port is obtained. Then, we saw the CT
samples according to the crack length e + a. The obtained natural incisive crack
samples would correspond well to the real situation in structural lumber. All there
experiments were taken under 23 ± 2 °C and the relative humidity is about
66 ± 3%.

6.1.3.2 Wood Fracture Toughness K TL


IC

The tests were performed on the computer-controlled mechanics of the material


testing machine and CT sample was connected by steel U hook. The test was loaded
with controlling displacement and loading speed was 2 mm/min, at the same time,
load–displacement curve (F-d) was drawn automatically by the computer during
testing. Figure 6.3 shows that the fracture behavior of air-dried wood parallel to the
6.1 The Fractal Features of Wood Fracture 107

Fig. 6.2 Schematic diagram Force


of CT sample with TL crack

Force B
a
e W

grains is quasi-brittle, except for a gentle curve in the initial loading stage because
of the gap between the sample and U hook, and then the F-d curve is nearly linear.
Once the crack began to craze along the grains, i.e., the unstable fracture expansion
happened. So the critical load FQ = Fmax. Then, taking the values of critical load
Fmax of each sample into the following equation recommended by ASTM E399-74
[17], we got average fracture toughness KIC parallel to the grains of the five kinds of
wood.

FQ a
KIC ¼ pffiffiffiffiffi  f ð6:7Þ
B W W

where
a rffiffiffiffiffi
a
f ¼ 29:6
W W
 a 3=2  a 5=2  a 7=2  a 9=2
 185:5 þ 655:7 1017:0 þ 638:9 :
W W W W

6.1.3.3 Fractal Dimension of Fracture Surface

After the experiment, the CT sample was not separated absolutely because of fiber
bridge, it was needed to be separated artificially. In order to measure DS, we put the
fracture surface at the digital microscope with 3CCD and its pixel is as high as
5,400,000. The digital microscope can show the 3D structure of the fracture surface
for its function of restructure and it can also measure the length of contour line
between any two points automatically. To collect a complete information of the
fracture surface, 4  5 = 20 testing surface elements (Fig. 6.4) were chosen from
the natural incisive crack surface (about 20  12.5 mm2). We obtained the 3D
image of the testing surface element under 60 times of microscopy magnification,
and drew 5 contour lines along longitudinal and tangential direction by electronic
scanning on the surface element (Fig. 6.5). Figure 6.6 shows an in log–log repre-
sentation curve of one C. lanceolata sample of the box-counting method which the
liner regression is used to measure the fractal dimension. Then DS of the fracture
108 6 Fractal Features and Acoustic Emission Characteristics of Wood Fracture

F/kN
F/kN

/mm /mm
Cunninghamia lanceolata Picea jezoensis var. microsperma

F/kN
F/kN

/mm /mm
Populus sp Larix gmelinii
F/kN

/mm
Castanopsis hystrix

Fig. 6.3 F-d curve of CT sample of five woods


6.1 The Fractal Features of Wood Fracture 109

Fig. 6.4 Testing surface


elements of fracture surface of
CT sample

surface in meso-field could be gained. The average values of DS of five kinds of tree
are shown in Table 6.1.

6.1.4 The Relationship Between Fracture Toughness


and Fractal Dimension

Generally, the morphology of fracture surfaces reflects the properties of materials. In


a macroscopic view, smooth and flat fracture surface belongs to brittle fracture while
roughness fracture surface is a ductile fracture when more fracture energy was
dissipated during the process of fracture. Different fractal dimensions of different
fracture surfaces reflect the intrinsic attributes of the material to some extent, such as
the component of the structure, microstructure, and so on. So, the fractal dimension
has a close relationship with the fracture toughness of the material [18, 19].
Wood is highly anisotropy biomaterial and tissues are mostly longitudinally
oriented. The first microstructure is multicell tubular structure and the secondary
microstructure is fiber-reinforced multilayer cell wall structure. These tissues were
connected together by hemicellulose and lignin in a quite but no absolutely effective
way to agglutinate together. Therefore, this causes the lower intensity on these
sections than that on longitudinally oriented. So, the fracture behavior of wood
perpendicular to grains is ductile fracture while the fracture behavior parallel to
grains is a brittle fracture. When the crack of Mode I propagates parallel to grains,
crack tip produces high stress-concentration perpendicular to the crack surface,
which makes the crack propagation swiftly among wood organizations. However,
the deformation of wood organizations develops incompletely and it results in the
smooth fracture surface. The resistance ability of fracture, that is, fracture tough-
ness, is different because of the diversified structure of different species. Such
110 6 Fractal Features and Acoustic Emission Characteristics of Wood Fracture

Fig. 6.5 The 3D image of the testing surface element and electron scanning microscope of
various woods fracture surface parallel to grains. a Cunninghamia lanceolata, b Picea jezoensis,
c Larix gmelinii, d Populus spp, e Castanopsis hystrix

Fig. 6.6 The curve of 4


box-counting method of one 3.5
Cunninghamia lanceolata 3
Sample
2.5

2
1.5
y = -1.2291x + 4.1402
1

0.5
0
0 0.5 1 1.5 2 2.5
log ε
6.1 The Fractal Features of Wood Fracture 111

TL
Table 6.1 The fracture toughness KIC and the corresponding DS of the five woods
pffiffiffiffiffiffiffiffi
Species Sample TL
KIC =MPa mm Fractal dimension
number DS
Cunninghamia lanceolata 30 8.0 2.19
Picea jezoensis var. 30 13.8 2.21
microsperma
Larix gmelinii 30 19.5 2.22
Populus spp 30 20.6 2.23
Castanopsis hystrix 30 21.6 2.24

differences are also shown on the morphology of fracture surface parallel to grains.
So, the fracture toughness parallel to grains is the correlation to the fractal
dimension.
The fracture surface of CT sample with TL crack is the radial section in this
work. The morphological characteristics of fracture surface are mainly determined
by radiation speckle on the radial section, growth ring pattern, and wood grains.
The former three kinds of softwood (C. lanceolata, P. jezoensis, L. gmelinii) have
fine to extremely fine xylem ray, and belong to straight texture wood. So, the
fracture surface should be more smooth then that of hardwood with broad xylem
ray. While, in terms of solftwood, C. lanceolata is the gradual transition,
P. jezoensis is a little abrupt transition, and L. gmelinii abrupt transition. The growth
ring pattern of the above three are distinct successively. So, the fracture surface
roughness of the three increases successively. Populus spp is diffuse-porous wood,
and its structure is far more exquisite than that of C. hystrix with the xylem ray of
oak type, meanwhile, C. hystrix has cross grains, so the fracture surface of Populus
spp is flatter than that of C. hystrix. Because of the difference of microstructure the
five types of woods, the fractal dimension of the fracture surface ranges from small
to big as shown in Table 6.1 and it has a significant direct proportion (Fig. 6.7) to
the fracture toughness parallel to the grain:
TL
KIC ¼ 269:77DS  582:65 ðR2 ¼ 0:98Þ ð6:8Þ

Fig. 6.7 The relationship 30


y = 269.77x - 582.65
between fracture toughness 25 R2 = 0.98
and fractal dimension
K TL /MPa mm

20

15

10
IC

0
2 2.1 2.2 2.3 2.4 2.5 2.6
DS
112 6 Fractal Features and Acoustic Emission Characteristics of Wood Fracture

Equation (6.8) shows that there exists a strong relationship between


microstructure and mechanical properties of wood. This makes it possible to for-
mulate the capability of wood to resist crack propagation by the fractal dimension,
i.e., the fractal geometry provides a new method for the analysis of wood fracture.

6.2 Acoustic Emission Characteristics and Felicity Effect


of Wood Fracture Perpendicular to the Grains

6.2.1 Introduction

Acoustic emission (AE) is a widely used nondestructive technique for detecting


damage evolution in various materials. It is defined as a transient elastic wave
generated by the rapid release of energy within a material. It was first applied to
fracture of metal by Kaiser [20] and then promoted and developed in the world.
Studies of wood using acoustic emission have increased over the past 30 years. It is
mainly used in two areas. First, acoustic emission is applied to monitor and control
the drying of wood in order to eliminate or minimize drying defects [21–24].
Second, acoustic emission has been used to monitor fracture behavior in wood
during loading.
Acoustic emission technique is sensitive to crack nucleation and growth and has
a special function in dynamic monitoring, it has been utilized in exploring damage
and fracture mechanism and strength performance of composite materials and made
great achievements. Ono et al. [25] studied the fracture mechanism of carbon
fiber-reinforced thermoplastic composites and successfully distinguished two
fracture modes fiber breakage and delamination. They found peak amplitude, signal
duration, and energy distributions of AE events are different in different fracture
modes. Bakuckas et al. [26] used acoustic emission technique to locate and monitor
damage growth in titanium matrix composites and correlations between the
observed damage growth mechanisms and the AE results in terms of the events
amplitude were established. Katsaga [27] employed acoustic emission techniques to
investigate the process of fracture formation in large, shear-critical, reinforced
concrete beams, and to gain improved insight into the mechanisms of shear failure
and demonstrated AE techniques were emerging as powerful tools in the study of
different aspects of the mechanisms of failure in reinforced concrete. Bucur [28]
showed the principles and a literature review of the AE technique of wood and
indicated that crack nucleation and growth resulted in a sudden change of energy
within a material and so acoustic emission could be used as an analytical tool for
monitoring crack nucleation and growth. Schniewind et al. [29] collected AE sig-
nals during mode I and mixed mode tests at different moisture contents and tem-
peratures and found that the AE activity in mixed mode tests was much higher than
that for Mode I. Aicher et al. [30] used AE to localize crack nucleation in glulam
loaded in tension perpendicular to the grains. Dill-Langer et al. [31] used AE
6.2 Acoustic Emission Characteristics and Felicity Effect … 113

technique to monitor the fracture of clear spruce wood under tensile loading and
found that there was an onset of AE prior to the first visible crack growth
step. Reiterer et al. [32, 33] used AE to monitor Mode I fracture of softwoods
(spruce and pine) and hardwoods (alder, oak, and ash) and stated that the AE counts
up to maximum force are much higher for the softwoods. Chen et al. [34] used AE
to monitor the failure process of hardwood and softwood test pieces under static and
fatigue torsion loading and found it was possible to monitor and analyze the failure
process in wood by AE techniques. Choi et al. [35] studied the fracture processes of
typical fiber-reinforced plastic composites laminates with continuous fiber rein-
forcement and the results showed that the AE characteristics might represent the
process of fiber breakages according to the various loading stages, which expressed
characteristic fracture processes for individual fiber-reinforced composite laminates.
The feature of the AE hit-event rate, in combination with AE amplitude classifi-
cations, could be utilized for nondestructive identification of different fracture
mechanisms.
But there are few studies on how to identify the different fracture modes with AE
on wood bending fracture. In this work, we study the AE characteristics during
wood three-point bending testing and try to identify or distinguish different modes
of wood fracture processes, and discuss the fracture evolution and mechanisms in
light of its microscopic structure with the help of double cantilever beam
(DCB) and compression tests.

6.2.2 Materials and Methods

6.2.2.1 Materials

In order to investigate the AE characteristics of different species in the process of


bending fracture, the air-dried softwood (P. jezoensis) and hardwood (C. hystrix)
were used for the test pieces and they were made of two groups: (1) standard
sample, 300(L)  20(T)  20(R) mm3 in size (Fig. 6.8a); (2) notched sample, 300
(L)  30(T)  20(R) mm3 in size, and 10 mm sharp crack was cut from the center
of the sample at the tangential direction so the section was also 20(T)  20
(R) mm2 in front of the crack (Fig. 6.8b). The total number of samples was 120, of
which 30 samples for each group. Three-point bending test was adopted in tan-
gential direction. The moisture content of the samples was about 13%.
114 6 Fractal Features and Acoustic Emission Characteristics of Wood Fracture

Fig. 6.8 Three-point bending


F
Load
test and AE sensor location:
AE
AE Sensor
Sensor
a standard sample and
b notched sample (a) 20mm

10mm
F
Load
AE
AE Sensor
Sensor

(b) 20mm 30mm


10mm
10mm

240mm
300mm

6.2.2.2 Methods

The bending tests were performed on the computer-controlled mechanics of the


material testing machine. All tests were loaded with controlling deflection and
loading speed was same, at the same time, the force–deflection curve was drawn
automatically by the computer during testing.
AE measurement data was recorded using a four-channel AE detection system.
AE waves were detected by one AE sensor and it was mounted on the sample using
ethyl a-cyanoacrylate glue. The distance between sensor and middle of the sample
was kept to 10.0 ± 0.1 mm (Fig. 6.8). The preamplifier transmission gain of the
AE sensor was 40 dB. The threshold voltage of the AE system was set between 35
and 55 dB depending on test requirements. And then, the AE signals of the wood
fracture process were analyzed automatically by AE detection system.
All these experiments were taken under 23 ± 2° and the relative humidity was
about 66 ± 3%.

6.2.3 Results and Discussions

6.2.3.1 Acoustic Emission Characteristics of Bending Test

When the wood sample was loaded at bending test, the load–deflection curve is
presented in three stages, i.e., (1) linear elastic deformation stage, (2) nonlinear
deformation stage, and (3) toughness fracture stage (Fig. 6.9a). Wood fracture is a
complex multilevel and multistage process. We will observe the molecular chain
rearrangement, slip, orientation, and fracture in the microscopic, at the same time,
the cracks of wood will growth and propagation and finally, fracture. All the
microscopic and macroscopic material structures changes come along with energy
saving and releasing. Therefore, apart from the AE cumulative counts and AE
6.2 Acoustic Emission Characteristics and Felicity Effect … 115

events peak amplitude Amax, the event energy and energy rate were adopted in this
analysis. The event energy was quantified by RMS2 where RMS was the root mean
square of the signal voltage calculated from
2 31=2
ZT
1
RMS ¼ 4 V 2 ðtÞdt5 ð6:9Þ
T
0

With t time, T duration of the AE event, and V the signal voltage value at
moment t.
Figures 6.9 and 6.10 show curves of force versus deflection and AE cumulative
counts versus Deflection compared with events RMS versus deflection and curves
of force versus deflection and AE counts versus deflection compared with energy
rate versus deflection for P. jezoensis and C. hystrix for the standard sample under
three-point bending tests. There was no AE count in the linear elastic deformation
stage for a standard sample, i.e., there was no damage and it could come back.
When it entered into the nonlinear deformation stage and because the compressive
strength was less than the tensile strength of wood, the compressed area started to
yield and the neutral area was offset to drawing area to maintain the overall balance
(Fig. 6.11), the curve of force versus deflection was nonlinear. There was only a
small number of AE signal in the front half of this stage (the difference depends on
the different tree species, the texture orientation, and the threshold value because the
helically wound cellulose cell wall reinforcement extends elastically within the
matrix of hemicellulose and lignin) and in the second half of the stage, the rapid
increases in AE signal due to the cell interface and interlaminar shear (and layer
resulted from the slip and shear within the molecular chain of cellulose). Then, it
entered into the toughness fracture stage and it came along with the high-energy
elastic wave produced by the fiber fracture and pull-out. Usually, the wood beam
sample kept on the integrality to certain load ability in the post-fracture period.
Along with the sample convex bending of fiber to the continuous tension, the
tension stress micro-fracture damage zone expanded and saved the energy. External

(a) (b)
AE cumulative counts

AE cumulative counts
Events RMS (mv)

Stage I Stage II Stage III


Force (kN)

Force (kN)

Energy rate

0 2 4 6 8 10
0 2 4 6 8 10
Deflection (mm)
Deflection (mm)

Fig. 6.9 a The curves of force versus deflection and AE cumulative counts versus deflection
compared with events RMS versus deflection and b the curves of force versus deflection and AE
cumulative counts versus deflection compared with energy rate versus deflection for Picea
jezoensis for standard sample under three-point bending tests
116 6 Fractal Features and Acoustic Emission Characteristics of Wood Fracture

(a) (b)

AE cumulative counts

AE cumulative counts
Events RMS (mv)
Force (kN)

Energy rate
Force (kN)

0 1.5 3.0 4.5 6.0 7.5


0 1.5 3.0 4.5 6.0 7.5
Deflection (mm)
Deflection (mm)

Fig. 6.10 a The curves of force versus deflection and AE cumulative counts versus deflection
compared with events RMS versus deflection and b the curves of force versus 0 deflection and AE
cumulative counts versus deflection compared with energy rate versus deflection for Castanopsis
hystrix for standard sample under three-point bending tests

power was absorbed before pull-out of the fiber cluster fracture, and which would
be released as the high-energy wave with the pull-out of the fiber cluster fracture.
We would find the interlocking form from the fracture surface in the tension stress
zone, but it was almost flush in the compression stress zone, similar to brittle
fracture (Fig. 6.12), which resulted from the reduction of the anti-break strength
due to the crushing loss of the wood cells in this zone.
Figures 6.13 and 6.14 show curves of force versus deflection and AE cumulative
counts versus deflection compared with events RMS versus deflection and curves of
force versus deflection and AE cumulative counts versus deflection compared with
energy rate versus deflection for P. jezoensis and C. hystrix for notched sample
under three-point bending tests. The notched wood sample was also presented as
three stages in bending failure process. But a large quantity of low-amplitude
low-energy AE was generated when the load increases to about 30–50% of the
maximum force Fmax (Figs. 6.13a and 6.14a), at the same time, lateral cracks would
be found around the visible crack tip on the sample surface and the force–deflection
curve would present a salient point due to the stiffness changes. Figure 6.15 shows
the cell interface and interlaminar shear of lateral cracks for P. jezoensis. The lateral
cracks expanded parallel to the wood grain in the interlayer and with the increase of
load the cracks expanded slowly and eventually stopped. And about a 20 mm high
new beam section was formed behind the original transverse crack, similar to the
standard sample (Fig. 6.16). And then the AE characteristics of the notched sample
fracture behavior were similar to the standard sample. It showed good toughness.
There were no significant differences of the bending strength in the statistical sense
between the standard sample and the notched sample after deducting the prefab-
ricated sharp crack (Table 6.2), and once again showed that sample containing
crack perpendicular to the wood grain would not produce low-stress rupture phe-
nomenon because of the crack tip stress singularity.
6.2 Acoustic Emission Characteristics and Felicity Effect … 117

σ− σ b− σ b−

Original
h/2 Compression neutral layer
stress Zone

h/2 Tension Neutral layer


stress zone move down

σ+ σ+ σ b+

Fig. 6.11 Sketch of center axial of beam was moved to pulling side on the bending process

10mm 350μm

Fig. 6.12 Scanning electron microscopy observation of the fracture surface of the standard
sample of Picea jezoensis: a the toughness fracture zone and b brittle fracture zone

6.2.3.2 Analysis of AE Characteristics and Source

There is a relationship between AE signal wave characteristics and AE source


(damage models). The experiments above show there are four kinds of typical
damage models that happened to the sample containing crack perpendicular to the
grain in the bending transversely process, which was demonstrated in the course:
cell wall damage and spallation, cell wall buckling and collapse, formation and
118 6 Fractal Features and Acoustic Emission Characteristics of Wood Fracture

(a) (b)

AE cumulative counts

AE cumulative counts
Events RMS (mv)
Force (kN)
Force (kN)

Energy rate
0 2.5 5.0 7.5 10.0 12.5
0 2.5 5.0 7.5 10 12.5
Deflection (mm) Deflection (mm)

Fig. 6.13 a The curves of force versus deflection and AE cumulative counts versus deflection
compared with events RMS versus deflection and b the curves of force versus deflection and AE
cumulative counts versus deflection compared with energy rate versus deflection for Picea
jezoensis for notched sample under three-point bending tests

(a) (b)

AE cumulative counts
AE cumulative counts

Events RMS (mv)


Force (kN)
Force (kN)

Energy rate
0 1.5 3.0 4.5 6.0 7.5 0 1.5 3.0 4.5 6.0 7.5
Deflection (mm) Deflection (mm)

Fig. 6.14 a The curves of force versus deflection and AE cumulative counts versus deflection
compared with events RMS versus deflection and b the curves of force versus deflection and AE
cumulative counts versus deflection compared with energy rate versus deflection for Castanopsis
hystrix for notched sample under three-point bending tests

Fig. 6.15 The cell interface


and interlaminar shear of
lateral cracks for Picea
jezoensis of SEM
6.2 Acoustic Emission Characteristics and Felicity Effect … 119

Force Force Force

(a) (b) (c)

Fig. 6.16 Sketch of bending process on notched sample: a the formation of lateral cracks at crack
tip; b the formation and expansion of collapse area; c the fiber fracture layer by layer of the tensile
zone brittle fracture of the collapsed zone

Table 6.2 Bending strength in the statistical sense between the standard sample and the notched
sample
Species Sample type Number of Average value SE CV
samples (MPa) (MPa) (%)
Picea asperata Standard 22 74.95 11.06 14.76
sample
Notched 22 78.23 9.62 12.30
sample
Castanopsis Standard 16 122.97 13.26 10.78
hystrix sample
Notched 16 118.08 10.88 9.21
sample
SE standard error; CV coefficient of variation

(a) (b)
Amplitude (mV)

Amplitude (mV)

0 100 200 300 400 0 100 200 300 400


Time (us) Time (us)

Fig. 6.17 The time-domain curve of different damage models: a interlaminar fracture; b cell wall
fracture
120 6 Fractal Features and Acoustic Emission Characteristics of Wood Fracture

expansion of micro-fracture damage area, and cell wall fracture. Figure 6.17 shows
the typical time-domain curve of interlaminar fracture and transverse fracture of the
notched sample for P. jezoensis combined with Figs. 6.13a and 6.14a, the char-
acteristics of AE signal in different damage and fracture models are summarized as
follows:
1. The characteristics of AE signal for interlaminar fracture stage (i.e.,: cell wall
damage evolution and spallation) is low amplitude, short duration, low-count,
and low energy;
2. The AE amplitude of wood cell compressive yields stage (i.e.,: cell wall
buckling and collapse) is lower or shows lower energy. And signal usually does
not count when a signal of the probe is lower threshold value because of the
attenuation of AE wave propagation;
3. The AE signal of wood macroscopic fracture stage (i.e.,: fiber bundle fracture
and pull-out) is high amplitude, long duration, high-energy, and high AE counts;
4. The AE signal characteristic with the correspondence of the formation and
expansion of the micro-fracture damage area is more complicated and it is the
prelude to damage fracture and also exists in the whole period. What makes it
differ from an interlaminar fracture is that the former mainly occurs in type I or
peeling cracking in the cell or the cell layer, while the latter is caused due to the
whole material damage by formation, expansion, and connection of the
micro-fracture caused by the cell wall tear or rupture. The AE characteristic,
which lies between interlaminar fracture and transverse fracture, is affected by
the relative density, cell wall thickness, microscopic structure, the number of
damage, etc.
Although we have the above understanding, it is still very difficult to distinguish
AE signal from different damage models in bending process. As wood is a multi-
level cell structure of biological composite materials, there is always a variety of
deformation and damage that change system energy in the same stage in the near
crack tip process zone. Therefore, double cantilever beam (DCB) experiment along
the grain of wood cracking and the compression experiment along the longitudinal
and transverse have been carried out (Fig. 6.18). And the DCB test will produce
Mode I interlaminar fracture and the compression test makes cell wall buckling and
collapse damage.

Force
(a) Force
(b) (c)
Force

Force
Force Force
AE Sensor AE Sensor

Fig. 6.18 Sketch of a DCB test and b, c compression test


6.2 Acoustic Emission Characteristics and Felicity Effect … 121

Table 6.3 AE characteristics corresponding to damage modes


Damage modes Vmax RMS
mV dB mV dB
Cell wall collapse <2 <65 <0.4 <52
Interlaminar fracture <10 <80 <1.0 <60
Cell wall laminate and cracks formation <35 <90 <10 <80
Fiber bundle fracture and pull-out >35 >90 >10 >80

The results confirmed that only low-amplitude and low-energy AE events


happened in the Mode I interlaminar fracture and lower AE signal energy in the
compression test. And the testing results had a good consistency with [29].
Based on a large number of AE signal analysis of different tree species damage
modes and the AE sensor was set in damage source of 10 cm range. We found both
the maximum of high amplitude AE events voltage Vmax and RMS value could be
distinguished from different damage modes and especially the RMS value was the
most effective. RMS value was direct related to the energy release of the AE events.
The Vmax and RMS values corresponding to different damage modes are listed in
Table 6.3.

