Principles For Design: H. J. Blaß C. Sandhaas
Principles For Design: H. J. Blaß C. Sandhaas
Blaß
C. Sandhaas
Timber Engineering
by
Hans Joachim Blaß
Carmen Sandhaas
Translation: Richard Mort
Impressum
ISBN 978-3-7315-0673-7
DOI 10.5445/KSP/1000069616
Foreword
Successfully designing timber structures requires an in‐depth knowledge and a funda‐
mental understanding of wood and wood‐based materials as modern and sustainable
construction materials, of structural elements, systems and joints using wood and wood‐
based materials. Although consideration of design rules set out in standards is still im‐
portant for structural design, successful and efficient constructions require more than
simply applying calculation standards alone. Accordingly, this book should contribute to a
more in‐depth understanding of the principles used to design timber structures.
The first step towards explaining the Eurocode 5 rules came in 1995 with the STEP an‐
thology − Structural Timber Education Programme. STEP, which was compiled by approx‐
imately 50 authors from 14 European countries, was intended both to educate those
studying timber engineering and provide further training for structural engineers in the
European countries concerned. Since it was released however, knowledge has advanced
significantly in areas such as construction materials, structural elements and joints, as is
also reflected in the relevant Eurocode 5 rules. This book embraces and builds on the
STEP concept, to establish a timber engineering textbook tailored to the national situa‐
tion. This involved expanding the timber engineering topic scope to accommodate addi‐
tional contributions, while also omitting some of the original STEP material, given its
limited relevance to timber construction in German‐speaking countries. Contributions on
individual topics were expanded, deepened and updated beyond the scope of the origi‐
nal STEP books, while derivations on individual topics were included in an annex and the
authors of the original STEP contributions are cited in the relevant places. The scientific
employees of the KIT Timber Structures and Building Construction section also provided
invaluable assistance to revise and supplement various contributions.
There will always be room to improve a book as comprehensive as this one. We kindly
request that anyone reading this who comes across an error informs us accordingly, so
that this can be rectified for future editions.
Hans Joachim Blass and Carmen Sandhaas
General note: Most standards and additional codes of practice are included without specifying the date, given
frequent changes. Exception: Excerpts from standards are dated, so that the reader knows which edition was
consulted. For this edition, DIN EN 1995‐1‐1:2010, DIN EN 1995‐1‐1/A2:2014 and DIN EN 1995‐1‐1/NA:2013
have all been considered.
Contents
A Introduction ............................................................................. 1
A1 Codes of practice in Germany .............................................................................. 3
A2 Sustainability ...................................................................................................... 11
B Material wood ........................................................................ 19
B1 Wood anatomy ................................................................................................... 21
B2 Wood physics ..................................................................................................... 37
B3 Wood formation ................................................................................................. 53
B4 Durability ............................................................................................................ 75
B5 Wood drying and strength grading .................................................................... 85
B6 Wood products ................................................................................................... 99
C Principles for design ............................................................. 133
C1 Safety concept .................................................................................................. 135
C2 Actions on structures ....................................................................................... 145
D Structural members and systems .......................................... 159
D1 Basic stresses .................................................................................................... 161
D2 Stability ............................................................................................................. 177
D3 Influence of volume and stress distribution .................................................... 195
D4 Members with variable cross‐sections or curved shapes ................................ 205
D5 Non‐reinforced notches and holes .................................................................. 219
D6 Glued composite members .............................................................................. 233
D7 Mechanically jointed members and CLT elements .......................................... 249
D8 Reinforcements ................................................................................................ 263
D9 Diaphragms and bracings ................................................................................. 279
D10 Timber‐concrete composite structures ........................................................... 309
D11 System strength ............................................................................................... 317
E Joints ....................................................................................323
E1 Joints in timber structures ............................................................................... 325
E2 Johansen model ............................................................................................... 347
E3 Joints with nails and staples ............................................................................. 365
E4 Joints with bolts and dowels ............................................................................ 375
E5 Joints with self‐tapping screws ........................................................................ 383
E6 Joints with connectors ..................................................................................... 393
E7 Punched metal plate fasteners ........................................................................ 407
E8 Cold‐formed steel connectors ......................................................................... 417
E9 Contact joints (Carpentry joints) ...................................................................... 423
E10 Glued‐in steel rods ........................................................................................... 435
E11 Joints loaded perpendicular to the grain ......................................................... 441
E12 Reinforced joints .............................................................................................. 459
E13 Joints with multiple fasteners .......................................................................... 477
E14 Moment‐resisting joints ................................................................................... 491
E15 Joints with multiple shear planes ..................................................................... 509
F Serviceability ........................................................................ 515
F1 Deformations .................................................................................................... 517
F2 Vibrations ......................................................................................................... 525
G Accidental loads and additions .............................................. 535
G1 Reaction to fire and fire design ........................................................................ 537
G2 Joints subject to seismic loads ......................................................................... 561
G3 Earthquake‐compliant structural details ......................................................... 573
G4 Damages in hall structures ............................................................................... 585
Annexes .................................................................................... 609
Annex 1: Dynamic modulus of elasticity ...................................................................... 611
Annex 2: Stress interactions ........................................................................................ 613
Annex 3: Elastic buckling load and buckling lengths ................................................... 619
Annex 4: Derivations lateral torsional buckling ........................................................... 623
Annex 5: Mechanically jointed beams ......................................................................... 629
Annex 6: Derivation Johansen equations .................................................................... 639
A
Introduction
A1 Codes of practice in Germany
Author: Rainer Görlacher
Above all, the building authority centres on preventing dangers to the public, which may
arise from the construction process. In Germany, the general conditions governing the
building, alteration or demolition of structures have been a federal responsibility since
the 19th century. Even so, individual tasks were transferred to the Deutsche Institut für
Bautechnik (DIBt) in an agreement and this allocation of responsibilities, alongside Euro‐
pean rules which have to be implemented in Germany make it far from easy for structural
engineers to decipher the current legal situation.
The following article will provide an overview (including online references) of the relevant
codes of practice for structural engineers. It is important to remember that regulations
on national and European levels are prone to change and since reflecting all such changes
in this article is impossible, only the basic context will be explained in this case. We advise
structural engineers to check the current status of regulations (particularly the current
list of Technical Building Rules, Section A1.2 and the current Construction Products Lists,
Section A1.3) on the specified web pages.
A1.1 Model Building Code and Federal Building Codes
The building codes of the individual German federal state dictate the form of construc‐
tion supervision imposed, although these State Building Codes (LBOs) tend to follow the
jointly compiled Model Building Code (MBO). The Model Building Code and the model
ordinances of ARGEBAU can be viewed and downloaded on the ARGEBAU homepage
(www.is‐argebau.de; there: Mustervorschriften (model rules) / Mustererlasse (model
decrees) / Bauaufsicht (construction supervision) / Bautechnik (construction technolo‐
gy)). However, the Model Building Code and model ordinances are not laws, but merely a
framework providing guidance for the legal building regulations of each state. With sim‐
plicity in mind, the model regulations are adopted more or less uniformly across different
federal states for construction products and types of construction. Even so, the regula‐
tions of individual federal states should still be taken into consideration.
3
Codes of practice in Germany
§ 3 of the LBOs sets out the general requirements for a construction:
(1) Built structures must be laid out, built, altered and maintained such as to ensure there
is no threat to public order and security; particularly life, health and natural resources.
(2) Construction products and types of construction may only be used if the built struc‐
tures in question are subject to proper maintenance for an appropriate duration com‐
mensurate with the purpose, meet requirements pursuant to or based on this law and
are fit for purpose.
(3) The technical rules stipulated by the highest building authority as publicly disclosed
Technical Building Rules must be complied with.
These general requirements imposed on construction result in the following provisions,
with which structural engineers must comply:
Compliance with technical rules is mandatory to verify structural safety. These are
included in the list of Technical Building Rules of German federal states (LTB).
The construction products and types of construction to be used in built structures,
are detailed in the so‐called Construction Products Lists A, B and C.
A1.2 Technical Building Rules
The list of Technical Building Rules (LTB) of federal states is set out separately based on a
sample list (M‐LTB) for each state, meaning the content may vary slightly between indi‐
vidual states. Moreover, the individual lists are not released at the same time. The
Deutsches Institut für Bautechnik (DIBt) is tasked with preparing the introduction, on
behalf of the states (https://www.dibt.de/de/Geschaeftsfelder/BRL‐TB.html):
Part I: Technical rules for planning, design and construction of built structures and
their parts.
Part II: Regulations governing the use of construction products and kits in accord‐
ance with European Technical Assessments and harmonised standards pursuant
to the Construction Products Regulation.
Part III: Regulations governing the use of construction products and kits in accord‐
ance with European Technical Assessments and harmonised standards pursuant to
the Construction Products Regulation and Sections § 17 para. 4 and § 21 para. 2 of
the Model Building Code. This part concerns the regulation to assess the suitability
of construction products and types of construction in accordance with water law
through verifications in accordance with the Model Building Code (WasBau‐PVO).
4
Codes of practice in Germany
Part I, as specified on the DIBt homepage, is a model application, from which the state
regulations may differ. Parts II and III, conversely, include valid usage regulations in Ger‐
many. The regulations for the state of Baden‐Württemberg can be retrieved from the
homepage of the Ministry of the Environment, Climate Protection and the Energy Sector.
(http://um.baden‐wuerttemberg.de/de/umwelt‐natur/berg‐und‐
baurechtsbehoerde/bautechnik‐und‐bauoekologie/technische‐baubestimmungen/).
Part I
The technical rules for planning, design and construction are generally set out as stand‐
ards (national or European) imposed as part of building regulations by inclusion in the
LTB. In certain cases (e.g. technical rules for health protection or to protect against fire),
directives may also be issued to meet building legislation requirements.
For timber structures, Part I of the LTB for EC 5 (DIN EN 1995) and its parts,
Part 1‐1: General – Common rules and rules for buildings,
Part 1‐2: General – Structural fire design,
Part 2: Bridges.
are introduced to meet building legislation requirements.
Regulation is also imposed through respective National Annexes (NA) and the Nationally
Determined Parameters (NDP, such as partial safety factors or load duration classes). The
NAs also contain non‐contradictory complementary information (NCI) and member states
cannot amend European standards.
However, the application of EC 5 also requires compliance with additional provisions, con‐
tained in the annexes of the LTB. The relevant annexes for EC 5 are 2.5/1E and 2.5/2, which
specify regulations to be complied with when using construction products in accordance
with harmonised standards for timber structures. Here, reference is made to application
standards (DIN V 20000‐X), which also apply when using specific construction products in
accordance with harmonised standards, although such use may also be excluded.
5
Codes of practice in Germany
Relevant examples:
Wood‐based panels are governed by the harmonised standard EN 13986:2004,
which is also set out in the Building Rules Lists of the DIBT (see next section).
However, when using wood materials in accordance with EN 13986, compliance
with the related application standard DIN V 20000‐1:2005 is also required.
Dowel‐type fasteners (nails, screws, dowels) are governed by the harmonised
standard EN 14592:2008, which is also set out in the Building Rules Lists of the
DIBT (see next section). However, Annex 2.5/1E specifies that this standard is only
applicable for bolts, dowels and smooth nails and other fasteners require addi‐
tional approvals in line with building legislation requirements (Note: In this case,
however, an application standard DIN V 20000‐6 is compiled).
Finally, the list of Technical Building Rules Part 1 for timber construction also introduces
DIN 1052‐10: Design of timber structures – Part 10: Additional provisions
to meet building legislation requirements. This standard was necessary, since EC 5 and
the related reference standards did not include all applications regulated to date in
DIN 1052 (including resined staples, glued‐in rods, steel rods with wood thread), or the
European product standards did not meet the statutory requirements imposed in Ger‐
many to meet building legislation requirements (e.g. proof of suitability for gluing from
an accredited test centre).
Part II
Since 01.07.2013, the European Construction Products Regulation (CPR), which regulates
the certification procedure for construction products and kits, has come into force and
differs in some aspects compared to the previous Construction Products Directive, which
applied up to 30.06.2013. For example, there are no more guidelines, so‐called ETAGs,
and without guidelines, via so‐called CUAPs, no further construction products and kits
can be approved. The old ETAGs have been transferred to the new European Assessment
Document (EAD), while existing European Technical Approvals (ETAs, whether or not a
guideline is included) remain valid until the date of expiry.
In the regulations governing the use of construction products and kits in accordance with
European Technical Assessments (ETAs) and harmonised standards in accordance with
the Construction Products Regulation, the LTB of federal states make direct reference to
the list of Technical Building Rules, as published in the “DIBt Official communication”
(www.bauministerkonferenz.de or www.dibt.de). The applicable rules in Part II are classi‐
fied in accordance with the procedure when creating the rules:
6
Codes of practice in Germany
Regulations governing the use of construction products within the applicable
scope of European Assessment Documents (EAD) for European Technical Assess‐
ments or for construction products with European Technical Approval, which were
compiled in accordance with a guideline prior to 01.07.2013.
Regulations governing the use of construction kits within the applicable scope of
EADs for European Technical Assessments or for construction kits with European
Technical Approval, which were compiled in accordance with a guideline prior to
01.07.2013.
Regulations governing the use of construction products, for which European
Technical Approvals were compiled without a guideline prior to 01.07.2013.
Regulations governing the use of construction kits, for which European Technical
Approvals were compiled without a guideline prior to 01.07.2013.
Regulations governing the use of construction products in accordance with
harmonised standards.
This means any user wishing to identify applicable regulations in the Technical Building
Rules must know whether the relevant object is a construction product or a construction
kit and also whether the ETA was created in accordance with an EAD, prior to 01.07.2013
based on a guideline (ETAG) or without a guideline (CUAP). Finally, there are also con‐
struction products which correspond to harmonised standards.
For construction products subject to European regulations (via ETAs or harmonised
standards), no additional regulations apply and they can be used, subject to inclusion in
the Construction Products Lists (see next section) and taking into consideration any exist‐
ing application standards (DIN 20000‐X), as specified in Part I.
A1.3 Construction Products Lists
In § 17, the Model Building Code (MBO) regulates the use of construction products,
which must correspond to one of the following specifications and be identified:
National
Construction products regulated in accordance with Construction Products
List A Part 1 (products for which technical rules (standards) exist.
Non‐regulated construction products, for which the following is required:
‐ A general (national) technical approval or
‐ A general (national) test certificate (Construction Products List A, Part 2) or
‐ Consent in an individual case (“Zustimmung im Einzelfall”).
7
Codes of practice in Germany
Excluded are construction products of only minor importance for meeting the require‐
ments of or based on this law and which the Deutsches Institut für Bautechnik (DIBt) has
disclosed in a list C, by agreement with the highest building authority.
Technical rules (standards) apply to regulated construction products, with which the
construction product must comply. If a construction product deviates significantly from
the technical rules, or if no technical rule exists for the same, a general technical approval
can be issued for this product (List of approvals for timber construction:
https://www.dibt.de/de/Fachbereiche/Referat_I5.html).
If the construction products used do not meet key structural safety requirements or if the
requirements can be assessed by generally accepted test procedures, instead of the
general technical approval, a general test certificate (allgemeines bauaufsichtliches Prüf‐
zeugnis) applies. These construction products are set out in the Construction Products
List A, Part 2 and for construction kits in Part 3. Finally, consent in an individual case
(Zustimmung im Einzelfall) can also be issued for the use of construction products in a
specific construction, even if not all the above requirements are met.
European
Construction products in accordance with Construction Products List B:
Construction products within the applicable scope of harmonised standards
in accordance with the Construction Products Regulation (CPR).
Construction products within the applicable scope of European Assessment
Documents (EAD) for European Technical Assessments (ETA).
Construction products, for which European Technical Approvals (ETA) were
compiled prior to 01.07.2013 with or without a guideline.
The following note in Annex 01 of the Construction Products List is important for all con‐
struction products subject to European regulations: “The levels, classes and conditions of
use specified in the Federal Building Codes (LBO) and in the provisions set out under the
same shall apply.” This means additional regulations often apply for a construction prod‐
uct in Germany subject to European regulations, which must be taken from LBOs or LTBs
(or annexes thereof). The current Construction Products Lists are published on the DIBt
homepage: https://www.dibt.de/de/Geschaeftsfelder/BRL‐TB.html
If there are no additional regulations, construction products for which an ETA has been
compiled can be used in Germany, even when not explicitly stated in the Construction
Products List B. An overview of granted and valid ETAs is included on the EOTA page:
http://valideta.eota.eu/pages/valideta/
8
Codes of practice in Germany
A1.4 Marking
According to Federal Building Code § 22 Attestation of Conformity, construction products
require an attestation of conformity with the technical rules, the general technical ap‐
provals, the general test certificates or consent in an individual case; any minor devia‐
tions are also classed as compliance. In this case, national regulated (A, Part 1) and non‐
regulated (A, Part 2) construction products are marked with an Ü‐symbol (compliance
mark), see Figure A1‐1 on the left. This mark indicates the technical specification (stand‐
ard or approval) with which the construction product complies and the nature of the
Attestation of Conformity (declaration of the manufacturer or certificate of a certification
body). A CE‐mark, as shown in Figure A1‐1 on the right, indicates the harmonised Euro‐
pean standard or ETA with which the construction product complies and the reference
number of the Declaration of Performance (DOP). This declaration of performance
includes all values required for the planner, e.g. characteristic stiffness and strength
properties of wood products, but also formaldehyde emission values for glued products.
Figure A1‐2 and Figure A1‐3, respectively, show examples of Ü‐ and CE‐marks.
Ü‐mark in accordance with the CE‐mark in accordance with
Hersteller
Herstellwerk x
Figure A1‐1 Right: Ü‐mark. Left: CE‐mark.
9
Codes of practice in Germany
Hersteller
Manufacturer ABC Hersteller
Manufacturer ABC
Manufacturing
Herstellwerk x Herstellwerk x
plant 123 DIN 1052:2004‐8
Timber with
Z Z‐9.1‐XYZ
- 9.1 - xyz Zfinger joint
- 9.1 - xyz
C24
Label of
certification body
Figure A1‐2 Example of Ü‐mark, technical specification is the national technical approval (left) or national
standard (right).
Timber worm PLC, 12345 Woodtown
13
HW OSB/4
DOP‐0745/31‐40
OSB/4‐E1
EN 13986:2004+A1:2015
0765
Heavy‐duty panels for internal use as structural components
in humid conditions
Figure A1‐3 Example of CE‐mark.
10
A2 Sustainability
Original article: T. Vihavainen
An important point to consider when using timber in constructions is the sustainability of
wood and wood products. As a renewable resource, wood has a key advantage compared
to alternative construction materials. For example, when biomass is formed (e.g. leaves,
wood), CO2 is extracted from air via photosynthesis, carbon (C) is incorporated in the
biomass and oxygen (O) is released into the air. Thanks to this effect, timber is consid‐
ered a CO2‐neutral construction material and it is only when biomass breaks down when
wood is burnt or when wood rots in the forest that the same amount of CO2 is released
as was originally absorbed. One further advantage is the fact that wood requires far less
primary energy than other materials when being processed into construction products.
Wood grows naturally, need not be manufactured and is far easier to work with than e.g.
steel or aluminium. Figure A2‐1 shows the ratio of primary energy consumption required
to produce a cubic metre of construction material from various raw materials. However,
different mechanical properties and larger cross‐sections mean larger volumes may be
required to create wooden structures compared to those made of steel or concrete.
Conversely, when it comes to other physical features like heat conduction, wood outper‐
forms steel with a lower heat conductivity between 0.13 and 0.20 W/(m∙K) (steel:
= 60 W/(m∙K)).
126
24
4 6
1
11
Sustainability
A2.1 Life Cycle Analysis (LCA)
Sustainability is defined in numerous ways, but the original definition came from Hans
Carl von Carlowitz, who requested as early as 1713 that “We should only take as much
wood out of the forest, as will grow back.” Nowadays, various aspects have to be consid‐
ered when assessing the sustainability of a system. The forest example includes focusing
on sociocultural aspects, such as its recreational value, economic aspects like the mone‐
tary value of the wood and ecological aspects like the need to preserve biodiversity. If we
assess the sustainability of a building, we soon realise just how complex a task it really is.
As well as the need to take the energy consumption and emissions of all materials into
consideration, including their production and the building construction themselves, the
assessment also includes expenditures during the useful life, for example the heat energy
required. At the end of its useful life, the building is demolished and the left‐over materi‐
al is either recycled or used for energy recovery. Balance sheets, including sustainability
assessments for these aspects, also have to be compiled. All the above points refer solely
to the ecological aspects, so an ecological balance sheet for the building in question is
compiled. To perform a complete sustainability assessment however, criteria for eco‐
nomic and sociocultural sustainability also need to be defined. Here, feasible economic
criteria would include, for example, how well the building retains its value or the costs of
the products used. Living comfort, the urban integration of the building or infrastructural
criteria meanwhile, like public transport connections or distance from the supermarket,
would constitute typical sociocultural aspects of a sustainability assessment.
Within the scope of this chapter, only the first section of a sustainability analysis men‐
tioned, the ecological balance, is relevant and this kind of analysis, featuring the devel‐
opment of a system right up to the end of its life, is known as a life cycle analysis, LCA.
Figure A2‐2 schematically illustrates the life cycle of raw material over production, con‐
struction and usage phases right up to demolition and recycling or energy recovery from
the waste material.
12
Sustainability
Construction
Recycling
(Building)
A life cycle analysis comprises four steps:
Definition of system limits (goal and scope definition):
The goal and frame of reference have to be defined. Is the goal to consider an in‐
dividual product or the ecological balance of the interaction of multiple individual
products? When there is a need to compare products made of various materials,
consideration of a functional unit is often more significant. For example, stat‐
ic/dynamic reasons dictate that over a one‐kilometre stretch of power line, more
wooden power poles are used than those made of steel or concrete. Another ex‐
ample is the ecological assessment to determine the origin of the wood; do you
start with the distribution of the seeds, the harvesting of the logs or only after the
cutting? In response, the system limits and assessment scope must be configured
in a way that facilitates comparison of the data determined.
Life cycle inventory (LCI):
The life cycle analysis determines all flows of materials and energy. The resource
consumption (raw materials, auxiliary materials, energy, water,…) and emissions
(end products, by‐products, waste, air emissions,…) of the system observed are
arranged, while a life cycle inventory is normally performed for the functional unit,
namely per cubic metre or kilogram of product.
Impact assessment (LCIA):
The material and energy flows described in the previous step are evaluated in
terms of their environmental impact, which involves considering a range of impact
categories introduced after the following step.
13
Sustainability
Interpretation:
The data obtained in steps 2 and 3 is evaluated and recommendations are made
for the client. This is mainly dictated by the quality of the available data and in‐
volves assessing the data quality itself. The parameters used may include compa‐
rability, completeness and the accuracy or consistency of the obtained data.
Some selected impact categories are:
Global warming potential GWP (in kg CO2‐equivalent):
The GWP specifies the extent to which a specific quantity of a greenhouse gas
contributes to the greenhouse effect, where the comparable figure is carbon diox‐
ide, namely the CO2‐equivalent. The scope usually involves observing the average
heating effect over a century. All emissions recorded in the life cycle inventory are
evaluated for their GWP and converted to a CO2‐equivalent figure.
Acidification potential AP (in kg SO2‐equivalent):
All emissions which exacerbate acidification are expressed in terms of
sulphur dioxide equivalent.
Human toxicity potential HTP (in kg of body weight):
Recording all emissions impacting on health.
Photochemical ozone formation potential POFP (in kg C2H4‐equivalent):
Emissions contributing to the formation of ozone are expressed in terms of
ethylene equivalent.
Eutrophication potential EP (in kg PO4‐equivalent):
Expressing the contributing emissions as a phosphate equivalent.
The site http://www.oekobaudat.de lets you retrieve complete life cycle analyses for
various construction materials. Key here is to ensure comparability of data at all times,
given the variance in system limits used.
14
Sustainability
A2.2 Ecological significance of wood
Around 30% of global land is forested, most of which comprises tropical rainforests and
boreal coniferous forests. Figure A2‐3 shows an overview of the world’s forests, with a
global total of 4 billion hectares, more than half of which is distributed over the five most
forested countries (the CIS, Brazil, Canada, the USA and China. Source: FAO 2010). The
biomass of these forests also holds around 289 gigatons of CO2 (FAO, 2010) and since CO2
is the most significant greenhouse gas generated by human activity, this means forests
perform a key carbon storage role. Such forests are known as carbon pools. Virgin forests
establish an equilibrium and only emit as much carbon as they absorb. The volume of
storage depends on the type of forest involved, the soil condition (CO2 storage in the soil)
and management. Forests can also function as carbon sinks if the timber stock rises,
namely, where more forest is planted than felled. Forests where harvesting outstrips
planting, namely where the net carbon release exceeds the figure incorporated into the
growing wood, are deemed carbon sources.
Development of the global forested area (see Figure A2‐4) shows a rising volume of for‐
est in Europe (→ carbon sink), albeit at a slow rate, while those in South America and Africa
are declining rapidly (→ carbon source). The high growth rate in Asia is due to the large‐
scale reforestation programs in China, in contrast to the shrinking forest elsewhere in Asia.
This also underlines the need for continued vigilance in checking the origin of tropical
woods, to reconfirm that they have been sourced from sustainably managed forests.
Sustainable management means not only minimising the ecological impact of the timber
harvest as far as possible (no surface clearance, preserving the ecological function of the
forest, including all animals and other plant varieties, preserving the most valuable areas
of forest from an ecological perspective, control tools for monitoring), but also ensuring
the long‐term sustainability of forest management by getting forest farmers involved and
clearly defining their claims of ownership and use. As well as these social aspects of sus‐
tainable forestry, however, economic aspects also play a part, such as ensuring an equal
share of proceeds from the timber trade and the fact that forest management generally
has to be economically self‐sustaining. Establishing a sustainable forestry can therefore
retain the function of the forest as a carbon reservoir and provided the forest is used
effectively and sustainably, can help it fulfil its role as a carbon sink, if more biomass is
produced than harvested, as applies in Europe, for example.
15
Sustainability
[%]
No data 30 – 50
0 – 10 50 – 70
10 – 30 70 – 100
Figure A2‐3 Ratio of forest area as a proportion of the overall land area. (FAO, 2010)
Scale
1 million ha
Africa Asia Europe North and Central America Oceania South America
Figure A2‐4 Change in forest area per region. (FAO, 2010)
16
Sustainability
A2.3 Sustainable forest management
Deforestation and silvicultural exploitation are becoming increasingly key aspects of the
international environmental debate and the fact that our forests are vanishing threatens
the Earth like never before. In response, most countries have committed to using forests
sustainably and for a long while, efforts to preserve production capacity as a measure of
sustainability came to the fore. Equally important, however, is focusing on wider ecologi‐
cal goals, such as protecting forest ecosystems as a whole, preserving diversity and nur‐
turing cultural, aesthetic and recreational values. This reflects how logging is changing
ecological conditions for forests in many areas. Numerous forest plants and animal spe‐
cies require specific conditions in which to flourish. Nowadays in Europe, truly natural for‐
ests are very rare and the natural forests or special biotopes that do exist are not generally
commercially exploited, or timber production is strictly regulated. Meanwhile, debate con‐
tinues to rage as to whether sufficient areas or natural parks have been established.
These are all reasons why timber consumers need to verify that the timber they use
comes from sustainable forestry, the most well‐known certificates for which are the label
of the Forest Stewardship Council (www.fsc.org) and PEFC (www.pefc.de). Both prioritise
sustainable forestry, which preserves the ecological, social and cultural values of forests
and an unbroken “chain of custody” allowing the trajectory of the wood, from felling
right up to arrival at the end consumer, to be properly traced.
A2.4 Literature
T. Vihavainen, original Article A16, STEP 1995.
FAO (2010). Global forest resources assessment 2010. FAO Forestry Paper 163, Rome.
Frühwald A. (2007). The ecology of timber utilization, life cycle assessment, carbon management etc..
Kick‐off meeting Probos Foundation, Doorn, the Netherlands.
17
B
Material wood
B1 Wood anatomy
Original article: P. Hoffmeyer
Only with a solid grounding in wood anatomy can mechanical and other physical charac‐
teristics of timber be understood and wood used in a way that ensures material compati‐
bility or optimally exploits its potential. Wood can also be observed on a range of levels,
as clearly shown in Figure B1‐1, while a tree trunk can be observed in terms of structural
timber, which involves evaluating knots and other growth‐related properties. Observing
at a level of detail beyond that for structural timber means examining wood on a macro‐
scopic level, including all features visible to the naked eye. The scope also includes flaw‐
less wood without fibre deviation (clear wood). When the mechanical properties depend
on the system size, meanwhile, we describe the wood on a mesoscopic level, e.g. in
terms of annual rings. The microscopic level includes the wood fibres within the vessels,
the large water‐conducting cells in hardwood and the smaller tracheids, key to hardwood
strength and performing both reinforcement and water‐transport functions in softwood.
Besides the cell wall structure, the chemical structure can also be examined.
Wood is a natural and organic material, comprising cells. It is also a compound chemical
complex of cellulose, hemicellulose, lignin and other constituents. Wood tends to be
anisotropic, due to the elongated structure of its cells and the orientation of the cell
walls. Anisotropy is also a factor of the variation in cell size during growth and partially
due to the preferred direction of specific cell types (e.g. rays). Meanwhile, the three key
structural variables impacting on the properties of wood as a construction material are its
fine cell wall structure, the collection of cells in clear wood and growth irregularities in
timber. For example, the fact that shrinking and swelling perpendicular to the grain is
generally 10 to 20 times as large as in the grain direction can be explained by the submi‐
croscopic structure of the cell walls. The microscopic structure of clear wood, meanwhile,
is why wood is 20 to 40 times more rigid longitudinally rather than perpendicular to the
grain, while its macrostructure (knots, fibre deviation) underlines its tensile strength in
the grain direction. This may range from over 100 N/mm² in clear wood to under
10 N/mm² for low‐quality timber, for the same wood species.
This chapter will initially explain the general stem structure and section planes of the
wood, before focusing on its anatomy, from the macroscopic level right up to the chemi‐
cal structure. Wood can be extracted from two main groups of plants; softwoods and
hardwoods, where the vast majority used for construction in Central Europe are spruce
and fir, two conifers. The differences between conifers (softwoods) and deciduous trees
(hardwoods) emerge in their growth pattern, different leaf shapes and also the structure
of the wood itself (Figure B1‐2).
21
Wood anatomy
Cell wall
ML, S1, S2, S3
Fibrils 10‐6 m
Microfibrils
10‐6 – 10‐7 m
Tracheids
10‐5 – 10‐4 m
Molecules Structural
Cellulose, Vessels timber
hemicellulose, (only hardwood) > 10‐1 m
lignin, pectin, 10‐4 – 10‐3 m
constituents
10‐10 – 10‐9 m
Clear wood
10‐3 – 10‐1 m
Figure B1‐1 Hierarchical levels of wood. (Mark Harrington, University of Canterbury, 1996)
0.3 mm
Figure B1‐2 Growth pattern and microscopic structure of oak (Quercus robur, left) and
spruce (Picea abies, right). (STEP 1995 Article A4)
22
Wood anatomy
B1.1 General structure of a tree stem
The role of a tree’s stem or trunk is to support the crown and supply it with water, while
at the same time, the nutrients formed in the crown are transported down the stem via
its bark. Nutrients not used to form wood are channelled towards the roots and stored
there. Accordingly, a tree stem meets the requirements imposed on it, as its structure
reflects. Most cell structures (water canals, fibres) follow the axial structure of the stem,
with only a small percentage (rays) oriented radially, to guarantee the radial transport of
nutrients and water. A tree thus grows both longitudinally as well as around its girth and
the concentric rings formed over the course of a year within the so‐called cambium are
known as annual rings. The annual rings themselves are, in turn, sub‐classified into early‐
and latewood. Wood is thus not homogenous or isotropic, but a naturally grown material,
which always adapts to the respective local conditions. Figure B1‐3 and Figure B1‐4 show
the stem structure of a tree.
Outer bark, dead
Bark
Inner bark, alive
Sapwood
Heartwood
Pith
Latewood
Annual ring
Earlywood
Cambium, not visible to the naked eye
Figure B1‐3 Cross‐section through a coniferous stem, Douglas fir. (Grosser, 1977)
23
Wood anatomy
B1.2 Section planes of wood
The natural section planes of the wood emerge as a result of the clearly axial symmetrical
and round cross‐section of the stem, as is clearly shown in Figure B1‐3. The main section
planes are cross‐, radial and tangential sections, as shown schematically in Figure B1‐4
and a complete picture of the three‐dimensional wood structure is only possible when all
three sections are combined.
The cross‐section is performed at right angles to the stem axis and substantially circular
as a rule; revealing heartwood, sapwood and annual rings. Since the dominant axial struc‐
ture has been severed, this cross‐section often reveals a reticular texture (see Figure B1‐5),
while the rays run from inside to outside.
The radial section splits the stem longitudinally through the pith, similar to a circle seg‐
ment and shows the axial and radial cell structures. Here, the rays run horizontally and
the annual rings can be clearly seen.
The tangential section, meanwhile, runs tangentially to the annual rings, like a circle
segment. The annual rings do not show up more easily in this case. The axial cells are
easily visible, while the radial rays are visible in the cross‐section.
Cross‐section
Heartwood
Sapwood
Pith
Earlywood
Annual ring
Latewood Radial section
Ray
Tangential section
Bark
Figure B1‐4 Macroscopic structure and section planes.
24
Wood anatomy
B1.3 Macroscopic structure
The following section will explain the macroscopic structure from inside to outside. We
start with the pith over the rays, to explain the heartwood and sapwood, then spotlight
the functions of the cambium and bark.
Pith
The pith comprises parenchymatous – accumulating – tissue comprising so‐called paren‐
chyma cells. Its cross‐section may be rounded, radial or square and its diameter is only in
the order of millimetres. Its main role is to supply the young and developing scion with
water and it perishes prematurely in a range of woods.
Rays
Rays, which exist in all deciduous trees and conifers, also include parenchyma cells and
facilitate the radial transport of water and nutrients, while their size and frequency vary
significantly by wood species. They form bright and fine lines, seldom more than 1 mm
wide, which run radially from the centre of the stem to the outside. Only the initially
formed and so‐called primary rays lead from the pith to the bark, which is why they are
called pith rays. All so‐called secondary rays are unconnected to the pith. The later they
emerge from the cambium, the further out their point of origin in the stem.
Heartwood and sapwood
As the secondary growth of the tree continues, the entire stem cross‐section is no longer
used to conduct water, but only the external portion and younger annual rings, between
5 and 100 or so, depending on the tree species. This external portion of the stem is
known as sapwood and also functions as a storage recipient, characterised by the pres‐
ence of living and physiologically active wood cells. The inner portion, meanwhile, known
as heartwood, generally no longer contains living cells and is solely tasked with stabilising
and strengthening the tree. Sapwood transforms into heartwood via biochemical pro‐
cesses, so‐called heartwood formation, which is synonymous with the death of paren‐
chyma cells and consumption of their starch deposits. As the heartwood forms, the gas
and water balance change, while the pits between the individual cells close off. In addi‐
tion, constituents are often stored in the cells. Heartwood is often darker in colour, with
differences in shade attributable to the chemical structure of the heartwood pulp. Inci‐
dentally though, lighter coloured heartwood does not necessarily indicate a lack of
heartwood pulp, but merely the fact that no pigmented heartwood pulp has been
formed. The fact that heartwood closes itself up and dies often makes it far more durable
than sapwood.
25
Wood anatomy
Cambium
The cambium comprises an area of dividable cells and is an extremely thin layer between
the wood and bark. It retains its divisibility up to the point at which the tree dies, forming
wood cells inwardly and inner bark cells outwardly and facilitating the secondary growth
(growth in girth) of the tree. It is not visible to the naked eye within the cross‐section.
Bark
Bark is further sub‐classified into outer and inner bark. The outer bark comprises dead
cells and functions to protect the stem, while the inner bark is made up of living cells and
transports nutrients produced in the crown to the cambium and storage cells.
Resin canals
Many wood species house resin canals, which are macroscopically visible. Resin canals
run parallel and perpendicular to the stem axis and transport resin, which is used as
sticky protection for the tree against wounds or invaders. The canals are lined with pa‐
renchyma cells capable of excreting resin, so‐called epithelial cells, although not all wood
species have resin canals. Firs and yews are resin‐free, although resin is contained in
spruce, pine, larch and Douglas fir trees.
B1.4 Mesoscopic structure, annual rings
The secondary growth of the tree is shown in the form of concentric annual rings. In
areas exposed to seasonal and temperate climates, the secondary growth is closely linked
to the prevailing climatic conditions, namely the water supply and temperature. When
the growing season gets underway in spring, the tree must conduct water effectively and
swiftly to ensure proper supply to the treetop. Reflecting this, the portion of the wood
structure in conifers which grows during spring is characterised by a large pore volume
and thin‐walled cells, since the tree is focusing on additional transfer of water. In hard‐
woods, conversely, particularly large and numerous vessels are the stand‐out features
and the elements responsible for the transfer of water. Wood formed in spring is known
as earlywood, but the priorities of the tree change when it comes to late summer and
autumn. Key at this time is the strength of the wood and the importance of the water
supply declines. Accordingly, conifers tend to form particularly thick‐walled cells with a
small cell lumen and small cell cavities, whereas deciduous trees form fewer and narrow‐
er vessels. This so‐called latewood is far thicker than earlywood and in softwood in par‐
ticular, these differences in thickness result in clear colour changes between early‐ and
latewood (see Figure B1‐3).
26
Wood anatomy
Figure B1‐5 Cross‐section through softwood, in which early‐ and latewood are clearly visible. (Schmid, 2002)
In winter, meanwhile, the tree stops growing. In softwood, Figure B1‐5, the prominent
disparity in cell wall thicknesses leads to characteristic variation in density, colour and
hardness within an annual ring. The wide‐lumen earlywood is lighter and the thicker
latewood is darker in colour.
Figure B1‐6b clearly shows that in coniferous trees, earlywood comprises the vast majori‐
ty of the annual rings. Even so, when exposed to stress, particularly dryness and cold, a
tree may only be capable of very limited secondary growth and accordingly only forms
minimal earlywood. In particularly bad years, the annual ring may be narrow to the point
of being virtually invisible to the naked eye and comprise no more than a few isolated
latewood cell lines. In Figure B1‐6, the differences in annual ring widths for various tree
species and within a single tree species can be clearly seen. Spruce (Figure B1‐6a) is a
softwood that grows far faster than yew (Figure B1‐6c); the same applies to poplar
(Figure B1‐6d) and robinia (Figure B1‐6e, in which the vessels present in hardwood are
also easily visible). The yew clearly shows the external impacts on secondary growth, as
reflected in the very varied width of annual rings. This leads us to conclude that the width
of the annual ring, the proportion of latewood and the regularity of annual growth (par‐
ticularly based on differences in respective densities) all have a key impact on the
strength of the wood.
27
Wood anatomy
a b c d e
Figure B1‐6 Annual ring widths (cross‐section): a. wide annual ring in spruce, b. medium‐wide annual rings in
fir, c. narrow annual rings in yew, d. wide annual rings in poplar, e. narrow annual rings in robinia.
(Wagenführ, 1999)
B1.5 Microscopic structure, cell types
Softwood
The wood structure of conifers comprises just two cell types; tracheids and parenchyma
cells. Tracheids are long (2 to 5 mm) and thin (10 to 50 μm) cells with tapered or curved
sealed ends. As shown in Figure B1‐7, the cells are arranged in two criss‐crossing systems,
so tracheids are distinguished in terms of longitudinal and perpendicular tracheids, while
parenchyma comprise longitudinal and ray parenchyma. The longitudinal tracheids take
up by far the most space in conifers at around 90 to 95%. Rays (= ray parenchyma), longi‐
tudinal parenchyma and resin canals, in contrast, occupy a relatively modest 5 to 10% in
the form of remaining cell elements or tissue systems. Some trees, however, such as the
yew, have no longitudinal parenchyma and resin canals at all. In firs, meanwhile, there
are no resin canals in normal wood and longitudinal parenchyma are only very seldom
present. The parenchyma excretion cells, which line the resin canals and from which the
resin emerges, are known as epithelial cells.
28
Wood anatomy
1) Latewood tracheids
2) Earlywood tracheids
3) Rays
4) Resin canals
Figure B1‐7 Schematic structure and cell types of softwood. (Nardi‐Berti, 1993)
Hardwood
Hardwood has evolved to a higher level than softwood and also includes distinctly more
cell types (depending on the species) and a more extensive division of labour, i.e. special‐
isation of cells, than softwood. See Figure B1‐8 for the schematic structure. Hardwood
includes a range of tracheids (= dead cells), which is why people tend to refer to hard‐
wood in terms of fibres rather than tracheids. The libriform fibres, for example, are solely
for strength, whereas the fibre tracheids also function to transport water as well as
providing reinforcement. The vessels or pores, which comprise dead cells and solely
conduct water, can be up to several metres long in some tree species. As in softwood,
the parenchyma cells store nutrients, but unlike softwoods, the rays in hardwoods may
also comprise multiple rows of parenchyma cells and are often clearly visible.
1) Latewood section
2) Earlywood section
3) Rays
4) Vessels
5) Longitudinal parenchyma
Figure B1‐8 Schematic structure and cell types of hardwood. (Nardi‐Berti, 1993)
29
Wood anatomy
Pits
The cell‐to‐cell material exchange takes place via small openings or gaps in the fibre wall,
which are known as pits. The main type in softwood is bordered pits, as shown in Fig‐
ure B1‐9. As well as letting water through, they also prevent any ingress of air into sap‐
filled cells; preventing the collapse of the water columns running from the roots to the
crown, which would ultimately mean the tree dying. Since closure of a pit is generally
irreversible, this must be borne in mind during wood drying or impregnation. For exam‐
ple, bordered pits in the spruce tree shut very quickly, which often means poor permea‐
bility for this species. (Impregnating agent cannot penetrate.)
Figure B1‐9 Bordered pits of conifers, schematic. Left: top view, centre and right: longitudinal section,
right: closure under one‐sided pressure. (Strasburger, 1998)
B1.6 Submicroscopic structure, cell wall
The wall of a wood cell generally comprises various cell wall layers, namely middle lamel‐
la and the primary, secondary and tertiary walls. The individual layers differ in terms of
thickness, chemical composition and orientation of the cellulose micro‐fibrils. The struc‐
ture is schematically shown in Figure B1‐10 and the cell wall structure itself dictates many
of the mechanical properties. The orientation of the cellulose micro‐fibrils, for example,
shows the direction where the tensile strength peaks. From an engineering perspective,
the cell wall is a particularly ingenious construction and the dominant S2 layer of axially
oriented micro‐fibril bundles can withstand tensile forces very effectively. When pressure
is exerted, long and slender columns emerge from the micro‐fibril bundles, but do not
buckle thanks to the reinforcing effect of the subtly sloping internal and external S1 and
S3 layers.