6.2.4 Felicity Effect

The energy release is irreversible in the process of material damage. And the same
is with AE, which is known as Kaiser effect. Kaiser effect has been described as
follows: AE only appeared obviously again when the material is loaded to the
previous unloading point. It is first considered that Kaiser effect exists in the metal
and rock. And then it is discovered that Felicity effect exists in the composite
materials and damage metal materials. Felicity effect refers to the phenomenon that
AE appears when reloading is lower than the previous loading. Felicity ratio is
defined as the ratio of repeated loading to the previous.
As a quantitative parameter, it can better reflect the extent of original suffered
material damage or the structure deficiencies. The smaller the Felicity ratio is, the
more serious of the damage or structure deficiencies is. Felicity ratio more than 1
means there exists Kaiser effect, on the contrary, Felicity effect.
Figure 6.19 is the curve of cyclic load bending test and the analysis curve of a
notched sample of P. jezoensis. There were eight times of loading and unloading in
the whole test. The first loading and unloading occurred the beginning of the
interlaminar fracture and the Felicity ratio was 0.99, which meant there existed
instability of the structure defect.
Felicity ratio was more than 1 in the second–sixth loading and unloading
because of the stress redistribution of samples as the lateral crack expansion and
gradually stable, but there was a decline trend close to maximum load. The seventh
122 6 Fractal Features and Acoustic Emission Characteristics of Wood Fracture

(a) (b)

AE cumulative (counts)
AE cumulative (counts)
Load (kN)

Time (s) Load (kN)

(c)
Felicity ratio

Loading-unloading times

Fig. 6.19 The curve of cyclic load bending test and the analysis curve of notched sample of Picea
jezoensis: a the curves of load versus time and AE cumulative counts versus time and b AE
cumulative counts versus load and c felicity rate versus loading–unloading times

loading and unloading process occurred nearby the maximum load and the eighth
loading and unloading process occurred in the third stage of toughness fracture, all
meant the material structure had been suffered serious damage. The repeated load
test behavior of the standard sample was almost similar to the notched sample. But
the Felicity ratio was more than 1 when the load was close to the maximum load in
the whole process. Therefore, the bending test of wood in low load or the most of
stage II showed Kaiser effect and in high load or toughness fracture stage presented
Felicity effect. As a result, AE technology could better monitor the wood damage in
the practical application.
It needed some clarification that the Felicity phenomenon would be different for
different species and threshold voltage. AE signal would also be produced due to
the friction of the microcracks section in the process of loading–unloading, so in
order to preclude the interference of “false damage”, the threshold voltage should
be set higher than normal. In the loading–unloading of this experiment threshold
voltage was set as 0.39 mV (about 52 dB).
References 123

References

1. Bouchaud E (1997) Scaling properties of cracks. J Phys: Condens Matter 9(4319):797–814


2. Kotowski P (1996) Fractal dimension of metallic fracture surface. Int J Fract 141:269–286
3. Yao L, Yang GW (1999) Measurement and calculation of fractal dimension of fractural
morphology and computer simulation. Phys Chem Test Phys Fasc 35(9):399–403 (in
Chinese)
4. Schmitt U, Richter HG (1996) Fracture morphology of hickory (Carya spp. juglandaceae)
under single-blow impact loading. IAWA J 17(2):151–160
5. Mandelbrot BB (1983) The fractal geometry of nature. W.H. Freeman and Company, New
York
6. Mandelbrot BB, Passoja DE, Paullay AJ (1984) Fractal character of fracture surfaces in
metals. Nature 308:721–722
7. Hatzikiriakos SG, Avramldis S (1994) Fractal dimension of wood surfaces from sorption
isotherms. Wood Sci Technol 28:275–284
8. Jose AR (1997) The fractal nature of wood revealed by water absorption. Wood Fib Sci
29(4):333–339
9. Qin DC, Guan N, Jiang XM (1999) Morphology of wood failure in relation to the variation in
tensile strength parallel to grain of three hard pines. J Inst Wood Sci 15(1):1–5 (in Chinese)
10. Liu J, Furuno T (2002) The fractal estimation of wood color variation by the triangular prism
surface area method. Wood Sci Technol 36(5):385–397
11. Severa L, Buchar J (2000) Methods of quantitative fractography usable for evaluation of
fracture surface of wood. Drevarsky Vyskum 45(4):9–17
12. Morel S, Bouchaud E, Schmittbuhl J et al (2002) R-curve behavior and roughness
development of fracture surfaces. Int J Fract 114:307–325
13. Ponson L, Bonamy D, Auradou H et al (2006) Anisotropic self-affine properties of
experimental fracture surfaces. Int J Fract 140:27–37
14. Wang H, Wang KQ, Bai XB (2007) The research of wood surface roughness based on fractal
dimension. J For Eng 23(2):13–15 (in Chinese)
15. Brown SR, Scholz CH (1985) Broad bandwidth study of the topography of natural rock
surfaces. J Geophys Res 90:572–575
16. Underwood EE (1987) Fractal in materiaes research. Acta Stereol 13:269
17. American Society of Testing Materials (1981) Standard methods for plain-strain fracture
toughness of metallic materials. ASTM Designation E399-74
18. Mecholsky JJ, Passoja DE, Feinberg-Ringel KS (1989) Quantitative analysis of brittle fracture
surfaces using fractal geometry. J Am Ceram Soc 72:60–65
19. Drummond JL, Thompson M, Super BJ (2005) Fracture surface examination of dental
ceramics using fractal analysis. Dent Mater 21:586
20. Kaiser J (1950) An investigation into the occurrence of noises in tensile tests or a study of
acoustic phenomena in tensile tests. Ph.D. thesis, Technische Hochschule – Munchen,
Munich, Germany
21. Noguchi M, Kitayama S, Satoyoshi K et al (1987) Feedback control for drying Zelkova
serrata using in-process acoustic emission monitoring. For Prod J 37:28–34
22. Quarles SL (1992) Acoustic emission associated with oak during drying. Wood Fib Sci
24:2–12
23. Booker JD (1994) Acoustic emission and surface checking in Eucalyptus regnans boards
during drying. Holz Roh-Werkstoff 52:383–388
24. Booker JD (1994) Acoustic emission related to instantaneous strain in Tasmanian eucalypt
timber during seasoning. Wood Sci Technol 28:249–259
25. Ono K, Jenh JS, YanK JM (1988) Fracture mechanism studies of carbon fiber-PEEK
composite by acoustic emission. In: Sachse W, Roget J, Yamaguchi K (eds) Acoustic
emission: current practice and future directions. American Society for Testing and Materials,
Philadelphia, pp 395–404
124 6 Fractal Features and Acoustic Emission Characteristics of Wood Fracture

26. Bakuckas JG, Prosser WS, Johnson WS (1994) Monitoring damage growth in titanium matrix
composites using acoustic emission. J Compos Mater 28(4):305–328
27. Katsaga T, Sherwood EG, Mp Collins et al (2007) Acoustic emission imaging of shear failure
in large reinforced concrete structures. Int J Fract 148:29–45
28. Bucur V (1995) Acoustics of wood. CRC Press, Boca Raton
29. Schniewind AP, Quarles SL, Lee H (1996) Wood fracture, acoustic emission, and the drying
process Part 1. Acoustic emission associated with fracture. Wood Sci Technol 30:273–281
30. Aicher S, Höfflin L, Dill-Langer G (2001) Damage evolution and acoustic emission of wood
at tension perpendicular to fiber. Holz als Roh Werkst 59:104–116
31. Dill-Langer G, Aicher S (2000) Monitoring of microfracture by microscopy and acoustic
emission. In: Proceedings international conference Wood and wood fiber composites,
Stuttgart, pp 93–104
32. Reiterer A, Stanzl-Tschegg SE, Tschegg EK (2000) Mode I fracture and acoustic emission of
softwood and hardwood. Wood Sci Technol 34(5):417–430
33. Reiterer A, Sinn G, Stanzl-Tschegg SE (2002) Fracture characteristics of different wood
species under mode I loading perpendicular to the grain. Mat Sci Eng A Struct 332:29–36
34. Chen Z, Gabbitas B, Hunt D (2006) Monitoring the fracture of wood in torsion using acoustic
emission. J Mater Sci 41:3645–3655
35. Choi NS, Woo SC, Rhee KY (2007) Effects of fiber orientation on the acoustic emission and
fracture characteristics of composite laminates. J Mater Sci 42:1162–1168
Chapter 7
Mechanical Characteristics of Bamboo
Structure and Its Components

Abstract In this chapter, first, the fine structure of bamboo is stated. Then the
strength and elastic modulus of bamboo fiber and ground tissue are analyzed by
mixture law and shear lag theory of meso-mechanics, and it is proved that bamboo
fibers are the main component determining the mechanical characteristic of bam-
boo. Furthermore, results of tensile tests on separated bamboo fiber bundles show
that the tensile strength of bamboo fiber obtained from the tests on bamboo blocks
is higher than that on separated fiber bundles. This might be due to the interaction
between components in bamboo in which parenchymatous ground tissue can pass
loads and uniform the stresses loaded on fibers. And the differences of structure
and strength between internodes part and node part of moso bamboo are studied.
The results indicate that either in the non-epidermis-planning samples or in
the epidermis-planning samples, the node does not take a declining effect on the
bending strength, the longitudinal shearing strength, and the compressive strength.
Instead, the node takes a reinforcing effect at different degrees. However, the node
has a significant decline effect on the longitudinal tensile strength. By analysis on
the structure of bamboo, it shows that though the vascular tissue passing bamboo
node is curved and discontinuous at different degrees, bamboo is able to increase
the bearing load area by expanding tissue at node, meanwhile, vascular bundles are
thickened and entwined circuitously. So, the ability to anti-bending and
anti-shearing of the culm under transverse load caused by snow or wind is increased
by bamboo nodes. However, the tensile load born by the culm is less than other
parts of culm during growth of bamboo. Therefore, the enlargement of node part
and the structure evolution of nodes cannot enhance the tensile strength in the
longitudinal direction.

7.1 Introduction

The macro-mechanical properties of bamboo depend on both the properties of its


components and the features of its meso-structure. There are various types of cells
in bamboo culm wall, and they can be classified into two groups generally from the

© Springer Nature Singapore Pte Ltd. 2018 125


Z. Shao and F. Wang, The Fracture Mechanics of Plant Materials,
https://doi.org/10.1007/978-981-10-9017-2_7
126 7 Mechanical Characteristics of Bamboo Structure and Its Components

mechanics’ angle: sclerenchymatous fibers that determine the mechanical properties


of bamboo and parenchymatous ground tissue that plays a role to buffer and transfer
load. The ground tissue and bundle sheaths account for about 55 and 40%
respectively of the whole bamboo tissue while the rest are vessel and primary xylem
[1, 2]. Thus, if vascular bundles are treated as reinforcements and ground tissue
matrix, bamboo culm wall can be seen as a biocomposite with strong structural
features including the volume fraction, distribution law and shape of reinforcement
and the interfacial property.
The unique structure and the corresponding properties of bamboo give useful
inspiration on artificial materials design. The composite structure optimizes the
contribution of vascular bundle to the perfectly mechanical characteristics of
bamboo. In the past 20 years, the structure and components of bamboo and its
corresponding mechanical properties have been widely studied [3–9]. Chuma et al.
[3] have reported on the relationships between fiber volume fraction and
mechanical properties. Inokuchi et al. [4–6] have reported on the effects of volume
fraction of bundle sheath on the vibration properties and bending creep behavior of
bamboo. In recent years, the development of artificial composites posed more
requirements on reinforcing components, and the natural plant fibers have drawn
more attention as reinforcement used for artificial composites. Due to the excellent
tensile strength and modulus of elasticity of bamboo fibers, bamboo flakes and fiber
bundles have been used in polymeric composites [10], and some products have
been used in wind turbine blade and boat industry [11, 12]. However, precise
studies on the mechanical properties of bamboo fiber and parenchymatous ground
tissue with respect to its fiber-reinforced composite structure were not reported
sufficiently as the previous results were estimated on a small quantity of specimens,
and no study was reported on the strength of separated bamboo fiber bundles.
In this chapter, the tensile tests on bamboo blocks and the corresponding volume
fractions of fiber and parenchymatous ground tissue were measured on a large
quantity of specimens. The tensile properties of bamboo components and the
macrographs of fractured bamboo blocks, as well as the micrographs of the fracture
surfaces obtained by scanning electron microscope, were analyzed. Further tensile
tests on separated bamboo fiber bundles were performed. The results were com-
pared with those on bamboo blocks and were discussed.
In addition, with great strength and stiffness, and also abrasion resistant, bamboo
is widely utilized as engineering structure materials in industry including as
high-quality composites [13–17]. Extensive research has been done on the struc-
ture, physical, and mechanical properties and also chemical composition of moso
bamboo [1, 2, 18–26]. However, most of these studies focused on the properties of
bamboo internodes, not the nodes. And the influence of node on the mechanical
property of bamboo is very controversial. Therefore, moso bamboo culm with node
and without node was studied to determine the influence of node on the strength of
the bamboo.
7.2 Mechanical Characteristics of Bamboo 127

7.2 Mechanical Characteristics of Bamboo

Through million years’ natural selection of the process of survival of the fittest, the
fine structure of bamboo has been evolved to adapt to the living environment. As
bamboo may encounter bending damage caused by wind or load during the process
of growth, various specific internal structures are evolved to intensify the ability to
resist damage. For example, the hollow circular section and stepwise changing
section of bamboo stem reflect the function selection of “equal strength bar” that
each part of bamboo stems has the same ability to resist bending deformation
(Fig. 7.1). Bamboo nodes appear at intervals of about 10 cm along the stem.
A connection between the longitudinal and the cross direction is made by the
crossing walls of nodes, which enables bamboo to achieve optimal stiffness and
stability with the least material and increases the capacity of bamboo to resist
transverse compression and shear. Although some discontinuity and different
extents in bending of the vascular tissue may affect the strength of bamboo node,

Fig. 7.1 Shape of bamboo


128 7 Mechanical Characteristics of Bamboo Structure and Its Components

(a) (b)

Fig. 7.2 a The cross section of bamboo stem; b distribution relation between the volume fraction
of fibers (Vf) and the location parameter (x/B) on cross section of moso bamboo

the vascular bundles in node are thickened, thus the thickness of bamboo node is
always greater than that of the adjacent internodes to avoid damage at node under
external load. What’s more, there are vascular bundles procumbent circumferen-
tially or interweave together in bamboo node, which will improve the transverse
property of node.
In bamboo internodes wall most of the cells are arranged along the longitudinal
direction, and there is no transverse ray cell like that in wood [27], so bamboo
internodes wall is a typical unidirectional long-fiber-reinforced biocomposite. The
distribution of reinforcements is sparse in the inner part then denser gradually in the
outer part of bamboo wall. As seen in Fig. 7.2, exponential distribution relation is
found between the volume fraction of fibers (Vf) and the location parameter (x/B) on
cross section of moso bamboo (Phyllostachys pubescens), which is corresponding
to the bending load caused by wind or snow and according to the theory of function
adaptability.
Bamboo fiber has a complex fine structure with multilayer. Just like other plant
fibers, cellulose chain molecule of bamboo fiber is composed of carbon, hydrogen,
and oxygen. Unit lattice is constructed by the cellulose chain molecule, then cel-
lulose basic fibril with a diameter of 35–50 nm constituted of unit lattices according
to Roelofsen structure model [28]. And it is easy for cellulose basic fibrils to gather
together to form microfibrils with larger diameter which will be embedded in the
matrix composed of hemicellulose and lignin to construct lamella structures, then
they are arranged concentrically to form bamboo cell wall alternating with broad
lamella and narrow lamella. The structures of bamboo fiber cell walls can be
divided into two types. One is multilayer structure consists of narrow lamella (4–5
lamellas) alternating with broad ones. The other type has thick cell wall with narrow
cell lumina, and there are three lamellas in the secondary wall, one narrow outer
lamella and two broad inner lamellas.
In broad lamella, the microfibrils are in 10–30° to the axial direction of fiber in
right-handed helix distribution, which determine the physical and mechanical
property along the axial direction. In the narrow lamella, the microfibrils are in
7.2 Mechanical Characteristics of Bamboo 129

40–90° to the axial direction, which affect the transverse physical and mechanical
property of fiber. The ground tissue of bamboo wall consists of parenchymal cells
that distribute among vascular bundles to buffer and transmit load. For parenchymal
cell, the cell wall is easy to buckle under pressure, while, as liquid and solid coupled
cell body material, the performance of the living ground tissue can be strengthened
by the internal hydrostatic pressure.
The structure and property of natural living body may provide human many
inspirations. Bamboo has possessed the macro- and micro-structure that adapts to
the living environment best and gives full play to its wonderful and complex
function. The fine structure of bamboo plays a very important role in the bionics
design of artificial composites.

7.3 The Mechanical Characteristics of the Components


Bamboo

Tensile tests parallel to grain were performed at two levels. One is at macroscopic
level, i.e., on bamboo blocks; the other is at mesoscopic level, i.e., on bamboo fiber
bundles. Meanwhile, fiber volume fraction was also measured on transverse sec-
tions. The moso bamboo (Phyllostachys pubescens) used in for the tests was 4 years
old and acquired from Lujiang, Anhui Province, China. The total height was about
15 m, and the diameter at breast height was 125 mm. The average air-dry density
was 0.71 g/cm3.

7.3.1 “Mixture Law” Method

7.3.1.1 Material

Specimens were taken from two adjacent internodes at the height of about 1.8 m
and the thickness of culm wall was about 10 mm. After air dried, the culm sections
were longitudinally split into 200 mm (L, longitudinal)  10 mm (T, tangential)
strips. Then, they were delaminated from the outer to the inner part of the culm
walls, after the cortex and pith periphery were removed. About six specimens were
obtained for each strip. The final shape and dimension of the specimens for tensile
tests are shown in Fig. 7.3. The dimension of the effective experiment part (middle
section) was 80 mm (L)  5 mm (T)  1.5 mm (R, radial). There are 60 speci-
mens in total with a moisture content of 12.5%.
130 7 Mechanical Characteristics of Bamboo Structure and Its Components

Displacement sensor
5mm R=1.5 mm
T=10mm
50mm
Marked length
35mm 25mm 80mm 25mm 35mm
Effective experiment part
L = 200 mm

Fig. 7.3 Shape and dimension of bamboo block used for tensile test

7.3.1.2 Theory

When the radial thickness of the specimen collected from the culm wall is not so
big, the vascular bundles can be considered as being uniformly distributed.
Normally, fibers have high strength and modulus of elasticity (MOE), while ground
tissue has low strength and MOE and high capability of deformation. It was found
that in the tensile stress (r)-strain (e) curve of bamboo block, the part from the
beginning of loading to the breaking point was almost linear. Therefore, in view of
mechanical behavior, the bamboo block can be simplified as a composite of parallel
connection model composed of two elements, i.e., fibers and ground tissue, as
shown in Fig. 7.4. It is considered that fibers and ground tissue have the same
deformation.
FC, Ff, Fm, rC, rf, and rm are noted as tensile force and stress on composite,
fibers, and matrix, and the corresponding strain and MOE are eC, ef, em EC, Ef, and
Em. The cross-sectional area of composite, fibers and matrix are AC, Af, and Am, so
the volume fractions of fibers and matrix can be calculated as Vf = Af/AC and
Vm = Am/AC, and Vf + Vm = 1. Thus, the following equations can be derived:

FC ¼ Ff þ Fm ð7:1Þ

Or

rC AC ¼ rf Af þ rm Am ð7:2Þ

Divide Eq. (7.2) by AC, then

Fig. 7.4 Schematic


representation of a simplified
parallel connection model of Force Parenchymatous ground tissue Force
bamboo internode culm wall Fiber
composed of fibers and
parenchymatous ground tissue
7.3 The Mechanical Characteristics of the Components Bamboo 131

rC ¼ rf Vf þ rm Vm ð7:3Þ

Fibers and matrix have the same strain:

eC ¼ ef ¼ em ð7:4Þ

Divide Eq. (7.2) by eCAC, then

rC AC rf Af rm Am
¼ þ ð7:5Þ
eC AC eC AC eC AC

In the range of linear elasticity, the relationships among EC and MOE, volume
fraction of fibers, and matrix are

EC ¼ Ef Vf þ Em Vm ¼ Ef Vf þ Em ð1  Vf Þ ð7:6Þ

Equations (7.3) and (7.6) are the “Mixture law” of composite meso-mechanics
[29, 30].

7.3.1.3 Test and Results

Tensile tests were performed on a computer-controlled testing machine. The


specimens were loaded at a constant crosshead speed of 2 mm/min until ruptured.
The stress–strain curve (Fig. 7.5) shows that there is tiny or no nonlinear defor-
mation. The test room temperature was 25 °C and the humidity was 62%.
Samples were taken near the fracture surface of the bamboo specimens after the
tensile tests. They were immersed in mixed solution of half 10% glacial acetic acid
and half 10% hydrogen peroxide and put in an oven for 2–3 days at 60 °C. Then
they were softened by hot and cold water alternately for 1–2 h. 15-lm-thick
transverse sections were cut with a microtome with disposable blades and stained
with safranine. Sections were observed under an optical microscope equipped with
a digital camera linked to an image analysis system. Sections’ images were taken
and treated to distinguish bundle sheaths from ground tissue. The values of Af/AC

Fig. 7.5 Stress–strain curves


of tensile test on bamboo V f = 0.493
blocks with different fiber
volume fraction (Vf)
σ/MPa

V f = 0.204
V f = 0.115
132 7 Mechanical Characteristics of Bamboo Structure and Its Components

(a) (b) (c)


100 100 100

(d) (e) (f)


100 100 100

Fig. 7.6 Images used for fiber volume fraction (Vf) measurements. a, b, c, d, e, and f show a
series of cross sections’ images of the samples located along radial direction from inner part to
outer part of a culm wall and the corresponding values of Vf were 10.65, 14.33, 21.13, 27.30,
34.02, and 44.29%, respectively

were then measured by the image analysis software. Figure 7.6 shows an example
of a series of treated sections’ images of the samples located along the radial
direction from the inner to the outer part of a culm wall.
The ultimate stress and MOE of each bamboo block can be calculated, and then
the relationships among ultimate stress, MOE, and the volume fraction of fibers can
be obtained according to Fig. 7.7, as follows:

rC ¼ 562:69Vf þ 19:042 ¼ 588:732  562:69Vm ðR ¼ 0:936Þ


ð7:7Þ
EC ¼ 40:129Vf þ 0:2219 ¼ 40:351  40:129Vm ðR ¼ 0:955Þ

According to Eq. (7.7), the tensile strength and MOE of fiber and parenchy-
matous ground tissue were estimated as rf = 581.7 Mpa, Ef = 40.4 GPa,
rm = 19.0 MPa and Em = 0.22 GPa, respectively. These results suggest that fiber
plays the determinant role in tensile properties of bamboo and the ground tissue is
the low-strength-and-modulus matrix material.

7.3.2 Test on Single Fiber Bundle

To verify the feasibility of “Mixture law” used to calculate the strength and MOE of
fibers and analyze the load transmission function of ground tissue, the tensile
strength and MOE of fiber bundles separated from ground tissues are tested.

7.3.2.1 Material

Specimens were collected from the same culm walls where samples had been taken
for tensile tests on bamboo blocks. The fresh culm walls were immersed in water
7.3 The Mechanical Characteristics of the Components Bamboo 133

(a) (b)
Tensile strength (MPa)

MOE (GPa)
Fiber volume fraction Fiber volume fraction

Fig. 7.7 Relationships between fiber volume fraction (Vf) and bamboo tensile strength (a) and
MOE (b)

for one day and then split along the fiber direction into small strips. During the
splitting, some fiber bundles were detached from ground tissues and then pulled out
of the bamboo strips at an angle of 10° (the fiber bundle will be damaged or broken
if the angle is too large) between the detached fiber bundle and the bamboo strip.
The fiber bundles more than 60 mm long were chosen and air dried. Due to the
fragility of the small specimens and the difficulty of installation for the tensile test, a
special method was used with great care. The fiber bundle was fixed with strong
glue in a paper casing made of coordinate grid paper with two quadratic holes in the
middle of the two paper halves, as shown in Fig. 7.8. The paper was folded in the
middle, so that the fiber bundle was fixed with glue inside the paper casing except
for the middle part. This method protected the fiber bundle from being damaged or
broken during the installation on the tensile test machine. After being installed in
the clamps of test machine and before being loaded, the paper casing was snipped
through the middle part. Details are shown in Fig. 7.8. The tensile test conditions of
fiber bundles were the same as those for bamboo blocks. In total, 71 specimens of
fiber bundles were prepared.

Force
Paper Glue Fiber bundle Glue

Paper
20mm
Fiber Clamp
bundle
20mm
Cutting
line Scissors
20mm

Fold line Force

Fig. 7.8 Schematic representation of tensile test on bamboo fiber bundle


134 7 Mechanical Characteristics of Bamboo Structure and Its Components

Fig. 7.9 Stress–strain curve


of tensile test on bamboo fiber
bundle

Stress (MPa)
Strain

7.3.2.2 Test and Results

After the tensile test, the broken fiber bundle was stained with safranine, and then
air dried and embedded in resin. Cross sections were cut perpendicular to axial
direction of the fiber bundle. The cross area of fiber bundle as measured under
microscope with image analysis system. Therefore, tensile stress can be calculated.
Figure 7.9 shows an example of stress–strain curve where the part from the
beginning of loading to the breaking point was entirely linear. High rupture strength
and high MOE were observed.
However, there were two possibilities which might cause the fallibility of MOE
value calculated from the figure. First, the specimen, including the paper casing,
might have some movement inside the clamps of the test machine. Second, the
MOE of fiber bundle was much higher than that of the glue inside the paper casing
used for fixing fiber bundle. Thus, under tensile loading, the part of fiber bundle
covered by glue might have slight but not negligible strain, although no fiber bundle
was pulled out of the paper casing during the tests. Therefore, in order to estimate
the MOE of fiber bundle, according to “shear lag theory” of composites mechanics
[30], it was assumed that no movement of the specimen happened inside the clamps
of the test machine and the fiber bundle was covered uniformly by glue so that the
interface shearing stresses loaded on fiber bundle were distributed uniformly. As
shown in Fig. 7.10, Lg and L0 were noted as the lengths of the part of fiber bundle
inside glue and the middle part without glue, respectively. Af and Sf were noted as
the cross area and perimeter of fiber bundle, respectively, s as the shearing stress on
the interface of fiber bundle inside glue, r0 as the axial tensile stress on the middle
part of fiber bundle without glue. In order to keep the balance inside the glue, s and
tensile stress r(x) should meet the following equation:

Sf  x  s ¼ Af  rðxÞ ð7:8Þ

For the middle part of fiber bundle without glue, s and r0 should meet the
following equation:
7.3 The Mechanical Characteristics of the Components Bamboo 135

τ Fiber bundle Glue


σo

x dx
σ
σo
σ(x)

x
Lg Lo Lg

Fig. 7.10 Schematic representation of stresses loaded on bamboo fiber bundle during the tensile
test

Sf  Lg  s ¼ Af  r0 ð7:9Þ

Thus, the tensile stress r(x) loaded on position x far from the left end of the fiber
bundle should be
8 S xs
< Af ¼ Lg  r0
> ð0  x  Lg Þ
f x

rðxÞ ¼ r0 ðLg  x  Lg þ L0 Þ ð7:10Þ


>
: ð2Lg þ L0 Þx  r
Lg 0 ðLg þ L0  x  2Lg þ L0 Þ

The total length of fiber bundle under loading L = 2Lg + L0. Therefore, the total
length change of fiber bundle (DL) under the tensile stress should be

ZL ZLg ZL0
rf ðxÞdx rf xdx rf dx rf
DL ¼ ¼2 þ ¼ ðLg þ L0 Þ ð7:11Þ
E Lg E E E
0 0 0

where DL was measured as the displacement of the crosshead of the test machine,
Lg = L0 = 20 mm. Thus, according to Eq. (7.11), the MOE of fiber bundle can be
calculated.
The results of the tensile tests on fiber bundles are shown in Table 7.1. The
average tensile strength of bamboo fiber was 482.77 MPa, which was 18% less than
the value calculated by the tests on bamboo blocks. This is mainly due to the fibers
in bamboo blocks are embedded in ground tissue, which can pass loads and dis-
tribute the stresses loaded on vascular bundles. Such effectiveness is not possessed
by a single tissue. Besides, the damage on the surfaces of some fiber bundles may
be produced during separation process of fiber bundles. This could cause the low
136 7 Mechanical Characteristics of Bamboo Structure and Its Components

Table 7.1 The results of tensile tests on the separated bamboo fiber bundles
Mean Number of Standard Coefficient of
samples deviation variation (%)
Tensile strength 482.179 71 108.32 24.11
(MPa)
MOE (GPa) 33.853 71 8.16 22.44

tensile strength measured on bamboo fibers. The average MOE of bamboo fibers
was 33.85 GPa, which was also 16% less than the value measured by the test on
bamboo blocks. Such a low value may be caused by the discrepancy between the
hypothesis and the practice. During the tensile tests on fiber bundles, some slight
movements of the specimens and some slight shearing deformations of the glue
might have happened inside the clamps of the test machine. These could influence
the result of MOE of bamboo fiber measured on fiber bundles.