30
Wood anatomy
Middle lamella
The middle lamella interconnects neighbouring cells and adjacent cells possess a com‐
mon middle lamella; mainly comprising lignin and pectin. The high lignin content gives
the middle lamella very high compressive strength, while the pectin acts as a binder. The
middle lamella is 0.5 to 1.5 m thick and appears thinner in earlywood than latewood.
Tertiary wall
Cell lumen: cavity
Middle lamella: pectin and lignin,
Thickness: 0.5 to 1.5 m
Primary wall: cellulose, lignin, pectin and
hemicellulose
Thickness: 0.1 to 0.2 m
Middle lamella + primary wall = middle layer
Secondary wall
S1, = 60° to 80°, Thickness: 0.25 m
S2, = 10° to 30°, Thickness: 1 to 10 m
S3, = 60° to 90°, Thickness: 0.5 to 1.0 m
Figure B1‐10 Structure of the cell wall. (Booker and Sell, 1998)
Primary wall
The middle lamella forms the middle layer together with the primary wall. The diffuse
texture (see Figure B1‐11) of the cellulose micro‐fibrils, namely their amorphous non‐
crystalline arrangement, helps ensure the high dimensional stability of the cell. The pri‐
mary wall is 0.1 to 0.2 m thick.
Figure B1‐11 Diffuse texture of the primary wall.
31
Wood anatomy
Secondary wall
The secondary wall connects to the middle layer and is by far the thickest cell layer, with
around 90% of its volume comprising cellulose micro‐fibrils. The secondary wall includes
three distinct wall layers, which differ in terms of the orientation of the cellulose micro‐
fibrils and thicknesses. The external secondary wall (S1) is also known as the transition
lamella; located on the primary wall and around 0.25 m thick. The cellulose micro‐fibrils
feature a parallel texture with a fibril angle of 60° to 80°, meaning they are virtually per‐
pendicular to the cell axis, see Figure B1‐10. The central secondary wall (S2) forms the
main portion of the cell wall in the early‐ and latewood with a thickness of 1 to 10 m.
The fibril angle is at around 10° to 30° to the cell axis and therefore basically in axial
direction, as also shown in Figure B1‐10. The micro‐fibrils are tightly packed and run
parallel to each other in a spiral shape (screw‐texture) in the direction of the cell axis. The
internal secondary wall (S3) is around 0.5 to 1.0 m thick and the micro‐fibrils have a
parallel texture. The fibril angle to the cell axis is 60° to 90°; and as in the S1 layer, the
micro‐fibrils run virtually perpendicular to the cell axis.
Tertiary wall
The tertiary wall divides the cell wall from the cell lumen and like the middle layer, con‐
tains a high proportion of lignin.
Cell lumen
The internal hollow cell space, which is filled with air or water, is known as the lumen,
which is why you will often hear the term wide‐lumen cells used to describe cells with a
large lumen as opposed to narrow‐lumen cells. Earlywood and latewood typically include
wide‐ and narrow‐lumen tracheids respectively.
B1.7 Chemical structure
Biomass is formed through the process of photosynthesis, whereby CO2 is extracted from
the air, carbon is incorporated into the biomass (C6H12O6, carbohydrates) using water
(H2O) and the excess oxygen is discharged back into the air:
32
tŽŽĚĂŶĂƚŽŵLJ
tŽŽĚŵĂŝŶůLJĐŽŵƉƌŝƐĞƐŽƌŐĂŶŝĐĐŽŵƉŽƵŶĚƐ͗
ϱϬй ĐĂƌďŽŶ
ϰϯй ŽdžLJŐĞŶ K
ϲй ŚLJĚƌŽŐĞŶ ,
фϭй ŶŝƚƌŽŐĞŶ E
фϭй ŵŝŶĞƌĂůƐ
dŚĞƐĞŵĂŝŶĐŽŵƉŽŶĞŶƚƐĐůƵƐƚĞƌƚŽĨŽƌŵĐŚĂŝŶĂŶĚŵĂĐƌŽŵŽůĞĐƵůĞƐ͕ǁŚŝĐŚƚŚĞŶĨŽƌŵƚŚĞ
ǁŽŽĚƐƚƌƵĐƚƵƌĞ͘dŚĞŵŽůĞĐƵůĞƐĨŽƌŵĞĚŝŶƚŚŝƐĐĂƐĞĂƌĞĐĞůůƵůŽƐĞ͕ŚĞŵŝĐĞůůƵůŽƐĞĂŶĚůŝŐŶŝŶ͘
tŽŽĚĂůƐŽĐŽŶƚĂŝŶƐŽƚŚĞƌĐŽŶƐƚŝƚƵĞŶƚƐůŝŬĞƚĂŶŶŝŶŐĂŐĞŶƚƐŽƌĐŽůŽƵƌĂŶƚƐĂŶĚƌĞƐŝŶƐ͘dĂŬͲ
ŝŶŐĂůůƚƌĞĞƐƉĞĐŝĞƐ͕ƚŚĞƉƌŽƉŽƌƚŝŽŶŽĨǁŽŽĚŵĂĚĞƵƉďLJĐĞůůƵůŽƐĞŝƐĂƌŽƵŶĚϰϬй͘,ŽǁĞǀĞƌ͕
ƚŚĞ ĐŽŶƐƚŝƚƵĞŶƚƐ ǀĂƌLJ ƐŝŐŶŝĨŝĐĂŶƚůLJ͕ ĚĞƉĞŶĚŝŶŐ ŽŶ ƚŚĞ ƚƌĞĞ ƐƉĞĐŝĞƐ ŝŶ ƋƵĞƐƚŝŽŶ ĂŶĚ ĂůƐŽ
ƌĞƐƵůƚŝŶĚĂƌŬͲĐŽůŽƵƌĞĚŚĞĂƌƚǁŽŽĚŝŶĐĞƌƚĂŝŶƚƌĞĞƐƉĞĐŝĞƐ͘
ĞůůƵůŽƐĞ
OH
O
O O
HO OH
ɴͲ1,4ͲŐluĐose
Cellulose
HO HO
HO OH O HO OH O
HO O O O OH
O HO OH O HO OH
HO HO
ĞůůƵůŽƐĞŝƐĂŶĞdžƚƌĞŵĞůLJůŽŶŐͲĐŚĂŝŶƵŶďƌĂŶĐŚĞĚƉŽůLJƐĂĐĐŚĂƌŝĚĞ͕ǁŚŝĐŚŝƐĨŽƌŵĞĚĨƌŽŵĂƚ
ůĞĂƐƚϭϬ͕ϬϬϬŐůƵĐŽƐĞĐŽŵƉŽŶĞŶƚƐ͕ƐŽͲĐĂůůĞĚEͲϭ͕ϰͲŐůƵĐŽƐĞ͘ĞůůƵůŽƐĞŝƐ;ŝŶƚŚĞĐŚĂŝŶĚŝƌĞĐͲ
ƚŝŽŶͿ Ă ŚŝŐŚͲƉŽůLJŵĞƌ ĐŚĂŝŶ ŵŽůĞĐƵůĞ ǁŝƚŚ ǀĞƌLJ ŚŝŐŚ ƚĞŶƐŝůĞ ƐƚƌĞŶŐƚŚ͕ ǁŚŝĐŚ ĐĂŶ ĞdžŝƐƚ ŝŶ
ĐƌLJƐƚĂůůŝŶĞĂŶĚĂŵŽƌƉŚŽƵƐĨŽƌŵ͘dŚĞĨŝďƌŝůƐŽĨƚŚĞĐĞůůǁĂůůĂƌĞĨŽƌŵĞĚŽĨĐĞůůƵůŽƐĞŵŽůĞͲ
ĐƵůĞƐĂŶĚĐĞůůƵůŽƐĞĞŶƐƵƌĞƐƚŚĞĂdžŝĂůƐƚƌĞŶŐƚŚŽĨƚŚĞĐĞůů͘ĞůůƵůŽƐĞŝƐƉĂƌƚŝĐƵůĂƌůLJŚLJĚƌŽͲ
ƉŚŝůŝĐ͕ĚƵĞƚŽƚŚĞŶƵŵĞƌŽƵƐĨƌĞĞŚLJĚƌŽdžLJůŐƌŽƵƉƐ;വK,Ϳ͘
,ĞŵŝĐĞůůƵůŽƐĞƐ
KK,
K
K,
,ϯK
K,
K K,
K K K K K
KĐ K, K, KĐ K,
K K K K K
K, K, KĐ
ϯϯ
tŽŽĚĂŶĂƚŽŵLJ
hŶůŝŬĞĐĞůůƵůŽƐĞ͕ƐŽͲĐĂůůĞĚŚĞŵŝĐĞůůƵůŽƐĞƐĐŽŵƉƌŝƐĞĨŝǀĞĚŝĨĨĞƌĞŶƚƐƵŐĂƌĐŽŵƉŽŶĞŶƚƐ;ŐůƵͲ
ĐŽƐĞ͕ ŐĂůĂĐƚŽƐĞ͕ ŵĂŶŶŽƐĞ͕ ĂƌĂďŝŶŽƐĞ ĂŶĚ džLJůŽƐĞͿ͕ ĨŽƌ ƉŽůLJƐĂĐĐŚĂƌŝĚĞƐ ƚŚĂƚ ĂƌĞ ƐŚŽƌƚͲ
ĐŚĂŝŶ͕ ďƌĂŶĐŚĞĚ ĂŶĚ ʹ ƵŶůŝŬĞ ĐĞůůƵůŽƐĞ വ ƚŚƌĞĞͲĚŝŵĞŶƐŝŽŶĂů͘ ,ĞŵŝĐĞůůƵůŽƐĞƐ ŚĂǀĞ ŵĂŶLJ
ĨƌĞĞ ŚLJĚƌŽƉŚŝůŝĐ ŐƌŽƵƉƐ ;വK,͕ വ,K͕ വKK,Ϳ ĂŶĚ ĂƌĞ ŚŝŐŚůLJ ƌĞĂĐƚŝǀĞ͕ ĚƵĞ ƚŽ ƚŚĞŝƌ
ďƌĂŶĐŚĞĚƐƚƌƵĐƚƵƌĞ͘ůŽŶŐǁŝƚŚĐĞůůƵůŽƐĞ͕ŚĞŵŝĐĞůůƵůŽƐĞƐĨŽƌŵƚŚĞĨƌĂŵĞǁŽƌŬŽĨĂǁŽŽĚ
ĐĞůů;ĂůƚŚŽƵŐŚĨŝďƌŝůƐĂƌĞĞdžĐůƵƐŝǀĞůLJŵĂĚĞŽĨĐĞůůƵůŽƐĞͿ͕ ď Ƶƚ ƚŚĞ ŚLJĚƌŽƉŚŝůŝĐ ŐƌŽƵƉƐŵĞĂŶ
ŚĞŵŝĐĞůůƵůŽƐĞƐŚĂǀĞŽƚŚĞƌĨƵŶĐƚŝŽŶƐ͘dŚĞLJĂďƐŽƌďŽƌĚŝƐĐŚĂƌŐĞǁĂƚĞƌƚŽĐŽŶƚƌŽůƚŚĞƉĞƌͲ
ŵĞĂďŝůŝƚLJŽĨƚŚĞĐĞůůŵĞŵďƌĂŶĞŽƌŵĂLJĂůƐŽĂĐƚĂƐŵĂƚƌŝdžĐŽŵƉŽŶĞŶƚƐĂŶĚƐĞĂůĂŶƚ͘
>ŝŐŶŝŶ
ŽŶƐƚŝƚƵĞŶƚƐ
Ɛ ƉƌĞǀŝŽƵƐůLJ ĞdžƉůĂŝŶĞĚ͕ ƚŚĞ ĐŽŶƐƚŝƚƵĞŶƚƐ ĚĞƉĞŶĚ ŽŶ ƚŚĞ ƚƌĞĞ ƐƉĞĐŝĞƐ ĐŽŶĐĞƌŶĞĚ ĂŶĚ
ĚĞƐƉŝƚĞĐŽŵƉƌŝƐŝŶŐŽŶůLJĂŵŽĚĞƐƚƉƌŽƉŽƌƚŝŽŶŽĨƚŚĞŽǀĞƌĂůůǁŽŽĚŵĂƐƐ͕ĐĂŶĞdžĞƌƚĂŵĂũŽƌ
ŝŵƉĂĐƚ ŽŶ ǀĂƌŝŽƵƐ ƉƌŽƉĞƌƚŝĞƐ͘ ,ŽǁĞǀĞƌ͕ ƚŚĞ ƉƌŽƉŽƌƚŝŽŶ ŽĨ ĐŽŶƐƚŝƚƵĞŶƚƐ ŶŽƚ ŽŶůLJ ǀĂƌŝĞƐ
ĂĐĐŽƌĚŝŶŐƚŽƚŚĞƚƌĞĞƐƉĞĐŝĞƐ͕ďƵƚĂůƐŽǁŝƚŚŝŶĂƚƌĞĞƐƉĞĐŝĞƐŝƚƐĞůĨ͕ŽƌĞǀĞŶĂƐŝŶŐůĞƚƌĞĞ͘
&Žƌ ĞdžĂŵƉůĞ͕ ƚŚĞ ƚƌĞĞ ĚŝƌĞĐƚƐ ĨĂƌ ŵŽƌĞ ĐŽŶƐƚŝƚƵĞŶƚƐ ůŝŬĞ ƌĞƐŝŶ ƚŽ ĚĂŵĂŐĞĚ ĂƌĞĂƐ ƚŚĂŶ
ƚŚŽƐĞ ǁŚŝĐŚ ĂƌĞ ƵŶƐĐĂƚŚĞĚ ĂŶĚ ŚĞĂƌƚǁŽŽĚ ŐĞŶĞƌĂůůLJ ĐŽŶƚĂŝŶƐ ŵŽƌĞ ĐŽŶƐƚŝƚƵĞŶƚƐ ƚŚĂŶ
ƐĂƉǁŽŽĚ͘dŚĞƐĞĐŽŶƐƚŝƚƵĞŶƚƐ͕ŝŶƚƵƌŶ͕ŵĂLJĞdžĞƌƚĂŶŝŵƉĂĐƚŽŶǀĂƌŝŽƵƐĐŚĞŵŝĐĂů͕ďŝŽůŽŐŝĐĂů
ĂŶĚƉŚLJƐŝĐĂůĨĞĂƚƵƌĞƐ͕ƐƵĐŚĂƐƚŚĞƉ,ǀĂůƵĞ͕ƉĞƐƚƌĞƐŝƐƚĂŶĐĞ͕ŽĚŽƵƌŽƌĐŽůŽƵƌ͘ZĞƐŝŶƐ͕ĨŽƌ
ŝŶƐƚĂŶĐĞ͕ĂƌĞƵƐĞĚďLJƚŚĞƚƌĞĞƚŽƐĞĂůƵƉǁŽƵŶĚƐĂŶĚĐŽŵďĂƚƉĞƐƚƐ͖ŝŶǀĂĚŝŶŐŝŶƐĞĐƚƐĂƌĞ
͞ƐƚƵĐŬƚŽŐĞƚŚĞƌ͘͟'ƌĞĂƐĞƐƌĞƉĞůǁĂƚĞƌ͖ĐŽůŽƵƌĂŶƚƐĂůƚĞƌƚŚĞĐŽůŽƵƌĂŶĚƐĞŶƐŝƚŝǀŝƚLJƚŽůŝŐŚƚ͘
dĞƌƉĞŶĞƐĂŶĚƉŚĞŶŽůƐŚĂǀĞŵŝĐƌŽĐŝĚĂůĞĨĨĞĐƚƐ͘tŝƚŚƚŚŝƐŝŶŵŝŶĚ͕ŽŶĞŝŶƚĞƌĞƐƚŝŶŐĨŝŶĚŝŶŐ
ϯϰ
Wood anatomy
is that trees from areas with a moderate climate generally have fewer constituents,
around 1 to 10% by volume, than those from the tropics, where the equivalent figure
ranges between 2 to 30% depending on the tree species.
B1.8 Literature
P. Hoffmeyer, original Article A4, STEP 1995.
Booker R.E. and Sell J. (1998). The nanostructure of the cell wall of softwoods and its functions in a living tree.
Holz als Roh‐ und Werkstoff 56:1‐8.
Grosser D. (1977). Die Hölzer Mitteleuropas. Ein mikrofotografischer Lehratlas. Springer Verlag, Berlin, 208 p.
Nardi‐Berti R. (1993). La struttura anatomica del legno ed il riconoscimento dei legnami italiani di più corrente
impiego. Istituto del Legno, Consiglio Nazionale delle Ricerche, Florenz, 155 p.
Schmid M. (2002). Anwendung der Bruchmechanik auf Verbindungen mit Holz. Dissertation,
Universität Karlsruhe (TH).
Strasburger E. (1998). Lehrbuch der Botanik. Gustav Fischer Verlag, Stuttgart, 1007 p.
Wagenführ R. (1999). Anatomie des Holzes. DRW‐Verlag, Leinfelden‐Echterdingen, 188 p.
35
B2 Wood physics
Original articles: P. Hoffmeyer, L. D. Andriamitantsoa
Wood physics, a key element of wood sciences in general, draws on findings from wood
chemistry, wood anatomy and biology, as well as classical chemistry, physics and me‐
chanics and can be defined as the “Science of physical‐mechanical properties of wood
and wood‐based materials”. Fields of wood physics meriting further discussion include
how wood behaves when exposed to moisture, wood density and the rheological (visco‐
elastic) properties of the wood.
B2.1 Wood moisture content
The way wood is arranged structurally means its moisture content determines almost all
its other properties, which becomes particularly evident when going below the fibre
saturation point. With increasing moisture:
Stiffness and strength decrease,
Sustained loading increases creep deformation,
The thermal conductivity of wood rises,
Wood becomes increasingly prone to fungal infection,
particularly when the moisture content exceeds 20%.
Table B2‐1 indicates the impact of humidity on certain mechanical properties of wood,
whereby the more moisture is absorbed, the lower the strength observed, as the molecular
binding forces decline. Namely, the ingress of water into the intermicellar and interfibril‐
lar cavities forces the cellulose chains far apart. In an oven‐dry state, the close proximity
of the cellulose chains in the wood helps generate strong intermolecular binding forces,
which render the structural and bonding substance stiff and brittle. However, as water
penetrates the cell wall and more moisture is absorbed, the forces of attraction decline
and the hydrogen bonds holding the cell wall together are weakened. Changes in mois‐
ture content exceeding the fibre saturation point have no further effect on mechanical
properties, but simply lead to excess water accumulating in the cell cavities, while a com‐
bination of high temperature and humidity render the wood far more malleable and
prone to permanent deformation. This is precisely the property exploited when making
so‐called (Thonet) coffeehouse stools from steam‐bent beech wood.
37
Wood physics
The impact of changing levels of moisture content on the various mechanical properties
varies. For example, the wood may break down when compressive stress is applied along
the grain due to the fibres buckling, a process in which the moisture‐sensitive hydrogen
bonds play a key role. Conversely, tensile breakdown along the grain signals a break in
the covalent bonds as the micro‐fibrils of the cell wall are torn apart. Accordingly, com‐
pressive strength is more sensitive to humidity than tensile strength.
Table B2‐1 Change in the properties of clear wood with a one percent change in wood moisture.
The benchmark figure is the properties at a moisture content of 12%.
Property Change
Compressive strength parallel to the grain 6%
Compressive strength perpendicular to the grain 5%
Bending strength 4%
Tensile strength parallel to the grain 2.5%
Tensile strength perpendicular to the grain 2%
Shear 2.5%
Modulus of elasticity (MOE) parallel to the grain 1.5%
It is crucial to ensure appropriate moisture content with subsequent construction use in
mind. This also helps avoid unwanted swelling and shrinking movements of the wood,
which result in timber members cracking or warping of panels in the absence of expan‐
sion gaps.
Wood is a capillary‐porous material, with porosity ranging from around 50 to 70% de‐
pending on density and a correspondingly huge internal surface area. Its cavity system is
hygroscopic, capable of absorbing airborne moisture, while capillary transport processes
in the cell lumen also allow liquid water or other liquids (e.g. wood preservative or adhe‐
sives) to be absorbed. Depending on the water content of the wood, three boundary
conditions are distinguished:
Oven‐dry
The wood contains no water, meaning 0% moisture content.
Fibre saturation point
The entire microsystem of the wood is filled with water. The fibre saturation point
occurs at around 28% and differs slightly depending on the type of wood concerned.
Water saturation
The microsystem and macrosystem (cell lumen) of the wood are filled with water.
38
Wood physics
The term used for the proportion of water in the wood up to the fibre saturation point is
bound water, since the stored water is bound in cellulose hydroxyl groups within the micro
system, by hydrogen bonds in the cell walls. The remaining water in the macro system up to
the water saturation point is termed free water and found in the cell lumen. The wood
moisture content u is defined as the ratio of the mass of water contained in the wood
(mu ‒ mdtr) and the mass of the dry wood (mdtr), equation (B2‐1), where the oven‐dry state
is the reference variable, meaning moisture content exceeding 100% is possible.
m u m dtr
u 100 (B2‐1)
m dtr
where
u Moisture content of the wood
mu Mass of the moist wood
mdtr Mass of the oven‐dry wood
When freshly cut timber is dried, the water initially emerges from the cell cavities. This
water has no molecular link to the wood and is known as free water, as opposed to the
water within the cell walls, which is linked to the cell wall via hydrogen bonds (to hydroxyl
groups) and Van der Waals forces and thus known as bound water. This is why far more
energy is required to drive out water from cell walls compared to eliminating free water.
The moisture content at which the cell walls are deemed saturated with water, but prior
to the presence of any free water in the cell cavities, is known as the fibre saturation
point, which means all “free (–OH) groups are occupied” and the only further water that
can be absorbed is free water. For most wood species, the fibre saturation point is be‐
tween 25 and 35%, making 28% a reasonable mean for most practical applications. For
engineers, knowing the fibre saturation point is crucial, given the significant changes to
most physical and mechanical properties before it is attained. Above the fibre saturation
point, however, most of the properties are virtually constant.
As a hygroscopic material, wood constantly discharges and absorbs moisture from its
surroundings. At any one combination of temperature T and humidity of the surrounding
air , a corresponding level of moisture content applies, where the moisture diffusing
into the wood balances out the moisture escaping from the same. This moisture content
is known as the equilibrium moisture content , although prevailing climatic changes
mean wood is seldom actually in this state. The moisture content and even the extent
and speed of the moisture transport have a key impact on almost all relevant engineering
properties of the wood.
39
Wood physics
The way the relative humidity at constant temperature dictates the self‐adjusting equilib‐
rium moisture content via so‐called sorption isotherms is shown using a typical S‐shaped
course in Figure B2‐1. The incongruent nature of the isotherms for moisture absorption
(adsorption) and moisture release (desorption) is also clear and this hysteresis effect
explains why the adjusted moisture content of the wood for desorption exceeds that of
adsorption by around 1 to 2%. Accordingly, under uniform ambient conditions, if you dry
a piece of wood from a wet state to the equilibrium moisture content, the equilibrium
moisture content will exceed that for adsorption. Engineers can leverage the sorption
hysteresis of the wood: since wood exposed to an alternating climate shows fewer
changes in moisture content with given humidity changes than would otherwise be ex‐
pected without hysteresis, sorption hysteresis lowers the effective gradient of current
sorption isotherms. This means, in turn, that any changes in e.g. wood dimensions linked
to varying humidity levels may be smaller than expected.
Free
water
Fibre saturation
Desorption
Moisture Bound
content[%] water
Adsorption ' = 1 - 2%
0% 25 50 75 100%
Relative humidity
Figure B2‐1 Sorption isotherms with the hysteresis effect. Sorption and desorption isotherms reveal
a so‐called hysteresis loop.
40
Wood physics
a cross‐sectional piece of spruce wood measuring 50 mm x 100 mm and with a moisture
content of 20% to attain an internal equilibrium moisture content of 10%, at 20°C and
relative humidity of 54%. This means that the equilibrium moisture content of a timber
member will be obtained at an earlier stage with longer exposure to the relevant average
temperature and relative humidity than with shorter cycles involving higher or lower
humidity. Figure B2‐2 shows how the equilibrium moisture content is determined for a
wood species under specific ambient conditions. In this example, the adjusted equilibri‐
um moisture content of around 11% is marked at a relative humidity of 50% and a tem‐
perature of 20°C.
Moisture content u [%]
30
20
10
0
0 20 40 60 80 100
Relative humidity \ [%]
Figure B2‐2 Schematic equilibrium moisture content u of 11% with T = 20°C and 50% relative humidity
Swelling and shrinking behaviour of wood
The water molecules stored in the intermicellar and interfibrillar spaces over the hydroxyl
groups lead to the cell walls expanding, as shown schematically in Figure B2‐3 in a pro‐
cess known as swelling. Conversely, the volume contraction that occurs when moisture is
released is termed shrinking. Since the expanding/contracting of cell walls dictates
whether water molecules are accumulated or released, swelling and shrinking is limited
to the hygroscopic area alone. There is no further swelling and shrinking above the fibre
saturation point, since only free water can be absorbed or discharged from this point.
Swelling
Shrinking
cellulose chains
cellulose chains
-OH HO-
H2O
Figure B2‐3 Swelling and shrinking. (www.holzfragen.de)
41
Wood physics
Wood is a roughly polar orthotropic material, although this is usually considered irrele‐
vant for practical construction purposes. The calculation only takes into account two
section planes of the wood; parallel and perpendicular to the grain respectively. Howev‐
er, this simplified summary of the radial and tangential section planes to the section
plane perpendicular to the grain is infeasible for swelling and shrinking processes. Even
so, construction practicalities mean tangential and radial shrinking is, in turn, collectively
considered as shrinking in a perpendicular direction, since the type of cut applied to a
piece of wood varies immensely and the engineer lacks such details. Figure B2‐4 shows
the swelling of the common beech as an example. Swelling varies considerably in radial
and tangential directions. Longitudinal swelling is very low, since the micro‐fibrils of the
layer are oriented in the fibre direction with very few fibrils perpendicular to the grain.
This also means that, conversely, numerous free hydroxyl groups of the cellulose chains
can be found perpendicular to the fibre direction, whereupon far more water can accu‐
mulate perpendicular to the grain, intensifying any resulting swelling/shrinking. Fig‐
ure B2‐4 also shows that swelling stabilises from the fibre saturation point (≈ 35% for
beech) onwards.
20
% Common beech V
18
16
14
T
12
Swelling
10
8
R
6
4
2
L
42
Wood physics
On average, the degree of longitudinal swelling for European woods is 0.4%, for radial
swelling 4.3 and for tangential swelling 8.3%; the level of longitudinal swelling is thus just
a tenth of that which occurs in radial or tangential directions. Key parameters influencing
the swelling and shrinking behaviour of various wood species include the wood density,
latewood proportion, anatomical structure and lignin proportion. The lignin content
influences the swelling and shrinking behaviour, since lignin is far more hydrophobic than
cellulose and strongly lignified wood species swell far less than more weakly lignified
species. However, the key influential parameter is the angle of the cellulose fibril in the
S2 layer (see Article B1). In compression wood fibres for example, the fibrils of the S2
layer are at a wider angle to the fibre direction than with normal fibres (compression
wood, see Article B3, section on reaction wood), which increases the degree of longitudi‐
nal swelling and reduces the level of swelling in radial and tangential directions. In com‐
pression wood, the angle of the cellulose fibril in the S2 layer may be up to 45°, in which
case equivalent humidity deformations occur in longitudinal and perpendicular directions.
The deformations of cut wood (Figure B2‐5) can be effectively clarified with the “axial
symmetrical” swelling and shrinking behaviour mentioned previously. The range of swell‐
ing and shrinking behaviour in three main directions results in significant deformation of
the wood and internal stresses and explains, for example, why the left side of a board,
namely that facing away from the pith, is always concave when drying out. If not dried
properly (if excessive humidity gradients are present between the outer and inner layers
of the wood), this may result in extensive cracking, particularly in wood with a higher
density (and particularly in a radial direction), so swift drying is best avoided.
Figure B2‐5 Distortion of wood due to varying swelling. (Kollmann and Coté, 1968)
43
Wood physics
The dimensions of the wood vary linearly with a wood moisture content ranging between 5
and 20%, within which range, humidity deformations can be calculated from the following:
h 1 100 u 2 u 1
h2 (B2‐2)
100
where h1 and h2 are the dimensions (thickness) for a wood moisture content of u1 or u2.
is the degree of swelling (positive) or shrinking (negative) in %/%.
In the absence of specific values for the wood species in question dictating the degree of
swelling and shrinking, an approximation equation may be used which means that the
degree of volume swelling and shrinking V corresponds to the numeric value for the
density in g/cm3. This means that the volume of wood with a density of 0.4 g/cm3 swells
by an additional 0.4% with each 1% increase in moisture content. The implicit and fun‐
damental consideration here is that the degree of volume swelling is equivalent to the
volume of the absorbed water. While the degree of swelling and shrinking in a longitudi‐
nal direction 0 may generally be discounted, the level in a perpendicular direction 90
equates to half the degree of volume swelling or shrinking.
For most wood species, such as spruce, fir, pine, larch, poplar and oak, engineers can use
values 0 = 0.01%/% and 90 = 0.24%/%. For denser wood species such as beech (Fagus
sylvatica) and ekki/bongossi/azobé (Lophira alata), there is more swelling in a perpendic‐
ular direction: 90 = 0.3%/% should be used for beech and 90 = 0.36%/% for ekki.
In plywood, the humidity deformations in the panel plane are of the same order of mag‐
nitude as those of wood in a longitudinal direction. For other wooden composites, how‐
ever, such as particleboards or fibreboards, these deformations depend very strongly on
the special panel type and production technique used. Perpendicular to the panel plane
meanwhile, the reversible humidity deformations are comparable to those for timber.
However, many panel products exposed to high compressive stresses during production
show additional, irreversible thickness swelling, which is also known as "spring back".
If the expansion of the wood is hampered (e.g. in joints employing mechanical fasteners),
the moisture absorption generates internal forces. The viscoelastic‐plastic behaviour of
the wood means such stresses decline over time, resulting in irreversible dimensional
changes. If the wood reverts to its former humidity, the dimensions will since have
shrunk, meaning the joint will no longer fit properly and will be less able to support any
load. This underlines the need for full access to any construction details if retightening
may be required.
44
Wood physics
B2.2 Wood density
Density, the ratio of mass to volume, is a key parameter when distinguishing wood. Wood
is a capillary‐porous, swellable material, which may contain water, water vapour, air or a
feeder liquid and the mass and volume of which vary according to the proportions of
such substances within the wood. Since the moisture content of wood varies depending
on the ambient climate and the volume also changes accordingly when under the fibre
saturation point, the density depends on the level of moisture. This is why the density
must always be specified for a specific climate (often an ambient temperature of 20°C
and 65% relative humidity (normal or indoor room environment) ‒ around 12% wood
moisture). The density u is the quotient from the mass of wood (including the water
contained in the pores) and the volume of a wooden body (including cavities) for a de‐
fined moisture content u:
mu
u (B2‐3)
Vu
where
u Density at moisture content u
mu Mass of the wood at moisture content u
Vu External volume of the wood at moisture content u
In wood technology and timber engineering, mainly oven‐dry density dtr and density 12
at 12% moisture content are used. The density values (= 12) in accordance with EC 5 or
EN 338 refer to mass and volume at the equilibrium moisture content (≈ 12%), which is
obtained at a temperature of 20°C and relative humidity of 65%. The density values in
accordance with EC 5 or EN 338 either refer to the mean density 12,mean or the character‐
istic density 12,k, which is defined as the 5% quantile. The assumption made for the
strength class of the timber generally involves a normally distributed density with a varia‐
tion coefficient (COV) of 10%. This results in the following: (factor 1.645 for normal distri‐
bution and 0.1 = 10% COV):
The density c of the cell wall is around 1500 kg/m3. Accordingly, the density of the wood
depends on the porosity, which is defined as a volume fraction of the cell lumen. Timber
generally has an oven‐dry density in the region of 300 to 550 kg/m3, which corresponds
to a void volume proportion of 0.8 to 0.63.
45
Wood physics
Influence of wood density on wood characteristics
Density is one of the main variables influencing virtually all wood characteristics. Most
mechanical properties correlate positively to density. Figure B2‐6 schematically shows
the effect of density on certain selected properties.
under compression
Max. moisture
Deformation
Strength
content
Density Density Density
Figure B2‐6 Schematic illustration of the influence of density.
Influences on the density and density distribution of wood
Wood species
The ratio between the cell wall and porosity varies considerably among individual wood
species and the same applies to density. However, density still varies considerably, even
within a single wood species.
Local conditions
The ground, climate and all other ambient conditions (including the interval distance
between trees) have a key effect on the growth of trees and with it, the density distribu‐
tion of wood (see also Article B3). This means, for example, that slower‐growing conifers
from colder areas of Europe have higher densities than faster‐growing conifers of the
same species from the Mediterranean region.
Early‐ and latewood proportion
The early wood of conifers has a lower density than latewood, since the cell walls are
considerably thinner and the cell lumen larger.
Annual ring width
Smaller annual rings mean a greater latewood proportion and hence a higher density of
softwood, while for deciduous trees, the opposite applies (see also Article B3, Figure B3‐2).
46
Wood physics
B2.3 Influence of temperature
Compared to swelling and shrinking of the wood, the thermal expansion of the wood is
far less significant. Any change in temperature, however, also alters the moisture content
and thus the swelling and shrinking deformations. The degree of swelling and shrinking
perpendicular to the grain is around tenfold greater than the thermal changes. Tempera‐
ture changes over a wide area affect the properties of the wood, while the strength
values of the wood decline with increasing temperature and vice versa (see Table B2‐2).
Further correlations and the fire behaviour of wood and wood‐based products are shown
in Article G1.
Table B2‐2 Influence of temperature on certain properties with an increase of 20°C to 100°C.
Property Change
Bending strength ‐28%
Tensile strength ‐8%
Compression strength ‐44%
Modulus of elasticity (MOE) in bending ‐17%
B2.4 Rheological properties
The flow and deformation properties of a material are referred to as rheological proper‐
ties and may apply to a solid material (→ elasticity, plasticity) or a liquid. The term used
when referring to liquids is viscosity; the higher the viscosity, the more sluggishly the
liquid flows. Wood exhibits both elastic and viscous, time‐dependent behaviour, which is
why it is known as a viscoelastic material. Its viscoelastic properties are classified into
categories including:
Creep
Relaxation
These two key aspects of the rheological behaviour of wood are briefly discussed in the
following section and creep behaviour is by far the most critical parameter of the two for
timber construction engineers. In design meanwhile, creep is taken into account using
kdef as a coefficient.
47
Wood physics
Creep
Creep is described as an increase in deformation over time under constant load. Fig‐
ure B2‐7 above shows a load scheme which may result in creep deformations in wood
shown subsequently in Figure B2‐7. From the time t0 at which the load is first applied, the
elastic deformation of the wood also progresses up to time tinst. When reaching tinst, the
elastic deformation uinst under applied load F is completed and immediately followed by
creep deformation. The creep curve can be subdivided into two areas. While creep de‐
formation increases rapidly at the start, it is then consolidated with a constant creep speed.
When load is removed at time tfin, this results in a recovery, which would revert to the
value u = 0 if ideal viscoelastic behaviour applied. However, this is usually not the case for
wood, depending on the load level and loading duration. A permanent deformation uplast
after the entire load is removed indicates that the microscopic area of the wood has
been damaged, so it is technically inaccurate to refer to wood as a viscoelastic material.
F
ufin
u
uinst
R
uplast
t0 tinst tfin
t
Figure B2‐7 Load (above) and creep deformation caused (below), horizontal axis time axis.
(STEP 1995 Article A19)
48
Wood physics
Variables influencing creep
The main parameters affecting creep behaviour are:
Moisture content and changes in moisture content,
Duration of load,
Temperature,
Level of stress.
The first two factors are what influence creep more than any other. The wood moisture is
included in the calculation via the selectable “service class” and the duration of load via
the “load‐duration class”. Creep behaviour also varies considerably among different
materials, meaning wood‐based panels are generally more prone to creep than solid
timber and the smaller the wooden components of the panels, the more pronounced the
actual degree of creep. This means that for hardboards in service class 2, a kdef value of
3.00 applies, 2.25 for OSB and 0.80 for solid timber, glued laminated timber and laminated
veneer lumber (LVL).
Figure B2‐8 shows creep curves for wood under various moisture conditions. The curves
shown also reveal an additional effect, namely the so‐called mechano‐absorptive creep,
which is a collective term referring to swelling, shrinking and creep. The creep defor‐
mation increases during the drying phase (shrinking, curve (b)) and declines in the mois‐
ture penetration phase (swelling, curve (d)). Figure B2‐8 clearly shows the great extent to
which this mechano‐absorptive effect affects the creep curves and the swifter and
stronger the changes in wood moisture, the greater the impact the wood moisture
changes will have. Meanwhile, the influence of the stress level is shown in Figure B2‐9.
The greater the load applied, the more rapidly and intensely creep progresses. The influ‐
ence of temperature on creep can also be shown using the chemical structure of wood,
where the polymer structure means creep deformations rise with increasing temperature.
49
Wood physics
u/uinst
(b)
3
2
(d)
(c)
1 (a)
0
0 8 16 24 32
t(d)
Figure B2‐8 Relative deformation‐time curves for beams at different moisture conditions. (a) green timber
kept green; (b) green timber drying to 12% moisture content; (c) timber kept at 12% moisture
content; (d) timber initially at 12% moisture content allowed to absorb moisture. t(d) is time in
days. Alpine ash, 24% of average short‐term strength, T = 25°C. (STEP 1995 Article A19)
σ5
u
σ4
σ3
σ2
σ1
t
Figure B2‐9 Influence of the stress level on creep, 1 < 2 < 3 < 4 < 5, u is deformation,
t is time. (STEP 1995 Article A19)
50
Wood physics
Relaxation
The decrease of stresses with constant deformations over time is described as relaxation
and particularly significant for prestressed structures. Figure B2‐10 shows relaxation
under compressive stress perpendicular to the grain over the course of a year. This phe‐
nomenon means that prestressed members must be retensioned, which was done twice
in Figure B2‐10. While the compressive stress perpendicular to the grain declines during
the drying phase due to shrinking of the wood, it increases during the moisture adsorp‐
tion phase. As the number of cycles increases, however, the stress declines dramatically.
1.0
0.8
Stress ratio (ρ)
0.6
0.4
ρ1 ρ2 ρ3
0.2
0.0
0 60 120 180 240 300 360
Elapsed time, t, (day)
Figure B2‐10 Stress curve over a year for a prestressed and twice re‐stressed connection.
(Awaludin et al. 1999)
B2.5 Literature
P. Hoffmeyer, L.D. Andriamitantsoa, original Articles A4, A19, STEP 1995.
Kollmann F.F.P and Coté W.A. (1968). Principles of wood science and technology. Volume I, Solid wood.
Springer Verlag, Berlin, 592 p.
Awaludin A., Hirai T., Hayashikawa T., Sasaki Y. and Oikawa A. (2008). One‐year stress relaxation of timber joints
assembled with pretensioned bolts. Journal of Wood Science 54(6):456‐463.
51
B3 Wood formation
Original articles: P. Hoffmeyer, L. M. R. Nunes, P. P. de Sousa
The life of the tree is the fundamental basis for understanding wood as a raw and con‐
struction material, since it exerts a greater impact on properties which affect the end
usage than many other raw materials. This also explains the strong focus on the individu‐
al details of a tree when assessing the raw material. The properties of the wood depend
on the tree species, the individual tree and the tree section, but the world of nature
encompasses innumerable structures, dimensions and shapes, which can even be a plus
when special requirements are imposed (e.g. formerly: curved tree stems for hulls).
Nevertheless, amid increasing industrialisation and the overwhelming trend to pursue
rational large‐scale serial production methods, variability in wood characteristics within a
particular timber variety is considered undesirable. Knowing details of how wood for‐
mation is influenced and possible changes in the wood can help mitigate such problems.
B3.1 Wood formation
The stem
The stem shape, Figure B3‐1, is determined by the number and width of annual rings in the
wood and the bark thickness. Depending on the yearly shoot lengths, meanwhile, the num‐
ber of existing annual rings may decline with increasing height. If the shoot lengths and
annual ring widths were the same at all stem heights, the debarked stem would be conical
in shape. However, since the shoots vary in length and the width of the annual rings is also
subject to change, tree stems deviate from the conical shape to a greater or lesser extent.
Over the lifetime of the tree, the distribution of annual ring widths may vary at various stem
heights, particularly following any change in the prevailing environmental conditions. If, for
example, older trees, which were cultivated in dense proximity, are released, the greatest
increase in diameter comes in the lower portion of the stem; a phenomenon particularly
prominent in coniferous rather than deciduous trees. In free‐standing or released trees
meanwhile, the annual ring width may decline from the bottom up.
The vertical classification of a tree into the roots, stem and crown or the roots, trunk and
branches changes throughout its life, which is why, in young trees for example, the crown
initially grazes the ground. The stem and crown first establish separation as the lower
branches die, in a phenomenon which intensifies as the branches fall away and the
stumps are overgrown with knot‐free stem wood. The vertical classification of trees,
particularly the stem length and crown size, is influenced by the distance between the
trees. Free‐standing trees retain abundant branches deep into old age and feature a
53
Wood formation
shorter stem full of branches and a larger treetop. When trees are in close proximity,
however, the shadowing of the lower branches tends to result in a proportionally longer
stem and smaller crown from a relatively early stage. The most valuable wood in terms of
length, diameter and wood quality is obtained from the stem. Varieties of wood which
fall from the crown area generally stand out due to the reduced dimensions as well as the
inferior internal quality (e.g. larger knots).
Figure B3‐1 Stem shapes: left prismatic, right tapered. Tapering tends to occur; particularly in free‐standing
trees and at leeward forest areas. A significant decline in diameter to the smaller stem end
(= tree taper) has a very detrimental effect on usability and yield. (Steuer, 1990)
The annual rings
Variation in ring width is dictated by the wood species and age of the tree, its vitality, its
positioning among others, environmental factors such as the supply of water, nutrients,
light and heat and silvicultural care measures. Forest fires and insect damage can also im‐
pact on the annual ring width. In addition to the change in annual ring width, change in the
latewood proportion in conifers and ring‐porous hardwoods is also a key pointer indicating
variation in annual rings. The water flow is limited to the areas of earlywood, meaning that
the stem cross‐section features concentric rings of water‐conducting tissue with corre‐
sponding dry zones. Curves c in Figure B3‐2, which indicate the way earlywood width rises
with increasing ring width, show that in highly developed ring‐porous hardwoods (ash,
Figure left), the tissue for the water flow forms independently of the growth rate, while
among conifers (pine, Figure right) the opposite applies. In ring‐porous hardwood types,
meanwhile, the earlywood width is virtually constant, while latewood width increases de‐
pending on growth. In contrast, the latewood width shows minimal change with increasing
ring width in conifers, which explains differences in annual ring widths as well as the late‐
wood proportion of fine ring‐patterned spruces from Northern Scandinavia and of coarse
ring‐patterned spruces containing a high proportion of earlywood from a good site in Ger‐
many. These connections impact on density and thus on mechanical properties.