7.3.3 Analysis on Fracture Surface

Figure 7.11 shows the tension fracture pictures of three specimens with different Vf.
The corresponding stress–strain curves of the three specimens are shown in

Fig. 7.11 Tension fracture


pictures of 3 specimens
with different Vf. The
corresponding stress–strain
cures of the three specimens
are shown in Fig. 7.5
7.3 The Mechanical Characteristics of the Components Bamboo 137

Fig. 7.5. The fracture surface of the specimen with low Vf (11.5%) was trimmed,
which indicates that the specimen was brittle. The specimen with medium Vf
(20.4%) showed an uneven fracture surface, which was caused by the interaction of
parenchymatous ground tissue and fibers. When the Vf (49.3%) was high, fracture
surface of the specimen was split along the longitudinal direction, meaning that the
specimen had high toughness.
The micrographs of the fracture surfaces obtained by scanning electron micro-
scope are shown in Fig. 7.12. The left part of Fig. 7.12a shows the ground tissue and
the right part shows the fibers. Figure 7.12b, c shows the ground tissue and fibers,
respectively, with bigger magnification. In Fig. 7.12b, the incompact multilayer
thin-walled structure of ground tissue can be seen clearly. By axial tensile stress
action, the cell walls got fractured along the cross section. Like the matrix in artificial
composites, ground tissue can pass and buffer loads among vascular bundles.
However, under loads, the cell walls of ground tissue are prone to buckle. However,
in live bamboo tissue, the buckling resistance of the cell walls would be reinforced
by the effect of the interior static hydraulic pressure caused by water. The fracture
surface of fibers was dense but rough in general (Fig. 7.12a). In the individual fibers,
different layers were also rough (Fig. 7.12c). This indicates the low strength of the
interface bond between different fiber layers and among different fibers

Fig. 7.12 Micrographs of the fracture surfaces obtained by scanning electron microscope:
a surface including parenchymatous tissue (left) and sclerenchyma fibers (right); b fracture surface
of parenchymatous tissue; c facture surface of sclerenchyma fibers
138 7 Mechanical Characteristics of Bamboo Structure and Its Components

notwithstanding the high tensile strength of bamboo fibers. Furthermore, for the
high-modulus bamboo fibers, the multilayer structure can improve the flexibility.
Bamboo internodes tissues are arranged along the longitudinal direction and are
conglutinated together by the non-cellulose component in an effective but somehow
not very effective manner, which makes the strength of the interface much weaker
than that of bamboo longitudinal direction. However, it is the weakened interface
that causes the high fracture toughness in transverse direction of bamboo [31],
making bamboo tolerant toward gnawing of wild animals and knife wound.
Meanwhile, the deficiency of the weak interface strength of culm wall can be offset
by the nodes. The unique macrostructure of a large cavity surrounded by culm wall
and the increasing distribution density of vascular bundles from the inner to the
outer part of the culm wall along the radial direction make bamboo adapt to bending
load caused by snow or wind. This reflects the optimization of nature selection, and
the structure formed is in accordance with the principle of most efficiency with least
materials.

7.4 Difference of Structure and Strength Between


Internodes Part and Node Part of Moso Bamboo

7.4.1 Material and Method

Six moso bamboos were selected from the Anhui Province in China. They were 3
years old and about 15 m high. The diameter at breast height (dbh) varied from 110
to 125 mm and the average thickness of the culm wall was 12 mm at dbh. After
cutting, the bamboo stems were air dried in the laboratory. The moisture content of
the dried samples was about 12% and the average air dry density of internodes of
moso bamboo was 0.712 g/cm3.
The structure differences between internode and node of moso bamboo were
compared and analyzed specifically on the macroscopic forms and distribution of
fiber bundles. Mechanical properties of bamboo were tested according to the China
National Standard [32]. Tests conducted were longitudinal tensile strength, bending
strength, longitudinal shear strength, and compressive strengths across and parallel
to grain. Two types of moso bamboo were used, namely, that with culm intact (i.e.,
intact sample) and the other, with the innermost and outermost parts of culm
removed by double planing (i.e., planed sample). Samples were then divided into
four groups which were intact samples with nodes and without nodes, and planed
samples with nodes and without nodes. When preparing materials for samples with
nodes, the nodes were positioned at the center of the samples. Samples with and
without nodes were taken from neighboring sections of the same culm (Fig. 7.13),
thus, giving an almost uniform thickness of samples. All experiments were carried
out by a mechanical testing machine controlled by a computer. The speed of
crosshead was 2 mm/min. The laboratory temperature was between 15 and 18 °C,
and the humidity of the room was 60–65%.
7.4 Difference of Structure and Strength Between Internodes … 139

Internode Node Tensile sample Compressive sample

Shearing sample Bending sample

Fig. 7.13 Sketch of the testing samples produced from bamboo culm

7.4.2 Structure Comparison Between Internodes Part


and Node Part Culm

For internodes of culm, it consisted of three parts, bamboo skin, middle part, and
pith ring tissue. In the culm wall, the numbers of vascular bundles increase from the
inner to the outer part of the culm, while the ground tissues decrease on the
contrary. Bamboo cell between two nodes are strictly oriented axially (Fig. 7.14),
and transverse ray cells that normally present in wood structure have not been found
here.
The node part of bamboo culm consists of a sheath scar, nodal ridge, diaphragm,
and the intra-node between nodal ridge and sheath scar [1, 2, 17]. They strengthen
to the culm erecting and conduct water. Except for the vascular tissues of the culm’s
outmost layer which are broken when bamboo primary leaf falls, most vascular
bundles are curving at different degrees when they thread a node. One part of
vascular bundles bends inward or outward, then spread along the original longi-
tudinal direction. Another part of vascular bundles changes its former direction to
transversely spread in node part; still some vascular bundles come into the dia-
phragm around periphery of the diaphragm or interweaved reticular formation. The
vascular bundles in node are thickened, thus the thickness of bamboo node is
always greater than that of the adjacent internodes (Fig. 7.14).

7.4.3 Comparison of Mechanical Properties Between


Internodes Part and Node Part of Bamboo

7.4.3.1 Tensile Test Parallel to Grain

Bamboo blocks with node and without node were obtained from neighboring culm
and cleaved into 10 mm width bar first. And then, tensile samples were processed at
numerical control milling machine according to China national standard GB/T
15780-1995 (Fig. 7.15). The tensile samplers were divided into two types, two
groups in each type. For one type of samples, the innermost and outermost parts
140 7 Mechanical Characteristics of Bamboo Structure and Its Components

Fig. 7.14 Comparison between internodes culm and culm with a node: a radial section; b cross
section of a bamboo internodes A—A; (c) cross section of a bamboo node B—B

were milled, namely epidermis-planning sample, and the other type keeping intact
clum, namely non-epidermis-planning sample. There are 21–26 samples in each
group, and node is at the middle of sample.
Breaking tensile loads of intact samples with node and without nodes were
2435.7 and 2987 N respectively (Table 7.2); the existence of node decreased the
load by 18%. The tensile strengths for planed samples with node and without node
were 102.7 and 154.24 MPa, respectively. The strengths of samples with nodes
were lower than samples without nodes in a range of 33%. The failure surface of
samples with nodes showed that the break started from both sides of the nodes
(Fig. 7.16). This could be due to the discontinuity of vascular tissues when they
cross the node of the bamboo. Samples without nodes split longitudinally and
showed a typical ductile fracture.

7.4.3.2 Bending Test

The size of bending sample was 160 mm (Longitudinal)  10 mm (Tangential) 


Thickness. For the samples with node, the node was at the middle of sample.
A three-point bending test was carried out, and the distance between the two bases was
120 mm, as seen in Fig. 7.17. Because local tissues augment at bamboo node of
non-epidermis-planning samples, section modulus of sample with node could not be
calculated. Therefore, only the breaking bending load Fmax could be compared
between groups with node and without node. The samples whose epidermis were
planed could be compared the breaking bending stress r = 3FmaxL/(2th2) between

Thickness
R120

2
10

80 30 60 30 80

Fig. 7.15 Dimension of tensile test sample


Table 7.2 Mechanical properties; culm with and without node
Characters Processing Unit Samples without node Samples with node Ratio (without Result of ANOVA
Average Standard Variation Average Standard Variation node/with df F value Sig.
distance coefficient distance coefficient node)
(%) (%)
Tensile strength Intact N 2987.3 457 15.30 2435.7 386 15.80 1:0.82 45 19.11 ***
parallel to grain Planed MPa 154.24 13.92 9.00 102.7 15.81 15.40 1:0.67 50 164.76 ***
epidermis
Bending strength Intact N 850.8 115.3 13.60 1047.9 142.5 13.60 1:1.23 47 27.74 ***
Planed MPa 150.96 14.3 9.50 155.7 16.7 10.70 1:1.03 47 1.55 is
epidermis
Shearing strength Intact N 1954.3 436.6 22.30 2516.2 546.6 21.70 1:1.29 57 19.83 ***
parallel to grain Planed MPa 18.33 3.12 17.00 18.92 2.62 13.80 1:1.03 66 0.71 is
epidermis
Compressive Intact N 13.71 1.03 7.50 14.7 1.04 7.10 1:1.07 59 14.18 ***
strength parallel to Planed MPa 56.4 4.97 8.80 59.8 5.65 9.50 1:1.06 40 6.37 *
grain epidermis
Compressive Planed MPa 18.8 3.37 17.90 24.9 6.27 25.20 1:1.32 40 16.69 ***
strength across epidermis
7.4 Difference of Structure and Strength Between Internodes …

grain
141
142 7 Mechanical Characteristics of Bamboo Structure and Its Components

Fig. 7.16 Tensile failure of a specimen without a node and a specimen with a node

groups with node and without node. The load bearing capacity is different along
different directions of sample, and generally the maximum force along radial direction
to inner Fmax,radial,inner is lower than that along radial direction to outer Fmax,radial,outer.
Prediction shows that for the intact sample with Tangential width 
Thickness = 10 mm  10 mm, the strength parameters of bending sample without
node meet the following relation:

rmax;radial;inner þ rmax;radial;outer
rmax;tangential  ð7:12Þ
2

The maximum loads of bending sample with node meet the following relation:

Fmax;radial;inner þ Fmax;radial;outer
Fmax;tan gential  ð7:13Þ
2

Here, the direction of bending load on the sample was along tangential according
to GB/T 15780-1995 [32].
The descriptive statistics and variance analysis of the test results are shown in
Table 7.2. Maximum bending loads of intact samples with nodes and without nodes
were 1047.9 and 850.8 N, respectively, a difference of 23%. The bending strengths
for planed samples with nodes (155.7 MPa) and without nodes (150.96 MPa) were
similar. This suggests that the anti-bending ability of the node part in the intact
culm is enforced by swelling of local tissues at the bamboo node. Thus, since nodes
do not cause any negative effects on bending strength of planed samples, they
should not be considered as defects in industrial utilization.

Fig. 7.17 Bending failures: a culm without nodes and a culm with a node
7.4 Difference of Structure and Strength Between Internodes … 143

7.4.3.3 Shearing Test Parallel to Grain

The dimension of shearing test sample was shown in Fig. 7.18. Because of augment
of local tissues at bamboo node of non-epidermis-planning samples, only the
maximum load Fmax could be compared between groups with node and without
node. The epidermis-planed sample could be compared with the shearing stress
r = Fmax/(bl) between groups with node and without node, while l and b were
length and width of shearing section, respectively.
Descriptive statistics and variance analysis of the shearing test data were listed in
Table 7.2. Maximum shearing loads of intact samples with and without nodes were
2516 and 1954 N, respectively; the shearing strength of samples with nodes was
higher by 29%. For planed samples, shearing strengths were similar, i.e., 18.9 and
18.3 MPa for nodes and without nodes, respectively. This showed that nodes could
enforce the anti-shearing ability of bamboo culm by swelling of local tissues at
nodes. For planed samples, nodes did not reduce the shearing strength. Tearing
failures of the samples are shown in Fig. 7.19. The periphery decumbent fiber
bundles were not found in the culm; only a few fiber bundles were detected at the
node. Failure plane in shear test is generally caused by slipping of the two planes
comprising ground tissues of fibers. The snipping of fiber bundles in the cross
direction did not contribute to the failure. Hence, effects of nodes on shear strength
are almost negligible for the processed samples.

7.4.3.4 Compressive Test Parallel to Grain

The dimensions of intact compressive sample parallel to grain are


30 mm (Longitudinal)  10 mm (Tangential)  Thickness. Because local tissues
augment at bamboo node of non-epidermis-planning samples, only the maximum
load Fmax when sample failures could be compared between groups with node and
without node. The dimensions of samples whose epidermises were planed are
20 mm (Longitudinal)  10 mm (Tangential)  Thickness, because of the smaller
thickness, about 6 mm. And the compressive strength, r = Fmax/(bt), can be
compared between groups with node and without node, while b and t were width
and thickness of samples, respectively.

Fig. 7.18 Dimension of


10

shearing test sample


9
15

20
10

11
144 7 Mechanical Characteristics of Bamboo Structure and Its Components

Fig. 7.19 Shear failure for culm a without node and b with node

Descriptive statistics and variance analysis of the experimental data were listed
in Table 7.2. Concerning the intact samples, the maximum compressive load par-
allel to grain of samples with node and without node was 14.7 and 13.7 N,
respectively. Compressive strength parallel to grain for planed epidermis samples
with node and without node was 59.8 and 56.4 MPa, respectively. Apparently, no
matter whether bamboo culm epidermis was planed or not, the compressive strength
or load of samples with node were slightly larger than that of samples without node.
The reason of this difference could be deduced from the higher proportion of
bamboo fibers in the node area (Fig. 7.14).

7.4.3.5 Compressive Test Across Grain

Because the load applied on the sample in the transverse compression test is along
radial direction of sample, the epidermis of bamboo samples must be planed. The
dimension of sample was 20 mm (Longitudinal)  20 mm (Tangential)  6 mm
(Thickness). The transverse compressive strength of samples with node and without
node was 24.9 and 18.8 MPa, respectively. The transverse compressive strength of
samples with node was 32% higher than that of samples without node. The dif-
ference between the two kinds of samples was significant. The reasons of this
difference could also be explained by a higher proportion of bamboo’s fibers in the
node area.
To sum up, the experiments of the bamboo mechanical properties have indicated
that nodes have not given any negative effects on tensile strength, bending strength,
and compressive strength parallel or transverse to grain. On the contrary, nodes
have the positive effect on the mechanical properties at different degrees. For intact
samples, nodes enhanced bending maximum load, shearing maximum load, and
compressive maximum load by 23, 19, and 7%, respectively. But the enhancement
effect initiated by nodes was not obvious for epidermis-planning samples. However,
negative effects of nodes on the tensile strength parallel to grain, for the both types
7.4 Difference of Structure and Strength Between Internodes … 145

of samples, i.e., intact and epidermis-planning samples, are significant, in which the
resistance decreased by 18 and 33%, respectively.
Bamboo is a biological compound material reinforced by long natural fibers
which give it a high resistance. Different extents in bending and some discontinuity
of the vascular tissue were observed at bamboo nodes. However, tumefaction of
bamboo nodes could increase the loading area and vascular bundles could be
thickened, circuitous, flexuose, and entwined. So, the ability to anti-bending and
anti-shearing of the culm under transverse load is increased by bamboo nodes.
However, the tensile load of the culm was relatively low during growth of bamboo.
Therefore, enlargement of nodes and the development of nodes could improve
tensile strength in the longitudinal direction.

References

1. Liese W (1998) The anatomy of bamboo culms. International network for bamboo and rattans
(INBAR), Beijing, Tech rep no. 18
2. Mohmod AL, Mustafa MT (1992) Variation in anatomical properties of three Malaysian
bamboos from natural stands. J Trop For Sci 5(1):90–96
3. Chuma S, Hirohashi M, Ohgama T et al (1990) Composite structure and tensile properties of
mousou bamboo. Zairyou 39:847–851 (in Japanese)
4. Inokuchi Y, Fushitani M, Chuma S et al (1997) Effects of volume fraction of bundle sheath on
the vibrational properties of bamboo. Mokuzai Gakkaishi 43(5):391–398 (in Japanese)
5. Inokuchi Y, Fushitani M, Kubo T et al (1999) Effects of water extractives on the
moisture-content dependence of vibrational properties of bamboo. Mokuzai Gakkaishi 45
(2):77–84 (in Japanese)
6. Inokuchi Y, Fushitani M, Kubo T et al (2002) Effects of volume fraction of bundle sheath and
water extractives on bending creep behavior of bamboo under changing moisture conditions.
Mokuzai Gakkaishi 48(6):413–424 (in Japanese)
7. Sui S, Li R (2003) A study of structure and performance of bamboo fibers. J Text Res 24
(6):535–537 (in Chinese)
8. Ahmad M, Kamke FA (2005) Analysis of calculate bamboo for structural composite
materials: physical and mechanical properties. Wood Sci Technol 39(6):448–459
9. Obataya E, Kitin P, Yamauchi H (2007) Bending characteristics of bamboo (Phyllostachys
pubescens) with respect to its fiber–foam composite structure. Wood Sci Technol 41:385–400
10. Deshpande AP, Rao MB, Rao CL (2000) Extraction of bamboo fibers and their use as
reinforcement in polymeric composites. J Appl Polym Sci 76:83–92
11. Jiang ZH, Sun ZJ, Ren HQ (2006) Application of advanced bio-composites in wind blades.
Acta Mater Compos Sin 23(3):127–129 (in Chinese)
12. Sun ZJ, Chen Q, Jiang ZH (2008) Processing and properties of engineering bamboo products.
Acta Mater Compos Sin 25(1):80–83 (in Chinese)
13. Jain S, Kumar R, Jindal UC (1992) Mechanical behaviour of bamboo and bamboo composite.
J Mater Sci 27:4598–4604
14. Zhao QS (1995) Industrial utilization of bamboo in China. China Forestry Press, Beijing, p 3
15. Nugroho N, Ando N (2000) Development of structural composite products made from
bamboo I: fundamental properties of bamboo zephyr board. J Wood Sci 46:68–74
16. Nugroho N, Ando N (2001) Development of structural composite products made from
bamboo II: fundamental properties of laminated bamboo lumber. J Wood Sci 47:237–242
17. Jiang ZH (2007) Bamboo and Rattan in the World. China Forestry Press, Bejing, pp 2–4
146 7 Mechanical Characteristics of Bamboo Structure and Its Components

18. Lakkad SC, Patel JM (1980) Mechanical properties of bamboo, a natural composite. Fiber Sci
Technol 14:319–322
19. Mohmod AL, Ariffin WTW, Ahmad F (1990) Anatomical features and mechanical properties
of three Malaysian bamboos. J Trop For Sci 2(3):227–234
20. Xian X, Xian D (1990) The relationship of microstructure and mechanical properties of
bamboo. J Bamboo Res 3:10–23
21. Mohmod AL (1993) Effects of age and height of three bamboo species on their machining
properties. J Trop For Sci 5(1):90–96
22. Yao X (1993) The microstructure of main Chinese bamboos. China Forestry Press, Beijing,
pp 6, 18
23. Suzuki K, Itoh T (2001) The changes in cell wall architecture during lignification of bamboo,
Phyllostachys aurea Carr. Trees 15:137–147
24. Yu H (2003) Study on property of bamboo culm. World Bamboo Rattan 1:5–10
25. Kamruzzaman M, Saha SK, Bose AK et al (2008) Effects of age and height on physical and
mechanical properties of bamboo. J Trop For Sci 20(3):211–217
26. Yu HQ, Jiang ZH, Hse CY et al (2008) Selected physical and mechanical properties of moso
bamboo (Phyllostachys pubescens). J Trop For Sci 20(4):258–263
27. Liese W (1998) The anatomy of bamboo culms. International network for bamboo and rattans
(INBAR), Beijing, Tech rep no. 18
28. Roelofsen PA (1959) The plant cell-wall. Gegruder Borntraeger, Berline-Nikolassee pp 126–
189
29. Jones RM (1975) Mechanics of composite materials. Scripta Book Company, USA, pp 355.
ISBN 0-07032790-4
30. Qiao SR (1997) Micromechanical properties of composites. Northwestern Polytechnical
China, University Press, p 187. ISBN 978-7-5612-0913-4
31. Gordon JE (1984) The new science of strong materials or why you don’t fall through the
floor. Princeton University Press, USA, p 287. ISBN 0-691-02380
32. China National standard (1995) Testing methods for physical and mechanical properties of
bamboo GB/T 15780. China National Institute of Standardization, Beijing
Chapter 8
Interlaminar Fracture Properties
of Bamboo

Abstract In this chapter, Mode I, II, III interlaminar fracture properties of bamboo
and the fracture mechanism are studied by double cantilever beam (DCB) method,
end-notched flexure beam (ENF) method, modified split cantilever beam (MSCB),
and split cantilever beam (SCB) method respectively. The results show that: (1) the
Mode I, Mode II, and Mode III interlaminar fracture toughness (GIC, GIIC, and
GIIIC) of bamboo are the basic attributes of bamboo that represent the capacity of
bamboo to resist the propagation of cracks; (2) On the Mode I fracture surfaces of
bamboo, smooth fibers, and plane ground tissue are found, which indicate that the
longitudinal interface strength was weak among bamboo cells. The Mode II fracture
surfaces is rougher, and ground tissue is characterized by hackle shearing defor-
mation, which indicates that a large amount of fracture energy would be absorbed
by the shear deformation of ground tissue, and GIIC  2.5 GIC; From the Mode III
fracture surface of bamboo, it can be seen that the resistance that hindered the
propagation of interlaminar crack is contributed by the transverse shear strength of
ground tissue cell wall and interface strength, and GIIIC  4.0 GIC; (3) The study on
the toughness contribution of bamboo node to the interlaminar fracture toughness of
bamboo and the mechanism show that GNode IC  2.9 GInternode
IC IIC  1.3
, GNode
GInternode
IIC , and G Node
IIIC  2.7 G Internode
IIIC , thus bamboo node can contribute a lot to
hinder the interlaminar fracture of bamboo.

8.1 Introduction

Through millions of years’ evolution, bamboo has formed a particular structure in


order to bear the bending load mainly caused by snow or wind. This structure
contributes to its high transverse bending strength and toughness. In contrast, the
anti-cleaving and anti-shearing strength along the grain of bamboo are relatively
low. Therefore, bamboo often cracks along the grain direction while drying. Many
studies have reported on the general mechanical properties of bamboo [1–3], but
researches on fracture characteristics of bamboo are few.

© Springer Nature Singapore Pte Ltd. 2018 147


Z. Shao and F. Wang, The Fracture Mechanics of Plant Materials,
https://doi.org/10.1007/978-981-10-9017-2_8
148 8 Interlaminar Fracture Properties of Bamboo

The unique tissue of bamboo culm is the material basis of perfect mechanical
characteristics. In view of macro-mechanical behavior, bamboo is a typical unidi-
rectional long-fiber-reinforced biocomposite. It exhibits significant anisotropy in
strength and stiffness corresponding to longitudinal, radial, and transverse direc-
tions. The tensile strength along the longitudinal direction can be as large as 150–
300 MPa, but the transverse tensile strength and shear strength parallel to grain are
much lower. Zeng et al. [1] reported that the tensile strength perpendicular to grain
was only about 2% of the one parallel to the grain. Therefore, once there is crack,
the propagation of delamination is controlled by the interlaminar fracture toughness
but not by strength. The interlaminar fracture toughness of bamboo represents the
ability of bamboo to resist the crack propagation along the grain. Even the crack or
lacuna perpendicular to the grain of bamboo may easily be transformed to develop
along the grain direction under loads, and thereby influencing the mechanical
behavior.
Bamboo and wood are two of the few natural composite materials that can be
used directly as structural materials. When used as structural materials, the inter-
laminar fracture toughness as well as the strength of bamboo and wood is an
important basis for the safety design of the structure. The interlaminar crack of
composites could be classified into three categories: Mode I––opening mode, Mode
II––sliding mode, and Mode III––tearing mode according to the load and fracture
characteristics.
In this chapter, the Mode I, Mode II, and Mode III interlaminar fracture
toughness of bamboo are tested, and the fracture mechanisms are studied.

8.2 Mode I Interlaminar Fracture of Bamboo

8.2.1 Theory

The Mode I interlaminar fracture toughness, i.e., the critical value of energy release
rate, GIC, for delamination growth of composite material is often measured by
symmetric bending tests using double cantilever beam (DCB) specimen and cal-
culated with the compliance of the specimen [4]. Triboulot et al. [5] measured wood
fracture toughness for TL crack opening mode using DCB method, and compared it
with finite element analysis (FEA). The results tallied well with each other. Now,
the DCB method has been recommended by American Society for Testing and
Materials (ASTM) as a standard measurement for Mode I interlaminar toughness of
unidirectional fiber-reinforced composites [6].
For DCB specimen under opening loads, the strain energy U and its release rate
GIC can be calculated as follows:
8.2 Mode I Interlaminar Fracture of Bamboo 149

1 1
U ¼ Fd ¼ F 2 C ð8:1Þ
2 2

@U Fcr2 @C
GIC ¼ ¼  ð8:2Þ
@A 2b @a

where d is load point deflection, C = d/F the compliance of DCB specimen.