54
Wood formation
100 100
% Ash a mm % Pine mm
80 4 80 4
Latewood proportion
Early‐/latewood width
c
Latewood proportion
Early‐/latwood width
b
60 3 60 3
40 2 40 2
a
20 1 20 1
c b
0 0 0 0
0 1 2 3 4 mm 0 1 2 3 4 mm
Annual ring with Annual ring width
Figure B3‐2 Early‐ and latewood width and latewood proportion depending on the width of annual rings in
Fraxinus excelsior (ash, left) and Pinus silvestris (pine, right); latewood proportion (a), latewood
width (b), earlywood width (c). (Knigge and Schulz, 1966)
The branches/knots
When it comes to structural use, the branches of a tree are a key feature of timber, since
each knot actually disrupts the wood structure. Most of this disruption involves localised
deviations from the straight grain, since the stem fibres have to grow around the knot. As
far as vertical branch distribution on the stem is concerned, though, differences emerge
between each tree species; namely some in which the branches are locally massed and
others in which they are scattered around. The most prominent accumulations of
branches occur in tree species in which the branches are arranged in whorl shapes
(Figure B3‐3) and which feature knot‐free or virtually knot‐free interim longitudinal sec‐
tions of annual shoots, such as the pine, spruce and fir. The impact of environmental
conditions often means branches are thicker on one stem side (e.g. trees skirting the
forest). The diameter of branches having formed at the stem tends to increase on aver‐
age from the bottom up to the crowning height and then decline within the crown space
itself. For coniferous species in particular, however, small branches still proliferate, even
in higher areas of the stem.
Figure B3‐3 Cross‐section with whorling branch.
55
Wood formation
Forestry measures are a highly significant factor influencing average branch thicknesses,
since they determine the growing space available to the individual plants via planting
distance. The more spaced out the trees and the greater the incidence of light, the more
growth is accelerated and branches at the lower edge of the crown live much longer,
which means they are larger in diameter. Pruning is another vital forestry measure, which
can allow thicker knot‐free layers to emerge as a result. Branch thickness directly impacts
on the quality of the wood, based on the extent of disruption to the wood structure and
indirectly via the periods of decay and overgrowth, which dictate the condition of the
snag. As vertical growth continues and new branches form, the lower branches are in‐
creasingly shadowed by younger crown parts and neighbouring trees and eventually die.
At the stem surface meanwhile, a protective barrier tends to form within the dying
branch, which segregates the healthy and living tree sections from the diseased branch
sections dying and dropping off. It is often also macroscopically visible thanks to the dark
colouring, as a narrow and sharply delineated zone. After dying, the branch is particularly
prone to attack and decay, initiated by fungi but also insects, in a process which usually
occurs more rapidly in deciduous trees than most coniferous varieties. The branches of
deciduous trees, for example, tend to break off completely, namely in one piece. In many
coniferous species, however, they break off in piecemeal fashion. Stumps emerging from
the stem thus vary considerably in length, or are preserved for a shorter or longer time
respectively which, in turn, influences how the timber is used, since the snag is encircled
by the secondary growth of the trunk. In fact, it lacks any form of bond to the surround‐
ing stem wood and resembles a foreign body in otherwise healthy timber (Figure B3‐4).
Even the healthy portion of the branch near the pith, which is connected to the wood, is
seen as a disruption, since the type and orientation of its cells and hence properties
clearly differ from those of the surrounding stem wood. The overgrowth of wood, which
encircles the snag, generally curves upwards, so that even after the wound has been
closed up, a bulging protrusion remains. Within this, the wood fibres deviate, initially to a
greater and then lesser extent, from the longitudinal orientation.
Figure B3‐4 Formation of the dead branches or knot holes: The dead branch (A) has not completely fallen off.
The subsequent annual rings have completely encircled the branch stub (B). The branch only re‐
emerges when the stem is cut open. (Steuer, 1990)
56
Wood formation
B3.2 Wood characteristics
Generically defining the term “wood characteristics” is no easy task. This is because many
features of tree growth, such as decreasing diameter, knots etc. are often considered
natural from a biological perspective and inevitably subject to specific limits. When it
comes to processing or using the wood in question, though, such characteristics fre‐
quently prove problematic. Accordingly, there is no one‐size‐fits‐all answer to the ques‐
tion of how exactly to define a wood characteristic; rather, it depends on the intended
end use. The growth irregularities and wood characteristics described in this article can
be roughly classified into the following categories:
Deviations from the ideal stem shape
For most usage purposes, the ideal stem shape would be as long a cylinder as possible.
The following are considered undesirable:
Unwanted deviations occurring longitudinally. These include:
curvatures and excessive decline in diameter (e.g. tree taper).
Unwanted deviations, which impact on the cross‐section. For example,
cross‐sections which deviate from the circular form (e.g. fluting).
Deviation from average wood formation
Unwanted deviations from the average chemical and anatomical
composition of the wood (e.g. reaction wood).
Unwanted deviations from the normal orientation of structural elements
relative to the longitudinal axis of the scion (e.g. spiral grain or knots).
Lack of bonding between successively formed wood layers (e.g. in the
event of wounded areas being overgrown).
Ingress of resin, bark etc. (e.g. in the overgrown areas of branches).
Unwanted subsequent changes
Discolourations, which can have a whole range of causes (e.g. action of light,
access to oxygen, fungal infection etc.).
Cracks, which either run radially or in the form of so‐called ring shakes of
annual rings, which are separated from each other, in whole or in part.
Fractures (e.g. treetop/stem breakages).
Holes, primarily caused by wood‐boring insects, as well as by white pocket rot.
Destruction by fungi.
57
Wood formation
Many of the so‐called fundamental features described in the following section (see Fig‐
ure B3‐5) do not occur in isolation, but alongside other characteristics, namely in the
form of combined wood characteristics. Such combined wood characteristics can be
found e.g. in the area of former wounds, where a greater or lesser area of the cambium
perished in the course of the tree’s life cycle. The diversity of the existing fundamental
features in the wound area, such as discolourations, ingrowths, structural changes etc.
can be seen in Figure B3‐6.
of limited use versatile applications
Figure B3‐5 Presenting certain features of trees offering limited use and versatile applications respectively.
(Knigge and Schulz, 1966)
Overgrown area
Bark ingrowths
Wound area
Dehydrated area
(discolouration and possible insect infestation)
Figure B3‐6 Wounds and wound overgrowth in beech wood. (Knigge and Schulz, 1966)
58
Wood formation
Taper
One key characteristic when assessing stems is the degree of tapering (above‐average
decrease in diameter with increasing tree height, Figure B3‐1. A decline in diameter of up
to 1 cm per metre of section is deemed prismatic, stems with any values exceeding this
figure are deemed tapered. The taper is particularly apparent when cutting the logs into
long sawn timber boards. The smaller stem diameter at the tapered end means a decline
in the yield and reduces the overall value. The degree of taper also determines the cut‐
ting losses in the sawmill. Goods produced with this wood have also lower strength, since
parallel cuts expose more annual rings and fibres than in prismatic sections. Tapering
depends significantly on the location and care of the tree, with peripheral trees, for ex‐
ample, particularly prone to tapering.
Curvatures
Curvature is defined as the deviation of the stem from the straight‐line axis and is also
normally always associated with other characteristics such as eccentric position of the
pith, reaction wood, etc.
Deviation from the circular form
Oval stem cross‐sections, as shown in Figure B3‐7, are usually attributable to the effects of
wind, whereby the diameter peaks in the wind direction and the radius for coniferous trees
peaks on the side sheltered from the wind. Inclined stems and those featuring curvatures
are also prone to deviations from the normal circular cross‐section. Additional causes in‐
clude one‐off stresses from the sun, snow push, crown shape and hillside location.
Figure B3‐7 Oval stem shape, eccentric growth.
59
Wood formation
Figure B3‐8 Fluting in robinia.
Another form of cross‐sectional change is fluting, as shown in Figure B3‐8. Here, the
annual rings run in undulatory manner in the cross‐section, resulting in a stem shape
penetrated by channels and meanders which significantly reduce the volume yield. There
is also the danger of warp and a loss of strength. However, the fluting may also surface
sporadically, e.g. in the form of so‐called recesses (proneness to grooves) under strong
branches and is attributable to irregular segmentation in the cambium. Triggers in this
case may include injuries, genetic disposition or butts.
Forked growth
A tree fork (Figure B3‐9) forms due to two trees converging at the stem base or due to
bud damage (e.g. in the ash tree, caused by the ash bud moth), browsing by game or
genetic causes. The need to cut out forked sections reduces the overall yield. Other prev‐
alent issues include sources of rot and discolourations, e.g. formation of false heartwood
in the beech, while forked growth is also often linked to bark ingrowths and intermingling
of fibres.
Figure B3‐9 Tree fork.
60
Wood formation
Spiral grain
When the grain does not rise in parallel to the stem axis but assumes a spiral form, this is
known as spiral grain and becomes apparent when observing the tangential surface, see
Figure B3‐10 on the left. Certain tree species (e.g. the horse chestnut) are known to
always have spiral grain associated with a specific rotation direction. Many coniferous
trees also exhibit a known tendency to rotate to the left while young, whereupon a por‐
tion of the stem is subject to straight grain or right spiral grain at an earlier or later stage.
Tropical tree species in particular show a rhythmic change in the rotation direction (inter‐
locked grain). What has also become apparent is that numerous tree species have very
specific regularities in terms of their spiral graining behaviour and virtually all wood has
some degree of spiral graining, Figure B3‐10 on the right. Spiral grain thus does not re‐
present any exceptional phenomenon, switching from left to right is generally associated
with indigenous conifers. It is also important to underline that the rotation angle of the
fibres normally undergoes numerous changes throughout the life cycle of the tree. Spiral
grain depends on the wood species, the main wind direction and one‐sided foliage.
Figure B3‐10 Left: Schematic presentation of the straight grain in all wood layers (A) and a change in the grain
from left to right (B). (Steuer, 1990) Right: Changes to the rotation angle with tree age in a beech.
(Knigge and Schulz, 1966)
61
Wood formation
In spiral‐grained wood layers, more fibres are diagonally bisected following an incision
made parallel to the stem than in straight‐grown sections, with a corresponding decline
in the strength of the processed sections. Changes in moisture trigger greater differences
in swelling and shrinking behaviour and spiral‐grained wooden bodies are prone to warp,
as dictated by the degree and direction of their rotation. This process has to be taken
into account during wood drying or in poles, for example, this leads to rotation of the
tapered end relative to the embedded base. Cracks in spiral‐grained sections look par‐
ticularly unattractive, since they result in a helical exposure of the section in line with the
grain, reducing the value of a larger area of wood than would otherwise apply to cracks in
straight‐grown stems. Spiral grain also disrupts surface processing, since on the same
wood surface, tools come into contact with the wood fibres at different angles.
Reaction wood
Reaction wood is an active form of direction tissue in trees, which involves the tree at‐
tempting to bring tree sections (trunk, branches) back into their original position, after
they have been moved due to e.g. wind pressure or landslide. The term reaction wood is
used for the “compression wood” of conifers and the “tension wood” of deciduous trees,
Figure B3‐11.
Compression wood forms on the underside of slanted stems and branches of softwood.
Increased lignin storage leads to reddish discolouration of the wood and increases the
density and hardness. The increased incline of the cellulose micro‐fibrils in the S2‐cell
wall layer causes the level of axial swelling and shrinking to soar (see Article B2). Given its
hard and brittle nature, there is limited scope to process compression wood and com‐
pression wood is also very prone to longitudinal shrinking. Tension wood develops on the
upper side of lop‐sided or unbalanced hardwood stems or branches, while the low lignin
content gives it a white to silvery colouring. Unlike the compression wood of conifers,
tension wood is no harder than the surrounding wood, but does show generally in‐
creased axial shrinkage and poorer processability.
62
Wood formation
Figure B3‐11 Eccentric growth and reaction wood. Compression wood of coniferous trees is formed on the
underside of drooping trunks or branches (left), while tension wood of deciduous trees forms
on the upper side of drooping trunks and branches (right). (Bosshard, 1984)
Growth stresses
Growth stresses are internal stresses that emerge during the tree growth due to the
interaction of growth processes (particularly secondary growth) and the weight of the
tree itself. External factors such as the amount of wind and the bending stresses it causes
may also impact on how internal stresses are generated. Such growth stresses are clearly
visible in a fresh felled and cut stem (undried). Felling or split cuts results in sections of
stems which previously held each other in place becoming relaxed and deforming,
whereupon cracking may emerge.
When a fresh stem is longitudinally bisected, both halves bend apart (comparable to cut
flower stems, Figure B3‐12 on the right). This shows that as newly formed wood layers
grow in the grain direction, longitudinal tensile stresses form, but pass over towards the
pith in longitudinal compressive stresses. This is confirmed if the fresh stem is cut per‐
pendicular to the grain, whereupon the cutting surfaces undulate due to compressive
stresses in the stem centre and the tensile stresses show convex movement at the stem
edge. These internal stresses have already formed by the time the tree is standing; inten‐
sifying with age and stem diameter and capable of exacerbating or triggering incidence of
heart shakes, star shakes, ring shakes and schilfer shakes (Figure B3‐12 on the left).
63
Wood formation
Frost crack
Ring shake
Small
heart shake
Large
heart shake
Superficial
shake
Ring shake
Figure B3‐12 Left: Presentation and description of different crack/shake types. Right: Beech planks,
subject to significant deformation right after being cut. (Knigge and Schulz, 1966)
Pith
As far as timber applications are concerned, the pith component is often unwanted. It is
seen as particularly disadvantageous when it shows a very uneven course. In older stems
meanwhile, short cracks often emerge from the pith, which exacerbate the level of defects.
False heartwood
Many older hardwood trees frequently respond to environmental factors by forming so‐
called false heartwood during the process of transforming sapwood to heartwood. Such
discolourations tend to be undesirable, because of the wide‐ranging colours produced
within the same core area, dark edge zones and irregular shapes. Figure B3‐13 shows
incidence of false heartwood in a stem slice taken from a pear tree.
Figure B3‐13 False heartwood in a pear tree.
64
Wood formation
Pitch pockets and pitch shakes
Pitch pockets are formed when the cambium sustains damage and remain visible as
localised wood discolourations following overgrowth. Pitch pockets occur in spruce, larch
and pine, but not in fir trees (since the latter contains no resin) and frequently lead to
incomplete heartwood formation, particularly in larch. During processing, they are cut
from a range of angles; the end product is considered substandard cut timber, which
stands out for its loss of strength, discolouration and the unappealing appearance when
installed in a visible area.
Irregularities in annual ring formation
Prominent wood characteristics associated with the anatomical structure of the wood
also include all irregularities in annual ring formation, which are attributable to wide‐
ranging causes, particularly the impact of weathering. The annual ring width reveals
pointers to certain wood characteristics, such as its internal knots, density, consistency of
quality, surface properties of timber products and geometrical and dimensional stability.
Many quality characteristics of the wood are closely linked to the annual ring width. In
softwood for instance, narrow annual rings point to more favourable physical features of
the wood, since they indicate a higher proportion of latewood. In hardwood, particularly
oak, narrow annual rings indicate a relatively soft wood. Wide annual rings in oak, for
example, point to hard wood, since wide annual rings indicate a higher latewood propor‐
tion, a marker of hardness. It follows therefore that changes in annual ring widths reflect
how the wood hardness varies, which impacts on aspects including surface processing.
The annual ring width also dictates density in particular, which, in turn, is the key deter‐
mining factor for many physical wood properties. Accordingly, wood in which the annual
ring widths differ is subject to varying degrees of shrinkage or moduli of elasticity. The
annual ring width shrinks to a minimum of 1 mm (e.g. in yew) and peaks at the level of a
few centimetres (e.g. for poplar or radiata pine).
65
Wood formation
Knots
We have already explored what may, at times, be a dramatic reduction in the strength
properties of timber members caused by knots, where quantity is less important than the
diameter and position of individual knots. The effect of an individual knot on defective‐
ness of the wood is influenced by multiple factors, the relevance of which may increase
or decline depending on the purpose of use:
By knot diameter. The thicker a knot, the greater its disruptive effect on the end
use of the wood in question.
By the length of the intergrown knot. Short snags, which end near the pith, allow
thicker knot‐free layers to form, which is less disruptive to the wood.
By branch angle. Assuming unchanged length, steeper branches traverse a longer
longitudinal, but shorter cross‐sectional area of the respective section of wood.
Depending on the purpose of use, either a larger or smaller branch angle may be
more advantageous.
By the proportion of various knot areas in intergrown snags. The section of branch
facing the pith, where the wood is tightly intergrown and healthy (living or sound
knot area) is more advantageous than the external stump portion, nearer to the
point where the branch has broken (dead or unsound knot area).
By the first cut of the knot. An incision through the centre of the stem often sees
knots released following a longitudinal cut in the form of so‐called spike knots. A
tangential incision, meanwhile, reveals round or even oval knot cross‐sections,
which tend to only reveal a specific area of the knot. For branches with a large
proportion of dead wood, cutting into planks often reveals so‐called dead knots,
which stand out due to their lack of bonding with the surrounding wood and
which easily fall off during the drying process.
Knots represent the key wood characteristic as far as mechanical properties are con‐
cerned and are classified based on their surface appearance, Figure B3‐14.
66
Wood formation
Wood characteristics influenced by forestry operations
Trees may also sustain damage due to forestry measures or forest visitors, which mainly
involves injuries to the cambium. Noteworthy examples also include felling damage to
felled tree stems:
Stems are particularly prone to rupture, if the upright stem is subject to one‐sided
pressure and the directional notch lacks sufficient depth. While in storage, crack
edges pave the way for secondary damage (e.g. fungal attack).
Breakages and cracks affect higher stem sections and result in either stem break‐
age or the branch stubs breaking off. The impact of strong and steep branches
(particularly when forked) often leaves the stem splitting downwards.
Cracks which go unnoticed and which are not taken into consideration at the time of
cutting can adversely affect timber strength. Moreover, the felling process can also result
in damage to neighbouring trees (damage to bark due to rubbing, impact damage) or
damage to the undergrowth. Damage akin to felling may also occur when transporting
the timber through the forest (skidding damage) and both these forms of damage leave
the timber prone to attack by fungi and insects. Felling at the wrong time (summer
felling) may result in discolourations, while incorrect storage may result in decay damage
(due to moisture or the lack of underlay), intensified crack formation (due to solar radia‐
tion) and discolouration.
Wood characteristics caused by extreme wheathering
The impact of weathering can damage a living tree in such a range of ways that potential‐
ly all the basic characteristics listed may be at risk. The most frequent examples are
breakages, cracks and cambium damage and their consequences, while weather‐related
structural characteristics and damage to living wood cells also play a role. Weather‐
related wood characteristics often come into play during extreme temperatures and
drought, sudden changes in temperature and under ambient conditions, where the tree
lacks sufficient time to adapt. Temperature and moisture influence wood formation to
such an extent that they are considered the main cause of any irregularities in the annual
ring formation. The irregularities themselves may vary very widely, but particularly in‐
clude rapidly changing annual ring widths, earlywood widths in coniferous trees and
latewood widths in deciduous trees. Additional weather‐related characteristics include
so‐called sunburn, which can lead to much of the cambium dying off due to overheating,
or frost cracks. Frost cracks occur radially, in a scope that often goes from the stem edge
up to the pith. Lightning can cause radial (lightning) cracks, which usually commence
under the crown and extend up to the stem base. These cracks may exacerbate the prob‐
lem of fungal and insect‐related damage. Among young and thin‐barked trees in particu‐
lar, hail can cause damage in the form of pitch pockets and bark ingrowths to the stem.
Figure B3‐16 shows a schematic diagram of the various basic characteristics.
67
Wood formation
Storm‐related damage can trigger breakages of various kinds. If they affect the stem, the
impact on the wood value is dictated by the height, size and area of breakage, the length
of the fragmented and cracked areas and the extent of secondary damage. The pace at
which the timber is prepared also has a key impact on secondary damage.
Treetop breakages as a wood characteristic are assessed in accordance with the size and
smoothness of the area of breakage. Smaller breakage areas in trees that are still rela‐
tively young can often be effectively overgrown. Provided no rot sets in, once a branch
has been straightened to support the vertical function, the result is a significant bend,
which emerges in a curve along the stem. Gradually however, the compensatory growth
renders the bend virtually imperceptible, meaning that ultimately, within and above the
old breakage area, the only features pointing to the former damage are the eccentric pith
position and multiple emergence of reaction wood. At the former breakage site, the
internal wood layers lack any fixed bond to the overgrowth, which may render this point
weaker. Some examples of treetop breakage are shown in Figure B3‐15.
Compression failures occur when the stem is subject to significant warpage, particularly
in spruces and firs. On the side facing away from the force application (pressure side), the
wood is compressed to such an extent at the point where the load peaks that compres‐
sion failures occur when elasticity limits are exceeded (moreover, compressive strength
in upright trees is lower than that of dry wood). They run perpendicular to the grain and
often on top of each other, from the stem edge to the pith, while the stem side under
tensile stress remains free of damage. The tree then proceeds to form thick bulges over
the compression failures, namely the wound supporting tissue. The damage to the wood
is caused by the loss of strength in the area of breakage and the heterogeneity of the
wood layers formed before and after the breakage. The damage is hardly discernible in
sawn timber, since the tree is not actually broken, but has simply formed pressure folds
in a specific area. These pressure folds are hardly visible, even on planed wood.
Figure B3‐15 Two examples of treetop breakage.
68
Wood formation
Reaction wood Lightning
Treetop breakage
Moon ring
Sunburn
Pitch pocket
Suction crack Compression Hail Mucilage
failure flow
Frost
heart
Hail
Snow
reflection Frost crack
Ring shake
Figure B3‐16 Wood characteristics caused by the impact of weathering. (Knigge and Schulz, 1966)
B3.3 Biodegradation
Insects
Insects, particularly beetles, mainly damage wood via the burrowing of their larvae
(“woodworms”), which may involve a dramatic reduction in the cross‐section and hence
load‐bearing capacity. A difference is established between greenwood insects, which
infest the wood of living trees or freshly felled green wood and dry‐wood insects, which
infest semi‐dry and constructional wood (stakes, fences) or dry timber (attics, furniture).
A third group includes decaying wood insects, which gravitate towards old and rotten
wood. Greenwood insects (e.g. bark beetles, Scolytinae) and decaying wood insects (e.g.
long‐horned beetle, Ergaster faber) can be disregarded in terms of their impact on pro‐
cessed wood.
While wood‐destroying fungi can no longer grow when the wood moisture goes below
20% or so, normal larval development in some species of insects is possible with wood
moisture of just 8 to 12%, which means purely constructive measures (keeping it dry) to
protect the wood against infestation by such insects is infeasible. When the relative hu‐
midity is 60% at 20°C, the wood moisture content is between 10 and 12%, which means
larval development is still possible. Within the forest, methods to prevent such beetle
infestation include prompt removal of the timber, preferably before warmer seasons and
before the beetles migrate. The close links between many pests to the average moisture
69
Wood formation
content of the wood and the egg deposition, most of which takes place on the bark,
suggest the need, however, to debark the wood promptly and either swiftly cut it up and
dry it or, conversely, use wet storage to keep it above the moisture region of danger.
House longhorn beetle (Hylotropus bajulus Linné)
The house long‐horn is native to the whole of Europe as well as Asia Minor and North
Africa and in Central Europe, is by far the most dangerous and economically devastating
destroyer of coniferous timber, although hardwood is spared. Since warm and humid
climatic conditions promote larval development, the house longhorn beetle thrives in
warm and sunlit attics or similarly exposed constructional timber. Typical entrance holes
and burrows can be seen in Figure B3‐17.
Figure B3‐17 House longhorn beetle. Left: entrance holes and destruction under the intact wood surface.
Right: Vein‐like bulges, which point to the presence of a house long‐horn beetle. (Sutter, 1997)
Common furniture beetle (Anobium punctatum)
The common furniture beetle is the most significant native beetle and found Europe‐
wide. However, A. punctatum has also migrated to South Africa, the USA, Brazil, South‐
East Australia and New Zealand. Coniferous and deciduous trees are both attacked, par‐
ticularly sapwood, while durable heartwoods are only affected after initial fungal growth
and tropical wood species like Abachi, Ilomba or Limba are spared. The common furni‐
ture beetle proliferates in buildings as a destroyer of processed and constructional wood,
whereas outdoor wood and wood exposed to rain remains almost unscathed. Given their
overwhelming proliferation as pests, after the house longhorn beetle, the common furni‐
ture beetle and the related Anobiidae‐species are the key wood pests of note. Their
philopatric infestation in particular exacerbates the damage they cause and the main
issue is often the destruction of furniture items, rather than actual structural timber.
70
Wood formation
Brown powderpost beetle (Lyctus brunneus)
Beetles from the powderpost family (Lyctidae) are among the most feared hardwood
pests worldwide, with most representatives of the over 60 species in this family populat‐
ing the warmer regions of the Earth. Having migrated to Germany via the timber trade
and since proliferated, the brown powderpost beetle is now one of the most significant
destroyers of dry wood. It infests large‐pored, high‐starch imported woods (Abachi, Lim‐
ba, Okoumé) and sapwood of indigenous deciduous species (oak, ash, elm, walnut and
chestnut). The powderpost beetle attacks only the sapwood portion of woods with dark
heartwood. While the light sapwood shows evidence of numerous burrows full of bore
dust, the dark heartwood is free of parasites.
Additional wood parasites
Termites are wood pests of paramount economic importance, far exceeding the signifi‐
cance of indigenous species. Our coasts are particularly plagued by Teredo navalis L., a
marine woodworm, and Limnoria lignorum R., a gribble, which threaten harbour structures.
Micro‐organisms
The microbial breakdown of wood is part of the overall course of life cycles and a natural
and necessary process. Using micro‐organisms, particularly bacteria and fungi, complex
macromolecules like wood cellulose are broken down into smaller and simpler molecules
and redirected into the materials cycle (the fungal enzyme attacks the hydroxyl groups of
the cellulose). The microbial breakdown takes place in any location where the ambient
temperature and humidity allow such micro‐organisms to develop. Soil, for example, is an
ideal habitat for bacteria and fungi, as reflected in the rotting of branches and tree
stumps in the forest or the deliberate composting of organic waste. But even structural
timber, which is not in contact with the ground, may be attacked and destroyed by mi‐
cro‐organisms (mainly fungi) since sufficient moisture content is the sole prerequisite.
This is why wood exposed to weather and unprotected or which stands moist within a
building, unable to dry out, is prone to attack from a range of fungi. Indeed, fungi are a
key cause of the breakdown of constructional wood and wooden objects in interior spac‐
es. The optimal range of moisture, in which the most common wood‐inhabiting fungi
flourish, is between 30 and 60%. During the so‐called dry phase, however, certain fungi
(mycelium and spores) can withstand extended dry periods and start regrowing when
living conditions improve. The temperature can also impact on fungal growth, which –
depending on the type of fungus involved – takes place within the range ‐2.5 and +40°C.
71
Wood formation
Classification
Two main groups of wood‐inhabiting fungi can be distinguished, wood‐staining and
wood‐destroying respectively. The problem with wood‐staining fungi is primarily visual,
while the presence of wood‐destroying fungi signals a widespread loss in the substance
and strength of the wood. The fungi are classified in accordance with the damage pattern
and nature of the wood components targeted, as shown in Figure B3‐18.
Moulds (Ascomycota, Fungi imperfecti)
Wood‐staining fungi
Blue‐stain fungi (Ascomycota, Fungi imperfecti)
Wood‐inhabiting fungi
Soft rot (Ascomycota, Fungi imperfecti)
Wood‐destroying fungi Brown rot (Basidiomycota)
White rot (Basidiomycota)
Figure B3‐18 Classification of wood‐inhabiting fungi in accordance with their damage pattern. (Sutter, 1997)
Moulds
Moulds grow on surfaces, without penetrating the wood to any significant extent and the
main giveaway is a discolouration of the surface itself. Moulds do not generally break
down the wood components (cellulose, lignin), so the physical features of the wood
remain intact. Nevertheless, mould infestation is generally a sign of conditions (optimal
temperature from 24 to 28°C, humidity ranging between 30 and 150%), which also pro‐
mote the growth of other, more dangerous fungi. Moulds may also prove a health risk in
interior spaces, particularly if they remain undetected (e.g. under parquet flooring).
Blue‐stain fungi
Blue‐stain fungi grow from within wood, where the dark inherent colour of the fungi
themselves stains the wood a bluish‐blackish colour, hence the name blue stain. Blue‐
stain fungi tend to affect softwood in particular and only target sapwood in trees forming
dark heartwood. Blue‐stain fungi do not generally break down cell walls, which means
they do not trigger rotting or any strength loss of the wood.
72
Wood formation
Soft rot
Soft rot fungi destroy the secondary walls of cells while growing, mainly breaking down
the cellulose, while leaving the lignin intact. Wood affected by soft rot shows tell‐tale
softness, even before any clear signals of destruction become visible to the naked eye.
Wood that has broken down even to a minor extent loses much of its strength. Moist
wood gives when thumb pressure is applied and can be depressed slightly with the fin‐
gernail. Moist wood is also dark in colour, with a musty and soft surface (name). Cracks in
dry wood only emerge at a significantly advanced stage and resemble the cube break
typical of brown rot, although not penetrating to the same depth. Deciduous trees are
more prone to soft rot fungi than conifers. Ideal conditions for soft rot to develop exist in
the soil, namely the ground‐air zone. Similarly, wood used in buildings or other objects
exposed to water and high levels of moisture (for an extended period), may be damaged
by soft rot.
Brown rot
Brown rot is the most significant form of fungal‐related wood destruction affecting inte‐
riors and is triggered by a number of basidiomycetes, which thrive at varying levels of
temperature and moisture. Brown‐rot fungi grow in the cell cavities (cell lumen, pith rays,
resin canals) and from there, proceed to break down the cellulose of the secondary wall
via enzymatic means, although the lignin remains more or less intact. As the cellulose
breaks down, the wood loses its strength and weight. The brown‐rot wood cracks into
deep cubical pieces, with one typical example shown in Figure B3‐19. The two key types
of brown‐rot fungi are true dry rot and cellar fungus.
Figure B3‐19 True dry rot with cubical cracks.
73
Wood formation
Gloeophyllum spp.
Gloeophyllum are brown‐rot fungi, which exhibit the unappealing property of growing
from inside to out. This means any outbreak is only detected at a very late stage and
when the fruiting bodies can be seen from outside, the affected wood will already have
been completely destroyed. Gloeophyllum tend to attack softwood and do not affect
interior spaces. Sites of affected wooden members include bridge constructions, balco‐
nies and windows. Gloeophyllum can tolerate high temperatures (thriving on sun‐
exposed wood and debarked wood) and sporadic dryness.
B3.4 Literature
P. Hoffmeyer, L.M.R. Nunes, P.P. de Sousa, original Articles A4, A15, STEP 1995.
Bosshard H.H. (1984). Holzkunde. Zur Biologie, Physik und Chemie des Holzes. Birkhäuser Verlag, Basel, 312 p.
Grosser D. (1985). Pflanzliche und tierische Bau‐ und Werkholzschädlinge. DRW‐Verlag,
Leinfelden‐Echterdingen, 159 p.
Knigge W. and Schulz H. (1966). Grundriss der Forstbenutzung. Entstehung, Eigenschaften, Verwertung und
Verwendung des Holzes und anderer Forstprodukte. Verlag Paul Parey, Hamburg.
Steuer W. (1990). Vom Baum zum Holz. Nutzholzarten, Holzschäden, Ausformung, Holzernte,
Rundholzsortierung, Verkauf. DRW‐Verlag, Leinfelden‐Echterdingen, 256 p.
Sutter H.‐P. (1997). Holzschädlinge an Kulturgütern erkennen und bekämpfen. Verlag Paul Haupt, Bern, 164 p.
74
B4 Durability
Original articles: G. Sagot, L. M. R. Nunes, P. P. de Sousa
Durability in the sense of resistance to destructive organisms and thus the ability to guaran‐
tee load‐bearing capacity and usability throughout the service life of an object is imperative
for wood as an organic material. The scope of durability generally also includes other as‐
pects such as corrosion resistance of metallic fasteners. However, our remit in this book is
limited to wood‐specific matters, namely the biological effects of elements like fungi and
insects. Frequently occurring damaging micro‐organisms were already covered in Article
B3. In the following section, we will focus on natural durability in particular.
B4.1 Assessing natural durability
As was explained in Article B1, wood incorporates other constituents, as well as the main
chemical components of cellulose, hemicellulose and lignin, which are either absorbed or
formed as sapwood transforms into heartwood and are specific to particular tree species.
Some also have microcidal qualities, like terpene or phenols. Trees in temperate climates
contain around 1 to 10% constituents, but this percentage can be much higher in the
case of tropical woods, ranging from 2 to 30%. The closing mechanisms of the heartwood
cells, which vary according to tree species, have a key influence on natural durability,
while isolating the heartwood in a virtually “hermetic” manner, via tyloses for example,
which is crucial help in warding off the attack of pests and damaging materials. Since
heartwood formation is often linked to the storage of constituents, heartwood is clearly
more resistant to pests than sapwood. Common tree species are assessed in terms of
their biological durability during experiments performed in a laboratory or in the open
air. Figure B4‐1 shows, for example, specimens from durability tests.
Figure B4‐1 Left: Specimens in the laboratory in accordance with EN 113. Right: Outdoor test bodies in
accordance with EN 252. (Rapp and Augusta, 2000)
75
Durability
Durability tests featuring various wood pests and untreated wood species show how the
resistance of the different wood species varies very considerably, which means they are
classified into a range of durability classes. The durability classes are specified in EN 350‐
2:1994 „Natural durability of solid wood“:
Against fungi
1 Very durable
2 Durable
3 Moderately durable
4 Slightly durable
5 Non‐durable
Against insects and wood pests in sea water
D Durable
M Moderately durable
S Vulnerable
SH Where heartwood is also classed as vulnerable
In EN 350‐2, numerous wood species are included in the various durability classes.
Spruce is included in durability class 4 and SH, while a tropical wood spieces Greenhart is
durability class 1 and D (termites). A European hardwood species with tyloses like robinia
achieves a higher durability class 1 to 2 against fungal attack. The classification specified
in EN 350‐2 of durability against fungi applies only to heartwood. Sapwood should
generally be classified in durability class 5, unless otherwise specified. The generally high
durability of tropical wood species compared to those in Europe is primarily attributable
to the constituents. These make up a higher proportion of tropical woods, namely 2 to
30%, a considerable portion of which also exhibit microcidal effects. The higher propor‐
tion depends on the tree species involved – since tropical woods tend to be more ex‐
posed to wood pests (e.g. termites), which explains why they have developed better
defence mechanisms through their evolution. Biological pests rely on specific environ‐
mental conditions, key among which by far is wood moisture content. Fungi cannot grow
optimally in wood with moisture content underneath the fibre saturation point (due to
the lack of free water in the cell lumen). Another key factor is whether the wood is per‐
manently moist or whether it can be dried again. Spruce members can resist fungal at‐
tack, provided the moisture content is kept under 20% at all times, with resistance declin‐
ing if the moisture content fluctuates and the wood becoming non‐durable in the event
of continual moisture. These fluctuating environmental conditions are taken into consid‐
eration by classifying into service classes, as stated in DIN 68800‐1 “Wood preservation –
Part 1: General”, Table B4‐1.
76
Durability
Table B4‐1 Classification in service classes (SC). (Table from DIN 68800‐1:2011)
77
Durability
Insect infestation is more difficult to assess. DIN 68800 points out that assuming a normal
interior climate, insect attack does not represent any risk except lyctus beetle infestation
of starchy hardwood species. The standard also cites constructive wood preservation
measures against insect attack, using sealed panelling or an effective means of inspecting
the wooden members. The second part of DIN 68800‐2 “Wood preservation – Part 2:
Preventive constructional measures in buildings” offers hints on preserving construction‐
al wood, since effective construction details can help significantly reduce the vulnerability
of wood members to biological pests, as well as the choice of wood species used. As a
general rule, wood should be installed with moisture content commensurate with subse‐
quent use and not permanently moist (in other words, covered or allowed to dry out
again and water can be effectively drained off). Preserving constructional wood against
fungal attack thus equates squarely to humidity protection. The scope of wood preserva‐
tion also encompasses positive silvicultural practices. Wood is felled in winter, when the
nutrient and water transfer processes in the tree are on hold and outside the flight and
mating seasons for insects that damage wood. This also means that the felled wood has
to be transported away and processed prior to spring (→ when the insects mate and lay
eggs). The wet storage or deliberate watering of stems is performed when stems cannot
be processed immediately and have to be placed in storage and offers two advantages.
Firstly, it imbues the wood with a moisture content which is excessive for wood‐
destroying fungi or insects. Secondly, it helps prevent shrinkage due to drying out and
hence shrinkage cracking, which reduces the access for insects due to fewer microscopic
cracks for laying eggs or fungal spores.
B4.2 Wood preservatives
Chemical wood preservatives are applied to wood using various impregnation methods.
Paints and coatings are considered physical forms of wood preservation and also repre‐
sent a means of weathering protection, since they can help mitigate precipitation, high
humidity, UV rays and mechanical stresses. Chemical wood preservation should always
be a last resort, since it inevitably means using biocides and should be avoided where
possible by selecting the right wood species as well as constructive wood preservation
measures. Chemical wood preservation may, however, make sense as remedial action
following an infestation. The prevailing wood species in our area, spruce, is very difficult
to treat, even when using pressure‐impregnation processes, since the sapwood can only
be penetrated a few millimetres. Heartwood, meanwhile, tends to be untreatable, since
its cell walls are saturated with core constituents and the pits close off as part of the
drying process. The unpopularity of biocide use notwithstanding, effective chemical
wood preservation, namely non‐stop treatment of the entire volume of wood, is infeasi‐
ble for conventional structural elements.
78
Durability
B4.3 Wood modification
Blocking Grafting
cellulose chains
cellulose chains
cellulose chains
cellulose chains
-OH MMF HO- -O -R-CH3 HO-
cellulose chains
-OH HO-
cellulose chains
cellulose chains
cellulose chains
-O -CH2- O-
Heat
HCHO
DMDHEU
Figure B4‐2 Starting points for wood modification. (www.holzfragen.de)
79
Durability
Both the most recently cited methods represent a chemical modification of the timber,
since a chemical reaction is triggered when the timber is impregnated with specific mate‐
rials, which involves the free hydroxyl groups being occupied. For the Belmadur® product,
the result is cross‐linking with the methylol already in use during the textile cross‐linking
(non‐iron laundry). Kebony®, conversely, uses furfur yl alcohol, an agricultural industry
waste product, to cross‐link the hydroxyl groups and "artificially soak" the wood in a
certain way.
In contrast, acetylation with acetic anhydride, which is the process used in Accoya®,
involves occupying (grafting) the hydroxyl group, so that no further water can be stored.
The waste product of acetylisation is acetic acid, which must then be recycled.
The final modification option is thermally treated wood, which incidentally has already
become widely established. The wood is heated to a temperature of between 150 and
240°C, which causes the –OH groups to “decompose”. Thermally treated beech may,
depending on the intensity of the thermal treatment applied, attain durability class 1, as
opposed to thermally treated pine, which can “only” achieve durability class 2 to 3 (see
also pre‐standard DIN CEN/TS 15679), while thermally untreated pine wood is classified
in durability class 3‐4 in accordance with EN 350‐2.
Up to Accoya, all modified wood products are only used in outdoor and garden areas.
Wood having undergone special thermal treatment is not usable for load‐bearing pur‐
poses, since the mechanical properties decline by up to 30%, while impact energy may
drop by up to 60%, rendering the material very brittle. Accoya alone has been used in
load‐bearing applications to date; the most well‐known examples of which include the
bridges in Sneek (NL) (see Figure B4‐3).
Figure B4‐3 Accoya bridge in Sneek, the Netherlands. (Blass and Eberhart, Karlsruhe)
80
Durability
All the chemical modification measures require impregnation, most often pressurised,
followed by drying and a reaction period. However, not all wood varieties lend them‐
selves to impregnation (→ pits). Spruce, for example, is very difficult to impregnate,
which rules out the possibility of chemical modification. Timber species open to the use
of acetylation include pine, beech, maple and particularly Radiata pine, a fast‐growing
species with large annual ring widths. Not all the workings behind such modification
methods are known or understood to date, but to put matters in a clearer chronological
context, it is worth noting that the first acetylation took place in 1928 and the first furfu‐
rylisation in the 1950s. Even so, several more decades elapsed before the modified wood
was first rendered commercially usable.
B4.4 Preventive constructional measures
Unlike metallic parts, which may corrode, wood is exposed to biotic damaging influences.
In terms of biotic damage, the two main causes identified are fungal and insect attack,
which means rules must be established for using wood members under various climatic
conditions without risking the load‐bearing capacity and durability of members made of
wood or wood‐based products. Practical examples should indicate which details are to be
avoided and which types of construction are advisable.
Fungal attack
Fungal attack occurs in wood with moisture content above 20%. Fungi require free water
and oxygen, to develop, although the optimal moisture content depends on the type of
fungus involved. Fungal attack can significantly reduce the load‐bearing capacity of
wooden constructions, although the strength loss varies and is dictated by the type of
fungus and the extent of the infestation. Wood may already be significantly weakened,
even if it looks virtually pristine, but merely ensuring a building is designed properly
should prevent wooden members from being exposed to excessive moisture and any
associated damage caused to the wood by fungi. Designs should take the following re‐
quirements into account:
Ensure humidification of wood is avoided as far as possible,
Ensure swift water drainage and ventilation, if occasional humidification
cannot be prevented,
Ensure woods with sufficient natural durability are used, if permanent
humidification cannot be prevented.
81
Durability
Insect attack
Heat tends to encourage insect attacks, since it promotes the development and repro‐
duction of insects. Those capable of completely destroying wood quicker than any other
are termites, which proliferate in warmer areas of the world, although their existence
and chances of survival are far slimmer in Central and Northern Europe (as things stand).
Central heating systems in buildings tend to establish conditions which favour the surviv‐
al and proliferation of insects, since they ensure optimal temperatures are reached, even
during colder seasons, while at lower temperatures, the insect larvae may perish. Cracks
and gaps which form in treated wood and penetrate the external treated wood layer, can
signal the basis for egg‐laying or the start of infestation, given the fact that they signifi‐
cantly impair the effectiveness of any wood preservative treatment. The natural durabil‐
ity of the individual wood species against insect attack varies considerably. Among most
woods, heartwood tends to be durable, although resistance among individual wood
species to termite infestations varies. The wood and insect species involved determine
whether or not wood from the sapwood range is durable.
Corrosion of metallic parts
Under normal conditions of use, wood is not attacked by acids and bases. However,
metallic parts must be protected against corrosion, where the conditions of use could
impact on their long‐term behaviour. To avoid surface discolouration of visible wood,
there may be a need to paint or coat metallic parts.
Preventive constructional measures
Many factors have to be taken into account when considering the durability of wood. As
far as possible, the wood should be installed at the equilibrium moisture content present
in the building, so that only additional seasonal variations in humidity need be taken into
account. If care is not taken here, however, the wood may be prone to cracking after
being installed. This risk of cracking peaks in wood members directly exposed to weather‐
ing or where there are significant seasonal variations in humidity. Cracks allow the ingress
of water and fungal spores and egg‐laying by insects within the internal cross‐section
beyond the protected outer area. The structural engineers must take into consideration
the risk of humidity changes due to:
Direct water influx,
The hygroscopic behaviour of the wood in response to the moisture and
temperature of the surrounding air.
82
Durability
In a liquid state, water is particularly liable to infiltrate the wood in the grain direction.
Accordingly, end grain surfaces should be arranged or covered such as to ensure that no
water can penetrate via the capillary effect. Certain cases which may cause moisture
content to increase are now listed below:
Warm and moist air e.g. in poorly ventilated attics, in which ventilation shafts are
installed, is conducive to moisture absorption of the wood.
Condensation problems can be prevented by additional heat insulation and a
vapour barrier. Condensation at the base of glass walls must also be taken into
consideration.
Wood is exposed to moisture when in contact with the ground, in doors and win‐
dow frames and in areas, in which snow accumulates due to snow drift. Particular
attention should be paid to areas liable to exposure to splashing water, i.e. show‐
ers, bathrooms or kitchens.
Water may accumulate in the soil, in walls behind watertight barriers and in con‐
nections, because the natural moisture release is prevented by evaporation in
these areas. Any features which may allow water to accumulate in the vicinity of
metallic plates must also be avoided.
There is an increased risk of humidification while wooden members are stored at
the building site and during assembly, while no roof covering is in place. Stacks of
wood should be covered and when there is no other alternative for assembly rea‐
sons, stored for a maximum of one week. Minimising the storage period is particu‐
larly important in the event of adverse weather conditions.
One of the most fundamental preventive constructional measures is to cover exposed
members, particularly those with end grain. DIN 68800‐2 explains precautionary con‐
struction measures used in building construction.
Precautionary fungal protection
The risk of any increase in moisture content can be reduced by careful planning of struc‐
tural details. Conversely however, high temperatures based on the geographical location
cannot be influenced and the risk of higher levels of moisture content rises with increas‐
ing average temperatures. If it is impossible to prevent the ingress of water into the
wood, swift drying must be ensured or the moistened area must be segregated, so that
the moisture content does not exceed the limit of 20%. The moisture content results
from the equilibrium between water absorption and water release and it can be lowered,
if the water absorption is delayed and efforts made to promote water release. One good
83
Durability
example is the use of metallic column footing, which allow a minimum distance between
the end grain of the column and the ground surface of 300 mm to be established. Even if
rain falls, the wood is not exposed to additional moisture from splashing water and can
dry out again afterwards.
Precautionary insect protection
The first step should be to determine the natural durability of the selected wood species
in terms of resistance to individual and potentially invasive insect species. There is also a
need to clarify whether the individual insects are likely to be present in the area in which
the wood is installed. As well as wood preservation using wood with natural durability,
structural barriers shielding the wood from the ground may also constitute effective
protection, in the event of any risk of infestation by termites which exist at ground level.
Termites often form tunnel‐like routes with canopies between the nest chambers and
wood sections attacked. Structural barriers or covers make it easier to detect the move‐
ment of such termites. There is a need for regular maintenance inspections to discover
and eliminate the relevant locations and care must be taken to ensure that the surfaces
between the ground and wood sections are both accessible and visible.
Corrosion protection for fasteners
EC 5 includes examples of minimum requirements or minimum corrosion protection
needed for various service classes. Enhanced corrosion protection measures are re‐
quired, e.g. in storage buildings for chemical products, salt and fertilisers, or also in phos‐
phoric acid factories, in which the use of bolts, dowels and steel plates made of special
stainless steels is imperative.
B4.5 Literature
G. Sagot, L.M.R. Nunes, P.P. de Sousa, original Articles A14, A15, STEP 1995.
Rapp A.O. and Augusta U. (2000). Dauerhaftigkeit in den Gefährdungsklassen unter besonderer
Berücksichtigung von Lärchen‐ und Douglasien Kernholz. 22. Holzschutz‐Tagung der Deutschen Gesellschaft
für Holzforschung, Bad Kissingen.
84
B5 Wood drying and
strength grading
Original article: P. Glos
B5.1 Principles of wood drying
Wood is hygroscopic when its moisture content is underneath the point of fibre satura‐
tion, namely its moisture content varies depending on the prevailing climate. In the pro‐
cess, the cross‐sectional dimensions also vary by swelling or shrinking. In addition, practi‐
cally all the technological properties such as strength and stiffness or heat insulation in
the hygroscopic area (below the fibre saturation point, see Article B2) depend on the
moisture content. Moreover, when its moisture content exceeds around 20%, wood
becomes particularly vulnerable to pests. Accordingly, avoiding subsequent damage
involves drying the wood before end processing of the same to ensure that it is already
adapted to the future prevailing climate as far as possible. In central Europe, the equilib‐
rium moisture content for wood placed outside is 12 to 14% in summer and at least 18%
in winter. An overview of wood moisture guidance values for installation is shown in
Table B5‐1.
Table B5‐1 Wood moisture guidance values.
Purpose of use Moisture in %
Structural timber 8‐20
External windows and doors 13‐16
Furniture, room doors, interior construction with permanent heating 6‐9
Plywood, LVL, musical instruments 5‐7
As a general rule, wood has to be dried; not only when producing structural and glued
laminated timber, but in all areas of woodworking and processing, for producing wood‐
based panels for example. During the construction period, it is important to strictly limit
increase in wood moisture during either transport or storage and deployment of the
dried wood, which can be done by storing the wood on covered underlayers and swift
covering or, where applicable, provisional covering of members. Absorption of moisture
during storage occurs particularly quickly and intensively via end grain surfaces, recesses,
boreholes and so on.
85
Wood drying and strength grading
Physical and technical principles of wood drying
The relative humidity
Relative humidity denotes the amount of water vapour absorbed by the air and is calcu‐
lated from the percentage ratio of the actual current quantity of water vapour in the air
relative to the maximum absorbable quantity, assuming unchanged air temperature and
pressure. If the air contains the maximum possible quantity of water vapour, this is re‐
ferred to as saturated steam or saturated humid air; whereupon the vapour starts con‐
densing on the surfaces of the space. Underneath the maximum quantity, the vapour is
described as unsaturated or overheated. As Figure B5‐1 shows, the relative humidity
clearly depends on air temperature. Warm air can absorb more humidity than cold air
and is relatively drier amid unchanged absolute water content. Increasing the tempera‐
ture in a specific room triggers a simultaneous decline in relative humidity. This link is
significant for wood drying, since falling relative humidity causes the equilibrium mois‐
ture content of wood to plummet, while the direct impact of air temperature on mois‐
ture content is far lower.
600
A = Maximum quantity of water in the air at
saturated state (relative humidity = 100%)
B = Quantity of water in the air at a relative
500 humidity of 40%
Absolute humidity [g/m3]
400
300
200
100
A
B
0
0 20 40 60 80 100
Air temperature [°C]
Figure B5‐1 Absolute humidity depending on relative humidity and air temperature.
86
Wood drying and strength grading
The equilibrium moisture content of wood
The resulting relative moisture content thus depends on the relative humidity, air tem‐
perature and pressure and wood species. Given storage for sufficient time under specific
climatic conditions, wood will reach the related equilibrium moisture content. The key
factor dictating the so‐called equilibrium moisture content of wood ugl is the relative
humidity of the surrounding air, whereas the direct impact of temperature is less im‐
portant. Figure B5‐2 shows the Keylwerth diagram, which can be used to determine the
connection between relative humidity, air temperature and the equilibrium moisture
content of wood (here: spruce).
130
°C
Dry temperature
120
110
100
90
80
70
60
50
40
30
20
10
Transport of moisture in the wood while drying
The transport of moisture in the wood is based on a complicated system of capillary
water movements and diffusion. At the beginning of the drying process, since levels of
moisture content exceeding the fibre saturation point and capillary transport processes
are accelerated for the drying process, free water is discharged more quickly than bound
water. The process of drying from u = 50% to u = 40% moisture content proceeds, with
the same cross‐sectional dimensions, namely far faster than u = 20% to u = 10%. The
speed of the moisture release as regards this capillary water movement is hardly direc‐
tion‐dependent.
87
Wood drying and strength grading
The moisture exchange between wood and the surrounding proceeds at different rates
for the individual wood species, while the drying speed clearly declines in the following
order: pine sapwood, spruce, beech, sipo, oak, iroko and teak. Since natural water
transport takes place in sapwood, it is easier to dry than heartwood, while wood with a
greater density generally dries more slowly than lightweight wood. Morphological fea‐
tures, such as storage of constituents during heartwood formation or the closure of
bordered pits in conifers may considerably hinder water transport and hence the drying
process. In such cases, the internal wood layers may still be very humid, while the exter‐
nal layers have already dried out, whereas for freshly sawn timber, the water contained
in the wood is still distributed virtually uniformly throughout the cross‐sectional portion.
With the release of moisture via the surfaces, the external layers are first to dry, hence
reducing the disparity in moisture content between the interior and surface, Figure B5‐3.
This disparity is the decisive factor controlling the movement of moisture within the
wood and thus the most important element for an optimal drying process. Excessive
wood moisture disparity, e.g. at the start of the drying process, hinders the capillary
water movement and means most of the free water from the interior has to travel via the
wood in the form of vapour, which is far slower than for liquid water. If the wood mois‐
ture disparity is very considerable, drying defects may occur (e.g. casehardening of
wood). Conversely, if the disparity is insufficient, the drying times required will be exces‐
sive and uneconomical. The term “casehardening” refers to the residual stress state of
the wood after the drying process, due to permanent localised tensile or compressive
deformations perpendicular to the grain. The external areas dry too quickly and shrink,
which exerts compressive stresses perpendicular to the grain accumulating on the inside
and tensile stresses perpendicular to the grain on the surface; ultimately resulting in
permanent tensile deformations perpendicular to the grain. If the drying process is con‐
tinued, the internal sections also dry out and shrink and the inability of the external areas
to accommodate such deformations leads to internal cracks forming.
Drying methods used include natural outdoor drying, while technical drying methods
(kiln‐drying) in drying chambers are particularly popular.
1
70
Moisture content [%]
60 2
50
40
30 3
20 4
10 5
0
0 1 2 3 4 5
Thickness 5 cm
Figure B5‐3 Change in wood moisture in a 50 mm thick plank during storage at the lumberyard,
Line 1: Green, Line 5: Wood moisture arising.
88
Wood drying and strength grading
B5.2 Principles of strength grading
As a natural raw material, wood is produced from a wide range of trees depending on
their type, genetic material, growth and environmental conditions. Wood characteristics
are also very diverse, from tree to tree, but also within the same stem, stem cross‐
section and longitudinal stem direction (see Articles B1 and B3). Cutting up logs by sawing
destroys the wood structure having developed, for example severing wood fibres in the
area of branches or in spiral‐grained stems. This also means greater variation in the
strength properties of sawn timber, particularly for smaller cross‐sectional dimensions,
which may vary much more than those of logs subject to no or minimal processing. In
fact, the strength properties of unsorted sawn timber within a single wood species may
vary so considerably that the strength of the strongest piece may be ten times as much
as the weakest, as in Figure B5‐4 above. Since the most important factor influencing the
possible end use of the wood is always its characteristic strength, namely the 5% quantile
of the basic population, using ungraded wood makes it impossible to exploit the high
strength of the majority of sawn timber. For economic reasons, however, the wood
needs to be classified into different strength classes by ensuring appropriate grading,
Figure B5‐4 below. However, where the strength of the individual piece is unknown and
can only be determined indirectly via visually recognisable approaches or non‐destruc‐
tively assessing the wood properties correlating to strength, the variation in strength
within a specific classification cannot be completely narrowed down. For this reason, the
classes overlap depending on the quality of grading (see Figure B5‐4 below) and the less
effective the grading method used, the greater the overlap concerned. This shows how
crucial it is to ensure effective strength grading to exploit the economic potential of the
wood. In addition, grading is also a prerequisite to ensure the availability of sawn timber
in the quality and quantity demanded by consumers and its compliance with all user
requirements, particularly meeting the criterion of being a predictable construction ma‐
terial with reliable properties.
89
Wood drying and strength grading
ft,0,k ft,0
(a)
h
(b)
(c)
Traditionally, visual assessment has been used as a means of grading wood, which in‐
volves determining its quality by assessing visible characteristics, particularly knots and
annual ring width. Up to the start of this century, visual grading was primarily conducted
based on traditions having been handed down and local experiences. The first detailed
standardisation of strength grading took place in 1923 in the USA and subsequently, from
the 30s onwards, the process established itself throughout Europe. Given the various
species and qualities of wood around Europe and the range of timber construction tradi‐
tions, for example using sturdy or slender cross‐sections, it is no surprise that the last
half‐century has seen variation in grading regulations for grading criteria; particularly in
terms of how knots are classified and classes divided. The common problem, however,
for these grading regulations is the visual grading process. Naturally, practical considera‐
tions dictate that only visually recognisable wood characteristics can be taken into con‐
sideration and simple combination rules defined as a result. Key factors determining
strength, such as wood density, cannot be assessed accurately via visual means alone, for
example via the annual ring width. The visual strength grading approach thus suffers
from inherent and unavoidable ambiguity, which limits the effectiveness of any grading.
Moreover, since the classification also depends on the alertness of the grader, it is not
completely objective.
90
Wood drying and strength grading
As part of efforts towards more reliable strength grading of sawn timber, enhanced usage
efficiency for the present timber stock and higher strength classes, machine grading
methods were developed in Australia, the USA and the UK from around 1960 onwards,
with other countries following suit later on. The increasing industrialisation in woodwork‐
ing and processing, the growing importance of quality assurance and the increasing de‐
mand for premium quality wood have seen interest in such machine grading rocket in
recent years, which has catalysed the development of new and more efficient grading
machines. During wood grading, there is generally a need to distinguish between so‐
called appearance grading and strength grading. The former process of appearance grad‐
ing involves assessing the wood in terms of optical appearance, namely aesthetic criteria.
This is particularly important whenever wood is deployed as a visible fixture, whether as
wall or ceiling panelling; for structural timber used in permanently visible construction or
in furniture‐making. Conversely, during strength grading, the wood is assessed solely in
accordance with relevant criteria dictating its load‐bearing capacity. It follows that under
certain circumstances, timber used in high‐level and permanently visible members must
be graded in accordance with both criteria and not just strength alone. However, the
remainder of this article only covers the strength grading of sawn timber.
General strength grading requirements
The strength grading process should ensure confirmed compliance with all key properties
of relevance when using timber for load‐bearing purposes. To do so, the grading rules for
each grade must include limit values for wood characteristics, which correlate sufficiently
to the actual strength and stiffness of the wood. For conventional visual grading, these
particularly include the annual ring width as a measure of the wood structure relevant for
strength and strength‐reducing wood characteristics such as knots, slope of grain, cracks,
reaction wood, fungi and insect attack and mechanical damage. During machine grading,
other properties, which are not visually measurable, such as the dynamic modulus of
elasticity (MOE), can be used and allow far more accurate estimation of the strength and
stiffness properties. In addition to relevant strength and stiffness grading criteria, limit
values for geometric properties, such as the wood wane, curvature and warp must also
be determined, given their relevance when the wood is used for construction. The fact
that the wood dimensions, as well as its curvature, warp and cracks, depend on the mois‐
ture content means the limit values have to be based on a reference moisture content,
which is fixed at 20% in Germany (see DIN 4074). In addition, the moisture content also
has to be taken into consideration during machine grading, if moisture‐related wood
parameters are measured.
91
Wood drying and strength grading
Visual strength grading
During visual grading, the appearance of the wood is considered and the characteristics
that determine its strength are assessed. Accordingly, the result depends on the grader
and the key visual grading parameter is knots. The fact that location and climate influ‐
ence the tree growth means the sample used to determine the correlation of the meas‐
ured grading parameters with the strength must be representative. The sample results
must also, without fail, be applicable to the basic population, since this is the only way to
guarantee that the actual non‐destructive strength grading leads to a wood piece being
reliably classified in a specific strength class using the correlated grading parameter. The
first step involves establishing a representative sample, which means the following influ‐
ences are to be taken into consideration:
Climate,
Growth area,
Moisture content,
Geometry of specimens,
Wood species.
The next step is to perform tests to correlate the grading parameter with the targeted
strength property, e.g. bending strength. A typical result of bending strength, depending
on the grading parameter knots, is shown for softwood in Figure B5‐5. Large knots mean
a drastic drop in bending strength, which means the wood has to be assigned to a lower
strength class. However, knots are only weakly correlated to strength and the end result
is a wide and diversified point cloud.
Subsequently, to use knots as a grading parameter, limit values are needed for each
strength class. Figure B5‐6 shows a schematic view of the process. The characteristic
bending strength of a strength class is set to e.g. 24 or 30 N/mm², while the horizontal
lines of the selected bending strengths intersect with the vertical lines of knot area ratios,
so that only 5% of all pieces of wood with equivalent or less knot area ratio have a small‐
er bending strength (black areas in Figure B5‐6). Accordingly, Figure B5‐6 shows knot
area ratios of 0.4 and 0.2, respectively, as decisive grading criteria for strength classes,
with characteristic bending strengths of 24 or 30 N/mm². EN 1912 links a visual grade,
which, in this case, equates to grade S10 “maximum knot area ratio = 0.4” in accordance
with DIN 4074 Part 1, to strength class C24 in accordance with EN 338, where fm,k =
24 N/mm². In EN 1912, national visual grades are assigned to a European strength class.
92
Wood drying and strength grading
80
Bending strength [MPa]
60
40
20
R² = 0.20
0
0.00 0.25 0.50 0.75 1.00
Knot area ratio
Figure B5‐5 Correlation between bending strength and knot area ratio (softwood).
80
70
Bending strength [N/mm2]
60
50
40
30
20
5%‐quantile
10 5%‐quantile
0
0 0.1 0.2 0.3 0.4 0.5 0.6
Knot area ratio
Figure B5‐6 Determining grading parameter “knot area ratio” per strength class (softwood).
Visual grading parameter and grading criteria
The key grading parameter “knot area ratio” is shown as an example in Figure B5‐7 for
squared timber (softwood). Knot parameter A is always the product of geometrical con‐
siderations concerning the ratio of knot area to the remaining area of wood. Evaluating
boards and planks, however, is somewhat more complex, since the knots are present as
individually interspersed on the narrow or wide side, through knots or knot clusters (see
also Figure B3‐14). Various limits (= grading criteria) are applied to the grading parameter
“knot area ratio” depending on the grade. In addition to the knot area ratio, other grad‐
ing parameters must also be added, such as slope of grain, wane, annual ring width,
curvatures, cracks, reaction wood or discolourations. For example, Table B5‐2 lists the
grading criteria for squared timber in accordance with DIN 4074.
93
Wood drying and strength grading
d3
d1
d2
h d4
d d d d
A max 1 ; 2 ; 3 ; 4
b h b h
b
Figure B5‐7 Knots in squared timber, Knot parameter A in accordance with DIN 4074‐1.
Machine strength grading
When machine strength grading is used, the key grading parameter is not knots, but
modulus of elasticity. The measurement accuracy exceeds that of visual grading and the
grading result is more easily reproducible. The modulus of elasticity correlates better
with strength than knots; Figure B5‐8 shows a clearly higher coefficient of determination
(R2) than Figure B5‐5.
80
Bending strength [MPa]
60
40
20
R2 = 0.51
0
0 5000 10000 15000 20000
Modulus of elasticity (MOE) [MPa]
Figure B5‐8 Correlation between bending strength and MOE (softwood).
94
Wood drying and strength grading
Table B5‐2 Grading criteria for square timber (softwood) during visual grading, DIN 4074‐1:2008.
Grading parameter Grade
S7 S10 S13
1. Knots up to 3/5 up to 2/5a up to 1/5
2. Slope of grain up to 12% up to 12% up to 7%
3. Pith permissible permissible not permissibleb
4. Annual ring width
‐ In general up to 6 mm up to 6 mm up to 4 mm
‐ For Douglas fir up to 8 mm up to 8 mm up to 6 mm
5. Cracks/shakes
‐ Shrinkage cracksc up to 1/2 up to 1/2 up to 2/5
‐ Lightning, ring shake not permissible not permissible not permissible
6. Wood wane up to 1/4 up to 1/4 up to 1/5
7. Curvaturec
‐ Longitudinal curvature up to 8 mm up to 8 mm up to 8 mm
‐ Warp 1 mm / 25 mm height 1 mm / 25 mm height 1 mm / 25 mm height
8. Discolouration, rot
‐ Blue stain permissible permissible permissible
‐ Brown and red stripes up to 2/5 up to 2/5 up to 1/5
‐ Brown, white rot not permissible not permissible not permissible
9. Compression wood up to 2/5 up to 2/5 up to 1/5
10. Insect damage burrows up to 2 mm in diameter by greenwood insects permissible
11. Other features are to be taken into consideration correspondingly, in line with the
other grading criteria
a For spruce and Douglas fir, up to ½ for annual ring widths of up to 4 mm for spruce and
5 mm for Douglas fir. The proportion within a delivery must not exceed 25%.
b Permissible for squared timber with width > 120 mm.
c These grading parameters are not taken into consideration for non‐dry woods.
Machine grading parameters
As well as the modulus of elasticity, additional parameters may also be mechanically
recorded and evaluated, including knots, density, moisture content and the cross‐
sectional dimensions. What all parameters have in common is the fact that they can be
measured via non‐destructive methods. The modulus of elasticity is measured via static
bending or dynamic measurement, while optical methods, such as surface scanning or
even X‐rays or gamma rays are used to measure knots. Wood moisture measurement
95
Wood drying and strength grading
devices reveal the moisture content, while modern strength grading machines measure
the dynamic modulus of elasticity via vibration. Provided the vibration frequency f, length
ℓ of the board and its density are known, the dynamic modulus of elasticity Edyn can be
calculated, see equation (B5‐1), featuring measurement of longitudinal oscillation as
shown in Figure B5‐9 (derivation in Annex 1). This measurement method is shown sche‐
matically in Figure B5‐9, while Figure B5‐10 shows the result of radiography with X‐rays
to measure knots.
E dyn 4 f
2 2
(B5‐1)
Figure B5‐9 Oscillation measurement.
Figure B5‐10 X‐ray image (bottom) of a board with knot (top).
The best results are obtained by combining several grading parameters. Such a combina‐
tion elicits improved correlation with strength and hence a better result when classifying
into a strength class. The grading parameters for “dynamic modulus of elasticity” and
“knots” are combined in virtually all modern grading machines, which may also involve
using density obtained through weighing (or via X‐rays) and the cross‐sectional dimen‐
sions. Figure B5‐11 shows a typical correlation between bending strength and a combina‐
tion of grading parameters. The coefficient of determination is even higher compared to
Figure B5‐8, which means a better correlation between bending strength and grading
parameters.
96
Wood drying and strength grading
80
Bending strength [MPa]
60
40
20
R2 = 0.62
0
0 20 40 60 80
Multiple grading parameter
Figure B5‐11 Correlation between bending strength and a combination of several grading parameters.
Four‐point bending tests
The correlation between grading parameters and strengths is normally determined by so‐
called four‐point bending tests (Figure B5‐12). The strength derived from grading criteria
is bending strength, whereupon all other strength properties of a strength class tend to
be determined using regression equations derived from the bending strength (EN 338).
The local modulus of elasticity between both force transmission points (see wlocal in Fig‐
ure B5‐12) does not include any shear strain portion, given the absence of shear stresses
in the relevant area. This means the local modulus of elasticity is determined based on
the measured force and deformation portions excluding contribution of shear.
F F
wglobal F F
F F
wlocal
Figure B5‐12 Four‐point bending tests with deflection curve, wglobal and wlocal.
97
Wood drying and strength grading
Calibration of grading machines
The correlation between grading criteria and bending strength must be set and checked.
The “settings” (adjustment values) of a grading machine must also be calibrated, so that
the machine is capable of assigning the tested sawn timbers to a strength class. This
calibration can be implemented in two different ways, namely “machine‐controlled” and
“output‐controlled“.
Machine controlled:
In this case, numerous pieces of sawn timber are tested using four‐point bending tests.
Grading models are then derived from the results and used to determine the settings for
the grading machines. During the production, no further sawn timber is tested.
Output controlled:
The initial settings of the grading machines are derived following a small number of tests,
based on which sawn timber is regularly taken from graded wood and tested and
checked to ensure it meets the selected criteria (characteristic values). If not, the settings
are adapted accordingly.
Log grading
Visual log grading involves checking logs with the naked eye, as is done for sawn timber.
Centring on the visibility of the characteristics for logs, the assessment encompasses
similar characteristics as when grading sawn timber, all of which impact on strength.
However, unlike sawn timber, logs are classified into quality classes.
B5.3 Literature
P. Glos, original Article A6, STEP 1995.
Kollmann F.F.P and Coté W.A. (1968). Principles of wood science and technology. Volume I, Solid wood.
Springer Verlag, Berlin, 592 p.
98
B6 Wood products
Original articles: P. Glos, F. Colling, A. Ranta‐Maunus, G. Steck, D. R. Griffiths, E. Raknes
At one time, the size of the tree in the forest used to dictate the dimensions of the struc‐
tural timber which it was used to make. A century ago, square timbers, with a cross‐
sectional area of 150 mm x 450 mm and up to 20 m long were generally available. Nowa‐
days, although timber with a cross‐section exceeding 75 mm x 225 mm and more than
5 m long is increasingly rare and thus costlier than ever, when such larger dimensions are
required, multiple pieces of wood can be combined to form a single member, e.g. of
glued laminated timber. Wood is inevitably non‐homogenous as a result of the natural
growth process of trees. Meanwhile, knots, pitch pockets and other growth‐related
properties have a key influence on strength, which explains the considerable variance in
strength properties within a member. If large pieces of wood are split up into smaller
parts, then re‐joined, any flaws within the materials are dispersed and the variance in
material properties also declines. The larger load‐bearing capacity of glued laminated
timber compared to solid timber is not due to a higher average load‐bearing capacity, but
a reduced variance in strength properties, which elicits a higher characteristic strength. In
general, the variance in strength of the wood products listed in Table B6‐1 declines with
increasing manufacturing cost and the homogenisation also increases. However, logs,
which are very rarely processed, represent an exception. They are particularly strong,
since the wood fibres remain uncut and the continuous fibres can convey the stresses
around the knots.
Table B6‐1 Wood construction products and their components.
Wood product Components
Logs Stems
Sawn timber Squared timber, planks, boards and battens
Glued laminated timber Boards
Variance declines
Laminated veneer lumber Veneers
Plywood Veneers or sawn timber
Parallel strand lumber Veneer strands
Particleboards Particles (chips)
Fibreboards Fibres
99
Wood products
As a natural material, wood displays different properties in various directions. Parallel to
the grain, namely in the longitudinal direction of the tree stem, wood is particularly
strong, but conversely far lower at right angles to the grain, for example the tensile
strength parallel to the grain is around 40 times higher than perpendicular to the grain.
This is reflected in the ease of chopping wood along the fibres with an axe, as opposed to
the far more difficult task of splitting a piece of wood at right angles to the grain. These
significant differences in strength and stiffness properties in various directions no longer
apply for most wood‐based panels. Since the wood particles and fibres are randomly
arranged in many types of panels, in panel planes, the loading direction is far less rele‐
vant to determine strength in the panel plane than in solid timber.
B6.1 Solid timber
The term solid timber is used for members completely made of wood in its natural struc‐
tural form. This is opposed to so‐called engineered wood products or wood‐based panels,
which are manufactured by disassembling and then reassembling parts.
Sawn timber
Sawn timber is produced by sawing logs longitudinally and refers collectively to products
produced from logs in sawmills. The main (product) groups, according to their cross‐
sectional dimensions, are boards, planks, battens and square timbers. The cross‐sectional
areas are standardised (e.g. in DIN 4074‐1), as preferred cross‐sections e.g. determined
for use as solid structural timber (KVH®) or separately determined depending on the
purpose of use. The various process technologies (frame saw, band saw, chipper‐canter)
allow various yields of sawn timber and help achieve the required quantities of industrial
residual wood. Figure B6‐1 shows cutting options for square timbers, where the free‐of‐
heart conversion is particularly relevant, as it helps ensure fewer shrinkage cracks.
Figure B6‐1 Cutting types. Left: Single‐stem cut; Middle: Dual‐stem split‐heart cut;
right: Dual‐stem free‐of‐heart cut.
100
Wood products
After the cut, the sawn timber is visually pre‐graded, before being kiln‐dried or dried
naturally outdoors. Smaller sawmills in particular often lack any kiln‐drying rooms, which
imbues the sawn timber with excessive moisture. Once the wood has been dried to the
desired moisture content, the sawn timber is visually or machine graded and assigned to
a strength class, while wet wood is cut based on the subsequent measure of the dry
wood. This involves cutting with a certain allowance for subsequent shrinkage of the
wood. An example of a piece of squared timber with pith after drying is shown in Fig‐
ure B6‐2. All additional processing steps used for sawn timber are known as finishing and
structural timber is often used rough‐sawn. Glued solid timber comprising two or three
bonded individual cross‐sections is available, while sawn timber can also be finger‐
jointed, to extend its dimensions. For this purpose, the manufacturer in Germany re‐
quires proof of suitability for gluing load‐bearing wooden members (DIN 1052‐10).
Figure B6‐2 Squared timber with pith, knots and drying cracks. (Studiengemeinschaft Holzleimbau, 1998)
Strength classes for solid timber
A system of strength classes has the advantage of aggregating the wide‐ranging combina‐
tions of wood species, grades and origins available in the European internal market into a
manageable number of groups, to simplify and improve the use of timber for load‐
bearing purposes. The system of strength classes as laid out in EN 338 is illustrated fur‐
ther in the following and comprises twelve classes for conifers (Table B6‐2) and eight for
deciduous species (Table B6‐3). The scope goes from the lowest softwood class C14 up to
the highest hardwood class D70 to cover all current strength classes in Europe. EN 338
includes details of the characteristic strength, stiffness and density parameters for each
strength class.
101
Wood products
It is possible to agree on strength classes, including a constant rate of strength and stiff‐
ness values in each case, because regardless of origin, practically all coniferous and de‐
ciduous species used for commercial applications have a similar ratio of strength and
stiffness values. Existing test data revealed that an effective approximation of all key
characteristic strength and stiffness values could be calculated from the bending
strength, the dynamic modulus of elasticity (MOE) and the density (see Article B5).
EN 1912 summarises which visual grades and wood species can be assigned to which
strength classes in EN 338, while Table B6‐4 sets out this assignment in EN 1912 for the
strength class C30. Using this table, all the various national grades can be assigned to a
strength class.
Table B6‐2 Strength classes and characteristic values in accordance with EN 338:2009; softwood.
N/mm2 C14 C16 C18 C20 C22 C24 C27 C30 C35 C40 C45 C50
fm,k 14 16 18 20 22 24 27 30 35 40 45 50
ft,0,k 8 10 11 12 13 14 16 18 21 24 27 30
ft,90,k 0.4 0.4 0.4 0.4 0.4 0.4 0.4 0.4 0.4 0.4 0.4 0.4
fc,0,k 16 17 18 19 20 21 22 23 25 26 27 29
fc,90,k 2.0 2.2 2.2 2.3 2.4 2.5 2.6 2.7 2.8 2.9 3.1 3.2
fv,k 3.0 3.2 3.4 3.6 3.8 4.0 4.0 4.0 4.0 4.0 4.0 4.0
2
kN/mm
E0,mean 7 8 9 9.5 10 11 11.5 12 13 14 15 16
E0,05 4.7 5.4 6.0 6.4 6.7 7.4 7.7 8.0 8.7 9.4 10 10.7
E90,mean 0.23 0.27 0.30 0.32 0.33 0.37 0.38 0.4 0.43 0.47 0.5 0.53
Gmean 0.44 0.50 0.56 0.59 0.63 0.69 0.72 0.75 0.81 0.88 0.94 1.00
3
kg/m
k 290 310 320 330 340 350 370 380 400 420 440 460
mean 350 370 380 390 410 420 450 460 480 500 520 550
102
Wood products
Table B6‐3 Strength classes and characteristic values in accordance with EN 338:2009; hardwood.
Table B6‐4 Strength class C30; assignment of visual grades, species and origins in accordance with
EN 1912:2013 (botanical identification is specified in EN 1912).
103
Wood products
B6.2 Glued laminated timber
Glued laminated timber (glulam) is a variety of laminated timber in which all components
(boards in this case) are arranged parallel to the grain. In contrast, components are or‐
thogonally arranged in plywood or cross laminated timber. One example of a glued lami‐
nated timber beam is shown in Figure B6‐3. For production purposes, the individual
components, so‐called lamellae, are glued over the entire contact surface with adhesives,
which results in rigid connections. The great advantage of using glued laminated timber is
the way it allows homogenisation of wood as a construction material.
The development of waterproof and mildew‐proof synthetic resin adhesives has made
this an increasingly important area in the timber construction field and also prompted a
notable upturn in the development of glued timber constructions in timber engineering.
This field allows far larger cross‐sections than with sawn timber, which is why glulam, like
other load‐bearing products such as cross‐laminated timber, is classed as an “engineered
wood product” and has revolutionised the timber construction world in recent decades.
The requirements for glued laminated timber are set out in EN 14080, which includes
details of both production requirements and strength classes. Rules are also set out for
use by each manufacturer, to establish its own strength class, which paves the way to
combine specific lamella and finger‐jointing qualities with each other.
Incidentally, it is also worth recapping that EN 14080 sets out a new generation of Euro‐
pean standards and unifies many originally disparate standards in the process. This
means EN 14080 replaces standards EN 385 to EN 387, EN 390 to EN 392 and EN 1194,
which were previously used to regulate strength classes (EN 1194) or delamination tests
of glue lines (EN 391) among others. This means that the regulations dispersed to date
over wide‐ranging of standards are now grouped together in a single standard.
Figure B6‐3 Glued laminated timber. (Studiengemeinschaft Holzleimbau e. V., 1998)
104
Wood products
Structure of glued laminated timber
The basic concept underpinning the development of wood‐based products involves re‐
ducing the variance in properties and homogenising wood as a material. The basic struc‐
ture of the tree is generally preserved when logs are used, which makes them stronger
than sawn timber, in which many fibres are cut during the processing. Given the wide‐
ranging strength values for wood, the strength of a board is determined by its weakest
point and the load‐bearing capacity can only be influenced by grading the boards or logs.
If you consider the average area of Table B6‐1 (glulam to parallel strand lumber), addi‐
tionally to grading, the wood characteristics can be boosted through homogenisation,
whereby individual solid timber components are bonded with each other (usually using
adhesive). If the engineered wood components are even smaller (particles or fibres),
grading in accordance with the strength of each individual component is no longer feasi‐
ble and this is when significant homogenisation is applied to the production process.
Within the compound structure of a glulam beam, weak points such as knots are less
significant, since they no longer influence the overall beam cross‐section. Areas with
knots have a lower modulus of elasticity than the clear wood above and below the same,
exert a load on the surrounding lamellae and are relieved. This systematic action involv‐
ing lamellae, which are glued together under pressure, is known as the lamination effect.
Dissecting and subsequently gluing the wood paves the way to homogenise the material,
since the variance in mechanical properties in the subsequent material is lower than in
individual lamellae. In addition, elements like larger knots, pitch pockets or bark in‐
growths are removed. This means that cutting out a flawed area from a piece of timber,
which must be classified in accordance with DIN 4074 in S10, means two timbers classi‐
fied in S13 are formed (Figure B6‐4). Figure B6‐4 shows a schematic view of how an area
with large knot size is removed, followed by finger‐jointing to produce a long board. As
standard, during glulam manufacturing, boards 3 to 6 m long are combined by finger‐
jointing to form an endless lamella. Cutting to length as required allows the finger‐jointed
structural timber in the glulam to be dispersed.
Figure B6‐4 Removal of low‐value sections and combining via finger‐jointing.
(Studiengemeinschaft Holzleimbau e.V., 1998)
105
Wood products
Glulam has further, clear advantages compared to squared timber. Squared timber with a
large cross‐section tends to include pith and large shrinkage cracks (see Figure B6‐5). As
well as hindering many applications, they also provide easier access for moisture or wood
pests, while the kiln‐drying of large cross‐sections can be particularly problematic when
significant lengths are involved. For glulam conversely, drying of individual boards is
problem‐free and during the drying process, any deformations arising can be planed off
in the production process (see Figure B6‐6).
Figure B6‐5 and Figure B6‐6 reveal another glulam production principle. According to
EN 14080, in glulam, the piths should generally be on the same side, but for glulam in
service class 3, the surfaces must be always on the right sides (facing the pith), which
results in a left‐on‐left side bond of both lower lamellae (away from the pith), as in Fig‐
ure B6‐5 on the right. Under adverse weather conditions, this helps prevent shrinkage
cracks on the surface. Since the left side of the wood is subject to stronger tangential
shrinkage, it is more prone to cracking. The shrinkage behaviour of lamellae or finished
glulam‐cross‐sections and their subsequent processing is set out in Figure B6‐6.
Figure B6‐5 Squared timber cross‐section with cracks compared to a glulam‐cross‐section.
(Studiengemeinschaft Holzleimbau e.V., 1998)
Figure B6‐6 From left to right: Shrinkage deformation of a board cross‐section in accordance with drying
with respect to the fresh sawn board. Cross‐section after planing the large sides with respect to
the fresh sawn state with parallel areas for bonding. Glulam cross‐section after gluing. Planed
and chamfered glulam cross‐section. Only the right sides are externally exposed.
(Studiengemeinschaft Holzleimbau e.V., 1998)
106
Wood products
The risk of cracking induced by changing humidity is greater for increasing lamellae thick‐
ness and it also becomes increasingly difficult to apply the required pressing power in the
glue line. For these reasons, EN 14080 defines limit values for the cross‐sections of lamel‐
lae. Glued laminated timber must be formed from lamellae between 6 and 35 mm thick
and for straight members not exposed to extreme and alternating climatic stress, the
maximum thickness may be increased to 45 mm. For members exposed to dramatically
fluctuating climates, however, board thicknesses below 35 mm are recommended. To
reduce the stresses generated within the boards due to changing humidity, the boards
can be equipped with a relief groove in the longitudinal direction, which also helps pre‐
vent any tendency towards warping. The end grain of the boards should be protected
regardless, since strongly fluctuating humidity may cause the glulam timber sections to
crack. Moreover, during transport, storage and assembly, it is also important to control
the humidity of the members e.g. caused by precipitation, ground moisture or drying.
Reference can be made to EN 14080 for all additional rules and requirements.
Production
Glued laminated timber can be produced from various wood species, although in Germa‐
ny, spruce is almost exclusively (up to 95%) the first choice. The use of pine, fir, larch,
Douglas fir and beech is negligible, although glulam produced from larch and Douglas fir
is suitable when increased durability requirements apply. Glulam is produced via the
following steps:
Pre‐grading of the boards,
Drying of the boards,
Strength‐grading of the boards,
Finger jointing to generate ”continuous lamellae“,
Capping of the continuous lamellae to the required length,
Planing the lamellae,
Applying adhesive,
Inserting into a pressing jig and pressing,
Planing to the final member size.
Experienced professionals using special equipment are required to produce glued load‐
bearing wood members. Requirements include covered and air‐conditioned working
spaces, an annex for kiln‐drying, reliable measurement devices to determine wood mois‐
ture, machines for processing adhesive surfaces (e.g. planing machines; the wood surfac‐
es to be bonded must have been freshly planed) and the board ends for which butt joints
107
Wood products
are envisaged (e.g. finger‐jointing systems), glue application devices and devices to apply
the required pressing power (presses). Plus, in Germany, the “Proof of suitability for
gluing load‐bearing wood members (“glue approval”)” (DIN 1052‐10) must have been
issued.
The sawn timber used to produce glulam is stored and normally kiln‐dried. To prevent
subsequent shrinkage cracks, the moisture content for gluing should be established at a
level commensurate with the subsequent average expected moisture content in con‐
struction. However, the moisture content should be set slightly below the subsequent
average moisture content, since any compressive stresses perpendicular to the grain with
swelling that follows are less damaging for glulam than tensile stresses perpendicular to
the grain while shrinking. When used indoors, prior to processing, the boards are dried to
a moisture content of 10 ± 2%. This is in the region of the equilibrium moisture content
when installed, normally prevents any further damaging shrinking and helps minimise
cracking. Unavoidable drying stresses which occur during conventional drying of sawn
timber must be minimised, to reduce problems for further processing (e.g. the glue line
warping and splitting open). The next step is to grade the boards visually or via machine
grading, normally the latter, and the wood is pre‐planed before grading. The quality
control also involves weeding out any excessively dry or moist boards and sending them
to be reconditioned. Glulam boards vary considerably in terms of their mechanical and
optical properties. Classification into (visual and mechanical) grades and reducing un‐
sightly natural wood characteristics are also vital when it comes to producing premium
glulam beams. Following the finger‐jointing and gluing (during which compliance with
specific environmental conditions, temperature and relative humidity must be ensured),
the final stage of production involves planing and chamfering (edge breaking), where‐
upon the desired cross‐section is attained for the glulam member. This is followed by
cosmetic repairs, cutting to length, packaging, delivery and erection.
Monitoring
Particular attention is paid to monitoring quality, which is established by factory produc‐
tion control (in‐house monitoring) and external supervision to verify the self‐monitoring
process (external monitoring). The in‐house monitoring mainly includes checking board
quality, the finger joint strength and the quality of glue lines. A glue log is also main‐
tained, which must include details of the production date and number, wood species,
strength class, dimensions of the member, moisture content, time of initial glue applica‐
tion, start and end time of the pressing process, laminating pressure, resin and hardener,
use of adhesive (g/m2), calibration of the moisture analyser and the temperature and
relative humidity of the various working spaces.
108
Wood products
Types of glued laminated timber
EN 14080 differentiates horizontal laminated glued laminated timber with one or more
adjacent lamellae (Figure B6‐7). Alternatively, multiple narrow glued laminated timber
members are bonded together to form a single wider member (block‐glued glulam).
In load‐bearing constructions, since glulam beams are particularly suited to accomodate
bending stresses, the main forces exerted on outer lamellae are tensile or compressive
forces. This explains the range of cross‐section composition using various strength classes
of lamellae, as shown in Figure B6‐8. In this case, a distinction is established between
homogenous and combined glued laminated timber (e.g. GL24h and GL24c respectively).