Therefore, it is only necessary to measure the differential increase in compliance
(@C) depending on the differential increase in crack length (@a), and then the GIC
can be calculated by Eq. (8.2).

8.2.2 Materials and Methods

The moso bamboo (Phyllostachs pubescens) used in the tests was 4 years old and
came from Lujiang, Anhui Province, China. The total height was about 15 m, and
the diameter at breast height was 125 mm. The specimens were collected from the
internodes sections located at a height between 0.6 and 6 m of the bamboo. After
air dried, the specimens were cut parallel to grain into small pieces. The moisture
content of the specimens was about 11% during the experiment.
The bamboo DCB specimens were taken from three different heights of the
culm, i.e., section 7 (height of 1.3 m), section 14 (height of 3 m), and section 19
(height of 5 m), respectively. Five or six DCB specimens per internode section
were prepared, total number  5 recommended by ASTM D5528 [6]. The
dimensions of the specimens are e = 20 mm, W = 200 mm, h = 20 mm, as shown
in Fig. 8.1. Generally, the initial crack of the wood specimen was first cut by a band
saw and then extended by a razor blade [7, 8]. In our case, to simulate a naturally
sharp crack, the crack was cleaved along the middle level parallel to the grain by a
knife as seen in Fig. 8.2. For bamboo internode specimen, the length of the initial
crack was controlled about 40 mm from the center of the loading holes; for a
bamboo specimen with node, the initial crack tip was controlled 2 mm from node to
ensure that the crack would spread through the node by on–off test. When the initial
crack was cleaved, a red dyeing reagent that had been prepared in advance was used
to dye the crack tip before the knife was pulled out, in order to obtain accurate
initial crack length.
Tests were performed on a universal metal material testing machine
(WDW-100). The test room temperature was 18 °C, and the humidity was 60%.
The DCB specimen was connected with steel U hook by steel pin (Fig. 8.3) with a
crosshead speed between 1 and 5 mm/min. At the beginning of the test, a low
crosshead speed was used, while the cantilever beams were short in order to
facilitate identifying the initial movement of the delamination. The crosshead speed
was increased when the cantilever beams were relatively long [4, 6]. A curve of the
applied load versus opening displacement (F-d) was automatically recorded by a
150 8 Interlaminar Fracture Properties of Bamboo

Force Loading hole Crack

e 0
w
Force

Fig. 8.1 Sketch of double cantilever beam (DCB) specimen. w = 200 mm (longitudinal
direction); e = 20 mm; h = 20 mm (tangential direction); b = 8–10 mm (natural thickness, radial
direction); diameter of loading hole = 5 mm; a0 = 40–50 mm

Fig. 8.2 Schematic diagram Force


of crack cleaving
Knife .

Specimen
of bamboo

computer during the test. From start loading to maximum load Fmax, the F-d curve
remained linear until the specimen started crazing. Once crazing, the bearing
capacity of the specimen decreased sharply and the crack propagation parallel to
grain was unstable. The top point of F-d curve represented the critical point of rapid
cracking and Fmax was taken as the critical load value Fcr. After the load went
down, the test machine was stopped immediately and the recorded data was stored.
The crack tips were marked on both the sides of the specimen by means of an
optical microscope. Then, the specimen was unloaded and reloaded and the same
procedure was repeated until the specimen was fractured completely. The specimen
was taken from the test machine and the crack length after each increment of
delamination crack growth was measured. As the resistance-arresting crack prop-
agation of the outer layer is lower than that of the inner layer of bamboo, the
delamination crack lengths on each side were unequal and the crack of the outer
8.2 Mode I Interlaminar Fracture of Bamboo 151

Fig. 8.3 Sketch of U hook


for testing Force

DCB specimen

Crack

Force

layer was 5–6 mm longer than that of the inner layer. Thus, the average crack
length of two sides was taken as the actual crack length.
F-d curves corresponding to different crack lengths obtained from one of the
DCB specimens of this study are shown in Fig. 8.4, which shows the typical F-d
curve of DCB specimen. With increasing crack length, the slope of the linear part of
F-d curve decreased. The reciprocal of the slope is the corresponding compliance
(Ci) of the DCB specimen with a certain crack length (ai). The relationship between

Fig. 8.4 Typical F-d curve F


of DCB specimen. a1
a1 = 45 mm, a2 = 60 mm,
a3 = 72 mm, a4 = 89 mm, a2
a5 = 114 mm, a6 = 141 mm, a3 a
a4 F
and a7 = 173 mm are crack
lengths a5
a6 a7
152 8 Interlaminar Fracture Properties of Bamboo

Fig. 8.5 Relationship 400


between compliance (C) and C=0.0016a2.3814
crack length (a)
R2 =0.996
300

C (mm/kN)
200

100

0
0 40 80 120 160 200
a (mm)

C and a of this specimen is shown in Fig. 8.5. A power law relationship between
C and a could be obtained as follows:

C ¼ qam ð8:3Þ

where q and m are the fitting coefficients of the a-C curve

8.2.3 Results and Analysis

The total mean value and standard deviation of GIC in this study were 358.08 ±
61.18 J/m2 (Table 8.1). The variance analysis of GIC in different sections, i.e., at
different heights, is shown in Table 8.2 and no significant difference was found
among different heights.
In general, along the longitudinal direction, bamboo density increases from the
bottom to the top [9]. In this study, the mean air-dry densities of the samples located
at section 7 (height of 1.3 m), section 14 (height of 3 m), and section 19 (Height of
5 m) were 0.636, 0.712, and 0.729 g/cm3, respectively. The increased density could

Table 8.1 Statistical analysis of Mode I interlaminar fracture toughness (GIC) for different
sections of the bamboo
Section Location Number Number of Mean Standard Coefficient
(height, of measurements GIC (J/ deviation of variation
m) samples m2) (%)
7 1.3 6 41 371.17 52.39 14.44
14 3.4 6 52 351.55 64.20 18.02
19 5.1 5 41 367.96 64.07 18.02
Total 17 134 358.08 61.18 16.88
8.2 Mode I Interlaminar Fracture of Bamboo 153

Table 8.2 Variance analysis of Mode I interlaminar fracture toughness (GIC) for different sections
of the bamboo
Sources of variation SS df MS F-value F0.05 Significance
Between 1414.9 2 707.5 0.1914 3.0653 not significant
Within 484189.1 131 3696.1
Sum 485,604 133

be explained by the increased bundle-sheath proportion as from the bottom to the


top of bamboo the culm diameter decreased, but the amount of bundle-sheath varied
little. Density is always important to characterize the property of cellular materials.
However, for Mode I interlaminar fracture toughness (GIC) of bamboo, no signif-
icant difference was found among different heights, i.e., among different densities.
This may be explained because the resistance-arresting crack propagation was
governed by the properties of the interface between bamboo cells or cell walls.
Similarly, along the radial direction, bamboo density increased from the inner to the
outer part due to the increasing bundle-sheath proportion.
Accordingly, mechanical strength should increase from the inner to the outer
part. Whereas, as mentioned above in the Sect. 8.2.2, during the tests it was found
that the delamination crack lengths on each side of the samples were unequal and
the crack of the outer layer was slightly longer than that of the inner layer. It
indicates that the resistance-arresting crack propagation of the outer layer was
slightly lower than that of the inner layer. As the ground tissue absorbs more energy
than the bundle-sheath due to its higher flexibility during the crack propagating,
higher proportion of bundle-sheath and less and incomplete continuous ground
tissue in the outer layer caused lower interlaminar strength. This further confirms
that the resistance-arresting crack propagation was controlled by the interlaminar
strength between fibers and ground tissue.
Figure 8.6 shows the relationship between Mode I interlaminar fracture tough-
ness (GIC) and crack length (a). The value of GIC decreased slightly as the crack
length increased, which is similar to the results found on polymer matrix com-
posites [10]. This could be mainly due to the nonlinear phenomenon caused by a
large deflection of the double cantilever beam, i.e., the so-called
geometry-nonlinearity, as crack increases. Therefore, for materials with
low-flexural modulus or high interlaminar fracture toughness, in order to minimize
the error caused by the geometry-nonlinearity, the thickness (h) of the specimens
should meet the following criteria recommended by American Society of Testing
Materials [6], to ensure the linearity of the response of load versus displacement.
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
h  8:28 3 GIC a2 =E11 ð8:4Þ

where E11 is the modulus of elasticity of the specimen in the fiber direction and in
this study this value was replaced by the elastic modulus measured on each section
of the bamboo by means of a bending test according to the National Standard of
China [11]. The specimen dimension in this study met the above condition of
154 8 Interlaminar Fracture Properties of Bamboo

800

GIC (J/m2 )
600
358.08
400
200
0
20 60 100 140 180
a (mm)

Fig. 8.6 Relationship between GIC and a

Eq. (8.4). Therefore, although the value of GIC measured on DCB specimen was
affected to a certain extent by the crack length, the statistical results show that the
change of GIC due to the different crack lengths was not significant in this study.
Thereby, the interlaminar fracture toughness of Mode I can be regarded as a basic
characteristic of bamboo material representing the resistance-arresting crack
propagation.

8.2.4 Analysis of Fracture Surface

Figure 8.7 shows the micrographs of Mode I interlaminar fracture surface obtained
by scanning electron microscope. It can be seen from Fig. 8.7a that in the Mode I
crack expanding area, there are the characteristics of smooth fiber and a matrix
(ground tissue) neatly parallel to the stem axis. Figure 8.7b shows the parenchyma
tissue of Zone A in Fig. 8.7a after 1489 magnification. Viewed from interlaminar
fracture surface, the contour of the ground tissue generally maintains integrity. The
surface is smooth or with few slice debris. Occasionally, tearing and stripping of the
whole cell wall happened. This indicates that the crack propagation among ground
tissues develops along the middle lamella or the primary cell wall. Figure 8.7c
shows the fibers of zone B in Fig. 8.7a after 14,649 magnification. It shows the
integrity of bamboo fiber and the smooth interlaminar fracture surface only with a
few traces of slight tearing in the middle lamella or primary cell wall. No tearing or
stripping was found on the secondary cell wall. It indicates that the Mode I crack
among the bamboo fibers develops through the interface between fibers, which
means that the interlaminar strength of bamboo is weak.
Bamboo is a kind of high anisotropic biomaterial with most of its tissues
arranged along the longitudinal direction and conglutinated together by the
non-cellulose component. This unique structure causes low interface strength.
However, it is just because of the weak interface strength that bamboo has high
toughness in the transverse direction, which makes the bamboo adapt to bending
load caused by snow or wind. This could be explained by the toughening mech-
anism due to the weak interface, which was first described by Cook et al. [12] and
has been widely used in designing composite material.
8.2 Mode I Interlaminar Fracture of Bamboo 155

Fig. 8.7 a Image of Mode I


interlaminar fracture surface
of bamboo obtained by
scanning electron microscope;
b 148magnification image
of zone A in (a);
c 1464magnification image
of zone B in (a)
156 8 Interlaminar Fracture Properties of Bamboo

8.3 Mode II Interlaminar Fracture of Bamboo

8.3.1 Test and Specimens

Mode II interlaminar fracture toughness GIIC of unidirectional fiber-reinforced


composites is generally tested by ENF (end-notched flexure) specimens. This
method was first proposed by Barlle and Foschi [13] to measure the fracture
toughness of wood. Later, ENF specimen was applied to test the fracture toughness
of carbon fiber-reinforced unidirectional composite by Rusell and Street [14].
Carsson et al. [15] researched stress field of ENF specimen by means of finite
element analysis (FEA) and the numerical analysis result showed that the test
results of ENF were pure Mode II fracture. In recent years, Hessein et al. [16],
Brunner et al. [17], and Sham et al. [18] also studied the ENF specimen by theo-
retical and experimental analysis. Figure 8.8 showed the ENF specimen and test
device.
The moso bamboo (P. pubescens) used in tests was 4 years old, from Lujiang,
Anhui Province, China. The total height was about 15 m, and the diameter at breast
height was 125 mm. The specimens were collected from the internode sections
located at the height between 0.6 and 6 m of the bamboo. After dried in the constant
temperature humidity chamber (27 °C, 60% humidity) about 20 days, the speci-
mens were cut into small pieces along the grain. The moisture content of the
specimens was about 11% during the experiment.
ENF specimens were chosen from the 8th, 11th, 14th, 15th, 18th, and 20th
section of bamboo culms respectively. Five–six pieces ENF specimen were made
from each internode section. In terms of the dimensions of specimens, such ENF
specimens were divided into two groups: (1) L = 60 mm, d = 2h = 15 mm
(T direction); (2) L = 75 mm, 2h = 17 mm (T direction). Width b was the natural

Fig. 8.8 ENF specimen and F


sketch of mechanics analysis L L
a
I = bh3 /12
I 0 = 8I 2h

A B C D

F/2 F/2
F/4
L L
MQ a

QM
F/2 F/4
8.3 Mode II Interlaminar Fracture of Bamboo 157

thickness of bamboo culm. Instead of being cut by a band saw then extending by a
razor, the pre-crack was cleaved along middle level parallel to grain by a knife to
form a nature-sharp (Fig. 8.2) and the crack body was TL system. The initial crack
length of each specimen was different and to avoid the influence of stress con-
centration caused by supports and loading roller on the crack tip, the initial crack
length/half span should be taken within the range of 0.25 < a/L < 0.75.
Mode II interlaminar crack was easy to propagate parallel to grain of the bamboo
outer layer. However in the inner layer, the resistance was larger and crack fragment
appeared at an angle of 45° to middle line at the surface of inner layer (Fig. 8.9). It
is because the surface of inner layer has a layer of pith ring constituted by stone
cells, which are crisp and isotropic, and it leads to the result that the pith-ring
fractures along the direction of maximum tension stress under the pure shearing
state (Fig. 8.10). Therefore, it is difficult to mark the position of crack tip in inner
layer. So different from Mode I interlaminar fracture tests that every specimen could
be loaded with repeated loading and the propagated length of crack could be
marked more than once, each specimen should be used only once in Mode II
interlaminar fracture test and a lot of specimens were needed during the test. In this
test, the number of specimen was 60 and there were 43 specimens that were
effective except the specimens diverging from the neutral layer by ±0.5 mm or
more.
Tests were performed by a computer-controlled testing machine. Temperature
was 18 °C and the humidity was 60% in laboratory. The load speed was 2 mm/min
and the curve of load–load point deflection (F-d) was recorded by computer
automatically during the test. The F-d curve of Mode II interlaminar fracture test
was shown in Fig. 8.11. At the initial stage of loading, the F-d curve kept straight,
then as the load increased, the initial crack began to crack. At this moment, because

Fig. 8.9 Crack trace of mode


II on the surface between out
layer (a) and inner layer (b)
158 8 Interlaminar Fracture Properties of Bamboo

Fig. 8.10 The fractures of


the pith ring along the
direction of maximum tensile
stress under the pure shearing
state

the rigidity of specimen changed, the slope of F-d curve changed as well and the
new crack surface was rougher. Because of the friction on the crack surfaces, the
crack propagated stably and slowly in the initial stage and until the maximum load,
the unstable propagation of crack began. Since in the ENF tests of bamboo, the
crack would always propagated slowly in different extent before the specimens
fractured, there was no obvious sign of unstable propagation. Thus, the maximum
load was not the critical load when unstable propagation of crack began. Then, the
matter was how to determine the critical load. To the Mode I interlaminar fracture
test, the critical load was the corresponding load when the propagated length
Da equaled to 2% of the initial crack length.
However, in practical test the curve we obtained was F-d curve but not F-D
a curve, so the corresponding load (Da/a = 2%) should be determined on F-d
curve. As theoretical analysis and approximation [19], it was found that the slope of
secant line of the corresponding load on the F-d curve decreased 5% compared with
the slope of the line segment of the F-d curve before the propagation began, so
graphing method was used to determine the critical load on the F-d curve. The
method to determine the critical load is approved by ESIS standard [20, 21]) and
applied by the industry standard in China [22] and used by Fonselius and Riipola
[23], Demorais et al. [24], Kuboki et al. [25] and Hiroshi [26] to analyze the ENF
specimen. In this section, the method was applied.
The detail method was: at first the maximum load Fmax was read; then a straight
line (tangential line) was drawn through the points of the load–displacement (F-d)
curve having F0.1 = 0.1 Fmax and F0.5 = Fmax; third, a secant line was drawn
corresponding to a 5% reduction in the slope; at last the critical load Fcr of initial
cracking was the point of intersection of the secant and the F-d curve. If the
intersection point was beyond the maximum load Fmax, then Fmax was taken as Fcr.
8.3 Mode II Interlaminar Fracture of Bamboo 159

Fig. 8.11 The load–load


point deflection curve (F-d
curve) of ENF test Fmax

Fcr

F (kN)

k=0.95tan

(mm)

8.3.2 The Data Processing Methods

8.3.2.1 Experiment Parameters Substitution Method

The displacement of support A and D to section B in the middle of span (which is


supposed to be fixed) in Fig. 8.8, due to the simple elastic beam theory and method
of superposition, can be calculated by the following formula:
F
 L3 FL3 FL3
dA ¼ 2 ¼ ¼ ð8:5Þ
3EI0 6EI0 48EI

dD ¼ dCD
D þ dC þ hC  a
BC BC
! !
 a3
F F
 ðL  aÞ3 Fa
 ðL  aÞ2 F
 ðL  aÞ2 Fa
 ðL  aÞ
¼ 4
þ 2
þ 2
þ 2
þ 2
a
3EI 3EI0 2EI0 2EI0 EI0
FðL3 þ 3a3 Þ
¼
48EI
ð8:6Þ

As the rigidity of the left half beam and the right half beam is different, the
displacement of B is not the maximum displacement and there exists a corner on
section B. So we can obtain the displacement at the loading point, d, as follows:
160 8 Interlaminar Fracture Properties of Bamboo

F
 L3 FL3 FL3
dA ¼ 2
¼ ¼ ð8:7Þ
3EI0 6EI0 48EI

dD ¼ dCD
D þ dC þ hC  a
BC BC
! !
 a3
F F
 ðL  aÞ3 Fa
 ðL  aÞ2 F
 ðL  aÞ2 Fa
 ðL  aÞ
¼ 4
þ 2
þ 2
þ 2
þ 2
a
3EI 3EI0 2EI0 2EI0 EI0
FðL3 þ 3a3 Þ
¼
48EI
ð8:8Þ

1 1
d ¼ ðdA þ hLÞ þ ðdD  hLÞ ¼ ðdA þ dD Þ
2 2
  ð8:9Þ
1 FL3 FðL3 þ 3a3 Þ Fð2L3 þ 3a3 Þ
¼ þ ¼
2 48EI 48EI 96EI

where I = bh3/12 is the axis moment of inertia of cross sectional in the crack region,
I0 = b(2h)3/12 = 8I is the axis moment of cross sectional in the crack free region,
E is the modulus of elasticity in the fiber direction, b is the thickness of specimen.
The loading point compliance C is defined as follows:

d 2L3 þ 3a3
C¼ ¼ ð8:10Þ
F 96EI

So strain energy U and strain energy release rate GII are represented as

1 F 2 ð2L3 þ 3a3 Þ
U ¼ Fd ¼ ð8:11Þ
2 192EI

@U 9Fcr a2 d
IIC ¼
GTest ¼ ð8:12Þ
b@a 2bð2L3 þ 3a3 Þ

where d is the displacement of the beam middle point corresponding to the critical
load Fcr when the initial propagation of interlaminar crack begins. As GTest IIC can be
calculated by the experiment parameters straightly and there is no need to test the
longitudinal modulus of elasticity E in advance, it is called experiment parameter
substitution method. What to be specified is that as unidirectional long-fiber-
reinforced composites, the crack propagation of bamboo ENF specimen will not
deviate from the initial crack direction during the test, therefore there is no need to
correct it by introducing additional crack length [27].
8.3 Mode II Interlaminar Fracture of Bamboo 161

8.3.2.2 Timoshenko Beam Theory Method

When the ratio of span (L) and width (h) is small, the influence of transverse shear
is significant. Due to Timoshenko beam theory [19–21], the shear modulus G13
is introduced in. The strain energy release rate GIIC can be expressed as follows
[28–30]:
"   #
9Fcr2 a2 E h 2
GTimo
IIC ¼ 1 þ 0:2 ð8:13Þ
16Eb2 h3 G13 a

As the propagation way of bamboo interlaminar crack is mainly type TL, the
ratio of L/h in this test ranged from 8 to 9 and the transverse shear influence on the
bending deflection. To calculate GTimo
IIC , axial elastic modulus EL should be test in
advance. In this test, according to the national standard “Testing methods for
physical and mechanical properties of bamboo” [11], the bending elastic modulus
was tested approximately as the axial elastic modulus. And the length of the
specimen was S = 160 mm, the height h = 10 mm, the span L = 120 m, and the
width was the natural thickness of bamboo culm. The variable span bending method
[31] was used to test shear modulus GTL and the length of the specimen was
S = 160 mm, the height h = 10 mm, the width was the natural thickness of bamboo
culm, and the ratio of L/h ranged from 6 to 19. Substituting the critical load Fcr and
the corresponding elastic modulus in Eq. (8.13), then the critical energy release rate
GTimo
IIC could be calculated respectively.

8.3.2.3 Compliance Calibration Method

Davies et al. [32] and Davidson et al. [33] proposed the compliance calibration
methods. By the method, the compliance of the same material with different crack
length need be tested. The compliance of ENF specimen is generally calculated by
third-order polynomial, as follows:

CL ¼ C0 þ ma3 ð8:14Þ

C0 and m are constants related to material that gained by test date fitting. So, the
Mode II interlaminar fracture toughness can be calculated by the following formula:

Fcr2 a2
GComp
IIC ¼ 1:5 m ð8:15Þ
b
162 8 Interlaminar Fracture Properties of Bamboo

8.3.3 Results and Analysis

The statistic analysis result of the Mode II interlaminar fracture toughness calcu-
lated by Eqs. (8.12), (8.13), and (8.15) was shown in Table 8.3. The average value
2 2
GTest
IIC was 1303.18 J/m (SD = 116.82 J/m ); the tested value of EL/GTL was 23.00,
then the average value GIIC was 1107.54 J/m2 (SD = 125.69 J/m2); the average
Timo

value GCompIIC was 1216.06 J/m2 (SD = 166.48 J/m2). The variance analysis of the
three calculated results was shown in Table 8.4, for GTest IIC , F-value = 0.0376; for
Comp
GTimo
IIC , F-value = 0.6144 and for G IIC , F-value = 3.1987. The variance analyses
Test Timo
results of GIIC and GIIC were both less than F0.05, and the variance analysis result
of GComp
IIC was a little greater than F0.05. It can be seen that in the condition of this
test, the influence of bamboo height and specimen size on Mode II interlaminar
fracture was small. And it indicated that the Mode II interlaminar fracture toughness
could be seen as constant.
Seen from the results, the result GTest IIC calculated by Eq. (8.12) was 15% higher
Timo
than the result GIIC calculated by Eq. (8.13). It was mainly because that under the
transverse bending loads, there existed friction between the crack faces of the ENF
specimen and with the propagation of Mode II crack, the rough crack face produced
bigger friction resistance, therefore, the propagation resistance of crack increased.
In Eq. (8.12), the critical load Fcr and the deflection of loading point d were
calculated according to the actual load–deflection curve, so the critical energy
release rate calculated by Eq. (8.12) was true. However, in Eq. (8.13), although the
critical load Fcr was calculated according to the actual load–deflection curve, the
3
þ 3a3
deflection of loading point was calculated value by d ¼ FC ¼ 2L96EI but not
Timo
actually measured value. So the result GIIC , calculated by Eq. (8.13), was the
resistance of crack propagation without considering the friction between crack faces
but not the true propagation resistance of bamboo Mode II crack under transverse
loading. Although modification had been done to correct the influence of shear
stress in Eq. (8.13), the calculated result was still lower than the result calculated by
the experiment parameter straightly. That is why Eq. (8.12) is always adopted as
national test standard to test the Mode II fracture toughness GIIC of fiber-reinforced
composites in some countries. The result GTest IIC calculated straightly by experiment
parameter was 6.68% higher than the result GComp IIC calculated by Eq. (8.15) and the
difference was very small. The main reason was that when calculated GIIC by
compliance calibration method, the calculated compliance from the test curve must
be fitted to the initial crack length, and then the experimental constant m was
gained. The m should be substituted into the Eq. (8.15) to calculate the GIIC . The
calculating process was complex relatively and it led to the lower calculated result
and the higher coefficient of variation of GComp IIC relatively. So the experiment
parameter substitution method was the most suitable method to calculate the
Mode II interlaminar fracture toughness.
Table 8.3 Statistical analysis of Mode II interlaminar fracture toughness GIIC processed by different methods in different sections of bamboo
2 2
Section Location Dimensions of Number of Mean GTest
IIC (J/m ) Mean GTimo
IIC (J/m ) Mean GComp
IIC (J/m2)
(height, m) specimens (mm) samples
8 1.70 L = 60, h = 7.5 6 1310.15 (SD = 155.22) 1060.92 (SD = 90.14) 1175.78 (SD = 168.70)
15 3.70 L = 60, h = 7.5 6 1344.24 (SD = 113.73) 1146.69 (SD = 191.25) 1271.31 (SD = 196.28)
20 5.40 L = 60, h = 7.5 6 1292.41 (SD = 130.39) 1178.61 (SD = 170.61) 1202.45 (SD = 159.81)
8 1.60 L = 75, h = 9 7 1292.39 (SD = 120.44) 1093.55 (SD = 65.53) 1047.37 (SD = 52.03)
11 2.40 L = 75, h = 9 6 1328.59 (SD = 84.35) 1072.90 (SD = 31.23) 1286.93 (SD = 116.13)
14 3.40 L = 75, h = 9 6 1307.06 (SD = 152.76) 1072.65 (SD = 212.14) 1187.19 (SD = 170.66)
18 4.70 L = 75, h = 9 6 1306.28 (SD = 109.78) 1114.63 (SD = 8.46) 1369.51 (SD = 116.05)
Total 43 1303.18 (SD = 116.82) 1107.54 (SD = 125.69) 1216.06 (SD = 166.48)
8.3 Mode II Interlaminar Fracture of Bamboo
163
164 8 Interlaminar Fracture Properties of Bamboo