In homogeneous glulam, all the lamellae comprise boards of a particular strength class. In
combined glulam, the middle lamellae may have a lower strength class than the outer
lamellae and this combination can be symmetrical or asymmetrical. For higher glulam
beams, the outer lamellae comprise at least two lamellae, while for smaller cross‐
sections with up to ten lamellae, at least one lamella. A combination may encompass
multiple wood species. This is appealing when the use of “lower quality” and hence more
economical wood is considered. However, producing hybrid and even combined glulam
makes production costlier, since the different boards have to be correctly arranged in the
production process. For asymmetrical cross‐sections, proper installation on the building
site must also be guaranteed with a label.
This is particularly applicable for certain members made of glued laminated timber and
shown in Figure B6‐9. If, as is done for three‐hinged frames and tapered beams, the
external lamellae are cut, particular attention is required for combined glulam structures,
since lamellae in the lower strength class tend to "shift" in the outer area.
Figure B6‐7 Horizontal laminated glulam with one (left) or multiple adjacent (right) lamellae.
109
Wood products
Figure B6‐8 From left to right: Symmetrical and asymmetrical combined and homogenous glued
laminated timber.
Curved beam
Three-hinged frame,
finger-jointed
Double-tapered beam
Figure B6‐9 Special glued laminated timber members, examples.
Influencing parameters that determine strength
Glued laminated timber is mainly used in constructions exposed to bending stress, which
is why this section focuses on the specific factors influencing the bending strength of
glued laminated timber members. The additional strength properties will be discussed in
the following section.
Systematic investigations (Colling, 1990; Frese, 2006) show that the strength of glued
laminated timber members is a factor of both the strength of the boards as well as that
of the finger joints. If a glued laminated timber beam is exposed to stress, the natural
result is a move to relieve the stress as soon as possible. The breakage thus tends to
occur at points where the ratio of stress over strength peaks and may be either a section
of the board with knots or a finger joint. In glued laminated timber members including
poor quality finger joints, the failure is usually due to the finger joints themselves, while
glued laminated timber members containing substandard planks are prone to break in
areas of knots. This means that efforts to attain premium glued laminated timber must
consider all of components, boards and finger joints alike. There is little sense in seeking
110
Wood products
to unilaterally reinforce only one of these components, because the load‐bearing capaci‐
ty of the glued laminated timber beam will increasingly be determined by the other
(weaker) component, making it impossible to exploit the targeted strength improvement.
The following conclusions can be made following the cited investigations:
Visual grading with stringent knot requirements for boards is not a suitable means
of effectively boosting the load‐bearing capacity of glued laminated timber mem‐
bers. This is because reducing the permitted degree of knots merely boosts the
board strength itself, taking no account of finger joint strength. Accordingly, the
failure of glued laminated timber members is increasingly triggered by the finger
joints, which prevents higher board strength from being exploited.
By using strength grading based on density and/or the modulus of elasticity of the
wood, the strength of both the boards and finger joints can be controlled, which
helps achieve clearly stronger glued laminated timber. The machine strength grad‐
ing of the boards can thus be considered key for high‐strength glued laminated
timber.
Since various finger‐jointing profiles are used and a range of production conditions
play a part (such as the age of the adhesives, pressing power applied or climatic
conditions in the production spaces), it is imperative to monitor the quality of the
finger‐jointing to ensure sufficient strength can be guaranteed. This applies all the
more because production‐related factors interact and are often very difficult to
control.
Background to the CEN provisions
EN 14080 specifies equations used to calculate the mechanical properties of homoge‐
nous glued laminated timber depending on the board and finger‐jointing properties.
Currently valid equations for certain important properties are specified in Table B6‐5.
Here, the specified equations apply for homogenous glued laminated timber, while for
combined glued laminated timber, the equations apply to calculate the properties of the
various cross‐sectional parts. The equations, in turn, are based on members with a
height/width of 600 mm for bending beams/tension members, or on a reference volume
of 0.01 m3 for beams exposed to tensile stresses perpendicular to the grain. This should
be taken into consideration when deviating from these reference dimensions. In this
case, reference is made to the so‐called volume effect, which takes account of the fact
that the strength of a brittle material declines with increasing member dimensions.
111
Wood products
Table B6‐5 Some mechanical properties of glued laminated timber.
Property Equation in accordance with EN 14080:2013
Bending strength [N/mm2] 0.65
0.75 f m,j,k
f m,g,k 2.2 2.5 f t,0,l,k 1.5 f t,0,l,k 6
1.4
if 1.4 f t,0,l,k f m,j,k 1.4 f t,0,l,k 12
Tensile strength [N/mm2]
parallel to the grain ft,0,g,k = 0.8 ∙ fm,g,k
perpendicular to the grain ft,90,g,k = 0.5
Compressive strength [N/mm2]
parallel to the grain fc,0,g,k = fm,g,k
perpendicular to the grain fc,90,g,k = 2.5
Shear strength [N/mm2] fv,g,k = 3.5
Rolling shear strength [N/mm2] fr,g,k = 1.2
MOE parallel to the grain [N/mm2] E0,g,mean = 1.05 ∙ Et,0,l,mean
MOE perpendicular [N/mm2] E90,g,mean = 300
Shear modulus [N/mm2] Gg,mean = 650
Rolling shear modulus [N/mm2] Gr,g,mean = 65
Density [kg/m3] g,k = 1.1 ∙ l,k
The manner in which the bending strength of the glued laminated timber depends on the
tensile strength of lamellae and the bending strength of the finger joints is recorded via
an empirical relationship, determined based on tests and numerical investigations. The
higher bending strength of glulam in comparison to the tensile strength of the lamellae is
attributable to various lamination effects, key examples of which are outlined below:
The load‐bearing behaviour of a board in a standard tensile test differs from that
in a glued laminated timber member. The test procedure outlined in EN 408 to de‐
termine tensile strength envisages a specific minimum test length and does not
provide any lateral support for the test pieces. Accordingly, eccentrically placed
knots or areas with asymmetrical density distribution may result in lateral defor‐
mations and thus additional bending moments, which reduce the tensile strength
of the boards. However, within a glued laminated timber beam, the boards are se‐
cured by bonded adjacent lamellae, meaning a board in a glued laminated timber
beam has seemingly higher tensile strength than in a free tensile test.
112
Wood products
The bonding of the lamellae means areas with lower stiffness may transmit forces
to more rigid neighbouring lamellae. This means, e.g. that the strain on board sec‐
tions with knots can be relieved, which, in turn, equates to a seeming increase in
tensile strength.
The way the structure of glued laminated timber includes individual lamellae results in a
more homogenous material with reduced variance in associated density. Accordingly, the
specified characteristic density values for glued laminated timber exceed those for indi‐
vidual lamellae.
Strength classes
EN 14080 specifies strength classes for homogenous glulam (see Table B6‐6).
Table B6‐6 Strength classes for homogenous glulam in accordance with EN 14080:2013.
113
Wood products
B6.3 Adhesives
Adhesives (glue) for load‐bearing wooden members are used to join two or more timbers
together in such a way that they behave as a single unit in structural terms. Here, the role
of the adhesive is to fill the joint between the wooden parts and establish adhesion be‐
tween the individual parts as strong and durable as the cohesion within the wood itself.
In addition, the adhesive layer itself must have sufficient strength and durability, to re‐
main effective in the service class during the expected service life of construction.
The binding forces between the adhesive and wood are, like the cohesive forces in wood,
electrical forces of attraction between the molecules. The resulting binding forces are
normally of the secondary type, namely hydrogen bonds and Van der Waals forces. In
some adhesives, primary bonds are also likely to form (real chemical bonds), such as
covalent forces. To ensure the effective contact required to establish these adhesive
bonds, the adhesive must be in a liquid form at a specific stage of the setting process and
the wood surfaces must have been freshly planed.
The binding process comprises two stages:
Applying a liquid adhesive, which impregnates the surfaces of both the parts
to be joined in such a way that forces of attraction between the adhesive and
wood molecules exceed the surface boundaries,
Transition of the liquid adhesives, which fill the gaps between the individual parts,
into a solid state, remaining sufficiently strong and durable and with unchanged
effectiveness during the construction service life.
The process above is known as curing and can be initiated in three ways:
Via a physical process, such as the transformation of a solution or the solidification
of a melt (thermoplastic adhesives like PVAc (polyvinyl acetate) and hot melts),
Via a chemical process, which involves the glue molecules interreacting
and polymer cross‐linking (epoxy resin and polyurethane adhesives),
Via a combination of solution transformation and chemical reaction
(urea, melamine, phenol and resorcinol‐formaldehyde resin).
Construction adhesives always involve a chemical reaction taking place. Purely physical
hardening adhesives are thermoplasts and the excessive creep they allow renders them
unsuitable for structural purposes.
114
Wood products
Classification of adhesives
In EN 301, adhesives are sub‐classified into:
Adhesive type I for use in unlimited weathering and for temperatures
exceeding 50°C,
Adhesive type II for use in heated and ventilated buildings, when protected
against external weathering, for short‐term weathering and for maximum
temperatures of 50°C. In Germany, since 2006, type II adhesives have no
longer been used for glulam production.
Adhesive types
Adhesives are non‐metallic and most often sub‐classified into natural and synthetic
groups. Synthetic adhesives are used for bonding wood, based on duroplastic resins:
Urea formaldehyde resin UF,
Urea melamine formaldehyde resin MUF,
Melamine formaldehyde resin MF,
Phenol formaldehyde resin PF,
Resorcinol formaldehyde resin RF,
Phenol resorcinol formaldehyde resin PRF,
Polyurethane PUR.
Additional synthetic adhesives use thermoplastic resins (polyvinyl acetate PVAc – the
most well‐known white glue for DIY enthusiasts), although this is not used for bonding
load‐bearing members. For thick‐layer constructive bonding (e.g. for glued‐in threaded
rods), epoxy resins (EP) are also used.
Formaldehyde‐based adhesives harden when the formaldehyde reacts with the added
resin. The environmental conditions required for this, including the pH level, temperature
and curing duration depend on the resins used or a combination thereof. To start the
curing process for UF, MF and RF adhesives, either salts have to be added or the pH
content must be changed. Conversely, all alkaline‐hardening PF‐adhesives need is the
addition of heat.
115
Wood products
Urea formaldehyde resin UF
Given the good workability, fast curing, colourlessness, low price and usability with vari‐
ous wood species, UF adhesives are in very widespread use. However, they are not re‐
sistant to moisture and temperature, although adding the far more expensive melamine
to MUF‐adhesives improves the moisture resistance.
Melamin formaldehyde resin MF
Compared to UF adhesives, MF adhesives are far more expensive, but also far more
resistant to moisture and temperature, which means they perform better in terms of the
swelling and shrinking response.
Phenol formaldehyde resin PF
PF‐adhesives are also moisture‐resistant and mainly used as hot‐curing PF‐adhesives to
produce particleboards and fibreboards, OSB, LVL and plywood. They possess good swell‐
ing and shrinking behaviour (low thickness swelling) and moreover, do not emit as much
formaldehyde. PF‐adhesives are economical and readily available.
Resorcinol formaldehyde resins RF and PRF
Adding resorcinol significantly increases the reactivity of the PF‐adhesives and allows for
cold curing, meaning heat is no longer needed for the adhesives to harden. They are
highly resistant to moisture and temperature and this, together with the cold‐curing
property, makes RF‐ and PRF‐adhesives popular conventional choices for glulam con‐
struction. The high costs of using resorcinol mean that in this case, as is done for MUF‐
adhesives, combinations of phenol‐ and resorcin adhesives (PRF) are often used. Phenol
and resorcinol are dark in colour, which may be a disadvantage for visible applications.
All the adhesives described to this point contain formaldehyde. Recent years, however,
have seen an increasing shift to the use of PUR‐adhesives, which contain no formalde‐
hyde. PUR‐adhesives usually function with a dual‐component approach, whereby the
addition of isocyanate triggers the start of the curing process. PUR‐adhesives often stand
out for their “foamy” character. Since the curing and adhesion largely depend on the
environmental conditions, temperature and moisture as well as the wood moisture con‐
tent itself, these parameters are also regulated in EN 14080. The various curing times of
adhesives (as specified by manufacturers) must also be taken into account, normally
specified in terms of “open time” (time between mixing and the start of the reaction) and
“pressing time”.
116
Wood products
B6.4 Cross‐laminated timber
All types of cross‐laminated timber (CLT) are classed as structural wood materials. Unlike
plywood, CLT is always used for load‐bearing members, which is why it merits a brief
introduction in a dedicated section. The advantages of this plate‐shaped CLT as opposed
to solid timber are the approximated isotropy in‐plane and lower variance in properties.
In addition, the CLT elements are far more dimensionally stable. The effective stiffness
values can be calculated using the “shear analogy method” (see Article D7).
Material
CLT comprises multiple cross‐wise arranged board layers. The compound structure is
usually formed via gluing, but hardwood dowels or aluminium nails may also be used as
mechanical joining means. The individual layers are completely (Figure B6‐10 on the left)
or partially (Figure B6‐10 on the right) filled with boards, with melamine resins or polyu‐
rethane (PUR) the most frequently used adhesives. All wood species used in practice to
date have been conifers, although first producers try to use beech. In the vast majority of
cases, CLT is made of spruce, but CLT panels are also made of fir, pine, larch or Douglas
fir. These boards, most of which are between 15 and 40 mm thick, are kiln‐dried before
the gluing and strength‐graded. Normally, the narrow sides of the boards are not sys‐
tematically glued together, while the size of the board elements produced depends on
the manufacturer. Finishing is performed in CNC facilities and there is also scope to ma‐
nufacture bent elements.
The layer‐by‐layer structure means a very wide range of panel thicknesses (between
around 60 to 400 mm) and compositions is possible. CLT is generally constructed sym‐
metrically, although partially filled CLT may also show asymmetrical cross‐sections
(Figure B6‐10 on the right). Numerous compositions are produced, the respective prop‐
erties of which are adapted to specific usage applications.
Figure B6‐10 CLT. Completely filled with boards (left), partially filled with gaps (right, photo: Lignotrend
productions GmbH).
117
Wood products
Types of loading acting on cross‐laminated timber:
Loading perpendicular to the plane: Bending parallel or perpendicular
to the grain of the external layers,
In‐plane loading: Bending / compression / tension in plane and parallel
or perpendicular to the grain of the external layers.
Since the modulus of elasticity of wood perpendicular to the grain only equates to
around 3% of the values in the grain direction, the stiffness of a CLT element is deter‐
mined in each case by the layers with the grain parallel to the stress direction. Based on
the variation in thicknesses among the individual layers or the variance in layer composi‐
tion (for example two longitudinal layers and only then another perpendicular layer), the
elasto‐mechanical properties can vary considerably, even among CLT elements with the
same overall thickness. One additional key aspect is rolling shear, since rolling shear
stiffness and strength are very low. When loads perpendicular to the plane generate a
non‐uniform bending moment, shear also acts in‐plane, resulting in rolling shear in the
layers perpendicular to the stress direction. In Figure B6‐11 and Figure B6‐12, the distribu‐
tion of bending stress and the related distribution of shear stress are shown schematically.
To sum up, we can conclude that numerous dimensions, compositions and materials are
available. The product standard for CLT is EN 16351 “Timber structures ‒ cross‐laminated
timber ‒ requirements” (currently not a harmonised version). Individual producers usual‐
ly also have German national technical approvals or European technical assessments, in
which the production requirements and material properties are specified. The approvals
also often regulate characteristic values (particularly effective bending stiffnesses, bend‐
ing and rolling shear strengths).
Longitudinal layer
Transverse layer Rolling shear
Longitudinal layer
Transverse layer Rolling shear
Longitudinal layer
Figure B6‐11 Distribution of bending stress (left) and distribution of shear stress (right) of a 5‐layer CLT
panel loaded perpendicular to the plane and parallel to the grain of the top layer.
118
Wood products
Constructive design
The CLT construction method is quick, easy and also advantageous from a building phys‐
ics perspective. The CLT elements are cut to size in the factory, thermal insulation and
windows are often installed immediately, whereupon the completed elements are trans‐
ported to the building site and installed. CLT are used as both wall and floor elements and
typically joined using self‐drilling screws and nails. Figure B6‐13 to Figure B6‐15 show
typical buildings and joints.
Figure B6‐13 Construction progress of a CLT building; after 2, 4 and 6 days respectively. (Rasom S.r.l.,
Pozza di Fassa)
Figure B6‐14 Left: Corner connection with screws. Right: Connection floor‐wall.
119
Wood products
A A
Section A-A:
LVL Partially threaded self-drilling screws
Figure B6‐15 Possible in‐plane joint between two wall panels.
B6.5 Wood‐based panels
120
Wood products
Solid wood panels (SWP), resin bonded
The panels comprise three or five board‐ or veneer layers of softwood, bonded together,
where the grain directions between adjacent layers run at right angles. One example is
shown in Figure B6‐17. Modified melamine resins and phenol resins are used for gluing.
The various thicknesses of the individual layers mean the elasto‐mechanical properties
may vary considerably, even among panels of the same overall thickness. The solid wood
panels are basically CLT elements, which were already presented. However, given that
they comprise thin board‐ or veneer layers, are not used as standalone elements and are
only deemed to be load‐bearing when used as sheathing for shear walls (like other wood‐
based panels in timber frame constructions), a brief explanation will be given here. Solid
wood panels are mainly employed in furniture‐making, divided into non load‐bearing (acro‐
nym NS non‐structural – SWP/1 NS) and load‐bearing (acronym S structural – SWP/3 S)
categories and regulated in EN 13986 and EN 13353. The technical class SWP/1 is usable in
service class 1, SWP/2 in service classes 1 and 2 and SWP/3 in service classes 1, 2 and 3.
Figure B6‐17 Triple‐layered solid wood panel.
121
Wood products
Veneer‐based panels, resin bonded
Plywood
Plywood comprises at least three layers (plies) bonded to each other, which generally run
at right angles, to ensure a barrier effect in‐plane (dimensional stability) (for example see
Figure B6‐18). The central plies may consist of veneers (plywood), wooden strips (block
board) or thin strips of wood (laminboard). In all three types, the top ply comprises ve‐
neers, which are thin sheets of wood, normally produced by peeling in a process that
yields strips 0.5 to 6.0 mm thick with a width determined by the length of the peeled log
portion. They are used to produce large‐scale panels. Veneers used in timber construc‐
tion are produced by rotary peeling (Figure B6‐19). For this purpose, the wood used must
be soft and as resistant to cracking as possible, which can be ensured by steaming the log
portions or storing them in hot water. The log portions, still hot from the steaming pro‐
cess, are then conveyed to a centring station, which scans the log shape mechanically or
visually to determine the optimal position for the spindles of the peeling machine; target‐
ing maximum yield. A knife carrier and pressure strip advance by the thickness of a single
veneer with each log rotation, while ensuring the adjustment angle of the knife conforms
to the log contour and with the remaining rolls used as raw materials for the parti‐
cleboard industry. After peeling, the veneers end up as a relatively regular platform, with
many defective spots at the edges and holes and other defects on the transport line. This
is where the veneers are scanned, a cutting device is used to eliminate defective portions
and plies of equivalent width are cut. The high moisture from the steaming process is
also reduced by 6 to 12%, whereupon adhesive is applied to the veneers, they are consol‐
idated to form panels and hot‐pressed together. Finally, the panels are climatised or
conditioned, to ensure they cool down and moisture is uniformly distributed throughout
the panel thickness. The final work process involves trimming the panels or cutting and
sanding them to fixed dimensions. For certain applications, e.g. formwork panels, a coat‐
ing process may follow. The production is shown schematically in Figure B6‐20.
Figure B6‐18 Plywood with longitudinal and perpendicular veneers of various thicknesses.
122
Wood products
Figure B6‐19 Rotary peeling (schematic illustration). (STEP 1995 Article A10)
Figure B6‐20 Production method for veneered materials. (Metsäwood, brochure „Ich bin Kerto“, 2014 version)
Table B6‐7 Technical plywood classes and usability.
Technical class Use in service class
For internal use in dry conditions 1
For internal use in humid conditions 1, 2
For external use 1, 2, 3
Plywood is created by arranging and bonding veneers in alternative crosswise layers,
whereby the veneers must be arranged symmetrically to the central plane. The adhesives
used in this case are urea resins (only of the “dry” class), alkaline‐hardening phenol res‐
ins, phenol resorcinol resins and resorcinol resins. Changing the number, thickness and
arrangement of the individual plies opens up a whole host of options concerning panel
composition, which means panels with targeted properties can be created. The type of
veneer wood used is also a key criterion dictating the plywood properties. Plywoods can
be constructed from various wood species, provided symmetry is maintained. Plywood is
mainly used as a sheathing material in horizontal (floors, roofs) or vertical (shear walls)
diaphragms. Since the modulus of elasticity of wood perpendicular to the grain only
equates to around 3% of the value in the grain direction, the stiffness of a plywood panel,
similarly to a cross‐laminated timber panel, is respectively determined by the layers with
grains running parallel to the stress direction. Accordingly, for plywood, a distinction is
made between both the main axes running parallel and at right angles to the grain of the
top plies.
123
Wood products
Plywood is regulated in EN 13986 and in the EN 636 product standard. Besides in tech‐
nical classes (Table B6‐7), plywood is also classified in terms of bending strength and
bending modulus. The technical classes govern the scope of use depending on the pre‐
vailing climate. The additional classification of plywood panels depends on the minimum
required values per length or width of panels for bending strength, letter “F”, or the
modulus of elasticity with the letter “E”, see Table B6‐8. This means that a plywood panel
classed as F 10/20 E 30/40 must have a minimum bending strength value along the
length of fm,0 = 15 N/mm² and widthwise, fm,90 = 30 N/mm² and a minimum value for the
bending modulus along the length of Em,0 = 2700 N/mm² and widthwise, Em,90 =
3600 N/mm².
Table B6‐8 Bending strength and bending modulus classes for plywood, EN 636:2012.
Bending strength in N/mm2 Bending modulus in N/mm2
Class Minimum value Class Minimum value
F 3 5 E 5 450
F 5 8 E 10 900
F 10 15 E 15 1350
F 15 23 E 20 1800
F 20 30 E 25 2250
F 25 38 E 30 2700
fm,0
F 30 45 E 35 3150
fm,90
F 35 52 Em,0 E 40 3600
F 40 60 Em,90 E 50 4500
F 50 75 E 60 5400
F 60 90 E 70 6300
F 70 105 E 80 7200
F 80 120 E 90 8100
E 100 9000
E 120 10800
E 140 12600
Laminated Veneer Lumber (LVL)
Laminated veneer lumber is produced using rotary cut softwood or hardwood veneers,
with phenol resin used for gluing. The fibre direction of the veneers is either generally
parallel to the longitudinal direction of the laminated veneer lumber or predominantly
parallel and slightly (up to around 25%) perpendicular to the longitudinal direction of the
laminated veneer lumber (Example see Figure B6‐21). The veneers within a layer are
generally combined using a scarf or overlapping process. Laminated veneer lumber is
regulated in product standards EN 14279 and EN 14374. Laminated veneer lumber is
124
Wood products
used as either load‐bearing sheathing or, like glulam, as a bar‐shaped member and classi‐
fied in accordance with the three technical classes of EN 14279:
LVL/1 for use in dry conditions (service class 1)
LVL/2 for use in humid conditions (service class 2)
LVL/3 for external use (service class 3)
Peeling cracks in veneers allow entire cross‐sections of laminated veneer lumber to be
impregnated, after which it can even be used in locations exposed to adverse weather
conditions. The strength values required for design are determined in accordance with
the method specified in EN 14374 and specified in the CE label for the individual prod‐
ucts. Although LVL centres on the concept of increasing the homogeneity of this wood‐
based product over solid timber, it generally encompasses a range of densities through‐
out the board thickness, see Figure B6‐22. This is often influenced by production factors
(pressing power and heat), but may also be a desirable feature, particularly in wood‐
based panels like OSB and particleboards, to impact on the end panel properties.
Figure B6‐21 Laminated veneer lumber made of beech veneers.
1100
BauBuche-Pk. (50-mm-Lamellen)
rho
1000
Rohdichte in kg/m³
3
900
Density in kg/m
800
700
600
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20
Layers (5, 10 and 15: with secondary bond from glulam production)
Schichtfolge (5, 10 und 15 mit Trägerfuge)
Figure B6‐22 Density profile of glulam made of beech LVL. The chart shows the minimum, maximum and
average values of 10 measurements on glulam, made of 4 laminated veneer lumber layers,
each encompassing 5 veneer layers.
125
Wood products
Parallel Strand Lumber (PSL)
PSL (Parallel Strand Lumber, see Figure B6‐23) is a special invention, which was originally
planned to facilitate the use of waste veneers from plywood production. For PSL, rotary
cut veneers, normally from Douglas fir or Southern Yellow Pine and around 3 mm thick
are used. The veneer sheets are cut into strips up to 2.50 m long and around 23 mm
wide, whereupon the “strands” are waxed, positioned longitudinally and made to trav‐
erse a through‐feed press in a process that can generate almost unlimited beam lengths.
A subsequent quality control process sees the beams checked visually and in terms of
weight before delivery. Parallel strand lumber is not regulated by product standards and
requires general technical approval in Germany. PSL (trade name Parallam) stands out,
thanks to its high strength values, increased stiffness and dimensional tolerance. The
main PSL applications include beams, structural members, purlins, columns and trusses.
Figure B6‐23 Parallel Strand Lumber, PSL.
Wood‐based panels made of chips, resin bonded
Laminated Strand Lumber (LSL)
LSL is so‐called laminated strand lumber and distributed under a range of trade names
(including Intrallam), see Figure B6‐24. LSL is generally technically approved in Germany.
LSL comprises trim strands of poplar or aspen wood, bonded to each other and with
dimensions of around 0.8 mm x 25 mm x 300 mm. The raw wood is debarked before the
machining process and any excessively short chips are discarded. The relatively high
densification and the type of adhesive allows high strengths to be attained, while there is
also scope to manufacture a range of strength classes by changing the distance between
the mat‐forming machine to the mat surface. The greater the distance, the higher the
proportion of trim strips in a perpendicular board direction, while for smaller distances,
the trim strips tend to be arranged in parallel.
126
Wood products
Figure B6‐24 Laminated Strand Lumber, LSL.
Oriented Strand Board (OSB)
OSB is made using longitudinal strands, preferably parallel to the board surface, which are
around 0.6 mm thick, 75 mm long and 35 mm wide, see Figure B6‐25. The strands can be
applied single or multiple‐layered, with multilayer panels offering additional scope to vary
the strand size in the individual layers, the degree of orientation, layer thicknesses, adhe‐
sives used and their proportion. Strands for top layers should preferably run parallel to the
production direction, while those in the middle layer should be random or perpendicular to
the production direction, which means OSB achieves different longitudinal and perpendicu‐
lar properties. The bending strength in the longitudinal direction of the board significantly
exceeds that in the perpendicular direction. Moreover, the density of OSB generally varies
through the board thickness (similar to LVL, Figure B6‐22). OSB is regulated in EN 13986
and in the EN 300 product standard, which includes four technical classes:
OSB/1 non load‐bearing boards for general purposes and interior fitments
(including furniture) for use in dry conditions (service class 1)
OSB/2 load‐bearing boards for use in dry conditions (service class 1)
OSB/3 load‐bearing boards for use in humid conditions (service class 2)
OSB/4 heavy‐duty load‐bearing boards for use in humid conditions
(service class 2)
Figure B6‐25 Oriented Strand Board, OSB.
127
Wood products
Particleboard (chipboard)
Particleboards are formed by spreading out relatively small wood chips sprayed with
adhesive on a level underlay and then pressing, as shown in Figure B6‐26. The pressed
boards are weighed, any defective portions are detected and discarded, using an ultra‐
sound device and the thickness is finally measured. The boards then undergo the final
production process, where they are conditioned, trimmed, sorted, labelled, sanded and
stored. Particleboards are regulated in EN 13986 and in the EN 312 product standard.
The technical classes and their usability are specified in Table B6‐9, although technical
classes for non load‐bearing use (P1, P2, P3) are excluded.
Figure B6‐26 Particleboard.
Table B6‐9 Particleboard classes and usability for load‐bearing purposes.
128
Wood products
Fibreboards, resin bonded
Classification of fibreboards
Fibreboards are quasi‐isotropic wood‐based panels, meaning their in‐plane properties
apply independently of any direction. They comprise individual or bundled fibres which,
depending on the production method, contain a greater or lesser proportion of chemicals
and adhesives. Fibreboards produced using wet production methods have almost no or
minimal quantities of adhesive added, with mainly natural fibre bonding used. Converse‐
ly, the boards having undergone the dry production method contain larger quantities of
adhesive, to ensure they attain the properties determined in the product standards.
Fibreboards are regulated in EN 13986 and the EN 316 and EN 622 product standards.
Since the fibreboard designation is relatively unclear, all the relevant tables from EN 316
are shown at this point, with the first sub‐grouping made according to the production
method:
Wet process: Fibreboard with moisture content exceeding 20%
during the forming stage,
Dry process: Fibreboard with moisture content of less than 20%
in the forming stage.
A further distinction is applied to boards produced using the wet process based on their
density:
Hardboards HB: Density > 900 kg/m3, Figure B6‐27 on the right
Low‐density medium board MBL: 400 kg/m3 to < 560 kg/m3
High‐density medium board MBH: 560 kg/m3 to < 900 kg/m3
Softboard SB: Density > 230 kg/m3 to < 400 kg/m3, Figure B6‐27 on the left
Figure B6‐27 Left: softboard, right: hardboard.
The dry‐process fibreboards are abbreviated to the label “MDF” (medium‐density fibre‐
boards), while EN 622‐2 explains the special technical classes. The additional classifica‐
tion is implemented depending on the application conditions and purpose of use,
Table B6‐10.
129
Wood products
Table B6‐10 Classification of fibreboards.
Application conditions, purpose of use, structural use Abbreviation
Application conditions
For internal use, dry conditions No abbreviation
For internal use, humid conditions H
For external use E
Purpose of use
General purpose No abbreviation
Load‐bearing purpose L
For all load‐duration classes A
Only for instantaneous or short‐term loading S
Structural use
General structural use 1
Heavy‐duty 2
Examples:
MDF.HLS = dry‐process medium density fibreboard for internal use as a structural
component in humid conditions, only for for instantaneous or short‐term loading
HB.HLA2 = hardboard for internal use as a heavy‐duty structural component
in humid conditions, for all load‐duration classes
Wood‐based panels, mineral bonded (cement, gypsum)
All mineral bonded boards, excluding wood wool boards, can be used as sheathing mate‐
rial in timber frame constructions. As a general rule, they are used for dry lining and fire
protection in particular.
Cement‐bonded particleboard
The cement‐bonded particleboard, Figure B6‐28, comprises treated spruce and fir wood
particles, which function as reinforcement and cement acting as a binder. The boards are
regulated in EN 13986 and the EN 634 product standard. There are two technical classes
for cement‐bonded particleboards, which only differ in terms of the requirement for the
minimum bending modulus:
Technical class 1: 4500 N/mm²
Technical class 2: 4000 N/mm²
130
Wood products
Figure B6‐28 Cement‐bonded particleboard.
Gypsum‐bonded particleboard
Gypsum‐bonded particleboards comprise gypsum and spruce or aspen particles, which
function as reinforcement.
Gypsum fibreboard
Gypsum fibreboards (Figure B6‐29) comprise gypsum and paper fibres, which are ex‐
tracted in a recycling method and which function as reinforcement
Figure B6‐29 Gypsum fibreboard.
Gypsum plasterboard
Gypsum plasterboards (Figure B6‐30) comprise a gypsum core, which is surrounded by
paper surfacing on the face, back and long edges commensurate with the purpose of use.
The encased gypsum core may be porous and include additives intended to imbue it with
specific properties. The key properties of the panels are a factor of the composite effect
of the gypsum core and the paper encasing, whereby the paper encasing acts to rein‐
force the tensile area and provides the gypsum boards with the required strength and
bending strength, when combined with the gypsum core.
Figure B6‐30 Gypsum plasterboard.
131
Wood products
Wood wool board
Wood wool boards are lightweight boards made out of wood wool and mineral binders
(cement or magnesite). They are used in construction to insulate heat and sound and
protect against fire.
B6.6 Literature
P. Glos, F. Colling, A. Ranta‐Maunus, G. Steck, D.R. Griffiths, E. Raknes, original Articles A7 – A12,
STEP 1995.
Studiengemeinschaft Holzleimbau e.V. (1998). Argumente für BS‐Holz. Informationsblatt zum Brettschichtholz
des Informationsdienstes Holz, Düsseldorf.
Colling F. (1990). Tragfähigkeit von Biegeträgern aus Brettschichtholz in Abhängigkeit von den
festigkeitsrelevanten Einflußgrößen. Reports of the Versuchsanstalt für Stahl, Holz und Steine of
Karlsruhe University, Volume 4, Issue 22, 205 pp.
Frese M. (2006). Die Biegefestigkeit von Brettschichtholz aus Buche. Experimentelle und numerische
Untersuchungen zum Laminierungseffekt. Dissertation, Universität Karlsruhe (TH).
132
C
Principles for design
C1 Safety concept
Original article: H. J. Larsen
The Eurocode principles for safety and serviceability of structures and the basis of design
common to all materials are explained. Moreover, specific rules for timber structures,
required due to the impact of load duration, wood moisture content and the scatter of
material properties, are specified. Requirements for normal design situations and special
requirements for accidental design situations are also described.
C1.1 Calculation models
135
Safety concept
Figure C1‐1 Heel joint. (STEP 1995 Article A2)
C1.2 Standards based on limit states
Eurocodes are standards based on limit states, which means that the requirements con‐
cerning structural reliability are subject to clearly defined limit states. Limit states are
those which, when exceeded, render the structure no longer capable of meeting the
adopted design requirements. Within the Eurocodes system, only two limit states are
taken into consideration: the ultimate and the serviceability limit state.
The ultimate limit states are any states, which, due to any collapse or other failure affect‐
ing the structure, may put personal safety at risk. This may involve e.g. the loss of equilib‐
rium of a structure or one of its parts, while failure can also be initiated by excessive
deformations, rupture or loss of stability of a structural part.
The serviceablity limit states comprise deformations which would impact on the appear‐
ance or planned usage of a structure, vibrations, which cause unease or damage to con‐
struction or related equipment or which limit the functional scope of the construction and
damage (including cracks), which has a long‐term impact on the durability of the structure.
C1.3 Design in accordance with the partial factor method
Structural design in Eurocodes is based on the partial factor method. The key parameters
are the actions F, the material properties X and the geometrical data a1.
1
The effects of actions E are a function of the actions F and the geometrical data a: E = (F, a)
The resistance R is a function of the material properties X and the geometrical data a: R = (X, a)
136
Safety concept
Emean Ek Rk Rmean
Figure C1‐2 Idealised statistical distributions for the effect of actions E and resistance R.
(STEP 1995 Article A2)
As a general rule, these variables are random with distribution functions, as shown in
Figure C1‐2 for the effect of actions E and the corresponding resistance R: e.g. bending
stresses and bending strength or the normal force in a centrally loaded column and its
buckling load. The corresponding distributions reveal the average values Emean and Rmean,
whereby the characteristic values Ek and Rk are determined as distribution quantiles. As
regards the characteristic values of the effects of actions Ek, in this case, although an
upper quantile value is usually applied (e.g. 98% quantile), applying a lower value is also
often acceptable, e.g. when designing against uplift. A lower quantile value (e.g. 5%
quantile) or average value is generally used for the characteristic resistance Rk. In excep‐
tional cases, e.g. if a connection should fail earlier than the connected member, an upper
quantile value may also be necessary. The goal of the design is to ensure a low probability
of failure, namely minimising the possibility of the effect of action E exceeding the re‐
sistance R. The probability of failure is shown visually in Figure C1‐3 and corresponds to
the overlap area of the frequency curves. Using the method with partial safety factors
allows a low probability of failure to be attained, whereby design values are used (Index
“d” = ”design”), which are determined for actions by multiplying the characteristic values
(Index “k” = ”characteristic”) with partial safety factors and for the resistance by dividing
the characteristic values by partial safety factors, see Figure C1‐4.
Frequency
Action Resistance
Failure
Figure C1‐3 Frequency distributions for action and resistance.
137
Safety concept
Frequency
Action
Resistance
Ek Ed fd fk
∙ G, Q ∙ kmod ÷ M
Figure C1‐4 Reduction of the probability of failure through partial safety factors and kmod.
In all key design situations, it is important to ensure that the limit states are not attained,
when the design values of the actions, material properties and geometrical variables are
used in the calculation model. In particular, it is important to ensure that:
In the ultimate limit states, the design values of the action effects
do not exceed the design values of the resistances and
In the serviceability limit states, the design values of action effects
do not exceed key limit values.
In a symbolic form (for the ultimate limit state corresponding to failure), it shall be veri‐
fied that
Ed Rd (C1‐1)
For the serviceability limit state, it shall be verified that
Ed C d (C1‐2)
Here
Ed Design value of the effect of action such as e.g. normal force, bending moment
Rd Design resistance
Cd Limit design value of the relevant serviceability criterion such as e.g. a limiting
value for deflection
138
Safety concept
C1.4 Actions
Representative and characteristic values
This section only covers actions to the extent required to describe safety principles. Arti‐
cle C2 includes a detailed description of actions. Under normal design situations, a dis‐
tinction is made between permanent actions G and variable actions Q. There are direct
actions (e.g. external forces) and indirect actions (e.g. actions caused by shrinkage or
uneven settlements), while actions, or parts of actions, can either be stationary or mo‐
bile. Mobile actions may be quantified as arbitrary values between predefined limits at
any point of the structure. The variable actions are additionally sub‐classified into ‘lead‐
ing variable action‘ and ‘accompanying variable action‘.
Material or product properties
The material properties correspond to either an average value or the 5% quantile and are
determined by standardised tests under uniform test conditions: The test is conducted
within five minutes at a temperature of 20°C and relative humidity of 65%. The average
values are used for serviceability limit state verifications, while the ultimate limit state is
verified by using the 5% quantiles of the material properties (the 5% quantile of stiffness
is only taken into consideration when verifying the stability).
Geometrical data
The characteristic values of geometrical data, such as span, cross‐sectional dimensions or
initial deformation, generally correspond to the nominal values determined in the design.
Design values
The design values of the actions may cover a range of values, when verifying the various
limit states. The first step is to determine the load cases, namely compatible load config‐
urations or deformations and imperfections. A load configuration describes the position,
size and direction of an action, whereupon the actions corresponding to the following
symbolic expression are listed – in this case the fundamental combination of actions for
persistent or transient design situations:
139
Safety concept
The ‐values are partial safety factors for the corresponding action, which take into con‐
sideration the influence of possible unfavourable deviations and possible inaccuracies of
actions and uncertainties when determining the final effects of actions. Certain partial
safety factors for actions for verifications, which govern the strength of construction
materials, are specified in Table C1‐1 (other partial safety factors apply, for example, if
the strength of the soil is the decisive element determining failure).
Table C1‐1 Partial safety factor for ultimate limit states in buildings, EN 1990:2010.
Failure of the structure where the strength of construction materials governs
Unfavourable permanent actions G,sup 1.35
Favourable permanent actions G,inf 1.00
Unfavourable variable actions Q 1.50
The first term in Equation (C1‐3) expresses the design value of the permanent action
Gd = G ∙ Gk, while the second term is the design value of the leading variable action
Qd,1 = Q,1 ∙ Qk,1. The third term Qd,i = Q,i ∙ 0,i ∙ Qk,i is the design combination value of the
accompanying variable actions. The combination factor 0 takes into consideration the
fact that e.g. a simultaneous combination of maximum snow and wind loads is improba‐
ble, whereupon the lower value may be reduced by 0. Table C1‐2 specifies examples for
‐values, while a complete overview is included in EN 1990 and in the NA to EN 1990.
Table C1‐2 Some combination factors for buildings, NA to EN 1990:2010.
where AEd is the design value of the seismic action.
140
Safety concept
For such combinations of actions, no partial safety factors are applied and also lower
combination factors (0 → 2). Similarly, no partial safety factors are applied to verify
the serviceability limit state:
Finally, the effects of actions E are determined, e.g. internal forces, stresses, strains and
deformations, based on the design values of the actions F (Gd, Qd, etc.), the geometrical
data a and, where applicable, the material or product properties X:
E d E F1,d ,F 2,d ,...,a 1,d ,a 2,d ,..., X 1,d , X 2,d ,... (C1‐6)
C1.5 Resistance
Xk
X d k mod (C1‐7)
M
Where
M Partial safety factor for the material property, see Table C1‐3
kmod Modification factor, which takes into consideration the impact of the load
duration and moisture content on the strength properties. Examples for kmod
are specified in Table C1‐4.
Table C1‐3 Partial safety factors for the material properties, NA to EN 1995‐1‐1:2013.
Ultimate limit states:
‐ Wood and wood‐based products, joints 1.3
‐ Punched metal plate fasteners 1.25
Serviceability limit states 1.0
141
Safety concept
The factor kmod depends on the service class of the structure and the load‐duration class.
There are three service classes 1, 2 and 3. The average equilibrium moisture content in
most softwoods does not exceed 12% in service class 1 (internally) and 20% in service
class 2. For service class 3 (externally), however, no limit value is determined for wood
moisture content. The five various load‐duration classes are distinguished based on the
accumulated duration of the characteristic load (see Table C1‐4). The coefficient kmod
thus takes into consideration the effect of various moisture contents and load durations
on the mechanical properties of wood and wood‐based products (see also Article B2).
In general, a structural model is used, which describes the relationship between re‐
sistance R and the strength values f, the stiffness values E (= material properties X) and
the geometrical data a. This kind of model uses the corresponding design values to de‐
termine design resistance:
R d R f 1,d , f 2,d ,...,E 1,d ,E 2,d ,...,a 1,d ,a 2,d ,... (C1‐8)
Table C1‐4 Load‐duration classes and factors kmod for solid timber and glulam.
Rk
R d k mod (C1‐9)
M
Within a structure made of different structural materials, e.g. timber, steel and wood‐
based products, it may be difficult to select the correct factor kmod. However, the selec‐
tion of the lowest values for the construction materials used is always on the conserva‐
tive side.
142
Safety concept
The design values of geometrical data generally correspond to the characteristic values,
namely the nominal values determined in the design. Many cases involve the geometrical
data being determined as follows (e.g. if deviations from the nominal value due to initial
deformations must be taken into consideration → calculations in accordance with sec‐
ond order theory):
a d a k Δa (C1‐10)
The values of a are specified in the corresponding sections of EC 5, e.g. in EC 5 Sec‐
tion 5.4.4 for plane frames and arches.
C1.6 Literature
H.J. Larsen, original Article A2, STEP 1995.
143
C2 Actions on structures
Original article: P. Racher
When structural engineers plan a building, they start by conceptually designing the struc‐
tural system and the design depends on the type of construction and the construction
material used. Based on this, the design gets underway by determining the actions on the
selected construction. Both direct actions of external forces must be taken into consider‐
ation, as well as the indirect actions that result from imposed deformations (e.g. dimen‐
sional changes due to varying moisture or temperature levels or settlement of supports).
Regardless of which construction materials are used, the design must determine the
range of actions which could occur during the planned service life of the structure. These
actions depend on the form, construction type and how the construction is actually
erected. The type of actions or the consequences of an action, which can be either static
or dynamic, must be taken into account to achieve an accurate structural analysis. This
means, e.g. that using a quasi‐static assumption may be inapplicable in the following cases:
Floors, which are subject to vibrations caused by people or machines.
Flexible, plate‐like constructions such as suspension bridge decks,
which may be subject to flutter at critical wind speed.
Constructions subject to stresses caused by the accelerations of seismic action.
In these cases, a dynamic analysis should be performed, to determine the actions of the
force‐time history and taking the stiffness, mass and damping behaviour of the structural
members into consideration, whereby, however, the resonant component of these ac‐
tions is low in most structures. Accordingly, a structural static calculation is frequently
performed while applying an equivalent, dynamic magnification factor (vibration coeffi‐
cient) for the static value of the action.
For these reasons, the current article only covers the direct actions and their combina‐
tions for static calculations. These calculations must also take into consideration the
National Application Documents and the current provisions for countries within which
the construction should be erected.
145
Actions on structures
C2.1 General concepts
Structural classifications
The Eurocodes for design (EC 2 to EC 9) mainly centre on adapting successful traditional
design methods, but even so, it is important to measure the criteria which underpin the
reliability concept of EC 0. The required safety and serviceability depends on the service
life and the design situation of the structures; taking into consideration any danger to
people and economic losses (C.E.B., 1980).
The planned service life (= design working life) specified in Table C2‐1 encompasses the
specific period for which a structure may be used for its intended purpose, without re‐
quiring any significant maintenance (except planned maintenance measures). The design
situations are based on incidents which could occur during the service life of the con‐
struction, whereupon the actions for key design situations, classified as follows, are de‐
termined:
Persistent situations, which constitute normal conditions of use.
Transient situations, which involve temporary conditions, such as
during execution.
Accidental situations related to exceptional conditions such as
fire or impact.
Seismic actions.
Table C2‐1 Classification of design working life (planned service life).
146
Actions on structures
Classification of actions
As well as the above‐listed classifications, the actions must also be distinguished in terms
of the way they vary over a spatial scope and over time. A normal design defines the
actions or effects of actions as follows:
Permanent actions (G), e.g. self‐weight of the structure,
Variable actions (Q), e.g. imposed loads, wind and snow loads.
Accidental (A) and seismic actions (AE) are not covered in this article (for fire, see Article
G1, earthquake see Articles G2 and G3).
The size of permanent actions remains virtually unchanged over time, except in the event
of construction changes (see Figure C2‐1). When variable actions apply, the changes can
be modelled as a discrete process (e.g. snow or wind) or as a process from a permanently
applied proportion QL and a proportion QT applied for only a short time (e.g. imposed
loads) (Hendrickson et. al, 1987, Rackwitz, 1976).
For wood, where elapsed time plays a bigger role in determining strength than for other
construction materials, special attention must be paid to the chronological change in
actions as well as the duration of loads. Accordingly, the structural engineer must classify
the variable actions in the specified load‐duration classes (see also Table C1‐4).
As regards their spatial variability, actions are considered as either stationary or free.
Free actions can be dispersed arbitrarily over the structure or parts thereof in spatial
terms, in which case a design taking into account unfavourable load configurations of
free actions must be performed.
p (kN/m2)
QT
2,0
1,0 QL
G
0,0
0 10 20 30 40 50
T
Figure C2‐1 Chronological change in actions on a floor. (STEP 1995 Article A3)
147
Actions on structures
Representative values of actions
The basic values of actions are the characteristic values, which are designated as Gk or Qk
and as a general rule, the permanent actions Gk correspond to the nominal values. If the
construction responds sensitively to changes in G or if the coefficient of variation (COV,
COV is defined as the standard deviation divided by the mean value) of G exceeds 10%,
two characteristic values, one lower value Gk,inf and one upper value Gk,sup should be
used, which correspond to the 5% quantile or the 95% quantile. Assuming a Gaussian
distribution of G, these values are determined as:
The characteristic variable actions Qk correspond to a given return period of N years,
corresponding to the probability of exceedance of pN = 1/N per year. According to EC 0,
the actions are defined as Qk for N = 50 years or p50 = 0.02, which corresponds to a 98%‐
exceedance quantile.
In addition, the structural engineer has to take other representative values into consider‐
ation for the variable actions:
The combination value (0 ∙ Qk) (= characteristic combination), which is
exceeded two percent of the time,
The frequent value (1 ∙ Qk), which is exceeded five percent of the time,
The quasi‐permanent2 value (2 ∙ Qk), which equates to the average value
over a period of time.
In practice, the values Gk, Qk, (0 ∙ Qk) and (1 ∙ Qk) are to be taken into consideration
when verifying the ultimate limit state. However, when it comes to the serviceability limit
state, these values are only relevant for calculating the short‐term deformations. The
long‐term effects (e.g. creep) should be determined taking the values Gk and (2 ∙ Qk) on
the load side into consideration and the deformation factor kdef on the material side.
2
The designation “quasi‐permanent value” may briefly prove confusing, since 2 is used for accidental design
situations or earthquake actions, although these seldom occur. However, the designation actually relates to
the fact that average values (= quasi‐permanent values) of the variable actions are used in infrequent design
situations, which are lower than their quantile values (0 corresponds to the 98% quantile, 1 corresponds to
the 95% quantile).
148
Actions on structures
C2.2 Permanent actions
The permanent actions comprise the self‐weight of the structural members and the
weight of all components which must be permanently supported by the structural mem‐
bers, e.g. partition walls, insulations, panels and plasters. To determine permanent ac‐
tions, insights into structural configurations and the construction building materials used
are needed. The permanent actions are calculated from the nominal dimensions of com‐
ponents and the average weights of the construction materials used. If these are not
specified e.g. in EC 1, the structural engineers should use the weights specified in accord‐
ance with the manufacturer’s details. To simplify the calculations concerned, the self‐
weights of the bracings and lightweight partition walls should be applied as uniformly
distributed loads, while a reasonable estimate can be obtained by referencing similar
construction parts. The dead weight of floors (wooden joists with sheathing) or roofs
(sheathing, rafter and purlins) is generally between 0.25 and 0.45 kN/m2.
C2.3 Imposed actions
Imposed actions in buildings depend on their usage and correspond to the loads which
move in their own right (e.g. persons, vehicles) and moveable loads (e.g. furniture, light‐
weight partition walls, stored goods), while the building areas exposed to load are differ‐
entiated in accordance with their intended use. In common buildings, three classes are
taken into consideration:
Office, residential and retail premises,
Roofs and
Storage and production facilities.
For storage and production facilities, a design is performed with the imposed loads aris‐
ing during actual use of the building (category E). In other cases, the imposed loads take
into account the density and type of public traffic, with the first class subdivided into four
further categories A to D (Table C2‐2). Roofs, meanwhile, are classified as non‐accessible
except for maintenance and repair (category H), or accessible (categories I and K). Cate‐
gories F and G apply for roads and parking areas in buildings. The NA for EC 1 also defines
categories T for stairs and Z for accesses and balconies. According to this classification,
floors and roofs are designed with a uniformly distributed load qk or a concentrated load
Qk. The concentrated load Qk is exerted on a square area with 50 mm side length at the
least favourable spot.
149
Actions on structures
Table C2‐2 Classification of floor areas in buildings.
C2.4 Snow loads
Snow loads are based on snow depth measurements and snow density. Depending on
the surrounding area and the prevailing weather, the specific density of the snow is be‐
tween 0.1 (new snow) and 0.6 (old or wet snow). A statistical evaluation of these records
is used to define the characteristic snow load sk for a return period of 50 years. The val‐
ues sk are determined by national load standards, since they depend on geographical
locations and altitude. In addition, the structural engineer should also take into consider‐
ation any local effect caused by modifying the value sk. This means e.g. that the snow
load exerted on a member may increase considerably when snow is transformed into ice,
in the event of rain falling on the snow or areas of roof may heat up, causing localised
melting of snow. During the static calculation, the engineer must take into consideration
the load configurations on roofs as follows:
Uniform distribution from uniform snowfall,
Uneven loads due to snow drifts and snow slides.
150
Actions on structures
Sk i sk (C2‐2)
The shape parameter i takes the influence of roof geometry on the load configuration
into consideration, e.g. whether the roof shape is more or less likely to cause snow slides
or drifts. EC 1 defines two parameters i depending on the roof pitch (Figure C2‐2).
Both shape parameters apply, when the snow is allowed to slide unhindered, namely
when no snow guards etc. are attached.
1.6
2
1
0.8
0
0 15 30 45 60
°
Figure C2‐2 Shape parameters for snow on single pitch, double pitch and shed roofs.
For barrel‐shaped roofs, shape parameter 3 applies.
The shape parameter 1 applies to single and simple double pitch roofs, since no accu‐
mulation or pouching of snow is possible. Accordingly, with a roof pitch from 60° onward,
no snow load can accumulate, since all the snow slides off. For double pitch roofs, a
range of snow loads must be checked on both sides of the roof (albeit uniform), since
snow may slide off on one side of the roof, but not the other. The shape parameter 2 is
taken into consideration, if the formation of snow drifts is possible, namely in the event
of lined‐up double‐pitch or shed roofs. In this case, the value for 2 peaks at a roof pitch
of 60°, since this is when snow slides into the gusset and accumulates there.
151
Actions on structures
C2.5 Wind loads
Wind loads change over time and are initially classified based on the short‐term load‐
duration class. The action on the structure can be perceived as the combination of a
quasi‐static component and a resonance component and the latter component in par‐
ticular may be significant for tall and slender structures, in which case more accurate
wind verifications may be required. Despite this fact, the resonance component is less
important for most constructions, which is why, in this article, wind loads are defined
using the simplified method described in this article. Wind loads are represented in the
form of static compressive forces on the construction surfaces or by global compressive
forces and friction wind forces (E.C.C.S. 1987). The scope of this article is limited to wind
loads exerted on constructions impervious to vibration, which can be represented via a
static equivalent load. The wind load assumptions are based on a basic wind velocity vb
and a basic velocity pressure qb = ½ ∙ ∙ vb2 (where = air density). Based on a return
period of 50 years (= probability of occurrence of 2% per year), vb is defined as the aver‐
age ten‐minute wind velocity at a height of 10 m over category II terrain (see Table C2‐3).
Table C2‐3 Terrain categories in accordance with EC 1.
Terrain category
0 Lake or coastal areas exposed to the open sea
I Lakes or areas with little vegetation and free of hindrances
II Areas with little vegetation like grass and individual hindrances (trees, buildings)
III Areas with uniform vegetation or development (e.g. villages, forest areas)
IV Areas, in which at least 15% of the area is developed with buildings with an
average height exceeding 15 m
Based on the basic wind velocity vb and the resulting basic velocity pressure qb, the peak
velocity pressure qp(z) relevant for the design can be determined as a factor that depends
on the construction height z. For this purpose, the ground roughness and topography in
the vicinity of the building location (see also Table C2‐3) as well as any turbulences have
to be taken into consideration.
152
Actions on structures
Pressure coefficients
The aerodynamic pressure coefficients cpe/i define the wind pressure, which impacts at
right angles on building surfaces. The external and internal pressure coefficients (cpe and
cpi) are positively defined, provided the wind pressure acts in the direction of the surface,
whereas a negative value means suction on the member surfaces. The impact of wind
direction can be differentiated using coefficients into two separate directions, to be
examined with the gable ( = 90°) or the longitudinal building ( = 0 or 180°) on the
windward side. The external pressure coefficients correlate strongly with the building
shape, while wind tunnel tests have also shown higher wind suction forces in corner
areas of the building (Lusch, 1964). These observations spawned the distributions shown
in Figure C2‐3 and Figure C2‐4. As an example, Figure C2‐3 specifies external pressure
coefficients for a simple rectangular building. The values correspond to the upper value
for all wind directions acting in a range of ± 45° from the normal to the surface under
consideration. Figure C2‐3 shows the coefficient cpe,10 for wall surfaces exceeding 10 m²
and building dimensions of h/d( = 0°) or h/L( = 90°) ≤ 0.25. These pressure distributions
are based on the dimensional area facing the wind e = min (b, 2∙h). Higher pressure coef‐
ficients apply for smaller wall areas.
As well as the forces on walls, special consideration of wind forces on roofs is also re‐
quired, since uplift wind forces may have a decisive impact when it comes to designing
joints and members. For flat roofs, Figure C2‐4 shows the external pressure coefficients
for wind directions of = 0 or 90°with a roof area of A ≥ 10 m2. The specified pressure
coefficients must be increased if the roof areas are smaller. If sloping roofs facing the wind
are involved, compressive or suction forces are exerted at roof pitches of between 15°
and 30° and the less favourable effects need to be taken into account during design.
d
1e 4e
5 5
-1.2
b -0.8
-0.3 -0.5
-0.5 -0.5
0,7 -0.5
-0.8
Wind -1.2
T = 0°
Figure C2‐3 Pressure coefficients cpe,10 according to NA for vertical walls, h/d ≤ 0.25, A ≥ 10 m2.
153
Actions on structures
0.2 / -0.6
d
-0.7 1e
2
-1.8 -1.2 -1.8
1e 1e 1e
Wind 10
4 4
Figure C2‐4 Pressure coefficients cpe,10 according to NA for flat roofs with sharp‐edged eaves area, A ≥ 10 m2.
Design wind loads
When designing buildings, the effects of wind loads are generally determined using the
wind pressure distribution on the surfaces, with the result determined as a combination
of external (we) and internal (wi) pressure as follows:
w e q p z e c pe w i q p z i c pi (C2‐3)
Since wind pressure rises with increasing height above ground, structural engineers have
to consider the reference height ze of the external surfaces. Depending on the building
shape and dimensions of the width b on the windward side Figure C2‐5 reveals reference
heights for walls and roofs. Here, zi is the reference height of the walls for closed build‐
ings or the average height of the openings.
Alternatively, the wind pressure w can also be determined using aerodynamic force coef‐
ficients cf instead of aerodynamic pressure coefficients cpe/i. This is particularly important
for constructions with linear structural members such as truss bridges, since only force
and not pressure coefficients are given for trusses. For trusses, the fullness factor is
important and defined as the ratio of the projected surface of the members relative to
the overall area (product of length and width) of the truss. The fullness factor hence
takes into consideration the fact that a truss has no closed surface, but is predominantly
open, depending on the cross‐sectional dimensions of the members (a completely closed
wall has a fullness factor of 1). The force coefficient for trusses is determined based on
said fullness factor. The wind force exerted on a structure can thus be determined by
taking into consideration the tributary area or the projected reference surface of a build‐
ing or of individual members. For larger areas exposed to the wind (e.g. free‐standing
roofs), a frictional force Ffr must also be taken into consideration.
154
Actions on structures
b < h ≤ 2b b
ze = h qp(z) = qp(h)
h-b
ze = b
qp(z) = qp(b)
h
b
h ≥ 2b b
ze = h qp(z) = qp(h)
ze = zi
qp(z) = qp(zi)
h hi
ze = b qp(z) = qp(b)
Figure C2‐5 Reference height ze depending on h and b and wind pressure distribution in accordance with EC 1.
155
Actions on structures
C2.6 Combinations of actions
After the actions have been determined, the next design task is to analyse their effects,
which also involves selecting realistic load configurations for which the structure or struc‐
tural members are designed. The design values are then determined from the following
load combinations. In the ultimate limit state, the fundamental combination for persis‐
tent and transient situations is:
G,j is the partial safety factor for permanent actions (see Article C1) and Qk,1 is the lead‐
ing variable action.
In the serviceability limit state, the combination depends on the action effects, meaning
two combinations must be noted:
Characteristic combination
Quasi‐permanent combination
Certain numeric values for the combination factors for buildings are specified in
Table C1‐2.
In timber structures, structural engineers must focus particularly on determining the
critical load cases, since they depend on the modification factors kmod for the load dura‐
tion and the moisture content and thus are values that are not independent of the
strengths of construction materials. For all verifications in the ultimate limit state, the
resistance must be determined taking into consideration the coefficient kmod, whereby
the selection of kmod is made in accordance with the leading variable action Qk,1. In the
event of snow or wind as leading variable action, kmod according to the short‐term load‐
duration class must be used (see Table C1‐4), while for other imposed loads the load‐
duration class is “medium‐term”, meaning kmod is smaller. Conversely, for the load com‐
bination without variable actions considering only self‐weight, kmod must correspond to
the “permanent” load‐duration class to determine the design resistance.
156
Actions on structures
While respecting the various limit states, the combination of actions for each critical load
case is calculated, taking into consideration the various values for kmod. During this, the
structural engineer must also take into consideration any unfavourable load configura‐
tions. The uniformly distributed loads are generally governing when designing most struc‐
tural members, although it is non‐uniformly distributed loads that tend to have critical
impacts on joints or bracing systems.
C2.7 Literature
P. Racher, original Article A3, STEP 1995.
C.E.B. (1980). Structural safety. Bulletins d'information Nos. 127 and 128, Brussels.
E.C.C.S. (1987). Recommendations for calculating the effects of wind on constructions.
European Convention for Constructional Steelwork, Technical committee 12, Report No. 52, Brussels.
Hendrickson E.M., Ellingwood B. and Murphy J. (1987). Limit state probabilities for wood structural members.
ASCE Journal of Structural Engineering, 113(1): 88‐106.
Lusch G. (1964). Wind tunnel investigations on buildings with rectangular base and with flat
and duo‐pitched roofs. Bauforschung No. 41.
Rackwitz R. (1976). Practical probabilistic approach to design. C.E.B., Bulletin d'information No. 112, Brussels.
157
D
Structural members and systems
D1 Basic stresses
Original articles: B. Edlund, B. S. Choo, P. Aune
D1.1 Tension and compression
Wood is an anisotropic material, which means it demonstrates different properties when
stress is applied in different directions, e.g. parallel or at right angles to the grain. The
idealised form of a tree stem is deemed to be cylindrical and orthotropic, namely an
orthogonal anisotropic structure, see Article B1. Axes L, R and T reveal the longitudinal,
radial and tangential directions respectively, while the properties in R and T directions
are often cumulatively expressed as properties perpendicular to the grain with the Index
“90” (this is possible, since EL >> ET ≈ ER. Moreover, the type of cut and hence positioning
of the R and T directions of the timber members is normally unknown). The longitudinal
direction is specified with the Index “0”. The following sections mainly cover strength and
stiffness properties under short‐term loading. Unless otherwise specified, the specified
values apply for European softwood with a moisture content of 10 to 15%.
Timber is non‐homogenous based on its growth irregularities and dimensions for load‐
bearing purposes. The material properties (density, strength, modulus of elasticity (MOE)
etc.) vary widely, even within a single log cross‐section, as well as throughout the log
length. Properties also vary among individual trees of the same wood species and obvi‐
ously among individual species. Even within a single annual ring, properties vary since
early‐ and latewood have significantly different properties. This scatter is not covered in
further detail in the current article.
Initially, the properties of clear wood with small dimensions are addressed, before mov‐
ing on to structural timber with growth irregularities and structural timber dimensions,
since only the latter information is relevant for timber structures. For timber which has
been dried and strength graded, (Article B5) characteristic properties are established,
with which the member resistances are determined. In the ultimate limit state, the par‐
tial safety coefficient m and modification coefficient kmod must be used to determine the
design values of the member resistance, while in the serviceability limit state, creep must
be taken into consideration applying the coefficient kdef (see Articles C1 and C2).
161
Basic stresses
Clear wood exposed to tension and compression
Tensile stress
If small wood samples from material which is as flawless as possible are exposed to load,
stress‐strain curves can be derived, as shown in Figure D1‐1. The tensile strength in the
grain direction ft,0 (Index “t” = tension) accordingly exceeds the compressive strength in
the grain direction fc,0 (Index “c” = compression). The stress‐strain curve when tension is
applied runs virtually up to the linear failure point, which occurs abruptly and is designat‐
ed as brittle fracture. When compression is applied, instead, plastic behaviour is evident
when failure occurs. Wood is at its weakest when tensile stress is applied at right angles
to the grain. This strength ft,90 is in the region of 1 to 2 N/mm2 and depends very strongly
on the volume involved, since tensile stress is strongly identified as a cause of brittle
failure. This volume effect is less evident for other strength properties (for more on the
volume effect, see Article D3). The tensile strength ft,90 is significantly reduced when
there are pre‐existing cracks, particularly in earlywood. Likewise, the stiffness is also far
lower with a modulus of elasticity of E90 = 400 to 500 N/mm2 perpendicular to the grain
than the same figure parallel to the grain at E0 = 11000 to 15000 N/mm2.
Timber constructions should generally be designed such as to ensure that tensile stresses
perpendicular to the grain are avoided or minimised as far as possible. Knowing which
areas of a construction are exposed to tensile stresses perpendicular to the grain is im‐
portant, to instigate the required measures to reduce such stresses, such as reinforce‐
ments. Constructions exposed to such tensile stresses include curved beams, corner
areas of portal frames, notched beam supports and beams with openings.
ft,0
ft,90 ε
fc,90
fc,0
Figure D1‐1 Stress‐strain‐curve of clear wood exposed to tensile (t) and compressive stresses (c) parallel to
the grain (solid line) and perpendicular to the grain (dashed line) at constant strain increase.
Typical strength values for softwood are: ft,0 = 80 to 100 MPa, fc,0 = 40 to 50 MPa, E0 = 11000 to
15000 MPa, E90 = 400 to 500 MPa. (STEP 1995 arcticle B2)
162
Basic stresses
Compressive stress
The curve progression in Figure D1‐1 reveals that when compressive stress is applied in
the grain direction, wood fibres will successively yield until the load peaks, whereupon
the specimen fails, with localised buckling of wood fibres (see Figure D1‐2). This is known
as a local stability failure caused by shear along a sloping surface. As with tensile stress,
the modulus of elasticity Ec,0 is between 11000 and 15000 N/mm2. However, since the
stress‐strain curve levels off earlier than with tensile stress, the proportional limit must
be noted.
LW EW AR
CC
50-65˚
T
R
Figure D1‐2 Compressive failure of compression test specimen with fc,0 due to localised buckling of wood
fibres, after Hoffmeyer (1990). LW latewood, EW earlywood, AR annual rings, CC compression
creases. (STEP 1995 Article B2)
If the wood sample is completely exposed to compression perpendicular to the grain (as
in case a in Figure D1‐3), the wood fibres are practically squeezed, as in a bundle of
tubes, until a type of squash load is attained, at which the tangent modulus (inclination of
the tangent on the ‐‐line) becomes very small. Under the maximum load, meanwhile,
the strain levels become very high.
If only part of the area on the upper side of the wood sample is exposed to load, the
stiffness value for the wood will be higher compared to case a in Figure D1‐3 (cases b to e)
and will only start declining at higher stresses than in case a. The transition from the
initial stiffness to the plastic region is less pronounced in this case than in case a, because
the load is also distributed on unloaded areas outside the part directly exposed to loads.
In case b, the areas not directly exposed to loads are too small to ensure effective load
distribution. The maximum load is attained immediately after traversing the transition
area (from elastic to plastic). In cases c, d and e, meanwhile, the compressive load can
still be absorbed, even at strain levels exceeding those shown in Figure D1‐3, without
any sign of failure occurring. Nevertheless, the deformation level rises considerably in
this case, which means that for constructive reasons, it makes sense to limit the strain to
a specific value, e.g. 1% of the cross‐sectional height or 5 mm and to view the related
stress as a type of strength (or “proof stress”). In this case, the figure determined is
fc,90 = 2 to 4 N/mm2. However, the strengths defined in this manner depend on the posi‐
tion of the annual ring in the cross‐section, see Figure D1‐4.
163
Basic stresses
σc (N/mm2)
e
8 d
c
6
b
4
b,c,d a
0
0 5 10 15 20
εc (%)
Figure D1‐3 Compressive stresses c acting on the upper side of the wood samples 150 mm x 150 mm at right
angles to the grain depending on the strain levels. (according to Suenson, 1938, STEP 1995 Article
B2) The right sketch shows the types of failure or the involvement of the wood fibres adjacent to
the area exposed to load and the resulting increase in load‐bearing capacity.
E90 (kN/mm2)
f p (N/mm2)
0,8
E90
0,6
5
fp
0,4 4
fp
3
fp
0,2 2
1
0,0 0
R D T
Figure D1‐4 Modulus of elasticity and stress limit of wood under compression perpendicular to the grain. fc,90
is defined in this case as stress at the proportional limit. (according to Siimes and Liiri, 1952, STEP
1995 Article B2)
However, both compressive and tensile stresses can also occur under an angle to the
grain, whereby the angle between the load and grain directions is termed angle . Fig‐
ure D1‐5 shows this kind of stress. Hankinson (1921) proposed the following equation
(D1‐1) (Figure D1‐5 (b)) for the strength values fc, assuming linear interaction (see also
Figure D1‐7), which shows effective compliance with test results. (For the derivation of
the Hankinson equation, see Annex 2.)
164
Basic stresses
f c,0 f c,90
f c,α (D1‐1)
f c,0 sin f c,90 cos
2 2
To determine the tension strength at an angle to the grain, a corresponding expression
is used, when replacing fc with ft, see Figure D1‐5 (a).
For small angles the strength significantly depends on changes to the angle. Small
changes in the slope of grain result in a considerable change in strength, particularly
tensile strength. Conversely, in the area of ≈ 90°, only slight influences on ft, and fc,
are apparent (low gradient of the curves in Figure D1‐5 (a) and (b)). For uniaxial tension
under an angle to the 1‐axis, in accordance with Figure D1‐5 (c) and (d), the following
equilibrium conditions are valid (see also Annex 2, equations (II), (III) and (IV)):
1 α cos 2
2 α sin 2
12 α sin cos
For comparison, the limit curves for the three separate failure conditions 1 ≤ fc,0,
2 ≤ fc,90, 12 ≤ fv are set out in Figure D1‐5 (b) (dashed lines).
fc,0/cos2α
fc
ft
ft,0
fc,0 fv=fc,90
fv/sinα cosα
fv=0,5fc,90
fc,90/sin2α
ft,90 0,2fc,0
Hankinson
0 0
0˚ 90˚ α 0˚ 15˚ 30˚ 45˚ 60˚ 75˚ 90˚ α
(a) (b)
α 1 2
σ2 σc,α
σt,α
τ12 0 < α < 90°
α
(c) (d)
Figure D1‐5 Tensile strengths (a) and compressive strengths (b) depending on the angle . The Hankinson
equation (D1‐1) is shown as a solid line. The dashed lines indicate failure at stresses fc,0, fc,90 or fv.
(STEP 1995 Article B2)
165
Basic stresses
Structural timber exposed to tension and compression
When using timber in dimensions used in load‐bearing constructions, it is important to
ensure the impact of various unavoidable growth irregularities such as knots and slope of
grain are taken into consideration. A knot of “normal” size reduces the effective cross‐
section of a member and results in localised fibre disturbances, which often, in turn,
trigger eccentricity of loads and high localised stresses. In addition, in uniaxially loaded
sawn timber, stresses perpendicular to the grain are induced in the area where the fibres
change directions around a knot. This is particularly significant for wood subject to tensile
stress parallel to the grain, since this means that stresses perpendicular to the grain are
also produced.
Although in clear wood, the tensile strength parallel to the grain far exceeds the com‐
pressive strength, the opposite applies for structural timber. This is partly due to the
above‐mentioned dependency on the slope of grain (see Figure D1‐5 (a)) and partly due
to the brittle fracture behaviour and volume effect (Article D3), which also encompasses
all other influences. Given the significant roles played by the load duration (creep) and
the moisture content, reference is made to Article B2.
Tensile stress
Inhomogeneities and other deviations from an ideal orthotropic material, as typically
occur in structural timber, are often labelled as growth irregularities. As explained previ‐
ously, these influences result in a clear reduction in tensile strength parallel to the grain.
For Northern European softwood (spruce, fir), average values for ft,0 are in the region of
10 to 35 N/mm2, a significant reduction compared to the value cited in Figure D1‐1 of ft,0
from 80 to 100 N/mm2 for clear wood. Various investigations showed the mean value of
ft,0 declining proportionally to the increasing diameter of the largest knot. However, the
scatter is high and the correlation weak. The obtained values also depend on the test
method, since failure may be induced by stress concentrations at the grip end devices.
Moreover, the tensile strengths depend on the tested volume, which is why a height or
width correction factor kh is used in EC 5 (Article D3, Section D3.3).
Tension at an angle to the grain
EC 5 does not specify any means of verifying tensile capacity perpendicular to the grain
( = 90°), except that the member size has to be taken into consideration. Tensile stress‐
es perpendicular to the grain tend to occur in curved beams, notched beams or joints
loaded perpendicular to the grain and are explained in the relevant articles (see Articles
D4, D5, D8 and E11). EC 5 also excludes details on tension at an angle to the grain, alt‐
hough the NA proposes a means of verifying this type of stress. Stresses at an angle to
the grain result in combined stress conditions, under which both tensile stresses parallel
and perpendicular to the grain and shear stresses occur, which is why the Hankinson
166
Basic stresses
equation (D1‐1) was expanded to encompass shear strength (derivation see Annex 2) and
is then used to convert longitudinal tensile strength to tensile strength at an angle to
the grain:
1
t,α,d k α f t,0,d f t,0,d (D1‐2)
f t,0,d f
sin
2 t,0,d
sin cos cos
2
f t,90,d f v,d
To verify tension in the grain direction, EC 5 includes a simple normal stress criterion where
the design value of tensile stress t,0,d must be smaller than that of tensile strength ft,0,d.
Compressive stress
Unlike tensile strength, compressive strength parallel to the grain is only moderately
reduced by growth irregularities and averages from fc,0 = 25 to 40 N/mm2. Similarly to
when verifying tensile strength in the grain direction, compression in the grain direction
is checked via a simple normal stress criterion:
Compression perpendicular to the grain
When verifying compression perpendicular to the grain, the design value of strength is
increased by a factor of kc,90:
kc,90 takes into consideration the type of effect, the splitting risk and the extent of the
deformation. Although the value for kc,90 is generally 1.0, under specific support condi‐
tions, the design value for compression strength perpendicular to the grain fc,90,d may be
increased (the reason for this increase in compression strength depending on the sup‐
port type is clarified from Figure D1‐3):
kc,90 = 1.25 for solid timber or 1.5 for glulam with continuous support
if ℓ1 ≥ 2 ∙ h, see Figure D1‐6 (a) or
kc,90 = 1.5 for solid timber or 1.75 for glulam with discrete support,
if ℓ1 ≥ 2 ∙ h, see Figure D1‐6 (b).
167
Basic stresses
Moreover, the design value for compressive stress perpendicular to the grain c,90,d is
determined with an effective contact surface Aef, which takes a possible load distribution
into consideration, see equation (D1‐5). The actual contact length ℓ may be increased on
both sides in the grain direction by 30 mm, but by a maximum of a, ℓ or ℓ1/2.
This increase in contact length to an effective value of ℓef takes into consideration the
fact that compressive stress perpendicular to the grain spreads out in the grain direction
to extend to areas not directly subjected to stress (see also Figure D1‐3).
F c,90,d
c,90,d (D1‐5)
A ef
a ℓ ℓ1 ℓ1 ℓ
h
b
a ℓ
(a) (b)
Figure D1‐6 Member subject to (a) continuous support and (b) discrete support, distance ℓ1 of continuous or
discrete load from the support.
Compression at an angle to the grain
When verifying compression at an angle to the grain, equation (D1‐6) applies, which
corresponds to the Hankinson equation (D1‐1), although in this case, coefficient kc,90 is
also considered:
f c,0,d
c,α,d (D1‐6)
f c,0,d
sin 2 cos 2
k c,90 f c,90,d
Shear strength fv,d is not used in equation (D1‐6), but is used in equation (D1‐2) to verify
tensile stresses at an angle to the grain.
The verifications for compression stress cited here only apply to members not at risk of
buckling (see Article D2).
168
Basic stresses
D1.2 Bending
Beams often are horizontal structural members, which are supported on both sides and
which transmit loads mainly through bending stresses. The bending moments in the
beams are generated by loads acting perpendicular to the beam axis. When assessing a
timber beam, it is important to ensure the following in particular:
The design value of bending strength is not attained, the bending stresses do
not result in any lateral torsional buckling of the beam and thus do not lead
to a premature stability failure,
The design value of shear strength is not attained,
The design value of compressive strength perpendicular to the grain at the
supports and under concentrated loads is not attained (equation (D1‐4)),
The deflection of the beam does not exceed the limit values recommended
in EC 5,
The vibration behaviour does not prove problematic.
This chapter focuses particularly on simple beams; namely straight beams of constant
depth and without notches. The loss of strength in curved and tapered beams (Article D4)
and the impact of notches (Article D5) are covered in additional articles. The bending
stress is to be verified in the critical cross‐section (e.g. square, T‐ or L‐shaped), which, for
simple beams in accordance with the above definitions, is usually at the point of maxi‐
mum bending moment of the beam. EC 5 also requires that the influence of initial de‐
formation, eccentricity and induced deflections be taken into consideration. The stability
problem of lateral torsional buckling is covered in Article D2.
Beams not at risk of buckling
If the dimensions and support conditions of the beam are adequate to prevent instability,
then the bending stresses according to elastic beam theory are as follows:
My z
(D1‐7)
Iy
where
My Bending moment about the y‐axis
Iy Moment of inertia (second moment of area) about the y‐axis
z Distance from the neutral axis
Stress at distance z
169
Basic stresses
This equation for bending stress in the beam is generally applicable, if the cross‐section
only needs to withstand bending about its minor principal axis or, when bent about its
major principal axis, lateral support at narrow distances is provided so that the slender‐
ness for bending is low.
Since EC 5 allows for timber structures to be designed while assuming elastic material
behaviour, equation (D1‐7) can be used for the design. The design value of bending
strength fm,d is determined in this process by taking the partial safety coefficient M and
the modification coefficient kmod (see also Article C1) into consideration. In addition,
further factors, which impact on the bending strength, should be taken into considera‐
tion. The influence of member volume is considered e.g. by the coefficient kh (see Section
D1.1) and for beams of parallel structural systems including load distribution systems, the
bending strength may be increased with coefficient ksys (see Article D11). Accordingly, the
verification of bending stresses is a simple normal stress criterion:
Stress combinations
Equation (D1‐8) only considers beams, which are only stressed about one principal axis
by bending moments. In addition, stress combinations may also result due to
Biaxial bending (bending about both principal axes),
Uniaxial or biaxial bending and simultaneous axial tensile or compressive load.
Beams with biaxial bending must meet the following conditions:
m,y,d m,z,d
km 1 (D1‐9)
f m,y,d f m,z,d
m,y,d m,z,d
km 1 (D1‐10)
f m,y,d f m,z,d
where
m,y(z),d Design bending stress due to moment about the y(z)‐axis
fm,y(z),d Design bending strength for bending about the y(z)‐axis
km Coefficient, which takes into account that the bending capacity of the beam has
not yet been attained, if, pursuant to the Euler‐Bernoulli beam theory, the stress‐
es calculated at a cross‐sectional corner attain the bending strength. km is as‐
sumed for rectangular cross‐sections made of solid timber, glulam and LVL to
km = 0.7; otherwise km = 1.0 applies. It is advisable to only use km for compact
cross‐sections where h/b ≤ 4.
170
Basic stresses
The verifications for the combinations of bending and axial tension are:
For bending and axial compression:
2
c,0,d m,y,d
k m m,z,d 1 (D1‐13)
f c,0,d f m,y,d f m,z,d
2
c,0,d m,y,d m,z,d
k m 1 (D1‐14)
f c,0,d f m,y,d f m,z,d
The combinations of “biaxial bending” and “bending and tension” are linearly super‐
posed, while the compressive normal stresses are quadratically included when designing
for “bending and compression”. This is due to the plastic and thus favourable response of
wood when exposed to compressive stresses. Moreover, the tension zone of a beam
exposed to axial compressive stress attains lower resulting stresses, meaning it only
attains bending strength at larger bending moments. Figure D1‐7 clearly illustrates the
significance of linear and quadratic stress superposition (see also Figure D2‐5, which
shows bending moment and normal force interacting). Whereas a linear superposition
results in strengths declining linearly in a main direction, when a stress in the other main
direction is applied, quadratic superposition allows a far slower reduction in strength for
combined stresses. This is the case as explained, particularly for ductile types of failure.
V2
2 2
Quadratic superposition: 1 2
f2 f f 1
Linear superposition: 1 2
1 2
1 V1
f1 f2 f1
Figure D1‐7 Graphic representation of linear and quadratic stress superposition.
171
Basic stresses
D1.3 Shear and torsion
In accordance with the theory of elasticity, shear stresses are generated when bending is
produced by loads perpendicular to the beam axis. Shear stresses perpendicular to beam
axis and fibre direction always occur simultaneously with shear stresses of the same size,
which act parallel to the beam axis and fibre direction. In solid timber and glued laminat‐
ed timber, the shear strength parallel to the grain is considerably lower than that per‐
pendicular to the grain, in which case fibres would have to be severed in the event of a
shear fracture. Accordingly, shear parallel to the grain (longitudinal shear) is decisive for
design. This article only covers longitudinal shear in solid timber and glulam members,
while rolling shear, particularly relevant for cross‐laminated timber members, is covered
in Article D7.
Torsional stresses occur when a member is twisted by an external load. This can be
caused by e.g. eccentric forces applied perpendicular to the member axis. One example is
beams, which are curved due to geometric imperfections in the ground view and thus
subject to torsional stresses.
Shear
In accordance with elastic beam theory (Euler‐Bernoulli), shear stress can be specified
at any point in the cross‐section in generic form by:
V S
(D1‐15)
I b
where V is the shear force, I the moment of inertia for the entire cross‐section, b the
cross‐sectional width of the particular cross‐sectional part at which is determined and S
the first moment of area about the neutral axis of the considered cross‐sectional part.
For a rectangular cross‐section, the maximum value amounts to:
3 V
(D1‐16)
2 A
The shear stress distribution runs over the whole height of the cross‐section, in a para‐
bolic shape, with the figure peaking in the neutral fibre as shown in Figure D1‐8.
172
Basic stresses
b τ
Figure D1‐8 Shear stress distribution. (STEP 1995 Article B4)
In equation (D1‐16), A indicates the cross‐sectional area. The negative influence of cracks
on the shear strength of beams must be taken into consideration, which is why area A is
determined with an effective beam width bef:
b ef k cr b (D1‐17)
The factor kcr (cr = crack) is assumed in the NA for solid timber as kcr = 2.0/fv,k and for
glulam as kcr = 2.5/fv,k. The perpendicular layers included in cross‐laminated timber mean
that kcr need not be taken into consideration and is assumed to constitute 1.0.
The factor kcr was introduced since the characteristic shear strength values fv,k, which can
be taken from the corresponding applicable product standards (e.g. EN 338 for solid
timber), were determined at uncracked cross‐sections. However, from a practical con‐
struction perspective, this hardly ever happens, e.g. the formation of shrinkage cracks is
inevitable, due to variable ambient moisture over the lifetime of a solid timber or glulam
member. The specified values for fv,k were lower in earlier editions of product standards.
For C24, fv,k in the EN 338 2003 edition amounted to 2.5 N/mm2, while EN 338 in 2009
specified an fv,k value of 4.0 N/mm2 and these changed nominal strength values are taken
into account with kcr. In accordance with the current regulation in EC 5, kcr must always
be taken into consideration for shear stresses of beams or if shear stresses are exerted
in parallel to possible crack planes. As well as for beams, kcr must also be applied e.g. to
verify the loaded end distance of single step joints or block shear in joints. The question
remains, however, how to handle other verifications, which despite requiring shear
strength, do not stipulate the use of kcr. Examples here include notches or verifications
with stress interaction as in equation (D1‐2).
173
Basic stresses
Various researchers (e.g. Keenan, 1978) proved that shear stresses produced by concen‐
trated loads near supports are smaller than values determined in accordance with elastic
beam theory resp. that compressive stresses perpendicular to the grain generated by the
reaction forces at supports actually increase shear strength, which is why reduced shear
forces are used in EC 5. Beams which are supported at the lower edge and with the upper
edge exposed to load, can be verified using a reduced shear force or the proportion of a
concentrated load, applied near the support and on the upper side of the beam, on the
total shear force may be disregarded. Accordingly, instead of increasing shear strength
when compressive stress perpendicluar to the grain is present, as is done e.g. in the
interaction equation for compressive stresses of tapers (equation (D4‐7), see also Annex 2),
the shear force is reduced in this case.
For biaxial bending, the shear stresses must be quadratically superposed according to the
NA to EC 5. Here too, the factor kcr is applied when determining shear stresses in parallel
to possible crack planes:
2 2
y,d z,d
1 (D1‐18)
f v,d f v,d
Torsion
In accordance with the Euler‐Bernoulli beam theory, the maximum torsional stresses tor
caused by a torsional moment MT can be calculated for rectangular cross‐sections as:
MT MT
tor (D1‐19)
WT hb 2
where h ≥ b and is a factor depending on the ratio h/b. Timoshenko (1955) specifies
the following values for :
Table D1‐1 Values for .
h/b 1.00 1.50 1.75 2.00 2.50 3.00 4.00 6.00 8.00 10.00
0.208 0.231 0.239 0.246 0.258 0.267 0.282 0.299 0.307 0.313 0.333
The distribution of torsional stresses along the principal axes of rectangular cross‐
sections is shown schematically in Figure D1‐9, where the stress peaks in the middle of
the longest side.
174
Basic stresses
Figure D1‐9 Distribution of torsional stresses. (STEP 1995 Article B4)
Torsional stress should meet the following condition:
Combination of shear and torsion
In some cases, both shear and torsion are exerted on a cross‐section, although there
have been few investigations into this phenomenon and knowledge in this area remains
limited. EC 5 does not provide any pointers for this combined stress state, but the NA
includes the following design equation:
2 2
tor,d y,d z,d
1 (D1‐21)
k shape f v,d f
v,d f v,d
Here, the factor kcr is to be applied to determine the shear stresses y,d and z,d.
175
Basic stresses
D1.4 Literature
B. Edlund, B.S. Choo, P. Aune, original Articles B2, B3, B4, STEP 1995.
Hankinson R.L. (1921). Investigation of crushing strength of spruce at varying angles of grain.
Air Service In‐form Circular III, No. 259, US Air Service, Washington DC.
Hoffmeyer P. (1990). Failure of wood as influenced by moisture and duration of load. Dissertation,
State University of New York.
Keenan F.J. (1978). The distribution of shear stresses in timber beams. Paper 9‐10‐1, CIB‐W18 Meeting 9, Perth.