Table 8.4 Variance analysis of Mode I interlaminar fracture toughness GIIC of bamboo in
different sections
Sources of df F−0.05 F-value F-value F-value Significance
variation (GTest
IIC ) (GTimo
IIC ) (GComp
IIC )
Between 6 2.3637 0.0376 0.6144 3.1987 not
significant
Within 36
Sum 42

The distribution relationship of the initial crack length and the Mode II inter-
laminar fracture toughness calculated by Eqs. (8.12), (8.13), and (8.15) were shown
in Figs. 8.12, 8.13, and 8.14 respectively. It could be seen that the values of GIIC
nearly has no difference except some values dispersed for the higher a/L ratio which
were easily influenced by the stress concentration caused by the loading indenter.
So, it could be concluded that when the a/L was in the range of 0.25–0.75, the
Mode II interlaminar fracture toughness of bamboo was a constant independent to
the initial crack length basically.
GIIC(Test)(J/mm2)

a/L

Fig. 8.12 Relationship between GTest


IIC and initial crack length/half span (a/L)
GIIC(Timo)(J/mm2)

a/L

Fig. 8.13 Relationship between GTimo


IIC and initial crack length/half span (a/L)
8.3 Mode II Interlaminar Fracture of Bamboo 165

GIIC(Comp)(J/mm2)

a/L

Fig. 8.14 Relationship between GComp


IIC and initial crack length/half span (a/L)

Fig. 8.15 a Image of Mode II interlaminar fracture surface of bamboo obtained by scanning
electron microscope; b 853magnification image of zone A in (a); c 1224magnification image
of zone B in (a); d 1922magnification image of zone C in (a)
166 8 Interlaminar Fracture Properties of Bamboo

Fig. 8.16 The morphologies of cracks: a crack surface of mode I; b crack surface of mode II

8.3.4 The Fracture Surface Analysis

Figure 8.15a was the SEM photos of bamboo Mode II interlaminar fracture surface
after 55 magnification and it could be seen that the fiber bundle in the crack
propagation zone was flat, but the ground tissue was rough and uneven in surface,
and then the selected A, B, and C that located in different area in the propagating
zone were magnified. Figure 8.15b was the SEM photo of the ground tissue at A in
the concavity zone after 853 magnification and it could be seen that the ground
tissue of bamboo was thin-wall ringlike lamellar structure with obvious intercellular
space. The contour of the ground tissue generally maintained the integrity and the
surface was smooth or with few slice debris. This indicates that the crack propa-
gation among ground tissues developed along the middle lamella or the primary cell
wall. Figure 8.15c was the SEM photo of the ground tissue at B in the convexity
zone after 1227 magnification and it could be seen that multilevel parenchyma
cell was torn along crosswise and the intercellular space could be seen clearly.
Figure 8.15d was the SEM photo of the fiber at C after 1922 magnification and
the torn primary wall fragment on the surface of fiber could be seen. Thus, it was
known that the ground tissue was separated from fiber along middle lamella under
shear stress and as the increasing of the dislocation of upper and lower layer, the
thin-walled ground tissue would fracture transversely by tension, while to
thick-walled fiber cell, only middle lamella and primary wall were torn then
debonded, fragments remained.
At last, the photos of Mode I and Mode II interlaminar fracture surface taken by
common camera was shown in Fig. 8.16. By the photos, we could compare the
difference between the two kinds of interlaminar fracture of the bamboo.
Figure 8.16a showed the feature of the Mode I interlaminar fracture surface and it
could be seen that the crack propagation zone was flat. Although tearing trace could
be seen on the matrix (ground tissue surface), it was hard to identify the defor-
mation of matrix. It indicated that the interface strength of matrix was weak. The
ground tissue of bamboo is a thin-walled cell body material with low tensile
strength and there exist space among cells. High stress concentration occurred at the
8.3 Mode II Interlaminar Fracture of Bamboo 167

crack tip, when the Mode I crack propagated along grain, and it made the crack
propagate quickly. Therefore, the deformation of matrix did not develop fully and it
led to the flat fracture surface of Mode I interlaminar fracture. In Fig. 8.16b, the
photo of Mode II interlaminar fracture was shown and the ground tissue was
characteristic of hackle shearing deformation. The ground tissue of cell construction
would form large shear deformation under shear stress and then a large amount of
fracture energy would be absorbed. Generally, under static load, the speed of
Mode II crack propagation was not high, so the plastic deformation of ground tissue
would develop fully during the process of crack propagation and it led to the
laminate fracture surface.

8.4 Mode III Interlaminar Fracture of Bamboo

It was found that the study on the Mode III interlaminar fracture behaviors of
artificial composites started early and the methods and theories were various, such
as crack rail shear (CRS) [34], split cantilever beam (SCB) [35], and edge crack
torsion (ECT) [36] and so on. The reason for the existence of the various methods
was that there was a considerable content of Mode II component in the measured
Mode III interlaminar fracture energy release rate (EER) and the EER varied
unstably with the propagation of crack [37, 38]. Compared with artificial com-
posites, the investigation on the Mode III interlaminar fracture behaviors of bamboo
and wood might encounter more complex problems because of the specific orga-
nization structure of natural composites. It was also one of the main reasons why
there was no research on the Mode III interlaminar fracture toughness of bamboo
and wood so far.
As known, the cells in bamboo internodes stem are all arranged strictly along the
axial direction with no crosswise ray cell, so the structure of bamboo is simple
compared with that of wood and the internodes stem of bamboo could be seen as
unidirectional long-fiber-reinforced composite with fiber bundles as reinforcement
and ground tissue as matrix. Thus, the study methods of artificial composites could
be used for reference to the investigation on the Mode III interlaminar fracture
behavior of bamboo, especially the modified methods proposed in recent years. For
example, Szekrényes [38] proposed a Modified Split Cantilever Beam (MSCB)
method, by which the deflection (d) was revised then the revised corresponding
compliance (C) could be obtained. By the finite element verification, the content of
Mode III component could be as high as 98% in the recommended range of rack
length. Mohammad et al. [39] investigated the influence of reinforcement volume
fraction on the Mode III interlaminar fracture EER of unidirectional E-glass/epoxy
composite by MSCB method. Davidson and Sediles [40] and Johnston et al. [41]
tested the Mode III interlaminar fracture toughness of unidirectional carbon/epoxy
composites by Shear-torsion-bending (STB) method and split-shear-torsion
(SST) method. Here, MSCB method and SCB method was applied to investigate
the Mode III interlaminar fracture toughness of bamboo by referencing to study the
168 8 Interlaminar Fracture Properties of Bamboo

methods of artificial composites. And the test results were compared with the
Mode III interlaminar fracture toughness of artificial composites to investigate the
mechanism of the high toughness of natural composites from the aspects of material
properties and organization structures.

8.4.1 Material and Method

8.4.1.1 Material

The material used in this test was 4 years old moso bamboo (P. pubescens) col-
lected from Huoshan, Anhui Province, China. Three adjacent internodes with two
nodes were collected at the chest height of bamboo stem, and then the nodes were
got through. After forced dried for 7 days (60 °C) in the constant temperature
humidity chamber and naturally air dried for 90 days, the sample was cut into
pieces along the grain.
The specimens were made with a size of 160  20  b (Unit: mm) for SCB
method and 160  10  b (Unit: mm) for MSCB method, where b was the
thickness of bamboo stem, about 9 mm. To simulate a naturally sharp crack, the
pre-crack was cleaved along middle level parallel to the grain by a wedge.

8.4.1.2 Method

SCB Method
As seen in Fig. 8.17, the surface of the loading end was a cambered surface, whose
center was on the center lines of the holding devices and specimen. It was a line
contact between the cambered surface and the indenter of testing machine. When

1- Grip1
3
5 2- Grip 2
3- Specimen
1 6
4- Limited Rollers
2 5- Torsion Limited Column
4
6- Base

Fig. 8.17 The test device of split cantilever beam (SCB) method
8.4 Mode III Interlaminar Fracture of Bamboo 169

loaded, load is transferred along the center lines above. The two limited rollers
played a role in limiting the free rotation of the end of specimen and keeping the
loading line at the center lines. What is more, it was also line contact between grip 1
and cylindrical rollers, which would reduce friction to make the experiment data
more accurate compared with surface contact. The torsion-limited column could
limit the torsion of specimen in order to reduce or eliminate the influence of torsion
on the Mode III interlaminar fracture behavior in the process of loading.
Tests were performed on the computer-controlled testing machine, and the load–
load point deflection (F-d) curve was recorded by computer automatically during
the test. Single-specimen with multi-points method was applied, that is, on one
specimen, repetitive loading–unloading–loading was performed. The initial loading
speed was 2 mm/min, and then it was increased properly with the propagation of
crack. At the initial stage of loading, the F-d curve was straight until the crack
began, then the slope of F-d curve increased suddenly, meanwhile, the crack
propagated slowly along the grain. After the F-d curve rose nonlinearly to the
maximum load, it began to decline. And at this moment, the testing machine is
stopped, the curve is curved, and mark the crack tip. As shown in Fig. 8.18, it was
the F-d curves of one specimen and there were 8 curves in total which meant that
the specimen was loaded for 8 times. The critical load Fcr was the load when the
crack began, namely the point the slope of the F-d curve increased. After the test
finished, length between the crack tip was marked each time and the center of the
load point was measured. Since the crack propagation amount in the outer part of
the bamboo specimen was different with that in the inner part after loaded each
time, the average value of the crack length in the outer part and that in the inner part
was taken as the final crack length. The temperature was 27 °C and the humidity
was 60% in the laboratory.
According to Irwin–Kies expression [42], the propagation resistance of Mode III
interlaminar crack along grain, namely the interlaminar fracture toughness can be
calculated by the following formula:

0.35
0.3
0.25
Load (kN)

0.2
0.15
0.1
0.05
0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 5.5 6
Deflection (mm)

Fig. 8.18 The typical load–deflection curves of one bamboo by SCB method
170 8 Interlaminar Fracture Properties of Bamboo

@U @ðFcr2 C=2Þ Fcr2 @C


GIIICSCB ¼ ¼ ¼  ð8:16Þ
@A @ðbaÞ 2b @a

According to the typical F-d curve of the SCB specimens, the slope of the
straight portion on the curve was inversed to crack length. And the reciprocal of the
slope was the compliance Ci corresponding to different crack length ai. By curve
fitting, the exponential relationship between the compliance and crack length was
obtained:

C ¼ qema ð8:17Þ

where q and m were the fitting coefficients of the a-C curve.


MSCB Method
According to Szekrényes [38, 43, 44], as seen in Fig. 8.19, the diameter of the
loading holes was 10 mm, S2 = 18 mm, and S1 was 15, 30, 45, 60, 75, and 90 mm
respectively. As a/(S1 + S2) ranged from 1.02 to 1.09, the corresponding initial
crack length was about 35, 50, 65, 80, 95, and 110 mm respectively. Tests were
performed on the computer-controlled testing machine, and the load–load point
deflection (F-d) curve was recorded by computer automatically during the test.
Each specimen was loaded for 6 times corresponding to the 6 initial crack lengths,
and 6 test curves were obtained. The initial loading speed was 2 mm/min, and then
it was increased properly with the increase of initial crack length. Temperature was
27 °C and the humidity was 60% in laboratory.
According to Szekrényes [38], the Mode III interlaminar fracture EER could be
obtained by the following formula:

Fig. 8.19 Exploded view of the modified split cantilever beam. Reprinted from ref. [38],
Copyright 2018, with permission from Elsevier
8.4 Mode III Interlaminar Fracture of Bamboo 171

12P2 a2
GIIICMSCB ¼ ½fEB2 þ fTIM2 þ fFT2 þ fSV2  ð8:18Þ
b4 hE11
   2  2    2  
where fEB2 ¼ 1  2 s1 þa s2 þ s1 þa s2 , fTIM2 ¼ 0:1 ba GE1113 , fFT2 ¼ 0:06 11 ba GE1112 ,
   1=2  1=2
fSV2 ¼ 0:32 1  s1 þa s2 ba GE1113 , and 1 ¼ 1  0:63l hb, l ¼ G 13
G12 .
b and h were thickness and half height of specimen. a was the length of crack,
and E11, G12, and G13 were the bending modulus of elasticity and shearing modulus
of elasticity.
According to the national standard “Testing methods for physical and
mechanical properties of bamboo” [11], the bending elastic modulus was tested.
The length of the specimen was S = 160 mm, the height h = 10 mm, the span
L = 120 mm, and the width was the natural thickness of bamboo culm. The vari-
able span bending method [31] was used to test shear modulus. The length of the
specimen was S = 160 mm, the height h = 8 mm, the width was the natural
thickness of bamboo culm and the ratio of L/h ranged from 6 to 17.
To test the mechanical properties of the reinforcement-bamboo fiber and the
matrix-ground tissue, the bamboo block tensile specimens and the “rule of mixture”
were applied (Chap. 7, Sect. 7.3.1). The fresh bamboo billets were cut into sections
with a length of 160 mm, and then each section was cleaved into several strips to
make tensile specimens with different fiber contents. The size in the middle part of
specimens was 40  3  1.3 (Unit: mm) and there were 28 specimens in total.

8.4.2 Results and Discussion

8.4.2.1 Results

The average Mode III interlaminar fracture toughness of bamboo obtained by


Eq. (8.16) was GIIIC-SCB = 2170 J/m2 (SD = 350 J/m2), and variable coefficient
was 16%.
The average Mode III interlaminar fracture EER calculated by Eq. (8.18) was
GIIIC-MSCB = 2040 J/m2 (SD = 400 J/m2), when the crack length was longer than
40 mm.
By the tensile test of bamboo block tensile specimens, the relationship between
tensile strength (rc), tensile elastic modulus (Ec), and fiber volume fraction (Vf)
were rc = 485.28Vf + 51.67 (MPa) and Ec = 16.90Vf+ 1.62 (GPa). The tensile
properties of bamboo fiber and ground tissue could be obtained according to the
“rule of mixture”, as shown in Table 8.5. And other three elastic coefficients were
also shown in Table 8.5.
172 8 Interlaminar Fracture Properties of Bamboo

Table 8.5 Mechanical properties of bamboo, bamboo fiber, and ground tissue
Item r (MPa) E11 (GPa) G12 (GPa) G13 (GPa)
Bamboo 172.43 10.30 0.35 0.33
Fiber 536.95 18.52
Ground tissue 51.67 1.62

Table 8.6 One-way variance analysis of the Mode III interlaminar fracture toughness values
obtained by the two methods
Sum of squares df Mean square F Significance
Between Groups 0.307 1 0.307 2.28 0.135
Within Groups 9.971 74 0.135
Total 10.278 75
There is no significant difference between the two groups
G IIIC-SCB(N/mm)

3
2.5
2
1.5
1
0.5
0
40 45 50 55 60 65 70 75
a (mm)

Fig. 8.20 The distribution relationships of crack length a and GIIIC-SCB

8.4.2.2 Discussion

It could be seen that the results obtained by the two methods were very close. As
seen in Table 8.6, it could be seen that there was no significant difference between
the results by one-way analysis of variance. The results not only indicated that the
applicability of MSCB method to natural fiber-reinforced composite bamboo, but
also proved the accuracy of SCB method. What’s more, the physical dimension of
specimens had no influence on the Mode III interlaminar fracture toughness of
bamboo
The distribution relationships of crack length a and the Mode III interlaminar
fracture toughness was shown in Fig. 8.20, which indicated that in the range of the
crack length, 40–72 mm, the of bamboo was kept stable with the propagation of a
crack in principle.
As seen in Fig. 8.21, the distribution relationships of crack length a and the
Mode III interlaminar fracture toughness showed that had no significant change
with the increase of crack length when the crack length was longer than 40 mm.
8.4 Mode III Interlaminar Fracture of Bamboo 173

GIIIC -MSCB(N/mm)
2.5
2
1.5
1
0.5
0
40 50 60 70 80 90 100 110 120
a (mm)

Fig. 8.21 The distribution relationships of crack length a and GIIIC-MSCB

Both the distribution relationships above were in accordance with the results of
Jun [45] that when crack length was longer than 40 mm, the Mode III interlaminar
fracture toughness trended to constant.

8.4.3 Compared with Artificial Fiber-Reinforced Polymer


(FRP)

The Mode III interlaminar fracture of artificial FRP has been studied by many
researchers by literature retrieval. The artificial FRP are mainly epoxy or polyester
reinforced by carbon fiber or glass fiber. As seen in Table 8.7, the Mode III
interlaminar fracture results of several unidirectional fiber-reinforced composites
studied by many researchers was shown. It could be seen that, compared with
bamboo, the general mechanical properties of both artificial FRP and their rein-
forcements were superior, while the Mode III interlaminar fracture EER of bamboo
was higher than that of all the artificial FRP above. Then, how to interpret that the
Mode III interlaminar fracture EER of new type artificial FRP with high perfor-
mance was lower than that of natural composite bamboo? Maybe the explanation
could be given by the comparison analysis on the organization structures of the two
kinds of materials.
The glass/epoxy composite was taken as example, which was used most widely
based on mature research. The glass fiber as reinforcement had low compatibility
and wettability with resin matrix because of its smooth columned surface, which
resulted in low interface bonding force. As seen in Fig. 8.22, it was the SEM image
of Mode III interlaminar fracture surface of E-glass/epoxy composite with
Vf = 55% [39]. Due to the difference in physical and chemical structure between
glass fiber and epoxy, abruptly changed interface properties of E-glass/epoxy
composite, and the low interface bonding force led to the phenomenon that the
exposed surfaces of glass fibers were smooth and a little deformed matrix adhered
to glass fibers, when Mode III interlaminar fracture happened. Thus, the Mode III
Table 8.7 The Mode III interlaminar fracture results of several unidirectional fiber-reinforced composites
174

Reference Method Material E11 E22 G12 m12 GIII (J/ Reinforcement Matrix
(GPa) (GPa) (GPa) (GPa) m2)
P. Robinson (1994) ISCB Carbon/ 150 9.5 1.07 0.26 1660 Carbon fiber (60%) Epoxy
epoxy
Shun-Fa Hwang (2001) SCB Carbon/ 121 9.4 6.2 0.23 1000 Carbon fiber (60%) Epoxy
epoxy
Glass/epoxy 37.9 14.3 5.6 0.29 1270 Glass fiber (60%) Epoxy
András zekrényes (2009) MSCB Glass/ 33 7.2 3 0.27 115 Glass fiber (43%) Polyester
polyester
Barry D. Davidson (2011) STB Carbon/ 147.6 – 4.31 0.32 1697 Carbon fiber Epoxy
epoxy 163.8 – 4.95 0.27 1463
Mohammad Khoshravan MSCB Glass/epoxy 41.93 5.25 3.43 0.27 120–150 Glass fiber (55%) Epoxy
(2014) 33.17 4.13 2.64 0.31 175–210 Glass fiber (46%)
23.46 2.54 2.23 – 175–475 Glass fiber (30%)
This chapter MSCB Bamboo 10.30 – 0.33 – 2040 Bamboo fiber Ground
(32%) tissue
SCB split cantilever beam; ISCB improved SCB; MSCB modified SCB; STB shear-torsion-bending
8 Interlaminar Fracture Properties of Bamboo
8.4 Mode III Interlaminar Fracture of Bamboo 175

Fig. 8.22 The Mode III


interlaminar fracture surface
of glass/epoxy composite with
Vf = 55%. Reprinted from ref.
[39], Copyright 2018, with
permission from Elsevier

interlaminar fracture EER was mainly contributed by the deformation of matrix and
a small amount by the debonding of the interface between matrix and glass fibers.
However, for bamboo, the connection of interfaces and the bio-assembly way of
fibers and ground tissue were special. First, the structure of both fibers and ground
tissue cells would transit gradually to decrease the abrupt change of interface
properties, namely, the cell wall of peripheral fiber possessed more lays and the size
of ground tissue cells next to fibers were smaller (Fig. 8.23). Then, for interface,
cells were connected by intercellular layer. Hydroxyl and carboxyl between cell
walls in dried bamboo were assembled by hydrogen bond or chemical bond, which
resulted in the interconnection of pectin and lignin to form complicated colloidal
complex, namely intercellular layer, and the interconnection among intercellular
layer and the primary cell walls of adjacent cells to form compound middle lamella.
Although there were significant difference in the structural features and physical
and mechanical properties between fibers and ground tissue, they could be com-
bined to form a whole by the compound middle lamella to achieve the best
performance.
As seen in Fig. 8.24, it was the SEM images of bamboo Mode III interlaminar
fracture surface. The failure characteristics of fibers were that most fibers debonded
along the interface of fibers (Fig. 8.24a), and few were peeled and fractured. As
seen in Fig. 8.24b, after fibers were debonded, ribbon-like torn traces could be
observed on the surface of the damaged intercellular layer which has obvious
feature of shear failure. For ground tissue, the failure characteristics were that in the
inner part where the content of ground tissue was high, most of the ground tissue
cells were torn and a few were debonded along the interface of ground tissue cells
(Fig. 8.24c); the ground tissue cells were nearly all torn in the outer part. In bio-
composite, cell wall is composed of multilayers and plays a role of mechanical
176 8 Interlaminar Fracture Properties of Bamboo

Fig. 8.23 The cross section of bamboo stem and vascular bundle

Fig. 8.24 The SEM images of the Mode III interlaminar fracture surface of bamboo

support. As seen in Fig. 8.24d, the multilayers of damaged ground tissue cell wall
could be observed clearly and was in shear failure mode, which would absorb large
amounts of energy during the process of Mode III interlaminar fracture.
It could be seen that the Mode III interlaminar fracture mechanism of bamboo
was that the resistance that hindered the propagation of interlaminar crack was
contributed by the transverse shear strength of ground tissue cell wall and interface
strength. The interface concluded the interface between ground tissue cells, fibers as
well as ground tissue cell, and fiber. The differences in organization structures and
interface connection between bamboo and artificial FRP were the main reason for
the high Mode III interlaminar fracture EER of bamboo compared with that of
8.4 Mode III Interlaminar Fracture of Bamboo 177

artificial FRP. Thus, the specific inner construction and organization relevance skill
of natural composite evolved by millions of years, especially the construction of
interface are valuable reference for the design of fiber-reinforced composites.

8.5 The Toughness Contribution of Bamboo Node


to the Interlaminar Fracture Toughness of Bamboo

Through millions of years natural selection of the process of survival of the fittest,
the fine structure of all living beings has been evolved to go through their envi-
ronment. To overcome this disadvantage, the excellent structure—“internode +
node” is formed after millions years’ evolution. The existence of node changes the
damage pattern that when damage along the grain happens, the force to resist the
cleavage only comes from the weak interfacial resistance. Thus, the cleavage
strength along grain is greatly improved and this is also the perfect embodiment of
function adaptability of biological material. However, no study on the interlaminar
fracture of bamboo specimens with node was founded by the literature search.
Because of unique structure, there is great difference between bamboo internode
and node in structure which leads to the difference of physico-mechanical prop-
erties. Here, the Mode I, Mode II and Mode III interlaminar fracture toughness
difference between bamboo internode specimens and bamboo specimens with node
were studied aiming at exploring the influence of the existence of node on bamboo
interlaminar fracture.