Siimes F. and Liiri O. (1952). Investigations of the strength properties of wood I. Tests on small clear specimens
of Finnish Pine (Pinus Sylvestris). (In Finnish). Valtion Teknillinen Tutkimuslaitos, Tiedotus 103, Helsinki.
Suenson E. (1938). Zulässiger Druck auf Querholz. Holz als Roh‐ und Werkstoff 1(6): 213‐216.
Timoshenko S. (1955). Strength of materials ‒ Part 1. D. Van Nostrand, New Jersey, Third Edition.
176
D2 Stability
Original articles: B. S. Choo, H. J. Blass
This chapter shows stability analyses and their backgrounds, starting with the buckling of
columns, before covering the way buckling lengths are derived. In addition to compres‐
sive and bending stress imposed on columns, compressive and bending stresses on
beams also lead to stability problems. This is called lateral torsional buckling and is sub‐
sequently covered. It is important to note that stability analyses are performed with
the lower quantile values for stiffness and not the average values.
D2.1 Columns
If a slender column is axially loaded, there is a risk that it deflects laterally which is called
flexural buckling and is shown in Figure D2‐1. The load‐bearing capacity of a slender
structural member when exposed to compressive stress thus depends not only on the
strength of the construction material itself, but also particularly on its bending stiffness.
The key material properties determining the load‐bearing capacity of a wooden column
are thus the compressive and bending strength and the modulus of elasticity (MOE).
Since lateral deflection generates additional bending stresses in the column, these must
also be taken into consideration in a stability design.
N Nu
Figure D2‐1 Column hinged at both ends, EI = const. (STEP 1995 Article B6)
177
Stability
Two different design methods are distinguished for columns: the first of which comprises
a calculation of the column in accordance with second order theory, namely the equilib‐
rium of moments and forces is determined on the deformed system. The second method
uses buckling curves to take the lower load‐bearing capacity of an actual column into
consideration compared to a column of unlimited rigidity, which means stability design is
performed via compressive stress analysis with reduced compressive strength. The ex‐
tent to which the load‐bearing capacity declines particularly depends on the slenderness
of the corresponding member and was derived based on a column hinged at both ends
(Figure D2‐1). For members supported otherwise or structural systems such as portal
frames, this method can be used, providing the buckling length has been determined
beforehand (see Section D2.2) whereupon the analysis is performed as for a column
hinged at both ends of the same length. In this context, only the column design based on
buckling curves is covered.
Variables affecting the load‐bearing capacity of columns
The key variables affecting the load‐bearing capacity of wooden columns can be classified
into two groups. The first group comprises the column dimensions, the support condi‐
tions and the construction material properties, which, in turn, are dictated by the selec‐
tion of a strength class, the prevailing climate (→ service class) and the load duration
class of the governing load case. The parameters included in this first group are either
determined by the structural engineer or at least known to the same and the engineer
can determine the load‐bearing capacity of the column corresponding to the design
requirements by changing these parameters. A second group of variables includes geo‐
metric and structural imperfections and their variations. Given the fact that actual struc‐
tures are never perfect, these parameters must also be taken into consideration during
the design. Since structural engineers generally lack any information about the extent of
such imperfections for a given structure, their influence is already included in the design
equations for columns in EC 5.
The key geometrical imperfections of wooden columns are initial curvature, inclination of
members and tendency for actual cross‐sectional values to deviate from nominal values.
The extent of initial curvature for columns made of glulam or LVL is limited to 1/500,
while the figure for columns made of solid timber is limited to 1/300 of the length (EC 5
Section 10.2). The buckling coefficients were determined within these limits, whereby in
actual constructions, eccentricity was determined, with its statistic distribution used as
input. Deviations of the nominal cross‐sectional dimensions are limited for solid timber
members to tolerance class 1 in EN 336 and for glulam members to the values in EN 390.
Members with higher moisture content or within which moisture content varies (see
Article B2, mechano‐sorptive creep) show a clearly increased level of creep deformation
178
Stability
(see also Figure B2‐8) and hence larger deformations over time, hence for columns larger
eccentricities e. Accordingly, the NA specifies that creep should also be taken into con‐
sideration in the ultimate limit state, if members in service classes 2 and 3 are subject to
compression through high permanent loads.
Structural imperfections include growth irregularities and other properties like moisture
content, which collectively determine the stress‐strain behaviour of timber. As a general
rule, the stress‐strain curve of timber can be considered brittle when tensile and shear
stresses are applied, while compressive stresses trigger considerable plastic deformations
(see also Figure D1‐1). For European conifers, the shape of the stress‐strain curve de‐
pends particularly on the following properties (Glos, 1978): density, knot size, compres‐
sion wood proportion and moisture content. Glos (1978) put forward mathematical
relations between these properties and the shape of the stress‐strain curve, for both
laminations for glued laminated timber as well as solid timber cross‐sections. Provided
parameters such as density, knot size, compression wood proportion and moisture con‐
tent are known, the expected shape of the stress‐strain curve can be calculated via the
cited relationships.
Background to the buckling curves in EC 5
In general, buckling curves reveal the impact of slenderness on the characteristic load‐
bearing capacity of columns hinged at both ends. In this context, each value of a buckling
curve corresponds to the characteristic load‐bearing capacity of columns of the corre‐
sponding slenderness ratio. The slenderness ratio is defined as the ratio of the buckling
length to the radius of gyration and various approaches are available to determine the
characteristic load‐bearing capacity values of columns. One basic option here is to con‐
duct load‐bearing tests with representative selected columns, but the excessive number
of tests required and resulting high cost makes this method unfeasible, which is why an
alternative approach based on simulations of tests via computer, was used to derive the
buckling curves in EC 5 (Blass, 1987; Blass, 1988a; Blass, 1988b). It involves modelling
columns by assigning them material properties and geometric imperfections that are
based on observations on real columns. This means that the strength and stiffness values
and the extent of initial curvature or deviations of the nominal cross‐sectional dimen‐
sions for a specific column are selected at random. Of course, the assigned properties
must be realistic, which also means that the existing correlations between the various
assigned properties must be taken into account in the simulation process. This also
means, for example, that columns with a high modulus of elasticity also tend to show
high strength values. Just like an actual column, the simulated column showcases a set of
properties, which determine its load‐bearing capacity.
179
Stability
The statistic distribution of the load‐bearing capacity of columns with a specific slender‐
ness ratio or strength class is determined by simulating numerous columns and calculat‐
ing their load‐bearing capacity. The load‐bearing capacity values vary due to the scatter
of strength and stiffness properties of the timber and the variance in geometrical imper‐
fections. The statistic distribution of load‐bearing capacity values for columns is used as
the basis with which to determine the 5% quantile as a characteristic value. This charac‐
teristic value corresponds to a point of the buckling curve (see Figure D2‐2). If corre‐
sponding simulations and load‐bearing capacity calculations are also performed for other
slenderness ratios, the resulting characteristic values reveal the course of the column
load‐bearing capacity depending on the slenderness ratio (buckling strength). Figure D2‐3
shows an example of the specific course of characteristic values for buckling strength
determined via simulations.
36
kc fc,0,d (N/mm2)
24
12
0
0 40 80 120 160 200
λ
Figure D2‐2 Distribution of the buckling strength and characteristic values for two different
slenderness ratios . The 5% quantiles are specified. (STEP 1995 Article B6)
36
kc fc,0,d (N/mm2)
24
12
0
0 40 80 120 160 200
λ
Figure D2‐3 Course of characteristic buckling strength values. (STEP 1995 Article B6)
180
Stability
Since an equation is easier to handle in practical design terms than a diagram, approxi‐
mation curves were adapted for the course of the characteristic load‐bearing capacities,
whereby the form of the design equations corresponds to that of the equations in EC 3
used to assess steel columns. Figure D2‐4 shows one example of the course of character‐
istic load‐bearing capacities with the corresponding approximation curve. The load‐
bearing capacity calculations for the simulated columns took place in accordance with
second order plasticity theory; taking the plastic deformation potential of the wood when
exposed to compressive stress into consideration. This approach spawned higher load‐
bearing capacities than when applying second order elasticity theory, where the load‐
bearing capacity is deemed already attained, when the compressive strength level of the
wood is attained at the edge of the critical cross‐section. However, when applying plasti‐
city theory, the numerous iterations mean the calculations take longer to perform. The
consideration of the plastic behaviour of the wood particularly increases the characteris‐
tic load‐bearing capacity under combined compressive and bending stress.
36
kc fc,0,d (N/mm2)
24
12
0
0 40 80 120 160 200
λ
Figure D2‐4 Approximation curve adapted to the course of characteristic values. (STEP 1995 Article B6)
Figure D2‐5 portrays a bending moment‐normal force interaction diagram for rectangular
cross‐sections. The straight line shows the load‐bearing capacity assuming elastic and
brittle behaviour, while the solid curve represents the characteristic load‐bearing capaci‐
ty, taking the plastic deformation capacity of the wood under compressive stress into
consideration. The dashed line is the design rule in accordance with EC 5 for a combined
stress comprising normal force and bending moment, when there is no risk of buckling or
if the internal forces and moments have already been determined in accordance with
second order theory (cf. equations (D1‐13) and (D1‐14)). For members at risk of buckling
due to combined compressive and bending stress, the course of the interaction relation‐
ship (solid curve in Figure D2‐5 for very compact members) gradually becomes a virtually
linear relationship for very slender members. When assessing compact columns with
slenderness ratios of up to around 20 (corresponding to rel = 0.3), the dashed line in
Figure D2‐5 applies, whereas the simple linear interaction applies in all other cases.
181
Stability
1,0
0,8
N/Nu
0,6
0,4
0,2
0
0 0,4 0,8 1,2
M/Mu
Figure D2‐5 Bending moment‐normal force interaction. (STEP 1995 Article B6)
Buckling curves in accordance with EC 5
The approximation curve shown in Figure D2‐4 elicits the factors kc,y and kc,z required for
the design, although they depend on slenderness, which is why limit values for the slen‐
derness ratios must first be determined. The relative slenderness ratios applied are de‐
fined by:
y f
rel,y c,0,k (D2‐1)
E 0,05
z f
rel,z c,0,k (D2‐2)
E 0,05
ef
(D2‐3)
i
182
Stability
2
c,0,d m,y,d
k m m,z,d 1 (D2‐4)
f c,0,d f m,y,d f m,z,d
2
c,0,d m,y,d m,z,d
k m 1 (D2‐5)
f c,0,d f m,y,d f m,z,d
In all remaining cases, namely for larger slenderness ratios, stresses should meet the
following conditions, whereby the reduced compressive strength now used is in accord‐
ance with Figure D2‐4 and linear interaction is assumed:
where:
m Bending stress in accordance with first order theory
1
k c,y/z
k y/z k y/z rel,y/z
2 2
k y/z
0.5 1 c rel,y/z 0.3 rel,y/z
2
c Coefficient for members, where the initial curvature does not exceed the above‐
specified limit values (columns made of glulam/LVL 1/500 and those made of
solid timber 1/300 of the length):
For solid timber: c = 0.2
For glulam and LVL: c = 0.1
The difference between solid timber and glulam or LVL is mainly due to the smaller degrees
of initial curvature in glulam and LVL members and the smaller deviations in cross‐sectional
dimensions from nominal dimensions. Moreover, in the case of columns made of glulam or
LVL, both the average value as well as the coefficient of variation of moisture content are
lower than with solid timber columns. Higher moisture content causes compressive
183
Stability
strength to decline, which then reduces the load‐bearing capacity of columns with low to
medium slenderness. In contrast, the modulus of elasticity and hence the load‐bearing
capacity of very slender columns is only influenced to a very minor extent by higher levels
of moisture content.
D2.2 Buckling lengths
Buckling curves used to design timber columns generally describe the course of the load‐
bearing capacity for columns with hinges at both ends depending on the slenderness. In
the case of the column shown in Figure D2‐1 the buckling length corresponds to the
actual length. However, the support conditions of actual columns frequently differ from
what is shown in Figure D2‐1. To continue using the buckling curves specified in EC 5 for
such cases, which involve deviating support conditions, the concept of an effective buck‐
ling length is used. Verifications performed in accordance with buckling curves are in‐
cluded in Section 6.3.2 of EC 5, while buckling length coefficients for cases frequently
occurring in practice are specified in NCI NA.13.
Example members in which the actual buckling length differs include e.g. truss webs. The
chords often have their outer edges braced to resist buckling. This means that the webs
are not laterally secured at the centre of gravity of the chords and the buckling length of
the webs approximately corresponds to the distance between the bracings of the upper
or lower chord and hence exceeds the distance between the theoretical truss nodes.
The buckling length or effective length ℓef of a member exposed to compressive loads is
defined as the length of a column hinged at both ends with the same stiffness properties
and elastic buckling load as the member in question. The buckling length can be shown as
the distance between two adjacent inflexion points in the deflection curve of the buckled
compression member (see Figure D2‐6). For practical applications, an effective length
factor = ℓef/L is often used, which indicates the ratio between buckling length ℓef and
the actual length L of the member. Figure D2‐7 shows the four Euler buckling cases,
whereby the buckling length ℓef is specified for the various ideal support conditions. In
Annex 3, the second Euler case (column hinged at both ends) is derived for example.
The NA, meanwhile, specifies approximate solutions for buckling lengths of various struc‐
tural systems. In cases in which the applicable limits of approximate solutions are ex‐
ceeded (e.g. for large initial curvatures), the corresponding structural system should be
calculated in accordance with second order theory (and not in accordance with buckling
curves). In other words, the equilibrium conditions should be met for the deformed system.
184
Stability
Ncrit
l
Kr
lef > 2l
Ncrit
M
Figure D2‐6 Buckling length ℓef of a clamped column with a semi‐rigid base connection. (STEP 1995 Article B7)
N N N N
√2
lef = l
2
lef = l
lef = l
l
l
l
lef = 2l
I II III IV
Figure D2‐7 Buckling length ℓef for various ideal support conditions (Euler cases I to IV). (STEP 1995 Article B7)
Influence of rotational stiffness in moment resisting joints
Moment resisting joints are typically classified by stiffness as rigid, semi‐rigid and (nomi‐
nally) pinned. Since rigid joints are almost impossible in timber structures, rotations in
semi‐rigid joints should be taken into account when determining buckling lengths. Semi‐
rigidity in moment resisting joints with mechanical connections means a greater buckling
length compared to a system with rigid joints. Here, the rotational stiffness Kr of a semi‐
rigid joint is defined as the moment necessary to cause an angle of rotation of 1 rad in the
joint. Using the slip modulus Ku of the fastener, the rotational stiffness Kr is calculated to:
n 2
K r K u ri (D2‐8)
i 1
where ri is the distance between the fastener i and the centre of gravity of the joint.
185
Stability
As an example, the buckling length of the column shown in Figure D2‐8 is derived consid‐
ering the influence of rotation in the semi‐rigid joint at the base of the column. The ap‐
proximate solutions for buckling lengths, which incorporate the impact of semi‐rigidity in
joints, remain valid, provided the semi‐rigidity does not reduce the elastic buckling load
by more than around 20%.
y Ncrit
x
l
Kr
Ncrit
M
Figure D2‐8 Deformed state of a column with a semi‐rigid base joint. (STEP 1995 Article B7)
Using the notation of Figure D2‐8, the bending moment M is determined as:
M( x ) N y( x ) (D2‐9)
and the differential equation of the deflection curve as:
E I y M( x ) (D2‐10)
which results in the differential equation
N
y y 0 (D2‐11)
E I
with the solution:
y A sin( x) (D2‐12)
Here
N
(D2‐13)
E I
186
Stability
With the boundary conditions
M( x ) N y( x ) K r y ( x ) (D2‐14)
the following buckling condition results:
E I
( ) tan( ) 1 (D2‐15)
K r
There is no analytical solution to equation (D2‐15). However, for
(D2‐16)
2 2
the following approximate solution can be established:
tan( ) (D2‐17)
4 2 2
1
2
If the approximate solution to the equation (D2‐17) is inserted in the buckling condition
(D2‐15), the resulting elastic buckling load can be calculated as:
1
N crit 2
(D2‐18)
4
E I
2
K r
Compared with the elastic buckling load of the Euler buckling case II
E I
N crit (D2‐19)
2ef
the resulting effective length factor is
ef 2 E I
4 (D2‐20)
K r
The same procedure can also be used to determine effective length factors for other
structural systems such as interconnected columns, arches or frames. Some effective
length factors for such systems are included in Annex 3 (without derivation).
187
Stability
D2.3 Lateral torsional buckling
Background
The first step when designing beams is to ensure sufficient load‐bearing capacity and
stiffness for bending about its major principal axis, usually on the vertical plane. This
often results in cross‐sectional forms in which the stiffness on the vertical plane may far
exceed the value on a horizontal plane. For columns, meanwhile, it was shown that buck‐
ling failure could occur when axial compressive force is applied to a slender load‐bearing
element (loaded in its stiff plane), resulting in lateral deviation of the element (deflecting
in its weaker planes). Figure D2‐9 portrays the behaviour of a slender simply supported
beam when subjected to bending stress on the vertical plane. This behaviour, which
involves the beam laterally deviating and twisting in the process, is known as lateral tor‐
sional buckling. This type of instability is comparable to a column buckling when subject
to axial (compression) force, since the stress imposed on the beam in its stiffer vertical
plane causes failure due to buckling from this plane, in a less stiff direction.
The bending moment, under which the ideal beam is rendered unstable, is known as the
critical moment, equations for which can be taken from textbooks, such as those from
Timoshenko and Gere (1961). The critical moment used in EC 5 is derived in Annex 4,
whereby, as a general rule, a beam of ideal elastic and isotropic material is assumed.
Hooley and Madsen (1964) proved that this theory also applies to beams made of wood
as an anisotropic material.
A – A:
M
A
y
M
z
(a) (b)
Figure D2‐9 Lateral torsional buckling of a simply supported beam subject to constant moment.
(a) simply supported beam, (b) buckled beam. (STEP 1995 Article B3)
188
Stability
The critical moment for the ideal beam in accordance with Figure D2‐9 with E∙I = const.,
simply supported in y‐ and z‐directions and torsional restraints about the x‐axis at both
supports is as follows (derivation of Mcrit shown in Annex 4):
M y,crit E I z G I tor (D2‐21)
ef
where
Iz Moment of inertia (second moment of area) about the corresponding axis
E Modulus of elasticity of the material
G Shear modulus of the material
ℓef Effective length, in this case the distance of the torsional restraints
Itor Torsional second moment of the beam cross‐sectional area
Since the lower quantile value of the critical moment remains to be determined, 5%
quantiles should be used for the stiffness properties E and G.
For a timber beam with rectangular cross‐section b x h, the critical stress can be calculat‐
ed from the critical moment divided by the section modulus as
For isotropic material, there is only one value for the modulus of elasticity E and the
shear modulus G, while the values for wood depend on the angle between the fibre and
stress directions. In general, the modulus of elasticity parallel to the grain should be used
and G is assumed to constitute G = E/16 for softwood. Accordingly, assuming rectangular
and slender beams with high h/b ratio ( = 0.333, see also Table D1‐1), the critical stress
for softwood is:
0,78 b 2
m,crit E 0,05 (D2‐23)
h ef
According to NCI NA 13.3, however, for simply supported beams with torsional restraints,
the influences of a semi‐rigid torsional restraint at the support, elastic foundation against
lateral deformation and against twisting may be taken into consideration. However, these
influences were not taken into consideration when deriving the equation for Mcrit used in
EC 5, see also Annex 4. The corresponding equations are specified in the NA. Moreover,
due to the homogenisation of glulam beams, in the event of torsion and bending about
the minor principal axis, the 5% quantile of the product of E and G may be increased by
40% for glulam beams in accordance with the NA.
189
Stability
Similar terms are also obtained for other load cases, load configurations and support
conditions. For a simply supported beam with a concentrated load at the centre that is
applied at the neutral axis of the beam, the equation for m,crit is very similar to equation
(D2‐22), merely replacing π with 4.24. The ratio π/4.24 is often referred to as the equiva‐
lent uniform moment or m‐factor. It is a measure denoting the influence of individual
moment diagrams compared to the basic case with a uniform moment diagram. The
values of the m‐factor for certain load cases are listed in Table D2‐1.
In general, the more the moment diagram deviates from the uniform case, the greater
the lateral stability. In EC 5, rounded m‐factors are used; m = 0.9 for simply supported
beams with a uniformly distributed load and 0.8 for a centrally placed concentrated load.
Cantilever beams with a uniformly distributed load have an m‐factor of m = 0.5, while for
cantilever beams with a concentrated load at free end, m = 0.8. The NA cites certain
additional factors for structural systems and load cases, with which effective lengths can
be determined (Table NA.25). The location of the load in the beam height is also im‐
portant, since loads on the upper edge of a slender beam have a destabilising effect,
while those on the lower edge of the beam act to stabilise it. This is not taken into con‐
sideration in the existing verifications in Europe.
Of course, the support conditions are also important. Torsional restraints, which prevent
lateral deviation and twisting of the beam, enhance its lateral torsional stability (or are
required when deriving Mcrit). Improving the lateral torsional stability through the support
conditions is generally reflected in lower values for the m‐factor. Lateral torsional stabil‐
ity of beams is a comprehensive topic, which goes beyond the scope of the current arti‐
cle. In textbooks however, e.g. by Timoshenko and Gere (1961), this topic is presented in
detail. The following main influences on lateral torsional stability can be deduced from
the comments to date:
The distance of the lateral supports (i.e. braces, the distance
between points at which lateral deflection is prevented),
The lateral bending stiffness EIz of the beam,
The torsional stiffness GItor of the beam,
The location of the load,
The torsional restraint at the beam ends.
The load‐bearing capacity of a beam at risk of lateral torsional buckling can be improved
by installing bracing members. The main requirements are sufficient stiffness of the
bracing to effectively prevent lateral deformations of the beam and sufficient load‐
bearing capacity of the bracing to accommodate forces transmitted by the beam (see
also Article D9).
190
Stability
Table D2‐1 Equivalent uniform moment factors according to Kirby and Nethercot (1979). The value of 0.74
for the load case „concentrated load on simply supported beam“ corresponds to /4.24.
(STEP 1995 Article B3)
Actual moment
Tatsächliche Equivalent uniform
Gleichwertige
Load case
Lastfall m
m
diagram
Momentenlinie moment diagram
Momentenlinie
M M
1.0
1,00
M
0.57
0,57
M M
0,43
0.43
F
0.74
0,74
q
0.88
0,88
L/4 F F L/4
0,96
0.96
F
0.69
0,69
L/4
F
0.59
0,59
q
0.39
0,39
Lateral torsional stability in EC 5
Similar to the kc values for columns, the strength of beams is also mathematically re‐
duced by a coefficient kcrit, which is determined from the relative slenderness rel,m and
the above‐defined critical bending stress m,crit. The coefficient kcrit was defined on a
completely different basis than the coefficient kc. The values for kcrit were derived for
elastic systems, homogenous material and fixed eccentricity e of ℓ/500 (see Annex 4),
while to derive kc, second order plastic theory assuming stochastically dispersed eccentri‐
cities, stiffness and strength properties of timber members was applied (see Sec‐
tion D2.1).
191
Stability
The verification thus reads as follows:
For rel,m ≤ 0.75, the beam is not at risk of lateral torsional buckling:
k crit 1 (D2‐25)
For 0.75 < rel,m ≤ 1.4:
For 1.4 < rel,m:
1
k crit (D2‐27)
rel,m
2
whereby the relative slenderness rel,m for bending is revealed by
f m,k
rel,m (D2‐28)
m,crit
kcrit 1,2
kinst
1,0
0,8
0,6
0,4
0,2
0,0
0 1 2 3 4 5
λrel, m
Figure D2‐10 Coefficient kcrit depending on the relative slenderness rel,m. (STEP 1995 Article B3)
192
Stability
Combined stresses
For beams exposed to a combined load of both a bending moment and a normal (com‐
pression) force, a combined verification (equations (D2‐7) and (D2‐29)) is provided:
2
c,0,d m,y,d
1 (D2‐29)
k c,z f c,0,d k crit f m,y,d
With a simultaneous occurrence of bending moments in y and z‐directions and aspect
ratios of h/b ≤ 4, the following equations apply according to the NA (the coefficient km
from equations (D1‐9) to (D1‐14) is not applied in this case):
2
c,0,d m,y,d m,z,d
1 (D2‐30)
k c,z f c,0,d k crit f m,y,d f m,z,d
2
c,0,d m,y,d
m,z,d 1 (D2‐31)
k c,y f c,0,d k crit f m,y,d f
m,z,d
D2.4 Literature
B.S. Choo, H.J. Blass, original Articles B3, B6, B7, STEP 1995.
Blass H.J. (1987). Tragfähigkeit von Druckstäben aus Brettschichtholz unter Berücksichtigung streuender
Einflußgrößen. Dissertation Universität Karlsruhe.
Blass H.J. (1988)a. Traglastberechnung von Druckstäben aus Brettschichtholz. Bauingenieur 63:245‐251.
Blass H.J. (1988)b. Einfluß des Kriechens auf die Tragfähigkeit von Holzdruckstäben.
Holz als Roh‐ und Werkstoff 46:405‐411.
Glos P. (1978). Zur Bestimmung des Festigkeitsverhaltens von Brettschichtholz bei Druckbeanspruchung aus
Werkstoff‐ und Einwirkungskenngrößen. Dissertation, Technische Universität München.
Hooley R.F. and Madsen B. (1964). Lateral stability of glued laminated beams.
Journal of the Structural Division ASCE, 3:201‐218.
Kirby P.A. and Nethercot D.A. (1979). Design for structural stability. Constrado Monographs,
Crosby Lockwood Staples, Granada Publishing.
Timoshenko S. and Gere J.M. (1961). Theory of elastic stability. McGraw‐Hill Book Co. Inc. New York, NY.,
2nd Edition.
193
D3 Influence of volume and
stress distribution
Original article: F. Rouger
Observations from numerous tests show that the strength of a material declines as the size
of the test specimen increases. This so‐called volume or size effect can easily be explained
by the fact that as the size of a member increases, the number of defective points (knots,
finger joints…) follows suit. Accordingly, the number of potential failure points rises and the
point with the lowest ratio value in terms of strength against stress is where failure is likeli‐
est. The volume effect only comes into play for brittle failure mechanisms (tension, shear),
since load redistributions may occur as part of ductile (plastic) behaviour (compressive
stress), which considerably mitigate the impact of the defective point on strength. Weibull
(1939) put forward a theory describing the influence of the stressed member volume and
stress distribution within this volume on the strength of homogeneous and isotropic mate‐
rials with brittle failure. Although wood is neither homogenous nor isotropic, the Weibull
theory can also be applied to wood (Colling, 1986). Denzler (2007) also cites further theo‐
ries, concerning the shortcomings of Weibull theory for wood as a material. For timber
engineers, for example, the volume effect has to be taken into consideration in each case of
brittle failure, particularly tensile failure perpendicular to the grain. In practical terms, this
means the volume effect is crucial in double‐tapered beams, pitched cambered beams and
in curved beams, since such beam types are prone to high tensile stresses perpendicular to
the grain in the apex zone due to their geometry.
D3.1 Theory
The theory of the weakest link was developed by Pierce (1926), Tucker (1927) and Weibull
(1939), who examined brittle material such as concrete. This theory states that “A chain
exposed to tensile stress is as strong as its weakest link.” To explain this theory, we consider
a unit volume subject to tension, the failure probability Pf of which is given by:
P f F ( ) Probability(strength ) (D3‐1)
whereby F() is the distribution function of strength , Figure D3‐1 and Figure D3‐2 on
the left.
195
Influence of volume and stress distribution
F(σ )
1
0,8
0,6
0,4
0,2
0
σ 5 15 20 25 30 35 40
σ0 σ (N/mm2)
Figure D3‐1 Distribution function for the failure of a unit volume. (STEP 1995 Article B1)
If we now consider a chain comprising N links, whereby each link represents a unit vol‐
ume, this system will only be sustained, if each link survives, namely:
Ps is the probability of survival of the system and Ps (i) is the probability of survival of an
individual element i. Based on equation (D3‐2) and assuming that the unit volumes show
the same probabilities of failure and that the occurrences of failure are independent
among all unit volumes (the power of which can be expressed with the exponent N), the
resulting probability of failure of the system is:
P f 1 Ps 1 1 F ( ) 1 e Nln(1F ( )) 1 e NF ( )
N
(D3‐3)
The simplification applies, since F() is small.
We now assume that the lower tail of the strength distribution of a unit volume (F()
from Figure D3‐1) may be adapted to an exponential function with a lower limit value 0,
namely
k
0
F ( ) mit 0 (D3‐4)
m
The probability of failure of the system, namely a structural member comprising multiple
infinitesimally small unit volumes dV, can then be expressed as follows:
k
0
k 0
N m
dV
m
P f ( ) 1 e 1e V
(D3‐5)
196
Influence of volume and stress distribution
If we now assume constant stress distribution in the member volume, equation (D3‐5)
can be simplified as follows:
k
0
V
m
P f ( ) 1 e (D3‐6)
Equation (D3‐6) is known as a three‐parameter Weibull distribution. When 0 = 0, a two‐
parameter Weibull distribution results. Parameters m and k can be estimated from the
average value of , E() and the coefficient of variation of , COV(). Parameter k is also
known as a shape parameter and determines the “appearance” of the Weibull distribu‐
tion. When k = 1 for example, the exponential distribution is obtained and for k = 2 the
Rayleigh distribution, while the parameter m is the scale parameter. However, the three‐
parameter Weibull distribution (D3‐6) depends not only on the parameters k and m, but
also on volume V. Figure D3‐2 shows the distribution and density function of the two‐
parameter Weibull distribution.
1.2 2
k = 0.5
1 k = 1.0
1.5
0.8 k = 3.0
k = 5.0
F()
f()
0.6 k = 0.5 1
0.4 k = 1.0
k = 3.0 0.5
0.2
k = 5.0
0 0
0 1 2 3 4 5 0 1 2 3 4 5
Figure D3‐2 Distribution function (left) and density function (right) of a two‐parameter Weibull distribution,
0 = 0, V = m = 1.0.
The Weibull distribution can hence be used e.g. to determine the dependency of tensile
strength on stressed volume, assuming the equivalent probability of failures Pf of two test
specimens. For the test specimen of volume V1 and a given probability of failure Pf(1)
under tensile stress 1 and a second test specimen with volume V2 and Pf(2) under ten‐
sile stress 2, a comparison of the characteristic strength values of both volumes results in:
1
k k
V k
P f ( 1 ) P f ( 1 ) V 1 1 V 2 2 2 1 (D3‐7)
m m 1 V2
This equation is the basis for explaining the size effect.
197
Influence of volume and stress distribution
If there are no uniformly dispersed tensile stresses, these equations must be modified as
follows for other stress distributions:
( x , y , z) w ( x , y , z) (D3‐8)
k
V *
m
P f ( ) 1 e (D3‐9)
whereby V* is defined as:
V * w( x , y , z) dV
k
(D3‐10)
V
This method used to calculate the influence of stress distribution was applied by Larsen
(1986) and Colling (1986), to estimate the influence of volume and stress distribution on
shear and tensile strengths perpendicular to the grain in curved and double‐tapered
beams. Larsen used the so‐called “distribution coefficient”, to record the impact of stress
distribution on strength:
V
k dis 1
(D3‐11)
V
* k
Accordingly, the coefficient kdis is used to calculate design tension strengths perpendicu‐
lar to the grain for different load configurations:
1
* V k (D3‐12)
f t,90,d k vol k dis f t,90,d with k vol 0
V
198
Influence of volume and stress distribution
D3.2 Research results
Numerous published test data has emerged to explain the size effect for timber, but
some of the results are contradictory (Barrett and Lam, 1992; Madsen, 1992), which
could be for the following reasons:
The size effect is explained using the brittle failure theory, which is applicable to
tension parallel and perpendicular to the grain (Barrett, 1974; Colling, 1986) and
to shear (Foschi and Barrett, 1976; Foschi, 1985; Colling, 1986). For compression
and particularly for bending, whereby a combined failure mechanism comprising
both compressive and tensile stresses is involved, this theory is debatable.
The size effect assumes an equivalent probability of failure of the “unit volumes”.
This assumption may not apply to all wood species, however, particularlyif the
knots are not dispersed at random within the wood volume.
For visually graded timber in particular, the sizes of wood defects rise with in‐
creasing member sizes. In other words, the material itself changes in line with di‐
mensions, meaning a pure size effect is distorted. In particular, the size effect can
be superimposed when timber of various qualities is used by imposing a “grading
effect”. Denzler (2007) thus makes a distinction between direct size effects (di‐
mensions) and indirect (knot area sizes). The combination of both effects as an ef‐
fective and measurable size effect depends significantly on grading (grading influ‐
ences knot area sizes) and testing conditions.
If tests are conducted for constant span to depth ratios in bending, the size effect
is a combination of length and depth effect (Barrett and Fewell, 1990). Both ef‐
fects cannot be identified separately.
Above all, to take the strong anisotropy of the wood into consideration, the size effect
according to Weibull was subdivided into respective length and depth effects. In Ta‐
ble D3‐1, the following parameters are listed for the size effect when bending stress is
applied.
The length factor SL (namely members of the same height h1 = h2) results from:
1
SL
2 L1 kL L1
(D3‐13)
1 L2 L2
199
Influence of volume and stress distribution
The depth factor Sh (namely members of the same span L1 = L2) results from:
1
Sh
2 h1 k h h1
(D3‐14)
1 h2 h2
The size factor SR (namely members with a constant span to depth ratio L1/h1 = L2/h2)
results from:
SL Sh SL Sh SL S h SR
2 L1 h h h h h
1 1 1 1 1 (D3‐15)
1 L2 h2 h2 h2 h2 h2
Additional test results for glued laminated timber with a far lower sample size were pub‐
lished by Ehlbeck and Colling (1990). From simulations on the bending strength of glued
laminated timber, a value SR of 0.12 (Frese and Blass, 2009) was determined. The size
effect for glued laminated timber is smaller than for solid timber, since ultimately, the so‐
called “lamination effect” increases strength (lamination effect see Section B6.2).
The results for tension differ slightly from those for bending, since a brittle behaviour
prevails in the first case, see also Table D3‐2. Based on simulations, in turn, conducted on
glued laminated timber subject to tension, the resulting length factor is SL = 0.11. (Frese
et al., 2010).
Table D3‐1 Size factors for bending.
S L Sh SR
Barrett and Lam (1992), solid timber 0.17 0.23 0.40
Madsen (1992), solid timber 0.20 0.00 0.20
Ehlbeck and Colling (1990), glulam 0.15 0.15 0.30
Table D3‐2 Size factors for tension, solid timber.
S L Sh SR
Barrett and Lam (1992) 0.17 0.23 0.40
Madsen (1992) 0.20 0.10 0.30
200
Influence of volume and stress distribution
Table D3‐1, Table D3‐2 and the simulation results of Frese show that size factors of solid
timber have a value of around 0.2 (corresponding to a shape parameter k of 5) and are
somewhat lower for glulam, around 0.1. Both these values are echoed in the height
factor kh, which was already addressed in Section D1.1 and which is applied when estab‐
lishing strength values for tension parallel to the grain and bending.
Although the results differ, they still prove the size effect (→ kh, kvol) for many types of
stress with an additional stress distribution factor (→ kdis), which may be in the same order
of magnitude. For the purpose of practical application, these approximations, particularly
influences from stress distribution, were further simplified.
D3.3 Size and stress distribution effects
in accordance with EC 5
Size effects are taken into consideration by modifying the characteristic strength values
determined in EN 338. The characteristic values for bending and tensile strength are
based on a reference height of 150 mm for solid timber and 600 mm for glued laminated
timber3. For depths less than these reference values, strength values are multiplied by a
size factor, which is limited by an upper value. This means that the size effect need only
be used in one direction (Figure D3‐3) as was already explained in the previous section
and in Section D1.1.
For solid timber:
150 0.2
k h min h (D3‐16)
1.3
For glued laminated timber:
600 0.1
k h min h (D3‐17)
1.1
where h represents the beam depth.
3
Namely, the test specimens, which were used to determine the characteristic strengths, were 150 mm or
600 mm in height.
201
Influence of volume and stress distribution
Since the size effect depends significantly on the grading and test conditions, equations
(D3‐16) and (D3‐17) remain under discussion (in DIN 1052:2008, for example, there was
no kh for solid timber). Moreover, no purely brittle behaviour can be observed for bend‐
ing members. Accordingly, it is very difficult to determine a pure size effect in isolation.
Based on the type of grading and classification of solid timber used in Europe, the scope
of the size effect is already partially included in the characteristic values determined.
2,0
kh
1,5
1,3
1,0
0,5
0,0
0 50 100 150 200 250 300
h (mm)
Figure D3‐3 Size factor for solid timber in bending or tension in accordance with EC 5 (solid line) and with
theory (dashed line). (STEP 1995 Article B1)
When designing double‐tapered and curved beams above all, it is indisputably crucial to
take a volume and stress distribution effect in the apex zone into consideration, since this
is where, based on geometry, high tensile stresses perpendicular to the grain frequently
occur with a simultaneous large volume. The following requirement must be met in the
apex zone for beams made of glulam or LVL (see also equations (D3‐12) and (D4‐15)):
0.2
V0
t,90,d k dis k vol f t,90,d k dis f t,90,d (D3‐18)
V
where
kdis Stress distribution factor (see also equation (D3‐11)), which is determined as
follows:
kdis = 1.4 for double‐tapered and curved beams
kdis = 1.7 for pitched cambered beams
kvol Volume factor, kvol = 1.0 for solid timber
V0 Reference volume of 0.01 m3
V Volume subject to tension perpendicular to the grain in the apex zone,
for definition see Figure 6.9 in EC 5
202
Influence of volume and stress distribution
Another interesting point here is the fact that the verification in equation (D3‐18) adopt‐
ed a slightly different format in DIN 1052:
0.3
h
t,90,d k dis 0 f t,90,d (D3‐19)
h ap
In DIN 1052 therefore, with (h0/hap)0.3 a depth effect had been taken into consideration
and not a volume effect as in EC 5 with (V0/V)0.2, which also explains the different shape
parameter k = 10/3 in DIN 1052. In DIN 1052, the coefficient for the distribution of tensile
stresses perpendicular to the grain kdis, the reference depth h0, the exponent 0.3 and the
design value of tensile strength perpendicular to the grain ft,90,d were coordinated with
each other to ensure any probability of failure would be sufficiently small by comparison
with the maximum value t,90,d. The reasons why a depth effect was taken into considera‐
tion rather than a volume effect can be attributed to the easier handling, since the beam
depth hap in the apex zone is easy to establish, whereas the volume V has to be deter‐
mined first.
D3.4 Literature
F. Rouger, original Article B1, STEP 1995.
Barrett J.D. and Fewell A.R. (1990). Size factors for the bending and tension strength of structural lumber.
Paper 23‐10‐3, CIB‐W18 Meeting 23, Lisbon.
Barrett J.D. and Lam F. (1992). Size effects in visually graded softwood structural lumber. Paper 25‐6‐5,
CIB‐W18 Meeting 25, Ahus.
Barrett J.D. (1974). Effect of size on tension perpendicular to grain strength of Douglas fir.
Wood and Fiber 6(2):126‐143.
Colling F. (1986). Influence of volume and stress distribution on the shear strength and tensile strength
perpendicular to grain. Paper 19‐12‐3, CIB‐W18 Meeting 19, Florence.
Denzler J.K. (2007). Modellierung des Größeneffektes bei biegebeanspruchtem Fichtenschnittholz.
Dissertation, Technische Universität München.
Ehlbeck J. and Colling F. (1990). Bending strength of glulam beams, a design proposal. Paper 23‐12‐1,
CIB‐W18 Meeting 23, Lisbon.
Foschi R.O. (1985). Longitudinal shear design of glued laminated beams. Paper 18‐10‐2, CIB‐W18, Beit Oren.
Foschi R.O. and Barrett J.D. (1975). Longitudinal shear strength of Douglas Fir.
Canadian Journal of Civil Engineering, 3(2):198‐208.
Frese M. and Blass H.J. (2009). Bending strength of spruce glulam. European Journal of Wood
and Wood Products 67:277‐286.
Frese M., Chen Y. and Blass H.J. (2010). Tensile strength of spruce glulam. European Journal of Wood
and Wood Products 68:257‐265.
Larsen H.J. (1986). EC 5 and CIB structural timber design code. Paper 19‐102‐2, CIB‐W18 Meeting 19, Florence.
Madsen B. (1992). Structural behaviour of timber. Timber Engineering Ltd., Vancouver.
Pierce F.T. (1926). Tension tests for cotton yarn. Journal of the Textile Institute, S. T155‐T368.
Tucker J. (1927). A study of compressive strength dispersion of material with applications.
Journal of the Franklin Institute 204:751‐781.
Weibull W. (1939). A statistical theory of the strength of materials. Royal Swedish Institute
for Engineering Research N. 141, 45p.
203
D4 Members with variable cross‐
sections or curved shapes
Original articles: J. Ehlbeck, J. Kürth, H. J. Larsen
Glued laminated timber beams are often realised with varying depths and with or with‐
out curvature, in line with architectural perspectives, due to gabled roofs or the need to
maximise interior space or reduce the height of exterior walls at supports. The most
frequently used beam shapes are single and double‐tapered beams, curved beams of
constant depth, fish‐bellied beams and pitched cambered beams (see Figure D4‐1) and
here, fish‐bellied, double‐tapered and pitched cambered beams optimise the use of
materials, since their depth follows the moment diagram. Many beam shapes also offer
the benefit of allowing secondary components or roof diaphragms to be directly applied
while eliminating the need for woodblocks to make a minimum roof slope. The incline
should not generally exceed 10° even though the limit value specified in the NA is unreal‐
istically high at 24°.
(e)
hap
Figure D4‐1 (a) Single‐tapered beam, (b) curved beam of constant depth, (c) double‐tapered beam,
(d) pitched cambered beam, (e) fish‐bellied beam.
205
Members with variable cross‐sections or curved shapes
Single‐tapered beams
The distribution of bending stresses in areas with variable beam depth h(x) is non‐linear
already due to the beam geometry, since the slope of the beam edge results in additional
shear stresses zx and stresses perpendicular to the grain Z, see Figure D4‐2, which trig‐
ger an increase in bending stress at the edge parallel to the grain and thus an overall non‐
linear distribution as is also shown in Figure D4‐3. For small angles of inclination, up to
around ≤ 10°, however, this non‐linearity is small (for derivations, see Petersen (1993),
Section 26.7). Accordingly, the bending stresses in the outer fibres can be determined
more simplistically in accordance with the Euler‐Bernoulli beam theory.
h x Vz h x Vz
z z
D W zx W = 0 D W zx W xz V
V x
dx · tan D
W = 0
dx · sin D
V = 0
dx
dx · cos D
dx (a) (b)
Figure D4‐2 Stresses at tapered edge; (a) cut normal to edge and horizontal, (b) cut vertical and horizontal
(according to Petersen 1993, Figure 57)
A B
x D
Vm hap
hmin h
ℓ
Figure D4‐3 Single‐tapered beam with beam width b. (STEP 1995 Article E4)
In addition to the geometry, the anisotropy of the wood also has an impact on stress
distribution, although this impact cannot be properly recorded with the Euler‐Bernoulli
beam theory. Nevertheless, an approximate value can still be calculated for the same
based on anisotropic plate theory and taking into consideration the ratios E0/E90 and E0/G
and Poisson’s ratio (Blumer, 1979).