8.5.1 Mode I

8.5.1.1 Material and Test Method

The material used in this test was one 4 years old moso bamboo (P. pubescens)
from Huoshan, Anhui Province, China. Three adjacent internodes with two nodes
(from the 12th internode to the 14th internode) were collected as sample chosen at
the chest height of the bamboo (Fig. 8.25). Before drying, the diaphragms in

Fig. 8.25 Schematic diagram of specimen selection


178 8 Interlaminar Fracture Properties of Bamboo

bamboo sample were destroyed to make the sample ventilated during the process of
drying. After forced air drying for 7 days (60 °C) in the constant temperature
humidity chamber and natural air drying for 90 days, the sample was cut into pieces
along grain with a width of 22–25 cm.
As shown in Sect. 8.2, the DCB method is applied. WDW-100
computer-controlled universal mechanical testing machine was used with a test
force accuracy of ±5%, displacement resolution of 0.01 mm, and speed accuracy of
1%. 15 kN mechanical sensor, steel U hook, steel pins, etc., were also applied.
Because the bamboo culm has a hollow structure separated with nodes and the
length and diameter of each internode are limited, the number of specimens gained
was much less than that of normal composites. In this test, the DCB specimens were
collected from the 12th internode to the 14th internode, among which the internode
specimens were collected from the 12th internode, seven effective specimens (total
number  5 recommended by ASTM D5528) were obtained, and the specimens
with node were collected from the 12th–13th and 13th–14th internode, 11 effective
specimens were obtained, as seen in Fig. 8.26. Due to ASTM D5528 [6], the
dimension of the specimen was 220  20  b (mm) and b was the natural
thickness of the bamboo culm along the radial direction. The pre-crack: to simulate
a naturally sharp crack, the crack was cleaved along the middle level parallel to the
grain by a knife as seen in Fig. 8.2. The test room temperature was 25–27 °C, and
the humidity was 60%.
The DCB specimen was connected with a steel U hook by a steel pin loaded by
the computer-controlled testing machine with a crosshead speed between 1 and
4 mm/min. For bamboo internode specimens, single-specimen with multi-point
method was used, that is, for one single-internode specimen, during loading, the
crack propagated and the corresponding applied load versus load point deflection
(F-d) curve was recorded by a computer. Once the load (F) went down, the load
machine stopped, and the crack tip at both the inner part (a1inner) and the outer part
(a1outer) of the internode specimen was marked taking the average length of a1inner
and a1outer as the corresponding crack length a1. Then, it was unloaded and reloaded
and the same procedure was repeated until complete fracture appeared so the

Fig. 8.26 Schematic diagram 20


of Mode I interlaminar
fracture specimens with the Loading hole
node (Unit mm) (Diameter=5)
a1

a2

a3

Initial crack
ai

Bamboo node
220
8.5 The Toughness Contribution of Bamboo Node to the Interlaminar … 179

Fig. 8.27 Typical Mode I 0.16


interlaminar fracture F-d
curve of one internode 0.14
specimen
0.12

0.1

F (kN)
0.08

0.06

0.04

0.02

0
0 2 4 6 8 10 12
(mm)

corresponding crack length a2, a3, a4…an was obtained. The loading speed was
lifted with increasing length of crack [4]; for bamboo specimens with node,
multi-specimen with single-point method was used, that is, on one single specimen
with node, load was applied only once and the crack length a was obtained by
averaging the ainner and aouter correspondingly. For different specimens, the loading
speed was lifted properly with increasing length between node and the center of
loading holes. The curve of load–load point deflection (F-d) was automatically
recorded by a computer during the test. The typical F-d curve of one internode
specimen is shown in Fig. 8.27, and the crack length of each curve from test point
a1 to test point a29 was 43.80, 46.75, 49.10, 50.65, 53.10, 55.30, 56.50, 59.10,
61.80, 64.50, 67.55, 74.65, 76.65, 79.80, 84.95, 89.40, 94.00, 97.50, 101.35,
112.20, 115.85, 122.05, 129.55, 133.60, 138.00, and 141.70 (unit mm), respec-
tively, and for the internode specimen, 29 test points were collected; the F-d curves
of all specimens with node are shown in Fig. 8.28, and the corresponding crack
length was 45.00, 51.00, 62.50, 63.65, 70.30, 80.00, 80.75, 103.50, 117.50, 133.60,
and 136.25 (unit mm), respectively. From the two figures, it could be seen that in
the initial loading, the F-d curve kept linear until the specimen started crazing. Once
crazing, the bearing capacity of the specimen decreased sharply and then the crack
rapidly propagated along the grain. The maximum load Fmax of the F-d curve was
determined as the critical load Fcr.
After the test was finished, the specimens with node were cleaved to measure
initial crack length: at the front end of the dyeing area, five points were marked
from the outer bamboo wall to the inner bamboo wall uniformly, and the distance
between each point and the center of loading holes was measured. Finally, the
average length of the five points was gained as the initial crack length (Fig. 8.29).
180 8 Interlaminar Fracture Properties of Bamboo

F (kN)

(mm)

Fig. 8.28 Mode I interlaminar fracture F-d curves of all specimens with node

Fig. 8.29 Measure the initial


crack length of specimen with
node

8.5.1.2 Results and Analysis

Reliability Analysis of Data


In this test, the fitting coefficient q ranged from 0.0007 to 0.0017, the fitting
coefficient m from 2.34 to 2.52, and the regression coefficients R2 were all above
0.98. The fitting curve of a and C of one internode specimen is shown in Fig. 8.30.
To analyze the reliability of the data, the data analysis software SPSS was used.
According to SPSS, linear model was chosen to perform the linear-regression
analysis. So after taking logarithm of Eq. (8.3) in Sect. 8.2.2, it met the linear
model, as follows:

lg C ¼ lg q þ m lg a ð8:19Þ

The linear-regression analysis result of lga and lgC of the corresponding


internode specimen is shown in Table 8.8. The C and a fitting curve of all speci-
mens with the node is shown in Fig. 8.31. And the corresponding linear-regression
8.5 The Toughness Contribution of Bamboo Node to the Interlaminar … 181

200
2.51
180 C = 0.0007a
160 2
R = 0.99
C (mm/kN) 140
120
100
80
60
40
20
0
40 60 80 100 120 140
a (mm)

Fig. 8.30 Fitted curve of test curve compliance and crack length of one internode specimen

analysis result of lga and lgC is shown in Table 8.9. It can be seen from the
Table that lgC has a remarkable correlation to lga and the data are reliable.
Results and Discussion
According to Eq. (8.2), by DCB method, the Mode I interlaminar fracture tough-
ness of internode specimens is listed in Table 8.10 and the Mode I interlaminar
fracture toughness of specimens with node in Table 8.11. From the information
above, the average Mode I interlaminar fracture toughness of internode specimens
was 498 J/m2 (SD = 65 J/m2) with coefficient of variation of 12.9%; the average
Mode I interlaminar fracture toughness of specimens with node was 1431 J/m2
(SD = 198 J/m2) with coefficient of variation of 13.8%. The conclusion was drawn
that the Mode I interlaminar fracture toughness of bamboo specimens with node
was much higher than that of bamboo internode specimens. The relationship
between crack length (a) and Mode I interlaminar fracture toughness (GIC) of both
internode specimens and specimens with node is shown in Fig. 8.32, from which it
is obvious that the Mode I interlaminar fracture toughness of bamboo specimens
with node is much higher than that of bamboo internode specimens. The Mode I
interlaminar fracture toughness of bamboo internode specimens decreased slightly
with increasing crack length, which could be mainly due to the nonlinear phe-
nomenon caused by large deflection of DCB, i.e., so-called geometry-nonlinearity,

Table 8.8 The corresponding linear-regression analysis result of lga and lgC of one internode
specimen
Model Sum of squares df Mean square F Significance
Regression 4.31 1 4.31 15185.931 0.000
Residual 0.008 27 0.000
Total 4.317 28
Predictors: (constant), lga; dependent variable: lgC
182 8 Interlaminar Fracture Properties of Bamboo

200
2.43
180 C = 0.0012 a
2
160 R = 0.99
C (mm/kN) 140
120
100
80
60
40
20
0
40 60 80 100 120 140
a (mm)

Fig. 8.31 Fitted curve of test curve compliance and crack length of all specimens with node

Table 8.9 The corresponding linear-regression analysis result of lga and lgC of all specimens
with node
Model Sum of squares df Mean square F Significance
Regression 1.58 1 1.58 1573.66 0.000
Residual 0.009 9 0.001
Total
1.589 10
Predictors: (constant), lga; dependent variable: lgC

Table 8.10 Descriptive statistics of the Mode I interlaminar fracture toughness of all internode
specimens
Specimen Number of test GInternode
IC
Standard Coefficient of
No. points (J/m2) deviation variation (%)
1 7 523 56 10.7
2 11 538 71 13.1
3 14 509 58 11.4
4 19 547 82 14.9
5 29 475 43 9.0
6 19 473 59 12.5
7 18 470 38 8.1

as crack increases [10]; however, there was no obvious trend for the Mode I
interlaminar fracture toughness of specimens with node, which may result from the
small amount of specimens.
From the results above, it can be concluded that the Mode I interlaminar fracture
toughness of bamboo specimens with node is much higher than that of bamboo
internode specimens, which was mainly caused by the difference in structure
between bamboo internode and node (Table 8.12).
8.5 The Toughness Contribution of Bamboo Node to the Interlaminar … 183

Table 8.11 Mode I interlaminar fracture toughness of all specimens with node
Specimen No. a (mm) b (mm) Fmax (kN) C (mm/kN) dC/da GNode
IC (J/m2)
1 45.00 11.20 0.217 12.096 0.661 1396
2 51.00 11.10 0.178 17.983 0.79 1128
3 62.50 11.20 0.179 23.91 1.056 1518
4 63.65 11.05 0.181 27.377 1.084 1614
5 70.30 11.00 0.168 33.068 1.249 1599
6 80.00 10.85 0.155 42.976 1.501 1653
7 80.75 11.30 0.154 53.436 1.521 1603
8 103.50 11.80 0.126 92.945 2.167 1462
9 117.50 11.60 0.099 125.44 2.596 1105
10 133.60 12.00 0.097 164.18 3.118 1223
11 136.25 11.60 0.102 178.65 3.206 1445
Average 1431

2000
1800
1600 G ICNode =1431J/m2
1400
GIC (J/m2)

1200
1000
800 Internode
GIC =498J/m2
600
400
200
0
20 40 60 80 100 120 140 160 180
a(mm)

Fig. 8.32 Relationship between a and GIC of both internode specimens and specimens with node

Table 8.12 Statistics analysis of all test points of internode specimens and specimens with node
Specimen The number of all test GIC (mean) Standard Deviation Coefficient of
points (J/m2) (J/m2) variation (%)
Internode 117 498 65 12.9
With 11 1431 198 13.8
node

8.5.2 Mode II

8.5.2.1 Material and Method

The material used in this test was 4 years old moso bamboo (P. pubescens) from
Huoshan, Anhui Province, China. Three adjacent internodes were collected at the
184 8 Interlaminar Fracture Properties of Bamboo

Fig. 8.33 Schematic diagram of the selection of Mode II and Mode III specimens

chest height of the bamboo (Fig.8.33), and then got through the nodes. After forced
air dried for 7 days (60 °C) in the constant temperature humidity chamber and
naturally air dried for 90 days, the sample was cut into pieces along the grain.
ENF specimen was used, and the diagrams of specimen. Because bamboo clum
has hollow structure separated with nodes and the length and diameter of each
internode are limited, the number of specimens gained was much less than that of
normal composites. The internode specimens were collected from the 9th internode
and 17 effective specimens were gained with half span L = 60 mm, 2h = 15 mm;
the specimens with node were from 9th internode to the 10th internode and 7
effective specimens were gained with L = 55 mm, 2h = 15 mm (Fig. 8.33). Wide b
was the natural thickness of bamboo culm. And according to Yu [30], the test
results would not be influenced by the change of span.
The test method and data process methods were the same with that in Sect. 8.3.

8.5.2.2 Results and Analysis

By ENF method, the Mode II interlaminar fracture toughness values of internode


specimens according to formula Eqs. (8.12), (8.13), and (8.15) in Sect. 8.3.2 were
2 Comp 2,
GTest
IICðinternodeÞ = 1371.54 J/m , GIICðinternodeÞ = 1276.238 J/m and GTimo
IICðinternodeÞ =
2
1350.66 J/m respectively (Table 8.13). And according to national standard, the
bending elastic modulus was measured, EL = 9602.85 N/mm2 and the tested shear
modulus GTL = 269.68 N/mm2.
According to Eq. (8.12), the interlaminar fracture toughness of specimens with
2
node was GTest Timo
IICðnodeÞ = 1711.53 J/m ; according to Eq. (8.13), GIICðnodeÞ =
1628.71 J/m2 (Table 8.14). As it was difficult to test the bending elastic modulus
and shear modulus of specimen with node, the EL/GTL value of internode speci-
mens was used to estimate the GTimo IICðnodeÞ of specimens with node approximately.
The descriptive statistics of Mode II interlaminar fracture toughness GIIC of both
internode specimens and specimens with node calculated by different data process
methods were shown in Table 8.15 and it could be seen that the variable coeffi-
cients were all less than 20%. By the results above, conclusion was obtained that by
the two calculated methods, the Mode II interlaminar fracture toughness of
internode specimens were both less than that of specimens with node.
Table 8.13 Mode II interlaminar fracture toughness GIIC of internode specimens processed by different methods
a
No. h (mm) b (mm) a (mm) L (mm) L Fcr (kN) dcr (mm) GTest
IICðinternodeÞ GTimo
IICðinternodeÞ GComp
IICðinternodeÞ
(J/m2) (J/m2) (J/m2)
1 7.65 12.30 28.25 60 0.47 0.1136 1.8515 1228.64 1368.14 1255.00
2 7.89 12.65 27.20 60 0.45 0.11588 1.8934 1173.25 1199.45 1178.47
3 7.78 12.30 20.50 60 0.34 0.1258 1.9828 837.36 1123.04 810.50
4 8.06 12.45 22.55 60 0.38 0.1301 2.2046 1130.56 1204.01 1037.52
5 7.94 12.03 28.25 60 0.47 0.1187 2.0321 1441.85 1435.72 1402.75
6 7.82 12.37 30.00 60 0.5 0.1146 1.9211 1405.15 1418.42 1433.45
7 8.14 12.25 32.30 60 0.54 0.1146 1.9211 1583.47 1453.34 1678.63
8 7.84 12.34 37.15 60 0.62 0.9457 1.9592 1592.47 1307.95 1501.10
9 7.92 12.29 24.75 60 0.41 0.1330 2.4471 1529.77 1482.29 1323.61
10 7.92 11.73 21.15 60 0.35 0.1406 2.4475 1282.32 1535.39 1130.14
11 7.61 12.13 25.65 60 0.43 0.1147 2.0284 1176.75 1287.36 1070.16
12 7.84 12.03 26.85 60 0.45 0.1226 2.0691 1395.83 1477.53 1350.48
13 7.66 12.21 27.55 60 0.46 0.1149 2.0752 1348.98 1374.78 1231.99
14 8.00 11.86 33.65 60 0.56 0.9673 2.2037 1676.42 1219.59 1340.08
15 7.67 12.23 33.75 60 0.56 0.1018 2.3119 1802.95 1415.63 1448.40
16 7.75 12.14 33.80 60 0.56 0.1011 2.1155 1654.11 1383.63 1443.73
17 7.85 11.76 24.50 60 0.41 0.1177 1.8608 1056.24 1274.98 1060.03
8.5 The Toughness Contribution of Bamboo Node to the Interlaminar …
185
186 8 Interlaminar Fracture Properties of Bamboo

Table 8.14 Mode II interlaminar fracture toughness GIIC of specimens with node processed by
different methods
a
No. h (mm) b (mm) a (mm) L (mm) L Fcr dcr GTest
IICðnodeÞ GTimo
IICðnodeÞ
(kN) (mm) (J/m2) (J/m2)
1 7.47 12.14 22.75 55 0.41 0.1235 2.2205 1430.11 1350.59
2 7.68 12.31 23.00 55 0.42 0.1369 1.9898 1427.07 1539.29
3 7.79 12.61 30.75 55 0.56 0.1198 1.7344 1668.82 1552.41
4 7.73 12.29 28.95 55 0.53 0.1283 1.8213 1768.97 1766.41
5 7.77 12.29 30.00 55 0.55 0.1262 1.9380 1949.10 1770.80
6 7.78 12.3 31.25 55 0.57 0.1238 1.9416 2025.11 1792.75

Table 8.15 Descriptive statistics of the Mode II interlaminar fracture toughness of all specimens
Sample Method GIIC (mean) Standard deviation Coefficient of
(J/m2) (J/m2) variation (%)
Internode Experiment 1371.54 254.13 18.53
method
Timoshenko 1350.66 117.54 8.70
method
Compliance 1276.24 212.01 16.61
method
With Experiment 1711.54 252.98 14.78
node method
Timoshenko 1628.71 177.33 10.89
method

For the Mode II interlaminar fracture of bamboo internode specimens, three data
process methods were applied and the GIIC value calculated by experiment
parameter substitution method was higher than those calculated by both
Timoshenko beam theory method and compliance calibration method and this result
agreed with the conclusion in Sect. 8.3, thus the GTest
IICðinternodeÞ value calculated by
experiment parameter substitution method was considered as the most proper GIIC
value. For the Mode II interlaminar fracture of bamboo specimens with node, the
GTimo Test Test
IICðnodeÞ value was also less than the GIICðnodeÞ value, so the GIICðnodeÞ was con-
sidered as the proper GIIC value.
The distribution relationships of a/L and the Mode II interlaminar fracture
toughness GIIC calculated by experiment parameter substitution method and
Timoshenko beam theory method were shown in Figs. 8.34 and 8.35 respectively.
It could be seen that the GIIC calculated by experiment parameter substitution
method of both internode specimens and specimens with node increased with the
increasing of a/L which resulted from the stress concentration caused by the loading
indenter when the crack tip closed to the loading indenter. However, the GIIC of
8.5 The Toughness Contribution of Bamboo Node to the Interlaminar … 187

Experimental method G IIC(J/m2)

2500
- GIIC
Test
(node)

2000 Test
- G IIC (internode)
1500

1000

500

0
0.3 0.35 0.4 0.45 0.5 0.55 0.6 0.65
a/L

Fig. 8.34 Relationship between GTest


IIC and initial crack length/half span (a/L) of internode
specimens and specimens with node
Timoshenkon beam method GIIC (J/m2 )

Timo
- G IIC (node)
2000 Timo
- G IIC (internode)

1500

1000

500

0
0.3 0.35 0.4 0.45 0.5 0.55 0.6 0.65
a/L

Fig. 8.35 Relationship between GTimo


IIC and initial crack length/half span (a/L) of internode
specimens and specimens with node

internode specimens calculated by Timoshenko beam theory method had no


obvious trend with the increase of a/L; the calculated GIIC of specimens with the
node increased slightly with the increasing of a/L.

8.5.3 Mode III

8.5.3.1 Material and Test Method

The material used in this test was 4 years old moso bamboo (P. pubescens) from
Huoshan, Anhui Province, China. Three adjacent internodes were collected at the
chest height of the bamboo (Fig. 8.33), and then got through the nodes. After forced
188 8 Interlaminar Fracture Properties of Bamboo

Fig. 8.36 Mode III specimens: a internode specimens; b specimens with node

air dried for 7 days (60 °C) in the constant temperature humidity chamber and
naturally air dried for 90 days, the sample was cut into pieces along the grain.
The SCB specimens were applied and the diagram of modified SCB fixture
device was shown in Fig. 8.17. As seen in Fig. 8.33, the internode SCB specimens
were collected from the 10th internode and the size was 170  20  b (mm),
where b was the natural thickness, with 7 effective specimens; the SCB specimens
with node were from the 10th to 11th internode and the size S  20  b (mm),
where b was the natural thickness, with 9 effective specimens (Fig. 8.36). For
different SCB specimens with node, the length between node and loading points
ranged from 40 to 100 mm.
SCB method was applied. Double cantilever inversion symmetry bending load
method was used to test Mode III interlaminar fracture (Fig. 8.17). The specimen

0.5
0.45
0.4
0.35
Load (kN)

0.3
0.25
0.2
0.15
0.1
0.05
0
0 2 4 6 8

Deflection (mm)

Fig. 8.37 Tipical load–deflection curve of internode specimen


8.5 The Toughness Contribution of Bamboo Node to the Interlaminar … 189

24 y = 4.7283e
0.0183x

22 2
R = 0.9783
20
C 18
16
14
12
10
50 55 60 65 70 75 80 85 90
a (mm)

Fig. 8.38 Tipical fitted curve of test curve compliance and crack length of one internode
specimen

was placed in the fixture device, then loaded by a computer-controlled testing


machine and the load–load point deflection (F-d) curve was recorded by computer
automatically during the test (Fig. 8.37). Temperature was 27 °C and the humidity
was 60% in the laboratory. For bamboo internode specimens, single-specimen with
multi-point method was applied, that is, on one specimen, repetitive loading–un-
loading–loading was performed and the loading speed was 1 mm/min; for bamboo
specimens with node, multi-specimen with single-point method was used, that is,
on one specimen, load was performed only once and for different specimens, the
loading speed was lifted properly with the increasing length between node and the
center of loading holes and controlled in 1–2 mm/min.
At the loading initial, the F-d curve kept straight, and until the crack began, the
slope of F-d curve increased suddenly, meanwhile, the crack propagated slowly
along the grain. Then after the F-d curve rose nonlinearly to the maximum load, it
began decline and it was the time to stop the testing machine, save the curve and
mark the crack tip. The critical load Fcr was the load when the crack began, namely
the point the slope of the F-d curve increased. After the test finished, for internode
specimens, the length an between the marked lines and the centre of load point was
measured.
By Eq. (8.17), the exponential relationship between the compliance and crack
length is obtained, and the determination coefficients (R2) of the internode speci-
mens were all above 0.95. As shown in Fig. 8.38, the fitting curve of compliance
C and crack length a of one certain internode specimen was shown and in Fig. 8.39,
the fitting curve of compliance C and crack length a of all specimens with node was
shown.
So the propagation resistance of Mode III interlaminar crack along grain, namely
the interlaminar fracture toughness can be calculated by Eq. (8.16)
190 8 Interlaminar Fracture Properties of Bamboo

0.0383x
70 y = 1.6422e
2
60 R = 0.8001

50
40
C

30
20
10
0
0 20 40 60 80 100 120
a (mm)

Fig. 8.39 Fitted curve of test curve compliance and crack length of all specimens with node

Table 8.16 Descriptive statistics of the Mode III interlaminar fracture toughness of all internode
specimens
No. Number of GIIIC (internode) Standard deviation Coefficient of
measurements (J/m2) (J/m2) variation (%)
1 6 3015.92 566.72 18.79
2 4 2931.63 223.64 7.63
3 3 2475.14 149.91 6.06
4 4 2141.51 182.15 8.51
5 6 2564.89 314.87 12.28
6 4 2083.59 146.49 7.03
7 6 3275.80 525.33 16.04

8.5.3.2 Results and Analysis

As shown in Tables 8.16 and 8.17, by SCB method, the Mode III interlaminar
fracture toughness of both internode specimens and specimens with node were
calculated by Eq. (8.16) and by descriptive statistics, we got GIIIC
2 2
(internode) = 2702.79 J/m (SD = 549.76 J/m ) with variable coefficient 20.34%;
GIIIC(node) = 7387.09 J/m (SD = 2645.31 J/m2) with variable coefficient 35.81%
2

(Table 8.18). It could be seen that the Mode III interlaminar fracture toughness of
internode specimens was much less than that of specimens with node.
The distribution relationships of crack length a and the Mode III interlaminar
fracture toughness GIIIC was shown in Fig. 8.40. The effective crack length of
specimens with node varied from 38.25 to 95.90 mm; the effective crack length of
internode specimens, from 50 to 100 mm, and from Fig. 8.40, we got that the
Mode III interlaminar fracture toughness GIIIC of internode specimens had no
relationship with the increasing of crack length. According to Jun [45], when crack
length was longer than 40 mm, the Mode III interlaminar fracture toughness
8.5 The Toughness Contribution of Bamboo Node to the Interlaminar … 191

Table 8.17 Mode III interlaminar fracture toughness of all specimens with node
No. b (mm) a (mm) d1 F1 d2 F2 Fcr C (mm/ @C GIIIC(node)
@a
(mm) (kN) (mm) (kN) (kN) kN) (J/m2)
1 12.465 95.90 3.8258 0.1248 9.4913 0.2515 0.3018 44.7094 2.48 9046.74
2 12.73 71.50 4.2003 0.3662 8.6173 0.4839 0.5382 37.5206 0.97 11064.90
3 12.445 63.55 3.9051 0.4174 7.321 0.5593 0.6287 24.0829 0.72 11389.80
4 12.60 59.75 2.0493 0.178 4.7544 0.3259 0.4496 18.293 0.62 4975.14
5 12.525 42.25 1.4932 0.2988 3.4757 0.7062 0.8018 4.86575 0.32 8140.87
6 12.245 49.10 2.3717 0.2052 5.2147 0.4084 0.5603 13.9911 0.41 5286.28
7 13.025 38.25 1.5358 0.2606 2.7647 0.4054 0.6488 8.48467 0.27 4398.69
8 12.465 48.00 1.7783 0.2173 4.3631 0.4396 0.5915 11.6263 0.4 5548.68
9 13.00 45.25 0.9349 0.2042 2.7295 0.505 0.6961 5.96603 0.36 6632.62

Table 8.18 Descriptive statistics of the Mode III interlaminar fracture toughness of all specimens
Sample Number of GIIIC (mean) Standard Coefficient of
measurements (J/m2) deviation (J/m2) variation (%)
Internode 33 2702.79 549.76 20.34
With 9 7387.08 2645.31 35.81
node

- G IIIC (node)
12000
10000
- G IIIC (internode)
GIIIC (J/m2)

8000
6000
4000
2000
0
30 40 50 60 70 80 90 100
a (mm)

Fig. 8.40 Relationship between a and GIIIC of both internode specimens and specimens with node

trended to constant. However, that of specimens with node had no obvious trend
with the increasing of crack length which might be caused by the few amount of
specimens with node.

8.5.4 Discussion

For bamboo interlaminar fracture, no matter Mode I, Mode II, or Mode III, it was
obtained that the interlaminar fracture toughness of bamboo specimens with node
192 8 Interlaminar Fracture Properties of Bamboo

Table 8.19 The three modes Mode G(Node) (J/m2) G(Internode) (J/m2) GIIICðInternodeÞ
of interlaminar fracture GIIICðNodeÞ
toughness of bamboo I 1431.45 498.48 2.87
internode specimens and II 1711.53 1371.54 1.25
specimens with node
III 7387.08 2702.79 2.73

was higher than that of bamboo internode specimens (Table 8.19), which was
caused by the difference in structure between bamboo internode and node.
The bamboo internode wall consists of the outer wall, the middle of wall, and the
inner wall, and the distribution density of vascular bundles from the outer to the

(a) (b) (c)


a a

b b

a-a b-b

Fig. 8.41 Structure of internode and node: a bamboo diametral plane, b cross section of bamboo
internode, c cross section of bamboo node

Fig. 8.42 Reconstruction


graph of bamboo node
3D structure. From [46].
Reprinted with permission
from Journal of Bamboo
Research
8.5 The Toughness Contribution of Bamboo Node to the Interlaminar … 193

Fig. 8.43 The SEM image of the Mode I interlaminar fracture surface in node

inner decreased gradually; on the contrary, the proportion of ground tissue


increased gradually, and in the inner wall, there is only few vascular bundles. It is
known that the cells of bamboo internode are strictly axially arranged and there is
no transverse ray cell (Fig. 8.41). The node of moso bamboo consists of nodal
ridge, sheath scar, and diaphragm [46], which strengthens the bamboo culm to stand
upright and also play the role of transporting water transversely. When vascular
bundles pass through node from internode, they bend inordinately except for the
outermost vascular bundle that breaks off at where sheath scar falls off. Among the
bending of the vascular bundles, some bend slightly outwards or inwards in node,
but when they have passed through, they are recovered straight in internode; the
194 8 Interlaminar Fracture Properties of Bamboo

Fig. 8.44 The SEM image of the Mode II interlaminar fracture surface in node

other vascular bundles change their direction in node; the vascular bundles close to
the inner wall tend to the outer wall and the vascular bundles close to the outer wall
tend to the inner wall oppositely; meanwhile, there are some vascular bundles
entering into the diaphragm circuitously, which incline procumbent circumferen-
tially or interweave together through diaphragm to the opposite wall. The bent
vascular bundles are generally thickened in node; thus, the thickness of the bamboo
wall at the node is thicker than that at internode. As seen in Fig. 8.42, the recon-
struction 3D structure graph of bamboo node reflects the structure feature of vas-
cular bundles vividly [46].
It is known that the strength of cell body biomaterial depends on the thickness of
the cell wall; however, the Mode I interlaminar fracture resistance of bamboo is
determined by the interface property between cells and wall layers. For bamboo
internode, the resistance comes from the interface resistance between ground tissue
and ground tissue, ground tissue and vascular bundle, vascular bundles and vascular
bundles; for bamboo node, the resistance not only comes from the interface
resistance between ground tissue and ground tissue, ground tissue and vascular
bundle, vascular bundles and vascular bundles but also from the contribution of the
fracture of transverse vascular bundles.
The features could be seen from the SEM images of the interlaminar fracture
surface of node. As shown in Figs. 8.43, 8.44, and 8.45, the fracture features of
8.5 The Toughness Contribution of Bamboo Node to the Interlaminar … 195

Fig. 8.45 The SEM image of the Mode III interlaminar fracture surface in node

procumbent fibers and ground tissue on Mode I, Mode II, and Mode III interlaminar
fracture surfaces of the node were shown. It could be seen that the fracture and
peeling of procumbent fibers occurred on all the three types of fracture, and it could
be found that the Mode III interlaminar fracture surface was rougher than that of
Mode II and Mode I. Especially, the damage of ground tissue cells on Mode III
interlaminar fracture surface was obviously more serious than those on Mode II and
Mode I interlaminar fracture surface. The phenomena above coincided well with the
experimental results that the Mode III interlaminar fracture toughness was higher
than Mode II and Mode I interlaminar fracture toughness.

References

1. Zeng QY, Li SH, Bao XR (1992) Effect of bamboo nodal on mechanical properties of
bamboo wood. Sci Silvae Sin 28(3):247–252 (in Chinese)
2. Ahmad M, Kamke FA (2005) Analysis of Calcutta bamboo for structural composite materials:
physical and mechanical properties. Wood Sci Technol 39(6):448–459
3. Obataya E, Kitin P, Yamauchi H (2007) Bending characteristics of bamboo (Phyllostachys
pubescens) with respect to its fiber–foam composite structure. Wood Sci Technol 41:385–400
4. Hodgkinson JM (2000) Mechanical testing of advanced fibre composites. Woodhead
Publishing and CRC Press, Cambridge
196 8 Interlaminar Fracture Properties of Bamboo

5. Triboulot P, Jodin P, Pluvinage G (1984) Validity of fracture mechanics concept applied to


wood by finite element calculation. Wood Sci Technol 18(6):448–459
6. American Society of Testing Materials (ASTM) (2001) Standard test method for Mode I
interlaminar fracture toughness of unidirectional fiber-reinforced polymer matrix composites.
Annual book of ASTM standards. Philadelphia, PA, D 5528-01
7. Yoshihara H, Ohta M (2000) Measurement of Mode II fracture toughness of wood by the
end-notched flexure test. J Wood Sci 46:273–278
8. Reiterer A, Tschegg S (2002) The influence of moisture content on the mode I fracture
behaviour of sprucewood. J Mater Sci 37:4487–4491
9. Ma LF, Ma NX (1997) Study on variation in bamboo wood properties of Phyllostachys
heterocycla var. pubescens. Sci Silvae Sin 33(4):357–364 (in Chinese)
10. Yu ZC, Jiao GQ (1996) The size effects of crack in DCB test of composite materials.
J Aeronaut Mater 16(4):46–53 (in Chinese)
11. National Technical Monitoring Bureau (1995) National standard GB/T 15780–1995: testing
methods for physical and mechanical properties of wood. China Standard Press, Beijing
12. Cook J, Gordon JE, Evans CC, Marsh DM (1964) A mechanism for the control of crack
propagation in all-brittle systems. Proc R Soc 282A:508–520
13. Barette JD, Foschi RO (1977) Mode stress intensity factors for cracked wood. Eng Fract
Mech 9:371–378
14. Russell AJ, Street KN (1985) Moisture and temperature effects on the mixed-mode
delamination fracture of materials. ASTM STP876, p 349
15. Carsson LA, Gillespie JW, Pipes RP (1986) On the analysis and design of the end notched
flexure (ENF) specimen for mode testing. J Compos Mater 6:20
16. Hossein Sadpour, Mehdi Barikani, Mutlu Sezen (2003) Mode-II interlaminar fracture
toughness of carbon/epoxy laminates. Iran Polym J 12(5):389–400
17. Brunner AJ, Blackman BRK, Davies P (2008) A status report on delamination resistance
testing of polymer-matrix composites. Eng Fract Mech 75(9):2779–2794
18. Sham Prasad MS, Venkatesha CS, Jayaraju T (2011) Experimental methods of determining
fracture toughness of fiber reinforced polymer composites under various loading conditions.
J Miner Mater Charact Eng 10(13):1263–1275
19. Li Q, Hu S, Zhu S (1998) Fracture mechanics and its engineering application. Published by
Harbin Engineering University pp 103–104
20. Davies P (ed) (1993) Protocols for interlaminar fracture testing of composites. ESIS Polymers
& Composites Task, Group
21. Davies P, Blackman BRK, Brunner AJ (2001) Mode II delamination. In: Moore DR,
Pavan A, Williams JG (eds) Fracture mechanics testing methods for polymers adhesives and
composites, ESIS publication 28. Elsevier, Amsterdam, pp 307–334
22. Aeronautical Department Standard (2002) Test method of interlaminar fracture toughness of
carbon fiber composite laminates in hot-wet environments. Part II: test method of Mode II
interlaminar fracture toughness. HB 7718-2 (in Chinese)
23. Fonselius M, Riipola R (1992) Determination of fracture toughness for wood. J Struct Eng
118(7):1727–1740
24. Demorais AB, Silva JF, Marques AT et al (2002) Mode II interlaminar fracture of filament
wound angle-ply specimens. Appl Compos Mater 9:117–129
25. Kuboki T, Jar P-YB, Cheng JJR (2004) Interlaminar fracture toughness and the associated
fracture behavior for glass fiber reinforced polymers (GFRP). J Mater Sci 39:1419–1423
26. Hiroshi Y (2005) Mode II initiation fracture toughness analysis for wood obtained by 3-ENF
test. Compos Sci Technol 65:2198–2207
27. Silva MAL, de Moura MFSF, Morais JJL (2006) Numerical analysis of the ENF test for mode
II wood fracture. Compos Part A 37:1334–1344
28. Yao W, Zhong W, Lim CW (2009) Symplectic elasticity. Published by World Scientific
Publishing Co. Pte. Ltd., pp 63–94
29. Fan C, Ben Jar P-Y, Roger Cheng J-J (2006) Revisit the analysis of end-notched flexure
(ENF) specimen. Compos Sci Technol 66:1497–1498
References 197

30. Zhicheng Y (1997) Study on Mode II interlaminar fracture test method of composites.
J Aeronaut Mater 17(4):54–61
31. Xu MQ, Lu ZY (2003) Determination of Loblolly’s shear modulus. Mech Eng 25:57–60
32. Davies P, Kausch HH, Williams JG (1992) Round robin interlaminar fracture testing of
carbon fiber reinforced epoxy and PEEK composites. Compos Sci Technol 43:129–136
33. Davidson BD, Teller SS (2010) Recommendations for an ASTM standardized test for
determining GIIC of unidirectional laminated polymeric matrix composites. J ASTM Int 7
(2):1–11
34. Becht G, Gillespie JW (1988) Design and analysis of the crack rail shear specimen for
Mode III interlaminar fracture. Compos Sci Technol 31:143–157
35. Donaldson SL (1988) Mode III interlaminar fracture characterization of composite material.
Compos Sci Technol 32:225–249
36. Mehrabadi FA, Khoshravan M (2013) Mode III interlaminar fracture and damage
characterization in woven fabric-reinforced glass/epoxy composite laminates. J Compos
Mater 47(13):1583–1592
37. Rizov V, Shindo Y, Horiguchi K et al (2006) Mode III interlaminar fracture behavior of glass
fiber reinforced polymer woven laminates at 293 to 4 K. Appl Compos Mater 13:287–304
38. Szekrényes A (2009) Improved analysis of the modified split-cantilever beam for mode-III
fracture. Int J Mech Sci 51:682–693
39. Khoshravan MR, Moslemi M (2014) Investigation on Mode III interlaminar fracture of glass/
epoxy laminates using a modified split cantilever beam test. Eng Fract Mech 127:267–279
40. Davidson BD, Sediles FO (2011) Mixed-mode I-II-III delamination toughness determination
via a shear-torsion-bending test. Compos Part A: Appl S 42:589–603
41. Johnston AL, Davidson BD, Simon KK (2014) Assessment of split-beam-type tests for mode
III delamination toughness determination. Int J Fract 185:31–48
42. Anderson TL (2005) Fracture mechanics—fundamentals and applications, 3rd edn. CRC
Press, Taylor & Francis Group, Boca Raton, London
43. Szekrényes A (2010) Fracture analysis in the modified split-cantilever beam using the
classical theories of strength of materials. JPCS 240:012030
44. Szekrényes A (2010) Development of an opening-tearing mode fracture system for composite
materials. EPJ Web Conf 6:42008
45. Jun X (1996) Experimental study on Mode III fracture toughness of multi-directional interface
of composite materials. J Nanjing Univ Aeronaut Astronaut 28(2):267–270
46. Yulong D, Liese W (1995) On the nodal structure of bamboo. J Bamboo Res 14(1):24–32
Chapter 9
Modeling on the Toughness Fracture
and Energy-Absorbing Mechanism
of Biomaterial—Bamboo (Phyllostachys
pubescens)

Abstract Studies showed that the whole process of the transverse bending fracture
of bamboo (Phyllostachys pubescens) involved various damage patterns: ground
tissue cracking, interface debonding, fracturing, and pulling-out of fiber bundles.
Different organization structures would contribute different toughness for different
energy wastages in the evolution of damage. In this chapter, the mesomechanics
method was applied to study the energy-absorbing mechanism of the four damage
patterns, the analysis equations of strain energy released rate were obtained, and
theoretical calculation was completed using theoretical model. The results showed
that among the four damage patterns, the work done by pulling fiber bundles out
contributed most to the fracture work, the second contribution was made by the
fracture of fiber bundles, while the energy absorbed by ground tissue cracking and
the initial interface debonding was low. The phenomena above corresponded well
with the characteristic of transverse bending fracture surface obtained by SEM
(scanning electron microscope) technique. In addition, the theoretical result was
very close to the test result of energy consumed in the whole transverse bending
fracture of bamboo. It would have great significance for the design and manufacture
of bionic material to find out the main structural factors that lead to the fine strength
and toughness properties of bamboo.

9.1 Introduction

Bamboo is studied more and more as an example of bionic material, and efforts
have been put into designing and making bionic material due to bamboo’s high
strength and toughness, as well as its regular and simpler construction [1–4].
Nowadays, the study that takes bamboo as a bionic-optimized model is mainly
derived from the following two aspects: one comes from the understanding of the
macroscopic morphology of bamboo, like hollow cylinder and the gradient distri-
bution of reinforced phase [5, 6]; the other comes from the knowledge about the

© Springer Nature Singapore Pte Ltd. 2018 199


Z. Shao and F. Wang, The Fracture Mechanics of Plant Materials,
https://doi.org/10.1007/978-981-10-9017-2_9
200 9 Modeling on the Toughness Fracture and Energy-Absorbing …

microstructure of cell wall of bamboo fiber, just like the spiral distribution of
microfibril and multistory structure [7]. In fact, the hierarchical structure of fiber cell
wall is a common feature of plant fibers, but high strength and toughness of bamboo
depend on both the excellent property of fibers and the perfect organization
structure of bamboo.
Shao et al. [8] and Wang et al. [9] studied the fracture mechanism of Mode I and
Mode II interlaminar fracture toughness of bamboo that was treated as unidirec-
tional fibers reinforced composite. Tan et al. [10] researched the propagation law
and mechanism of crack in bamboo stem by four-point bending test with
single-edge-notched bend (SENB) specimens combined with finite element model.
Habibi et al. [11] applied environmental electron microscopy (ESEM) technique in
the tensile and compress tests of bamboo to study crack path in the longitudinal
direction and radial direction. Habibi et al. [11] gave that the interface zone of
ground tissue cells and fiber cells was the preferred path of crack propagation.
References [12–16] showed that the energy absorption mechanism of the failure of
fiber-reinforced composite with the crack involved matrix cracking, interface
debonding, post-debonding friction, fiber pull-out, fiber fracture and stress redis-
tribution and so on. Kim et al. [12] pointed out that among the failure patterns
above, the toughness contributed by interface debonding and the frictional sliding
during fiber pulling-out played main contribution role for most fiber-reinforced
polymer composites.
Similarly, bamboo internodes stem could be treated as natural polymer com-
posite reinforced by unidirectional fibers that present dense distribution to sparse
distribution from outer part to the inner part of the stem. The strong anisotropy
caused by low interface strength [8] and the unequal distribution of fibers leads to
the complex process of bamboo transverse fracture. And the fracture properties
depend both on the properties of constituent materials and the meso-structure
features including the volume fraction, the distribution law and morphology of
fibers, and the properties of the interface between fiber bundles and ground tissue.
Although Tan et al. [10] have studied the toughness mechanism of crack propa-
gation in bamboo stem by both theoretical analysis and computer model, the work
was only on the energy release rate of interface cracking. Careful observation has
been done by Habibi et al. [11] on crack propagation path in the bamboo stem,
which involved no energy absorption question. As the lack of the basic data on the
mechanical properties of bamboo cells and their interfaces, quantitative analysis on
toughness contribution of bamboo organization structures has not been proposed so
far, and deep understanding on the bionic principle of cell body material
strengthened by fibers is hindered.
In this chapter, theoretical analysis model was built combined with experimental
verification to explore the toughness contribution of bamboo with different orga-
nization structures under different damage patterns. The work will be conducive not
only to the understanding of the toughness mechanism of bamboo, but also to the
research and development of advanced composite with special strength and
toughness.
9.2 Development of Theoretical Model 201

(b)
(a) F Inner part Outer part

H
e
B
S 2 mm
L

Fig. 9.1 Transverse fracture of bamboo: a diagram of bamboo transverse test; b fracture surface
of bamboo SENB specimen

9.2 Development of Theoretical Model

9.2.1 The Damage Patterns of Bamboo Transverse Fracture

Single-edge-notched bend (SENB) specimen with the crack perpendicular to grain


was applied to study the damage patterns in the process of bamboo (Phyllostachys
pubescens) transverse fracture, as seen in Fig. 9.1a, where L  B  H =
160  8.5  13 (mm), e = 3 mm and S = 120 mm. The side surface of the
fractured specimen was shown in Fig. 9.1b, and it could be seen that the fracture
process involved various damage patterns and the interactions of them. The evo-
lution of transverse failure could be described as follows: it began with the cracking
of matrix; when crack tip propagated to fiber bundles, it turned to interface zone
between fiber bundle and ground tissue, and the interface debonding began along
half periphery of vascular bundle to both sides; when the interface crack spread to a
certain length and ceased, fiber bundles fractured where defects or weakness
existed; meanwhile, the load on fiber bundles transferred to the material nearby and
caused the redistribution of internal stress, and then the fractured fiber bundles were
pulled-out from one side of the fracture surface.
It needed to be pointed out that when crack tip spread to fiber bundles, the crack
would not spread forward across the fiber bundles to form “fiber bridge”, but spread
to both sides to seek the access with the lowest energy where the weakest interface
existed. So the whole process of bamboo transverse fracture included at least four
patterns: ground tissue cracking, interface debonding, the fracture of fiber bundles,
and the pulling-out of fiber bundles.

9.2.2 The Simplification of Bamboo Mechanical Model

In order to analyze and calculate the energy absorbed by the four damage patterns
above from the angle of mesomechanics, simplification was done to bamboo
components. As seen in Fig. 9.2, in a vascular bundle, the outer, side, and inner
202 9 Modeling on the Toughness Fracture and Energy-Absorbing …

G G

V V
Simplify
V V V V

100um 100um

Fig. 9.2 The simplification of bamboo vascular bundle. G Ground tissue; V vessel

Fig. 9.3 The cross section


(10 mm  8.5 mm) of
bamboo stem

2mm

fibrous sheaths, numbered as fiber bundle ①, ②, and ③, were simplified as


semicircles and circle in shape with radius of R1, R2 (it was assumed the two side
fibrous sheaths were of the same size) and R3 respectively. According to the sta-
tistical analysis results of the area of each fiber bundle, the average areas were A1,
pR2 pR2
A2, and A3. As A1 ¼ 2 1 , A2 ¼ 2 2 and A3 ¼ pR23 , the corresponding R1, R2, and R3
could be obtained. For example, as seen in Fig. 9.3, the average areas of fiber
bundle ①, ②, and ③ were A1 = 19531.82 um2, A2 = 17698.55 um2, and
A3 = 25972.89 um2 on the cross section in the area A = 85 mm2, so we got
R1 = 111.54 um, R2 = 106.17 um, and R3 = 90.95 um.

9.3 Theoretical Analysis

9.3.1 Ground Tissue Cracking

For the deforming and cracking of ground tissue, the fracture work could be cal-
culated by a two-dimensional model, as seen in Fig. 9.4. It was assumed that the
average line densities of ground tissue and vascular bundles were km and DV
respectively. The volume fraction of ground tissue and vascular bundles were Vm
9.3 Theoretical Analysis 203

Fig. 9.4 Cracking model of


ground tissue
Vascular bundle

Crack

Ground tissue

Dv m Dv m Dv m
Dv

and (Vf + VT), and the average outer diameter of vascular bundle was DV = 2RV,
where Vf and VT were the volume fraction of fibers and conducting tissues. The
geometrical relationship of the parameters above was km =Vm ¼ Dv =ðVf þ VT Þ. As
ground tissue cell was a porous mass with thin wall, when loaded with tensile force,
the stress and strain of cell wall presented an approximately linear relation. Thus,
the energy absorbed by per unit volume of ground tissue was emu rmu =2 ¼
r2mu =ð2Em Þ. So, in the process of bamboo transverse fracture, the energy absorbed
by ground tissue to form unit fracture area was

r2mu r2mu Vm2 Dv r2mu Vm2 Rv


Um ¼  V m km ¼ ¼ ð9:1Þ
2Em 2Em ðVf þ VT Þ Em ðVf þ VT Þ

where Em was the elastic modulus of ground tissue and rmu, emu were the ultimate
stress and strain.

9.3.2 Interface Debonding

When crack tip propagated to fiber bundles, it turned to interface zone of fiber and
ground tissue cells as shown in Fig. 9.5. According to the energy release rate
criterion of fracture mechanics, the interface debonding region was treated as
interface crack, and the propagation of crack depended on the energy balance of
interface fracture energy [16–18]. The researches of Charalambides et al. [19] and
Tan et al. [10] indicated that the energy release rate of interface crack depended on
load and had no relation with crack length. According to Williams [20, 21], Wang
and Williams [22] and Reeder [23, 24], the strain energy release rate formula of
interlaminar crack propagation of laminate composite was derived as (Chap. 4)
204 9 Modeling on the Toughness Fracture and Energy-Absorbing …

Fig. 9.5 Interface debonding


model of bamboo
Ground tissue

Ld

Vessel

" #
1 M12 M22 2
Gi ¼ þ  ðM1 þ M2 Þ ð9:2Þ
2BEI0 ðh1 =HÞ3 ð1  h1 =HÞ3

where M1 and M2 were the moment at the tip of interlaminar crack, and I0 = BH3/
12.
For bamboo transverse fracture specimen (Fig. 9.6), M1 = 0, the bending
moment at crack tip was M = F(S − Ld)/4. So, the energy release rate of bamboo
interlaminar crack could be calculated by the following formula:
" #
3F 2 ðS  Ld Þ2 1
Gi ¼ 1 ð9:3Þ
8EB2 H 3 ð1  e=HÞ3

From Eq. (9.3), it was derived that strain energy release rate decreased with the
increasing of debonding length Ld. When Gi was less than the critical strain energy
release rate of bamboo, the interlaminar crack would stop propagating. Especially,
when Ld = 0, namely the moment the transverse crack converted to side crack along
the grain, the critical strain energy released rate could be calculated by the fol-
lowing formula:
" #
3Fcr2 S2 1
Gi0 ¼ 1 ð9:4Þ
8EB2 H 3 ð1  e=HÞ3

Fig. 9.6 Diagram of F


interlaminar fracture of
bamboo beam with the crack
perpendicular to the grain Lcd H

S
L
9.3 Theoretical Analysis 205

where Fcr was the load when the crack along the grain began, E was the bending
elastic modulus of bamboo SENB specimen. In general, the duration of the prop-
agation of side crack was very short, so the load during the process of side crack
propagation was taken as F  Fcr.

9.3.3 The Fracture of Fiber Bundle

In the process of bamboo transverse fracture, when crack tip turned to the interface,
about half periphery of vascular bundle detached from the cluster of ground tissue,
so the homogenizing transfer of the load on fiber bundle was weakened. Under the
bending load, fiber bundle was elongated which caused the deformation of ground
tissue around the other half periphery of the vascular bundle. When the interface
crack propagated to a certain length, fiber bundles would fracture and recoil to
release elastic deformation energy. As the elastic modulus (Ef) of the fiber bundle is
much higher than that (Em) of ground tissue, when the load on fiber bundle
transferred to ground tissue after the sudden fracture of the fiber bundle, the
redistribution of strain in ground tissue was caused. Strain energy is an another
energy-absorbed mechanism, namely the fracture work of fiber bundle [25, 26].
To calculate the fracture work of fiber bundle, it was supposed that the fracture
surface of fiber bundle happened where Lp/2 far from the main crack surface (as
seen in Fig. 9.7). The elastic potential energy in a length of dx was
2
1 rfðxÞ
dUf ¼ rfðxÞ efðxÞ  Af dx ¼  Af dx ð9:5Þ
2 2Ef

where Af was the cross-sectional area of fiber bundle.

Fig. 9.7 The fracturing and


pulling-out model of the fiber
bundle

dx
Lp /2
u(x) x
206 9 Modeling on the Toughness Fracture and Energy-Absorbing …

After fiber bundle fractured, the work done by the diastrophism between fiber
bundle element and ground tissue that undetached from fiber bundle was

dUfm ¼ pRf sdx  uðxÞ ð9:6Þ

where Rf was the radius of fiber bundle, u(x) was the slipping distance between fiber
bundle element and ground tissue, and it was

Zx Zx
rfðxÞ  
uðxÞ ¼ efðxÞ dx ¼ dx 0  x  Lp =2 ð9:7Þ
Ef
L L
 2p  2p

The total fracture work of fiber bundle element should be the sum of dUf and
dUfm, so the energy consumed by the fracture of a single fiber bundle was

Ufu ¼ Uf þ Ufm ð9:8Þ


 
L pR2
For fiber bundle ①, as pR1 p
2 þ x ss ¼ 2 1 rfðxÞ , we obtained

1 r2fðxÞ pR21 ps2s ðL


p þ 2xÞ
2
dUf ¼ rfðxÞ  efðxÞ  A
f  dx ¼   dx ¼  dx ð9:9Þ
2 2Ef 2 4Ef


ps2s ðL
p þ 2xÞ
2
dUfm ¼ pR1 ss dx  uðxÞ ¼  dx ð9:10Þ
4Ef

So,

L
p
Z2

ps2s ðL
p þ 2xÞ
2
2ps2s ðL

3
Ufu ¼ Uf þ Ufm

¼2  dx ¼ ð9:11Þ
4Ef 3Ef
L
p
 2

In the same way, the energy consumed by the fracture of fiber bundles ② and ③
was

L`
p
Z2
`
ps2s ðL`
p þ 2xÞ
2
Ufu ¼ Uf` þ Ufm
`
¼2  dx ð9:12Þ
4Ef
L`
p
 2
9.3 Theoretical Analysis 207


p
Z2
´
ps2s ðL´
p þ 2xÞ
2
Ufu ¼ Uf´ þ Ufm
´
¼2  dx ð9:13Þ
4Ef

p
 2

Hence, for the fracture of a vascular bundle, the fracture work was
 ` ´
v
Ufu ¼ Ufu þ 2Ufu þ Ufu ð9:14Þ

9.3.4 The Pulling-Out of Fiber Bundle

After the bamboo fiber bundles were fractured, as the bending continued, the
fractured fiber bundles were pulled-out from the crack surface. As seen in Fig. 9.7,
when fiber bundle element was pulled-out from the cluster of ground tissue, the
resistance was pRf ðLp =2  xÞsu , where Rf was the radius of fiber bundle. So, the
work done to pull a single fiber bundle out was

L
p
Z2
 1 pR1 ss ðL

2
Ufp ¼ pR1 xss  dx ¼ ð9:15Þ
2 16
0

Z L`
pR2 ss ðL`
p 2
`
2

Ufp ¼ pR2 xss  dx ¼ ð9:16Þ
0 8

Z L´
pR3 ss ðL´
p 2
´
2

Ufp ¼ pR3 xss  dx ¼ ð9:17Þ
0 8

Hence, for the whole vascular bundle, the work done to pull fiber bundles
out was
 ` ´
v
Ufp ¼ Ufp þ Ufp þ Ufp ð9:18Þ

9.3.5 On the Calculation of Lp

By literature retrieval, for artificial fiber-reinforced composites, there are many


analyses and discussion on the interface debonding length, pull-out length and the
volume fraction of fiber pulled-out, and relationship between the length above and
the embedded length of fiber [12, 14, 15, 27–30]. However, as bamboo is a
208 9 Modeling on the Toughness Fracture and Energy-Absorbing …

Fig. 9.8 The macroscopic


fracture surface of bamboo
transverse fracture specimen
(Unit cm)

functional gradient biomaterial reinforced by natural fiber, some of the analysis


models for artificial fiber-reinforced composites are not applicative for bamboo [12,
31, 32] for the unique morphology and distribution of bamboo fibers. It is still
cheering that the abundant research background will provide valuable reference for
the study of bamboo.
In the process of bamboo-bending fracture, with the propagation of interface
crack, the half periphery of vascular bundle was detached from the cluster of ground
tissue. When the interface crack propagated to a certain length, fiber bundles
fractured at that moment the interface shear resistance pRf (LC − x)s was in equi-
librium with the tensile force on fiber bundle element [33]. When x = 0, the fracture
length of fiber bundle was maximum.
So, for fiber bundle ①, ②, and ③, the maximum fracture length was
pR1 L pR2 R1 rf
L
p
 ss  ¼ 1  rf ; p ¼ ð9:19Þ
2 2 4 ss

L` pR22 R2 rf
L`
p
pR2 ss  ¼  rf ; p ¼ ð9:20Þ
2 2 ss

L´ pR23 2R3 rf

p
pR3 ss  ¼  rf ; p ¼ ð9:21Þ
2 2 ss

It could be seen from the formula above that the fracture length was in relation to
the size of fiber bundle. Because the size and morphology of fiber bundle varied
gradually on the radial direction of bamboo stem, the fracture length of fiber
bundles in the outer part of bamboo stem was longer than that of fiber bundles in the
inner part (Fig. 9.8).

9.4 Results

9.4.1 Result of Experiment

Moso bamboo (P. pubescens) was taken as a sample, 4 years old, collected from Lu
Jiang, Anhui Province. After removing the cortex and pith periphery, SENB
specimens with the crack perpendicular to grain were made (Fig. 9.1a), and there
9.4 Results 209

Fig. 9.9 Load-deflection F

curve of bamboo transverse 800 Fmax


H
bending fracture test 700 e

600 S
B

L
500

Load/N
400 F0cr
300
200
100
0
0 5 10 15 20 25
Deflection/mm

Table 9.1 The size and No. H-e (mm) B (mm) UText (J)
fracture energy of SENB
specimens 1 9.44 8.59 6.4409
2 9.89 8.63 6.9065
3 9.58 8.38 6.6264
4 10.16 8.53 6.9101
5 10.17 8.87 5.2684
6 9.8 8.72 5.3467
7 9.49 8.44 7.2042
8 10.08 8.38 6.4788
9 9.62 8.84 5.6438
10 9.9 8.69 5.8968

were 10 specimens in total with a moisture content of 11%. The temperature was
15–18 °C and the humidity was 60–65% in the laboratory. Three-point bending
tests were performed on a computer-controlled testing machine with a loading
speed of 2 mm/min, and the load–deflection (F-d) curve was recorded by computer
automatically during the test (Fig. 9.9). When the interface debonding began, the
average critical load was Fcr = 400 N. The total energy consumed in the whole
bending failure process could be calculated by the following formula. The size and
test value of each specimen were shown in Table 9.1. And the average test value
was Utest = 6.272 J (SD = 0.689 J).

ZSu
Utest ¼ FðsÞds
0

9.4.2 Theoretical Result

The parameters in the theoretical model were actually measured according to Shao
et al. [34, 35], as seen in Table 9.2. The fracture length of fiber bundle ①, ②, and
210 9 Modeling on the Toughness Fracture and Energy-Absorbing …

Table 9.2 Parameters of bamboo and its components


A Vf Vm Vt Rf rf N rm Ef Em s E
(mm2) (%) (%) (%) (mm) (MPa) (MPa) (MPa) (MPa) (MPa) (MPa)
85 0.307 0.613 0.08 0.2275 523 138 51.67 40,351 1620 18 10,742

③ were L ` ´
p = 3.24 mm, Lp = 3.09 mm, and Lp = 5.28 mm due to Eqs. (9.19)–
(9.21). The critical load was Fcr = 400 N, when interface debonding began in the
transverse fracture process of bamboo SENB specimen.
Due to Eq. (9.1), the energy consumed by the cracking of ground tissue was

Umtotal ¼ A  Gm ¼ 0:038 J ð9:23Þ

The interface debonding could be treated as interface crack, and here it was
treated as Mode I interlaminar crack. According to Shao et al. [8], the Mode I
interlaminar fracture toughness of bamboo was GIC = 0.385 N/mm, so the fol-
lowing equation was derived:
" #
3F 2 ðS  Ld Þ2 1
Gi ¼  1 ¼ GIC ð9:24Þ
8EB2 H 3 ð1  e=HÞ3

It was obtained that Ld = 27.81 mm by substituting the parameters shown in


Table 9.2 into Eq. (9.24). According to Eq. (9.4), the energy consumed by number
i crack face in the process of transverse bending fracture could be calculated as
follows:

ZLd =2 ZLd =2 " #


3 1
Uii ¼ 2 Gi Bdd ¼  1 F 2 ðS  2dÞ2 dd
4EBH 3 ð1  e=HÞ3
0 0
" # ð9:25Þ
F 2 ½S3  ðS  Ld Þ3  1
 1
8EBH 3 ð1  e=HÞ3
¼ 0:113 J

The energy consumed by the cracking of a single layer was not high, thus no
significant change would be found on the load–deflection (F-d) curve of
transverse-bending fracture. However, multilayer cracking would happen, so much
energy would be consumed in the whole process of transverse-bending failure.
Because of the discrete distribution of fiber bundles, the laminated crack was not
straight or continuous, which brought difficulty to the calculation of the energy
consumed by multilayer cracking. As seen in Fig. 9.4, the distribution of vascular
bundles on the cross section upon the crack was shown. If it was possible that
during crack propagation, each vascular, the crack encountered, would cause a new
layering, thus it was supposed that there were about 14 layers that laminated
9.4 Results 211

cracking happened, because there were about 14 vascular bundles through the
center line of the cross section of bamboo specimen. The energy consumed by the
cracking of all layers in the process was

X
14
Uitotal ¼ Uii  1:578 J ð9:26Þ
i¼1

Due to Eqs. (9.11)–(9.14), the energy consumed by the fracture of fiber bundles
was
 ` ´
total
Ufu ¼ N  Ufu
v
¼ N  ðUfu þ 2Ufu þ Ufu Þ ¼ 1:628 J ð9:27Þ

And due to Eqs. (9.15)–(9.18), the energy consumed by the pulling-out of


fractured fiber bundles was
 ` ´
total
Ufp ¼ N  Ufp
v
¼ N  ðUfp þ Ufp þ Ufp Þ ¼ 4:029 J ð9:28Þ

So, the total energy consumed by four fracture patterns in the whole process was

Umodel ¼ Umtotal þ Uitotal þ Ufu


total
þ Ufp
total
¼ 7:273 J ð9:29Þ

Compared the theoretical result Umodel = 7.273 J with the result of test
Utest = 6.272 J, the theoretical result was 15.68% higher than the test result.

9.4.3 Analysis and Discussion

Among the four damage patterns in the process of bamboo transverse fracture, the
work done by pulling the fiber bundles out contributed most to the fracture work,
then it was work done by the fracture of fiber bundles and interface debonding, and
the energy absorbed by ground tissue cracking was low. The phenomena above
corresponded well with the characteristic of the transverse bending fracture surface.
The SEM images of the transverse bending fracture surface were shown in
Fig. 9.10. Figure 9.10a, b was the fracture surface of the ground tissue nearby the
inner part of bamboo, and it could be seen that ground tissue was torn transversely,
and the fracture surface was flat, so the energy absorbed was low. Figure 9.10c–e
was the SEM images of the fiber bundle pulled-out. Although single fiber cell was
presented at the fracture surface, the surface was flat relatively on the whole. It
could also be observed that there were a large amount of torn fragments on the side
face of the fiber bundle. Comparatively speaking, the fracture of fiber bundles with
high modulus did not absorb much energy and so to interface debonding for the
weak interface tensile strength, but the shearing failure caused by the pulling-out of
212 9 Modeling on the Toughness Fracture and Energy-Absorbing …

(a)
(b)

(c) (d)

(e)

Fig. 9.10 SEM images of bamboo transverse-bending fracture surface. a Ground tissue fractured
transversely; b the flat fracture surfaces of ground tissue; c fiber bundles were pulled out;
d fragments on the side of fiber bundle; e separated single fiber

the fractured fiber bundle from the ground tissue did absorb a considerable amount
of energy. The fracture surfaces demonstrated the validity of the theoretical model.

9.5 Verification

As mentioned in Sect. 4.2, the interface debonding could be treated as interface


crack, and it was treated as Mode I interlaminar crack. To verify the validity of the
assumption, namely Eq. (9.24), theoretical analysis and experiments were done.
9.5 Verification 213

y
F
r x

Fig. 9.11 The side crack of bamboo SENB specimen was assumed to be Mode I (opening mode)
crack

9.5.1 Theoretical Analysis

Finite element was applied to analyze the crack propagation model of bamboo
SENB specimen, and it was proved that there were both normal stress and shear
stress at the tip of side crack, so the propagation of side crack was I + II mixed
mode. In general, the Mode II interlaminar fracture toughness (GIIC) of composites
was much higher than its Mode I interlaminar fracture toughness (GIC) [36], so as to
bamboo [8, 9]. Thus, from the angle of energy, for bamboo, it was thought that
when the energy release rate reached the value of GIC which was much lower than
that of GIIC, crack would preferentially propagate in Mode I pattern, as seen in
Fig. 9.11.
To measure the Mode I interlaminar fracture toughness, tests were performed
according to Shao et al. [8]. The double cantilever beam (DCB) specimen was
connected with steel U hook by steel pin with a crosshead speed between 1 and
5 mm/min. The crosshead speed was between 1 and 5 mm/min, and curves of the
applied load versus opening displacement (F-d) were recorded automatically by the
computer during the test. And once load (F) went down, the load machine is
stopped, and the crack tip at both the inner part and the outer part of DCB specimen
was marked. Then, it was unloaded and reloaded and the same procedure was
repeated until the DCB specimen fractured completely. As seen in Fig. 9.12, it
showed the F-d curves of one DCB specimen and the loading procedure was
repeated eight times, so there were eight curves and each curve corresponded to
different crack lengths. The specimen from test machine was taken off, the crack
length was measured. As resistance-arresting crack propagation of out layer
(ai − outer) is lower than that of inner layer (ai − inner) of bamboo, the average crack
length of two sides was taken as the actual crack length, namely  ai ¼ aiinner þ2 aiouter .
Due to Shao et al. [8], the strain energy U and its release rate GIC can be
calculated as follows:
214 9 Modeling on the Toughness Fracture and Energy-Absorbing …

Fig. 9.12 The F-d curves of 0.1


one DCB specimen of 0.09
bamboo 0.08
0.07
0.06

F (kN)
0.05
0.04
0.03
0.02
0.01
0
0 2 4 6 8 10 12 14 16
δ (mm)

1 1
U ¼ Fd ¼ F 2 C ð9:30Þ
2 2

@U Fcr2 @C
GIC ¼ ¼  ð9:31Þ
@A 2b @a

where d is load point deflection, C = d/F the compliance of DCB specimen.


Due to GB/T15780-1995 [37], the elastic modulus of bamboo was tested on a
computer-controlled testing machine by three-point bending method, the span
length was l = 120 mm, the crosshead speed was 2 mm/min. The upper and lower
limit of the load was 160–60 N, and the load procedure was repeated two times, as
seen in Fig. 9.13.
According to GB/T15780-1995 [37], two points (x1, y1) and (x2, y2) in the linear
region on the F-deflection curve were collected, then the elastic modulus could be
calculated as

l 3 ð y2  y1 Þ
E¼ ð9:32Þ
4bh3 ðx2  x1 Þ

For bamboo SENB specimen, the size was 160(L)  13(H)  b (Thickness)
(mm). The average value of b was 6.56 mm, and e = 3 mm, and span length
S = 120 mm. According to Eq. (9.31), the Mode I interlaminar fracture toughness
of bamboo could be obtained, and the average value was GIC = 484.22 N/m
(SD = 74.34 N/m). The average value of elastic modulus was E = 8811.20 MPa
(SD = 1051.07 MPa). The average value of load when side crack initiated was
F0 = 310.73 N. Thus, substituted the parameters above into Eq. (9.3), then the
theoretical maximum length of side crack could be obtained, Ld = 23.53 mm.
Experiments were performed to verify the rationality of theoretical result.
9.5 Verification 215

Fig. 9.13 The F-deflection 0.18


curve of the bamboo 0.16
specimen that used to test
0.14
elastic modulus, where F is
force and deflection is the 0.12
displacement of the load point 0.1

F (kN)
0.08
0.06
0.04
0.02
0
0 0.2 0.4 0.6 0.8 1 1.2 1.4
Deflection (mm)

9.5.2 Experiments

The SENB specimens were tested on a computer-controlled testing machine by


three-point bending method with a loading speed of 2 mm/min, and the load–load
point deflection (F-deflection) curve was recorded by the computer automatically
during the test (Fig. 9.14). During testing, the propagation of side crack is observed
closely, and once it ceases, simultaneously the length of side crack by the tick mark
is read which is prepared in advance as seen in Fig. 9.15 and recorded.
As seen in Fig. 9.14, it could be seen from the F-d curve that there were three
stages in the whole fracture process of bamboo SEB specimen. In stage OA, the
deformation of specimen was linear elastic and crack did not propagate. Stage AB
included the initiation, propagation, and cease of side crack. The crack was initiated
at point A where was just the side crack began due to observation, thus load at that
point was the critical load F0. With the propagation of side crack, the stiffness of
specimen decreased until point B when side crack ceased. Then fiber bundles
fractured, crack turned along the contrary direction of loading direction. And once
fiber bundles are encountered, it would turn along the fiber direction, then a new
side crack is initiated, propagated, and ceased. The processes were repeated until
the specimen got fractured completely, just as many stepwise decreases presented in
stage BC. As seen in Fig. 9.16, it was the fracture surface of one failure SENB
specimen and the features of fracture surface were in accordance with F-d curve.
As seen in Fig. 9.17, the length of side crack on the outer surface was always
longer than that on the inner surface of specimen, which was caused by the gradient
distribution of fibers along the radial direction of the bamboo stem, dense in the
outer part then sparse gradually in the inner part. And the energy absorbed by the
deformation or damage of ground tissue cells would determine the resistance that
the propagation of side crack should overcome. By experimental measurement, the
average length of crack on the outer surface was 25.27 mm, the average length on
the inner surface was 17.50 mm, and the average of the length in both parts was
21.39 mm, as seen in Table 9.3.
216 9 Modeling on the Toughness Fracture and Energy-Absorbing …

0.6 B

0.5

0.4
Load (kN)

0.3

0.2

0.1
C

O
0
0 2 4 6 8 10 12 14 16 18
Deflection (mm)

Fig. 9.14 The transverse fracture F-deflection curve of one SENB specimen

Fig. 9.15 The propagation of side crack during the transverse fracture of bamboo SENB
specimen (Unit mm)

9.5.3 Analysis and Discussion

The theoretical length of side crack in the transverse fracture process of SENB
specimen was 23.53 mm, which was a little longer than the experimental measured
value 21.39 mm, about 10%, which could be explained by the comparison analysis
on the fracture surfaces.
As seen in Fig. 9.18, the scanning electron microscope (SEM) images of the
fracture surfaces of side crack and Mode I interlaminar crack were shown. It could
be seen that the surface of ground tissue presented to be hackle on the whole in the
side crack zone of bamboo SENB specimen, while the surface of most ground
tissue cells was flat and only a few ground tissue cells were damaged (Fig. 9.18a).
The failure characteristics of fibers in the side crack zone of bamboo SENB
specimen was shown in Fig. 9.18b and the ribbon-like traces on the surface of the
damaged intercellular layer of fiber cells could be observed obviously. The feature
9.5 Verification 217

15 10 15
10
5 5

Fig. 9.16 The fracture surface of failure bamboo SENB specimen (Unit mm)

35
Outer side
30 Inner side
25
C (mm)

20
15
10
5
0
1 2 3 4 5 6 7 8 9 10 11
Specimen

Fig. 9.17 The length of side crack on the outer and inner surfaces of bamboo SENB specimen

Table 9.3 The experimentally measured results of bamboo SENB specimens


No. H (mm) b (mm) e (mm) F0 (N) Fmax (N) Le (mm)
1 13.38 6.72 3 316.71 600.46 18.50
2 13.18 6.06 3 265.04 405.57 16.50
3 12.91 6.6 3 334.25 537.91 21.00
4 13.26 6.96 3 355.89 593.44 26.50
5 12.84 6.33 3 254.28 424.39 20.50
6 13.33 6.75 3 348.25 683.64 20.75
7 12.56 6.65 3 325.45 626.71 24.25
8 13.00 6.58 3 305.45 578.45 19.75
9 13.67 6.97 3 367.29 751.64 24.00
10 12.16 6.36 3 285.08 545.61 22.00
11 12.84 6.29 3 262.35 471.74 21.50

of ground tissue in the Mode I crack zone of bamboo DCB specimen was similar to
that in the side crack zone of bamboo SENB specimen (Fig. 9.18c). However, the
surface of fiber cells in the Mode I crack zone was much more smooth than that in
the side crack zone as seen in Fig. 9.18d that was in accordance with Shao et al. [8].
218 9 Modeling on the Toughness Fracture and Energy-Absorbing …

(a) (b)

Cracking direction
(c) (d)

Cracking direction
Fig. 9.18 SEM images of fracture surfaces: a, b The fracture surfaces of ground tissue and fibers
in the side crack zone of bamboo SENB specimen; c, d The fracture surfaces of ground tissue and
fibers in the Mode I zone of bamboo DCB specimen

By comparison analysis on the fracture surfaces, it could be seen that fracture


surface of side crack of SENB specimen was a little rougher than Mode I fracture
surface of DCB specimen, especially the fracture surface of fibers cells. Thus, it was
derived that the energy absorbed by the propagation of side crack was higher than
that absorbed by the propagation of Mode I crack. As the experimentally measured
average crack length of side crack was Le = 21.39 mm, when substituted Le into
Eq. (9.3), the energy absorbed by the propagation of side crack could be calculated
and it was Gi = 506.00 N/m, which was 4.30% higher than GIC = 484.22 N/m.
However, Gi was replaced by GIC to calculate the theoretical length (Ld) of side
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
crack in theoretical analysis, and due to Eq. (9.3), Ld ¼ S   8EB
Gi 2 H3
1
1 3F 2 and
ð1e=HÞ3

Gi > GIC, so the theoretical result (Ld) was slightly higher than the experimentally
measured result (Le).
However, in practice, it is difficult to determine the length of side crack, the
energy absorbed by the propagation of side crack and its fracture mechanism
because of the complicated transverse fracture process and jagged fracture surface
of bamboo or FRP. Due to the observation on the propagation behavior of side
crack of SENB specimen, it can be treated as the propagation behavior of Mode I
crack to evaluate the crack length of side crack and energy absorbed by the
9.5 Verification 219

propagation of side crack of bamboo or FRP. After all, for the Mode I interlaminar
fracture of both bamboo and FRP, the test method is quite mature [38–41].

9.6 Conclusion

The energy-absorbing mechanism of ground tissue cracking, interface debonding,


fracturing, and pulling-out of fiber bundles in the process of the transverse bending
fracture of bamboo was studied by the mesomechanics method, and the analysis
equations of strain energy released rate were obtained. The results indicated that
among the four damage patterns, the work done by pulling fiber bundles out
contributed most to the fracture work, the second was the work done by the fracture
of fiber bundles, and the energy absorbed by ground tissue cracking and the initial
interface debonding was low. The phenomena above corresponded well with the
characteristic of transverse bending fracture surface obtained by SEM technique. By
comparison, the theoretical result was very close to the test result of energy con-
sumed in the whole transverse bending fracture of bamboo. It indicated that the
theoretical model of transverse bending fracture suggested in this chapter was
feasible.

References

1. Ma JF, Chen WY, Zhao L et al (2008) Elastic buckling of bionic cylindrical shells based on
bamboo. J Bionic Eng 5:231–238
2. Xing DH, Chen WY, Xing DJ et al (2013) Lightweight design for thin-walled cylindrical shell
based on action mechanism of bamboo node. J Mech Design 135:014501
3. Cheng WM, Fu WG, Zhang ZQ et al (2012) The bionic lightweight design of the mid-rail box
girder based on the bamboo structure. Prz Elektrotechniczn 88:113–117
4. Xie F, Zhang ZP, Li X et al (2013) Bamboo section’s bionic design based on bamboo. AMM,
pp 385–386, 34–38. ISSN: 1660-9336
5. Ghavami K (2005) Bamboo as reinforcement in structural concrete elements. Cement
Concrete Comp 27:637–649
6. Ghavami K, Rodrigues CS, Pacionik S (2003) Bamboo: functionally graded composite
material. Asian J Civil Eng 4:1–10
7. Parameswaran N, Liese W (1976) On the fine structure of bamboo fibres. Wood Sci Technol
10:231–246
8. Shao ZP, Fang CH, Tian GL (2009) Mode I interlaminar fracture property of moso bamboo
(Phyllostachys pubescens). Wood Sci Technol 43:527–536
9. Wang FL, Shao ZP, Wu YJ (2013) Mode II interlaminar fracture properties of moso bamboo.
Compos Part B—Eng 44:242–247
10. Tan T, Rahbar N, Allameh SM et al (2011) Mechanical properties of functionally graded
hierarchical bamboo structures. Acta Biomater 7:3796–3803
11. Habibi MK, Lu Y (2014) Crack propagation in bamboo’s hierarchical cellular structure. Sci
Rep-UK 4:5598
12. Kim JK, Baillie C, Mai YW (1991) Interfacial debonding and fibre pull-out stresses: Part I
critical comparison of existing theories with experiments. J Mater Sci 27:3143–3154
220 9 Modeling on the Toughness Fracture and Energy-Absorbing …

13. Zhou LM, Kim JK, Mai YW (1992) Interfacial debonding and fibre pull-out stresses: Part II A
new model based on the fracture mechanics approach. J Mater Sci 27:3155–3166
14. Wang C (1997) Fracture mechanics of single-fibre pull-out test. J Mater Sci 32:483–490
15. Li LB (2015) Modeling the effect of interface wear on fatigue hysteresis behavior of carbon
fiber-reinforced ceramic-matrix composites. Appl Compos Mater 22:887–920
16. Meng QH, Wang ZQ (2015) Theoretical analysis of interfacial debonding and fiber pull-out in
fiber-reinforced polymer-matrix composites. Arch Appl Mech 85:745–759
17. Kim JK, Mai YW (1991) High strength, high fracture toughness fibre composites with
interface control a review. Compos Sci Technol 41:333–378
18. Chiang YC (2001) On fiber debonding and matrix cracking in fiber-reinforced ceramics.
Compos Sci Technol 61:1743–1756
19. Charalambides PG, Lund J, Evans AG et al (1989) A test specimen for determining the
fracture resistance of bimaterial interfaces. J Appl Mech 56:77–82
20. Williams JG (1988) On the calculation of energy release rates for crack laminates. Int J Fract
36:101–119
21. Williams JG (1989) End corrections for orthotropic DCB specimens. Compos Sci Technol
35:367–376
22. Wang Y, Williams JG (1992) Corrections for mode II fracture toughness specimens of
composite materials. Compos Sci Technol 43:251–256
23. Reeder JR, Crews JH (1990) Mixed-mode bending method for delamination testing. AIAA
28:1270–1276
24. Reeder JR and Crews JH (1991) Redesign of mixed-mode bending test for delamination
toughness. In: Proceedings of ICCM-8 Section, pp 13–36
25. Piggott MR (1970) Theoretical estimation of fracture toughness of fiber composites. J Mater
Sci 5:669
26. Beaumont PWR, Fitz-Randolph J, Phillips DC et al (1971) Acoustic emission studies of a
boron-epoxy composite. J Compos Mater 5:542–548
27. Thouless MD, Evans AG (1988) Effects of pull-out on the mechanical properties of
ceramic-matrix composites. Acta Metall 36:517–522
28. Hsueh CH (1994) Pull-out of a ductile fibre from a brittle matrix: PART II a simplified mode.
J Mater Sci 29:5135–5140
29. Guo SQ, Kagawa Y (2002) Tensile fracture behavior of continuous SiC fiber-reinforced SiC
matrix composites at elevated temperatures and correlation to in situ constituent properties.
J Eur Ceram Soc 22:2349–2356
30. Gao XL, Li K (2005) A shear-lag model for carbon nanotube-reinforced polymer composites.
Int J Solids Struct 42:1649–1667
31. Gao YC (1987) Debongding along the interface of composites. Mech Res Commun 14:67–72
32. Li LB, Song YD, Sun ZG (2008) Uniaxial tensile behavior of unidirectional fiber reinforced
ceramic matrix composites. Acta Materiae Compositae Sinica 4:154–160 (in Chinese)
33. Qiao RS (1997) The meso-mechanics property of composite. Northwestern Polytechnical
Press, pp 52–62 (in Chinese)
34. Shao ZP, Fang CH, Huang SX et al (2010) Tensile properties of Moso bamboo (Phyllostachys
pubescens) and its components with respect to its fiber-reinforced composite structure. Wood
Sci Technol 44:655–666
35. Shao ZP, Zhou L, Liu YM et al (2010) Difference of structure and strength between
internodes part and node part of Moso bamboo. J Trop For Sci 22:133–138
36. Jia PR, Jiao GQ (1996) On better test method of determining mixed-mode interlaminar
fracture toughness for composite materials. J Northwest Polytech Univ 14(1):105–109
37. National Technical Monitoring Bureau (1995) National standard GB/T 15780–1995: testing
methods for physical and mechanical properties of wood. China Standard Press, Beijing
38. Triboulot P, Jodin P, Pluvinage G (1984) Validity of fracture mechanics concept applied to
wood by finite element calculation. Wood Sci Technol 18(1):51–58
39. Hodgkinson JM (2000) Mechanical testing of advanced fibre composites. Woodhead
Publishing and CRC Press, Cambridge
References 221

40. American Society of Testing Materials (ASTM) (2001) Standard test method for Mode I
interlaminar fracture toughness of unidirectional fiber-reinforced polymer matrix composites.
Annual book of ASTM standards. Philadelphia, PA, D 5528-01
41. Wang FL, Shao ZP, Wu YJ, Wu D (2014) The toughness contribution of bamboo node to the
Mode I interlaminar fracture toughness of bamboo. Wood Sci Technol 48:1257–1268

Das könnte Ihnen auch gefallen