The third parameter is the fact that areas with cut wood fibres (at the tapered edge) exist
due to the beam shape and production methods. Since the area of the tapered edge is,
additionally to the axial stresses, also subject to tranverse and shear stresses Z and
zx depending on the taper angle (Figure D4‐2), this means the activation of very low
206
Members with variable cross‐sections or curved shapes
stiffness and strength values at the tapered edge when normal stress perpendicular to
the grain and shear are applied, see Figure D4‐4. Tensile stresses perpendicular to the
grain form at a tapered tensile edge and compressive stresses perpendicular to the grain
at a tapered compressive edge. As a general rule, cut wood fibres should preferably be
arranged in the compression zone of beams, so that the lamellae run parallel to the beam
edge in the tensile zone. During design, this stress combination is taken into considera‐
tion by reducing the design values for bending strength fm by a factor km, . Moreover, for
beams with variable cross‐sectional depth, it is important to identify the critical cross‐
section, in which bending stresses peak. The strength of the beam in Figure D4‐3 must
hence be verified in two cross‐sections: in cross‐section A, in which the shear stresses
peak and cross‐section B, in which the bending stresses peak.
(a)
α
τ σm,0
σc,90
(b)
α
τ σm,0
σt,90
Figure D4‐4 Axial stresses (parallel to the grain, longitudinal),stresses perpendicular to the grain and shear
stresses at the tapered edge with cut wood fibres with (a) compressive bending stresses and
(b) tensile bending stresses. (STEP 1995 Article B8)
Double‐tapered beams
Double‐tapered beams (Figure D4‐5) represent an expansion of single‐tapered beams,
but include an apex, at which normal stresses must be deflected. This results in high
tensile stresses perpendicular to the grain, which cannot be ignored. For double‐tapered
beams, this means that as for single‐tapered beams, the strength for two cross‐sections
must be checked (cross‐sections A and B, see Figure D4‐5). In addition, the tensile stress
perpendicular to the grain and bending stress in the apex cross‐section must also be
checked. This bending stress depends on the taper angle , as calculations using aniso‐
tropic plate theory have shown (Blumer, 1979). Both additional stress components must
also be verified for curved beams; the formation of which is explained for both beam
types in the following section.
207
Members with variable cross‐sections or curved shapes
A B
x
Vm h D
hmin V t,90 ap
0.5 ℓ 0.5 ℓ
Figure D4‐5 Double‐tapered beam with beam width b. (STEP 1995 Article E4)
Curved beams
Unlike single and double‐tapered beams, curved beams with a constant cross‐section
have no tapered edge (and thus no cut wood fibres) and no variable depth. However,
here also the bending stresses are non‐linear, regardless of the material used and, in
turn, ultimately based on the beam shape. For clarification, Figure D4‐6 shows the distri‐
bution of bending stress at a section of a curved beam. Since the internal fibres are con‐
siderably shorter than the external fibres, the strains at the edges are as follows, assum‐
ing the Bernoulli hypothesis that plane sections remain plane applies and that the neutral
axis is going through the centre of gravity:
Δ d i Δd o
i o (D4‐1)
d i d o
Δlo
σo
lo
M
h
M
li
Δli
b σ r
σi
Figure D4‐6 Distribution of bending stresses in a curved beam. (STEP 1995 Article B8)
208
Members with variable cross‐sections or curved shapes
Consequently, Hooke’s law states that the maximum bending stress |i| at the inner
edge exceeds the same at the outer edge |o|. To maintain the equilibrium of internal
forces in the cross‐section, the neutral axis must shift from the centre of gravity position
towards the inner edge resulting in a non‐linear, hyperbolic stress distribution. The max‐
imum bending stress in a symmetrical single‐span girder under a uniformly distributed
load occurs at the inner edge of the apex cross‐section, while the extent and distribution
of the maximum bending stress depends on the curvature (Blumer, 1979). As with beams
of varying depth, the distribution of bending stresses is also influenced by the anisotropy
of the wood, which means the bending stresses increase at the inner edge (Blumer,
1979). For curved beams, the rule is that the bending stresses in the apex cross‐section
must be checked, the size of which varies depending on the radius of curvature. The
bending stress in the apex can be approximately determined (Blumer, 1975 and 1979)
from Map/Wap („ap“ = apex) multiplied by a coefficient kℓ (kℓ > 1, see Figure D4‐7 on the
left), which depends on the ratio of the apex height hap to the radius of curvature r and
for beams where the depth is not constant (double‐tapered beams), additionally on the
taper angle . For curved beams of constant depth, = 0.
However, bending moments in curved and apex areas of double‐tapered beams also
generate stresses perpendicular to the grain. Figure D4‐8 shows the apex area of a
curved beam subject to a constant moment and specifying the tensile stresses perpen‐
dicular to the grain forming. These tensile stresses also form due to climate‐related
stresses, for example changes in moisture content caused by variable environmental
conditions.
kp
kl
α=25˚
6 Map 6M
σm,d=kl σt,90,d=kp 2ap
2,5 bh2ap 0,15 bh ap
α=20˚ α=25˚
2,0 0,10
α=15˚ α=20˚
α=15˚
1,5 α=10˚ 0,05 α=10˚
α=5˚
α=5˚ α=0˚ α=0˚
1,0 0,00
0 0,1 0,2 0,3 0,4 0,5 0 0,1 0,2 0,3 0,4 0,5
hap /r hap /r
Figure D4‐7 Coefficients kℓ (influence of curvature or taper angle on bending stress, left) and kp
(influence of curvature or taper angle on tension perpendicular to the grain, right) for
different curvatures hap/r and taper angles . (STEP 1995 Article B8)
209
Members with variable cross‐sections or curved shapes
1
Fc= - Ft= σ bh U max σt,90
4 m
Fc Fc
M M
h/2
h Ft Ft
h/2
U
b
Figure D4‐8 Tension perpendicular to the grain caused by a constant moment. (STEP 1995 Article B8)
Taking the simplified assumption of a linear longitudinal stress distribution, it is easy to
show that the resulting tensile and compressive forces, Ft and Fc, generate a deviation
force U in a radial direction. If the bending moment has a tendency to increase the radius
of curvature, this causes tensile stresses perpendicular to the grain to form. The extent of
these tensile stresses depends on the radius of curvature or the taper angle of the upper
chord and the stresses involved cannot be ignored. Given the generally large apex vol‐
ume in curved and double‐tapered beams and the fact that tensile stresses perpendicular
to the grain cause brittle failure, the influence of volume and stress distribution on
strength must be taken into consideration, see Article D3. The maximum tensile stress
perpendicular to the grain max t,90 in the apex cross‐section, can, like bending stress, be
approximately calculated from Map/Wap multiplied by a coefficient kp (kp < 1, see Fig‐
ure D4‐7 on the right). In addition to the stresses from external loads, for curved beams
with small radii of curvature, the stresses generated prior to bonding from the bending of
individual lamellae must be taken into consideration. In a board of thickness t, curvature
rin / t = 240 and a modulus of elasticity of E0 = 10000 N/mm2, the theoretical bending
stress would equate to:
E i t 10000 N
m 20.8 (D4‐2)
2 r in 2 240 mm 2
However, since such stresses are partially dissipated by plastic processes and relaxation,
they only need to be taken into consideration for small curvature radii. This is done by
reducing the design values of bending strength by kr.
Pitched cambered beams
This beam shape combines all the properties discussed to date. As well as a variable depth
and tapered edge with cut wood fibres, pitched cambered beams also include a curvature
of the bottom chord. All the verifications performed to date must therefore be applied.
210
Members with variable cross‐sections or curved shapes
Conclusion
The distribution of bending stress in beams with variable cross‐sections or a curved shape
is non‐linear; one of the reasons for which is the beam geometry, which is intensified by
the anisotropic properties of the wood. Moreover, stress interactions must also be taken
into consideration for tapered edges with cut wood fibres, since this is where additional
shear stresses and stresses perpendicular to the grain form and the respectively assigned
stiffnesses and strengths of wood are low. The following verifications must be made and
are shown in the following sections:
Verification in the critical cross‐section with reduction km, for the tapered edge
with cut wood fibres:
For beams with variable cross‐section; namely single and double‐tapered beams
and pitched cambered beams.
Verification of bending stress in the apex cross‐section:
For double‐tapered beams with reduction kℓ depending on the taper angle of
the upper chord, for curved beams with reductions kℓ and kr depending on the
curvature, pitched cambered beams like curved beams, although kℓ additionally
depends on the taper angle of the upper chord.
Verification of tensile stresses perpendicular to the grain in the apex cross‐section:
For double‐tapered beams with reduction kp depending on the taper angle of
the upper chord, for curved beams with reduction kp depending on the curvature,
for pitched cambered beams with reduction kp depending on the taper angle of
the upper chord and curvature. In addition, when verifying tensile stresses per‐
pendicular to the grain, the volume effect must be taken into consideration via kdis
and kvol, see also equation (D3‐18).
211
Members with variable cross‐sections or curved shapes
D4.1 Single‐tapered beams
Single‐tapered beams, like that shown in Figure D4‐9, must be verified at the cross‐
section where bending stress peaks, while the stresses perpendicular to the grain and
shear stresses which occur at a tapered edge are taken into consideration using a reduc‐
tion coefficient km,.
σm,α,d
I α < 10˚
hap h
hs
I
b σm,0,d
x I-I
l
Figure D4‐9 Single‐tapered beam with critical cross‐section I‐I. (STEP 1995 Article B8)
Verification of bending stresses in the critical cross‐section
The maximum stress acts at a point x, at which ∂/∂x = 0. In the case of a uniformly dis‐
tributed load, the following applies for x:
x (D4‐3)
h ap
1
hs
If other load configurations apply or if the characteristic bending strength value is in‐
creased by the factor kh for beam depths under 600 mm, the maximum value is deter‐
mined by calculating the stresses in various cross‐sections, by proceeding step by step
along the beam, from the point of maximum bending moment in the direction in which
the cross‐sectional area declines.
The stresses in cross‐section I‐I (Figure D4‐9) are determined as follows:
6M d
m,α,d m,0,d (D4‐4)
bh2
Equation (D4‐4) shows that EC 5 equates the bending stresses in the outer fibres at the
straight and tapered edges as M/W simplistically. As was explained in the previous sec‐
tion, this simplified verification is still permissible for the inclination angles encountered
in timber structures.
212
Members with variable cross‐sections or curved shapes
The following condition must be met for the edge with the cut wood fibres:
The stress combinations at the edge with cut wood fibres are taken into consideration
using the factor km, (see also the Hankinson equation (D1‐1) and equation (6) in Annex
2). For tensile stresses along the edge with cut wood fibres, the following applies for km,:
1
k m,α (D4‐6)
2 2
f f m,d
1 m,d tan 4 tan
2
f 0.75 f
t,90,d v,d
For compressive stresses along the edge with cut wood fibres, the following applies (ft,0,d
is replaced by fc,0,d and the shear strength is increased due to the simultaneous, favoura‐
bly acting compression perpendicular to the grain):
1
k m,α (D4‐7)
2 2
f f m,d
tan tan
4 2
1 m,d
f
c,90,d 1.5 f v,d
In addition, shear stress and compression perpendicular to the grain at the supports must
be verified in cross‐section A (see Figure D4‐3) (see Article D1).
D4.2 Curved, double‐tapered, pitched cambered
and fish‐bellied beams
The verifications shown for single‐tapered beams, namely shear stress and compression
perpendicular to the grain at the supports and bending stresses in the critical cross‐
section, must also be conducted for double‐tapered, pitched cambered, fish‐bellied and
curved beams, whereby the depth of curved beams tends not to vary, meaning the criti‐
cal cross‐section for verifying bending stress is usually at the point where the bending
moment peaks. As well as verifying bending stress at the critical cross‐section, for double‐
tapered and pitched cambered beams, the bending strength and tensile strength per‐
pendicular to the grain must also be verified in the apex cross‐section. The reason for the
tensile stress verification was already shown in Figure D4‐8, for which the volume effect
must also be taken into consideration. Figure D4‐10 sets out important geometric details,
which are used in the following.
213
Members with variable cross‐sections or curved shapes
α
α V
V hap
r t
rin
r=
8
0,5hap 0,5hap r = rin + 0,5 hap
hap hap
σt,90 σt,90
b σm b σm
Figure D4‐10 Geometric details and stress distribution at apex for (a) double‐tapered beams and (b) pitched
cambered beams. (STEP 1995 Article B8)
qd
h = hap
β
V
t hap
r σt,90
rin
σm
b
Verification of bending stresses at apex
In the apex cross‐section, the design value of the maximum bending stress can be calcu‐
lated as per:
M ap,d 6 M ap,d
m,d k k 2
(D4‐8)
W ap b h ap
214
Members with variable cross‐sections or curved shapes
Coefficient kℓ is (see also Figure D4‐7 on the left)
2 3
h ap h ap h ap
k k1 k 2 k 3 k 4 (D4‐9)
r r r
where
k 1 1 1.4 tan ap 5.4 tan 2 ap
k 2 0.35 8 tan ap
k 3 0.6 8.3 tan ap 7.8 tan 2 ap
k 4 6 tan ap
2
r r in 0.5 h ap
The coefficient kr required in the verification has already been addressed and takes into
consideration the stresses generated prior to bonding due to the bending of individual
lamellae for curved beams with small radii of curvature. It is determined as follows,
where t = lamellae thickness:
For rin/t ≥ 240:
kr 1 (D4‐11)
For 150 ≤ rin/t < 240:
r in
k r 0.76 0.001 (D4‐12)
t
215
Members with variable cross‐sections or curved shapes
Verification of tension perpendicular to the grain at apex
In the apex cross‐section, the design value of the maximum tensile stress perpendicular
to the grain can be calculated as per:
M ap,d 6 M ap,d
t,90,d k p kp 2
(D4‐13)
W ap b h ap
Coefficient kp is (see also Figure D4‐7 on the right):
2
h ap h ap
kp k 5 k 6 k 7 (D4‐14)
r r
where
k 5 0.2 tan ap
As with coefficient kℓ, the equation for double‐tapered beams is reduced to the factor k5,
while the relevant factors, in turn, are determined by regression equations based on the
work of Blumer (1979).
The verification is established as follows (see also equation (D3‐18)), whereby the factors
kdis and kvol to be used here take the influence of stress distribution (dis = distribution)
and volume into consideration, see Article D3:
The factor kdis takes into consideration the distribution of tensile stresses perpendicular
to the grain in the stressed volume. The factor kdis is assumed to be 1.4 for double‐
tapered and curved beams and 1.7 for pitched cambered beams.
The factor kvol is the quotient from the reference volume V0 = 0.01 m3 and volume V
subject to tensile stress perpendicular to the grain and takes into consideration the influ‐
ence of volume on tensile strength perpendicular to the grain. However, no more than
2/3 of the entire beam volume Vb must be mathematically applied. kvol for solid timber:
k vol 1 (D4‐16)
216
Members with variable cross‐sections or curved shapes
For glulam and LVL:
0.2
V
k vol 0 (D4‐17)
V
If shear and tensile stress perpendicular to the grain apply simultaneously, a combined
verification is performed with linear stress interaction:
d t,90,d
1 (D4‐18)
f v,d k dis k vol f t,90,d
The verifications for tensile stresses perpendicular to the grain in the apex only apply to
non‐reinforced beams, while reinforced beams are covered in Article D8.
D4.3 Climate action on curved, double‐tapered
and pitched cambered beams
The specified tensile stress verifications perpendicular to the grain (equations (D4‐15)
and (D4‐18)) for the apex area of curved, double‐tapered and pitched cambered beams
only apply for non‐reinforced beams. All verifications are performed with design loads;
stresses due to swelling and shrinking processes, which, in turn, are triggered by climatic
fluctuations are not taken into consideration. However, the beam shapes shown are very
vulnerable to tensile stresses perpendicular to the grain, which may also be triggered by
climate change and which may exert additional strain on the beams, additionally to the
pre‐existing external forces. Accordingly, the NA recommends that such beams should
always be reinforced to take account of additional climate‐related tensile stresses per‐
pendicular to the grain. For reinforced beams, the verifications for reinforced members,
Article D8, apply. Finally, according to the NA, double‐tapered beams should be rein‐
forced from a utilisation factor of 80% in equations (D4‐15) and (D4‐18).
D4.4 Literature
J. Ehlbeck, J. Kürth, H.J. Larsen, Original Articles B8, E4, STEP 1995.
Blumer H. (1975). Spannungsberechnung an Brettschichtholz mit gekrümmter Längsachse und veränderlicher
Trägerhöhe. Holzbau (Zürich), (6):158‐161; (7):191‐194; (8):235‐237.
Blumer H. (1979). Spannungsberechnung an anisotropen Kreisbogenscheiben und Satteldachträgern konstanter
Dicke. Veröffentlichung des Lehrstuhls für Ingenieurholzbau und Baukonstruktionen, Universität Karlsruhe.
Petersen C. (1993). Stahlbau: Grundlagen der Berechnung und baulichen Ausbildung von Stahlbauten.
3. überarbeitete und erweiterte Auflage. Vieweg Verlag Braunschweig Wiesbaden.
217
D5 Non‐reinforced notches
and holes
Original article: P. J. Gustafsson
Figure D5‐1 shows beams with various notches and holes. Notches or holes can consider‐
ably reduce the load‐bearing capacity of a member and should therefore be avoided in
the construction as early as the design stage. One way this can be done is using truss
structures instead of solid girders with large depths, although it is not always possible to
eliminate notches. For example, when timber beam flooring has to be brought to the
desired level, a clearance height or a fit between members has to be ensured. Old timber
constructions in particular tend to use many different notches and recesses for installing
joints, while large holes in glued laminated timber members are required, e.g. to accom‐
modate ventilation pipes.
A crack may propagate from a notch or a hole along the dashed lines and parallel to the
grain (Figure D5‐1). Such failure is brittle and happens suddenly, without any large de‐
formations or visible signs beforehand. Depending on the beam geometry, a crack which
progresses swiftly along the beam axis may trigger the complete failure of the beam in
question.
The cracks are generally initiated by a combination of shear and tensile stresses perpen‐
dicular to the grain, which may become excessive at the notch tip. According to the theo‐
ry of elasticity, the stresses at the tip of a sharp notch are even infinite (Figure D5‐2). In
this case, the extent of the stress remains undefined at the point of singularity. In reality,
however, infinite stress is not possible, since the strength of all material is finite. The
localised failure of the material results in a stress distribution at the time of crack growth
as shown schematically in Figure D5‐2 by the dashed line.
219
Non‐reinforced notches and holes
σt,90
ft,90
x
Figure D5‐2 Stress distribution at the notch tip in accordance with the (linear) theory of elasticity (solid line)
and assumed actual stress distribution (dashed line). (STEP 1995 Article B5)
D5.1 Fundamentals of fracture mechanics
Background
Since excessive stresses tend to be focused on a very small area, it is difficult – and also
irrelevant given the theoretically infinite stresses – to determine the load‐bearing capaci‐
ty of a member with a notch or a hole using a conventional stress criterion. Such a crite‐
rion would involve comparing the extent of stress at the critical cross‐section with frac‐
ture stress, which equates to the strength of the material itself. However, the theore‐
tically infinite stresses at the point where said stress peaks make this impossible. With
this in mind, determining the load‐bearing capacity involves either a combination of
analytical approaches and test results, or the use of fracture mechanics instead of a
conventional stress criterion.
220
Non‐reinforced notches and holes
Fracture mechanics – general
Fracture mechanics is a branch of the science of material strength. When exposed to high
levels of stress, a solid body responds with significant deformation or fracture. The inci‐
dence of fracture, involving separation and consequential loss of contact between two
parts of a body, is relevant to fracture mechanics. From an engineering perspective,
conversely, it is more important to calculate the extent of the load, which causes the
fracture. As a general rule, fracture mechanics differentiates three different kinds of
crack opening, see Figure D5‐3.
If there are no or only minimal stress concentrations, e.g. when uniform tensile or bend‐
ing stress is exerted on a member, the fracture load can be calculated using a conven‐
tional stress criterion. For high stress concentrations, meanwhile, for example at the tip
of a sharp notch or crack, another approach is required. The scope of linear‐elastic theory
allows appropriate calculation of the fracture load, either by evaluating the stress intensi‐
ty at the tip of the notch or evaluating the energy release rate as the crack propagates.
Although formal distinctions are established between both approaches, their basic pre‐
requisites coincide and at this point, we focus solely on the second method. Analysing
cracks as part of linear‐elastic theory is often referred to as linear‐elastic fracture me‐
chanics. Other models attempt to take the non‐linear behaviour of the material in the
vicinity of the crack tip into consideration, particularly the fracture process and crack
growth, which take place in the fracture process region immediately behind the crack
already having opened up. In linear‐elastic fracture mechanics, it is assumed that this
area of energy dissipating fracture process region is very small compared to the scale of
the actual structural details and thus mathematically a point, i.e. a region of zero size.
221
Non‐reinforced notches and holes
Energy release analysis – an example
h/2 h
h/2
a b
Figure D5‐4 Test specimen with an end crack. (STEP 1995 Article B5)
A beam with a large longitudinal crack is considered; exposed to load in accordance with
Figure D5‐4. It is assumed that for F = 0, no stresses and strains occur within the beam. In
accordance with linear‐elastic theory, the potential energy of the existing system, a com‐
bination of beams and load, equates to:
1
W F u (D5‐1)
2
whereby u represents the deformation at the point of loading.
The area of the longitudinal crack can now be considered a system of two cantilevers of
height h/2 and length a. In accordance with Euler‐Bernoulli beam theory, the total dis‐
placement u for both cantilevers amounts to:
F a 3
u 2 (D5‐2)
3E I
F 2 a 3
W (D5‐3)
3E I
The change in potential energy dW due to a small crack growth da is obtained from the
derivative:
F 2 a 2
dW da (D5‐4)
E I
222
Non‐reinforced notches and holes
This reduction in potential energy corresponds to a positive release of energy ‐dW and a
simultaneous increase in the fractured surface of b ∙ da. The released energy ‐dW applied
to the fractured surface b ∙ da is normally identified with G (after A. A. Griffith, who es‐
tablished the principles of fracture mechanics back in the 1920s):
‐dW F 2 a 2
G (D5‐5)
b da b E I
If the force F is so large that the crack begins to propagate, this signals that G has
reached its critical value Gc and this value corresponds to the potential of the material to
dissipate energy. For European conifers, the value Gc, depending on the density, is
around 150 to 600 J/m2 for mode I, tension perpendicular to the grain as shown in Fig‐
ure D5‐3 (Larsen and Gustafsson, 1990). Accordingly, the fracture criterion (the moment
at which the material fails) equates to
G Gc (D5‐6)
which elicits the fracture load Fc with G in accordance with equation (D5‐5) (where
I = b ∙ (h/2)3/12):
G c bE I G c E bh2
Fc (D5‐7)
a h 96 a
This equation includes two fundamentally important general results:
The material properties which dictate the resistance to crack propagation are
present in the modulus of elasticity E and fracture energy Gc. The tension strength
perpendicular to the grain of the material is not required to determine Fc.
The load‐bearing capacity is heavily dependent on the dimensions, namely the
computed strengths (e.g. Fc /(b ∙ h)) decline with increasing absolute size of the
specimen.
In this example, in applying equation (D5‐2), the implicit prerequisite was a slender sam‐
ple, namely the fact that the ratio h/a is small and the deformation u is solely attributable
to bending stress, while additional deformations caused by shear deformations and actu‐
al semi‐rigid clamping of the cantilever were not taken into consideration. However, this
method can also be applied to other crack geometries and fracture mechanisms (see
Figure D5‐3). In this case, naturally, equation (D5‐2) must be replaced with an equation
which applies to the actual geometric conditions.
223
Non‐reinforced notches and holes
D5.2 End‐notched members, theoretical
and experimental results
Using the above described method shown as an example, the force Fc during crack
growth can be calculated for a loaded member and with an end crack in accordance with
Figure D5‐5 (a) (Gustafsson, 1988):
hb G c h
Fc (D5‐8)
0.6
2
6 1
2
Gv E0
and are geometric conditions, as defined in Figure D5‐5. Gv and E0 indicate the shear
modulus and modulus of elasticity parallel to the grain.
Equation (D5‐8) also applies to rectangular notches (Figure D5‐5 (b)) and various notch
types in accordance with Figure D5‐1 (g, h and i), but not Figure D5‐5 (c). Compared to
equation (D5‐7), when deriving equation (D5‐8), various other geometries were also
taken into consideration via the factors and and the deformation u was determined,
not only due to bending stress (equation (D5‐2)), but also shear deformations and a semi‐
rigid clamping of the cantilever. Here, it was assumed that the individual deformation
components can be superposed, namely utotal = ubending + ushear + urotation. For small notch‐
es, namely for against 1.0, the resistance against failure is considerable. In this case,
the possibility of shear and bending failure of the net cross‐section ∙ h ∙ b must also be
considered.
αh FV h αh h
(a) (b) F
V
βh βh
αh h
(1−α)h
F
V
βh i(1−α)h
(c)
Figure D5‐5 Geometry of a beam with an end notch. (STEP 1995 Article B5)
224
Non‐reinforced notches and holes
Table D5‐1 lists the results of some short‐term tests (Gustafsson, 1988; Riberholt et al.,
1991). The values indicated are mean values; determined using dry wood with uniform
moisture content. The coefficient of variation of a test series is around 20. Table D5‐1
clarifies the actual low strengths of notched beams, even under favourable conditions.
Moreover, also clarified is the significant extent to which strengths are influenced by
dimensions, as could already theoretically be derived from equation (D5‐7). It is also
important to note that the mean values for Fc for specimens with knots in the vicinity of
the notch exceed those for specimens without knots (Larsen and Riberholt, 1972; Möhler
and Mistler, 1978).
Table D5‐1 Test results. Strength (mean values) for notched beams.
225
Non‐reinforced notches and holes
D5.3 Background to EC 5 equations for notched members
Equation (D5‐8) was modified to develop a simplified design equation for EC 5 (Larsen,
1992). The ratio E0/Gv was set to 16. In addition, the introduction of the “new” material
parameter Gc was avoided by assuming that E 0 G c is proportional to the shear
strength fv of the material. The corresponding proportionality constant kn was deter‐
mined by testing. A range of constants emerged for solid timber, glued laminated timber
and LVL and the final simplified form of the equation (D5‐8) elicited the factor kv shown
in equation (D5‐10). In addition, via tests (Riberholt et al., 1991) a function f(i) was de‐
termined, which allowed the influence of the chamfer i (Figure D5‐5 (c)) to be taken into
consideration. Through these considerations, in EC 5, the risk of a crack growth in a
notched member can be taken into consideration through a formal reduction by a factor
kv of the design shear strength fv,d of the net cross‐section ∙ h ∙ bef (with bef = kcr ∙ b
according to equation (D1‐17)):
3 Vd
d k v f v,d (D5‐9)
2 h b ef
The reduction factor kv (≤ 1.0) is calculated in accordance with:
kv
k n 1 1.1 i
1.5
h (D5‐10)
h 2 0.8
1
2
226
Non‐reinforced notches and holes
h β i
1,0 1 200 0,5 0
kv
0,9
2 100 0,5 0
0,8
3 200 2,0 0
0,7
4 200 0,5 4,0
0,6
0,5
0,4
0,3
0,2
0,1
0
1,0 0,9 0,8 0,7 0,6 0,5 0,4 0,3 0,2 0,1 0
α
Figure D5‐6 Factor kv depending on for solid timber with various h, and i. (STEP 1995 Article B5)
D5.4 Climate‐related tension perpendicular
to the grain in notched members
The NA regulations stipulate that non‐reinforced, notched beams may only be used in
service classes 1 and 2 and that notches in service class 3 must be reinforced (see Article
D8). The background to this rule corresponds to the regulation for curved, double‐
tapered and pitched cambered beams. In addition to external forces (e.g. dead weight,
wind or snow), climate fluctuations intensify tension perpendicular to the grain in the
area of the notch. These tensile stresses perpendicular to the grain, generated by swell‐
ing and shrinking processes, must either be taken into account by applying reinforce‐
ments or, as stipulated in the NA, the use of notched beams must be limited to service
classes 1 and 2.
D5.5 Holes in glued laminated timber members
ℓy a ℓz a a ℓz a ℓy
hro
d
d hd h
r = 15 mm hru
b
ℓA
ℓA
Figure D5‐7 Non‐reinforced holes.
227
Non‐reinforced notches and holes
EC 5 does not specify any design equations for beams with holes (holes see Figure D5‐7),
the only regulations for which are included in the NA. The regulations only apply to non‐
reinforced holes in glulam and LVL, which must not be subject to systematic tension
perpendicular to the grain in the hole area. Similar to the regulations for notched beams,
non‐reinforced holes must only be used in service classes 1 and 2. Reinforcements of
holes are covered in Article D8.
When reference is made to holes in beams of height h, this refers to holes of d > 50 mm,
hd is the height of the hole, hro is the remaining height above the hole and hru that under‐
neath the hole (Figure D5‐7). Unlike the fracture mechanics design format for notched
beams regulated in EC 5 (which was formally converted in a shear stress verification
there), verifications for beams with holes are based on the Euler‐Bernoulli beam theory,
extended with a volume effect. Comprehensive works have shown that holes in beams
under bending stress show crack propagation as illustrated in Figure D5‐1 (d) to (f), which
means tensile stresses perpendicular to the grain and therefore forces Ft,90 occur at the
corners of the holes, the extent of which can be calculated from a shear force component
Ft,V and moment component Ft,M (Kolb and Epple, 1985):
The shape of the hole also influences failure and must be taken into consideration in
terms of geometric factors (Kolb and Epple, 1985). Kolb and Epple (1985) were able to
use FE simulations to confirm the validity of the approach shown in equation (D5‐11) and
explained in the following.
Shear force component
The shear stresses occurring in a rectangular beam without a hole are shown in Fig‐
ure D5‐8 schematically. Via the crack propagation path, a portion of the shear force, Ft,V,
corresponding to the hatched area shown in Figure D5‐8 in the shear stress diagram,
must be transmitted to the remaining cross‐section above or underneath the hole.
hro
V
hd h y
z V z
hru
a b
Figure D5‐8 Shear stress diagram in a beam with rectangular hole. (according to Blass et al., 2005,
Figure 11/12)
228
Non‐reinforced notches and holes
This component Ft,V can be determined via the equations of the Euler‐Bernoulli beam
theory (geometric definitions see Figure D5‐8). The shear stress diagram of a rectangular
cross‐section can be calculated as follows:
V S ( z) 3 V
2
z
( z) 1 4 2 (D5‐12)
I b 2 b h h
The integral over the hatched area reveals the shear force component Ft,V:
0 3 V 0 z2 1 h h2
F t,V b ( z) dz 1 4 2 dz V d 3 d2
2 h hd h 4 h h (D5‐13)
h
d
2 2
The equation specified in the NA corresponds precisely to equation (D5‐13), Ft,V,d is the
design value component of the shear force Vd:
Vd hd h2
F t,V,d 3 d2 (D5‐14)
4h h
hro
V
hd h y
45° z z
V
hru
a
Figure D5‐9 Shear stress diagram in a beam with circular hole. (according to Blass et al., 2005, Figure 11/12)
In circular holes, the crack appears approximately at a point on the circumference, where
the diameter, at an angle of 45°, intersects the hole edge (Figure D5‐9) and the integral
for the shear stresses declines in this case below the figure for rectangular holes. In this
case, instead of hd, the value 0.7 ∙ hd is used.
However, the shear force component derived in this manner assumed the hole was pre‐
cisely centrally positioned, which renders it invalid for asymmetrically configured holes.
229
Non‐reinforced notches and holes
Moment component
Kolb and Epple (1985) could show with test results on beams with holes, that such beams
were prone to tension failure perpendicular to the grain in hole corners also in areas not
exposed to internal shear forces. Accordingly, they could confirm that the tensile force
perpendicular to the grain Ft,90 in equation (D5‐11) also had to include a moment compo‐
nent. This empirical moment component was iteratively adapted to the test results and
determined as follows:
M M
F t,M 0.008 (D5‐15)
hr hr
The format shown in equation (D5‐15) indicates that the moment component is deter‐
mined from the ratio of the bending moment M at the critical cross‐section to the edge
distance of the crack to be expected; the coefficient was iteratively determined as
0.008. In the NA, in turn, the design value Md of the critical bending moment is used, to
determine the design value of the moment component Ft,M,d.
For rectangular holes therefore, the required height hr is the smallest value of the re‐
maining height underneath or above the hole, h = min(hro; hru), since this is where the
tensile stresses perpendicular to the grain peak and hence where cracks emerge. This
regulation is excessively conservative for circular holes, since the maximum tensile
stresses perpendicular to the grain only occur at the point, at which the circle diameter is
at an angle of 45°. The following applies in this case:
Using the design value for tensile force perpendicular to the grain Ft,90,d, determined in
accordance with equation (D5‐11), the tensile strength perpendicular to the grain can
thus be verified:
where
b Beam width at hole
ft,90,d Design value of tensile strength perpendicular to the grain of the glulam or LVL
kt,90 = min(1; (450/h)0.5), h in mm
ℓt,90 = 0.5 ∙ (hd + h) for rectangular holes
ℓt,90 = 0.353 ∙ hd + 0.5 ∙ h for circular holes
230
Non‐reinforced notches and holes
b ∙ ℓt,90 in this case corresponds to the area subject to tensile stress perpendicular to the
grain, whereby ℓt,90 is the length starting from the hole corner. However, the stresses are
not uniformly dispersed over the area b ∙ ℓt,90, but the tensile stresses perpendicular to
the grain peak in the hole corners, similar to a notch corner, see Figure D5‐2. These stress
peaks are taken into consideration, whereby in equation (D5‐17) the tensile force per‐
pendicular to the grain Ft,90,d is dispersed in a triangular stress, which explains the 0.5
factor. The coefficient kt,90 takes into consideration the impact of height, which makes it a
coefficient in accordance with Article D3 and thus takes the impact of volume on strength
into consideration.
It is important to note the fact at this point that rounded corners in holes are stipulated,
to prevent areas of singularity and reduce the stress peaks forming in the corners, see
also details of the minimum radius in Figure D5‐7.
D5.6 Literature
P.J. Gustafsson, original Article B5, STEP 1995.
Blass H.J., Ehlbeck J., Kreuzinger H. and Steck G. (2005). Erläuterungen zur DIN 1052:08‐2004.
Herausgeber: Deutsche Gesellschaft für Holzforschung. Bruder‐Verlag, Karlsruhe.
Gustafsson P.J. (1988). A study of strength of notched beams. Paper 21‐10‐1, CIB‐W18, Parksville.
Kolb H. and Epple A. (1985).Verstärkung von durchbrochenen Brettschichtbindern.
Forschungsvorhaben
I.4‐34810 der MPA Stuttgart.
Larsen H.J. and Riberholt H. (1972). Tests with not classified structural timber. Rapport nr R 31 (in Danish),
Technical University of Denmark.
Larsen H.J. and Gustafsson P.J. (1990). The fracture energy of wood in tension perpendicular to the grain ‒
results from a joint testing project. Paper 23‐19‐2, CIB‐W18 Meeting 23, Lisbon.
Larsen H.J. (1992). Latest development of EC 5. Paper 25‐102‐1a, CIB‐W18 Meeting 25, Åhus.
Möhler K. and Mistler H.‐L. (1978). Untersuchungen über den Einfluß von Ausklinkungen im Auflagerbereich
von Holzbiegeträgern auf die Tragfestigkeit. Forschungsbericht, Lehrstuhl für Ingenieurholzbau und
Baukonstruktionen, Universität Karlsruhe.
Riberholt H., Enquist B., Gustafsson P.J. and Jensen R.B. (1991). Timber beams notched at the support.
Report TVSM‐7071, Lund University.
231
D6 Glued composite members
Original articles: K. H. Solli, H. J. Blass, J. G. M. Raadschelders
This article explains the design of glued composite components regulated in EC 5 with
thin webs or thin flanges. At this point, determining internal forces is crucial and ex‐
plained in detail. Glued components can be used in very wide‐ranging applications, fre‐
quent examples of which include glued thin‐webbed beams (I‐joists or box girders, Fig‐
ure D6‐1 on the left) and glued thin‐flanged beams (stressed skin panels, Figure D6‐1 on
the right).
bef b ef
1 1 bc,ef bc,ef bc,ef
2 2 2
h
h f,c
f,c
1 1 hw 1 1 hw
h h
f,t
bw bw f,t
bw bw
b b bf bf
Figure D6‐1 Thin‐webbed beams (left) and thin‐flanged beam (right). 1‐1 are glue lines.
While the webs are relatively thin compared to flanges in I‐joists and box girders,
stressed skin panels are frequently seen in the form of beams with thick webs and thin
flanges. For both component types, thin‐webbed and thin‐flanged beams, linear strain
distribution over beam depth is assumed, although shear deformations in the flanges of
stressed skin panels mean normal stresses in the centre plane of the flanges are non‐
linearly distributed. A separate explanation of both component types is given in the fol‐
lowing sections accordingly.
Glued components often comprise various materials. Since various stiffnesses affect the
stress distribution, the stiffness of individual components must also be taken into consid‐
eration during the design. Meanwhile, if the materials used also exhibit differing creep
behaviour (various creep coefficients kdef, see also Article B2), this should also be taken
into consideration during the design. In such cases, the stress distribution for the instan‐
taneous and final states must also be determined and for the instantaneous state, the
mean values of stiffness properties are normally used. The stress distribution in the final
state, meanwhile, can be calculated in accordance with EC 5 using the final mean values
of stiffness properties. Equation (D6‐1) shows how final values are determined, citing the
example of the modulus of elasticity.
233
Glued composite members
E mean
E mean,fin (D6‐1)
1 2 k def
The rule adopted is that 2 = 1.0 for permanent loads, otherwise the combination coeffi‐
cient should be used for the quasi‐permanent load component. For load combinations,
2 may be proportionally weighted.
D6.1 Glued thin‐webbed beams
A glued thin‐webbed beam (examples Figure D6‐1 on the left) comprises three main
components:
Flanges,
Web,
Glued joints between flanges and web.
Although the flanges are frequently made of finger‐jointed solid timber, they can also be
manufactured from other materials such as glued laminated timber or laminated veneer
lumber (LVL). Their main function is to absorb stresses generated by bending moments
and axial forces. Given the generally small dimensions of the flanges, it is imperative to
ensure that the material has few and minimal defects.
The web can be made of various types of wood materials, such as plywood, parti‐
cleboards, fibreboards or OSB. There are also beam shapes including truss‐like webs,
particularly formwork girders. The web primarily absorbs the stresses from shear forces,
while for longer beams, joints in the webs may be required. If the web splices are in an
area subject to lower shear forces, they can be configured as butt joints, while in other
cases, the web splice should be force‐locking. There is also often a need to reinforce the
web in the support area, which can be done via additional wood‐based panels, which are
nailed or preferably glued onto the web. The reinforced web at the load input position or
support must be configured such as to ensure that the shear forces exerted can be effec‐
tively transmitted.
234
Glued composite members
Production
Glued, thin‐webbed beams are normally manufactured on an industrial scale and it is
crucial to ensure the correct gluing temperature to optimise the joint between the web
and flanges. It is equally important to ensure that the surfaces of the glued joints are
planed or milled immediately before the glue is applied and that the moisture content of
the flange and web material is checked.
Use of glued thin‐webbed beams
Glued thin‐webbed beams, in relation to their dead weight, have exceptional load‐
bearing capacity and stiffness, while their low weight also facilitates transport and as‐
sembly. Moreover, using tools, they can be easily processed by hand. As a general rule,
these beams can be used instead of solid timber beams. Glued thin‐webbed beams are
often used to construct floor structures, when it is difficult to procure sufficiently large
cross‐sections made of solid timber or if the cost of using glued laminated timber (from 5
to 8 m long) would be prohibitive. When used in floor and wall structures, beam depths
of 300 to 500 mm are possible, which helps pave the way to install a range of technical
building equipment. The construction depth also guarantees sufficient room for any
insulation material that may be required. In countries subject to cold winters, the re‐
quired insulation thickness for the cross‐sectional dimensions of studs is a crucial param‐
eter and here, if a glued thin‐webbed cross‐section is used, the proportion of insulation
material can be maximised. In service class 3, however, the use of glued thin‐webbed
beams is infeasible due to the restrictions on using web materials in this service class.
Special production and transport aspects
Stiffness about the z‐axis is very low compared to that about the y‐axis (see Figure D6‐2
for axis definition), which must be noted during production and transport from the man‐
ufacturing company to the building site. Web materials are also prone to damage during
the transport and assembly processes and beams must be protected against moisture
during the construction phase. If the moisture content in the web soars, this considerably
increases the risk of permanent deformations in the finished construction.
Lateral stability
Flanges carrying compressive stresses must be supported, to avoid lateral deflection and
buckling. For beams used as simply supported beams in floors, meanwhile, the joint
between the compression flange and floor panel generally suffices to prevent lateral
deflection. It is also important to ensure that in the lower chord, compressive stresses
can be generated e.g. at the intermediate supports of continuous beams.
235
Glued composite members
Effective cross‐sectional values
For subsequent calculations, it is assumed that flanges and web are glued together such as
to form a single constructive unit. It is also assumed that any strains will run linearly over
the beam depth. According to Hooke’s law for linear‐elastic stress‐strain behaviour, stresses
at a particular point can be calculated from the product of strain and the modulus of elas‐
ticity. Since a beam may be made of materials including various moduli of elasticity, this
means that stresses over beam depth are non‐linearly distributed. Figure D6‐2 shows an
example of how the stresses due to bending moment vary in the beam cross‐section.
σf,c,max
σf,c
σw,c,max
0,5hf,c
hf,c
bw bw hw z
hf,t y
0,5hf,t
b b σw,t, max
σf,t
σf,t,max
Figure D6‐2 Example of the stress distribution in glued I‐joists and box girders, index w = web,
Index f = flange. (STEP 1995 Article B9)
Given the fact that the moduli of elasticity in the cross‐section vary, as a general rule,
effective cross‐sectional values are calculated. In this case for example, the entire profile
can be considered a cross‐section of homogenous material with properties equivalent to
the flange material. When adopting such an approach, the web width must be reduced
by the ratio of the average moduli of elasticity. All specified equations apply for perma‐
nent loads with 2 = 1.0; while for other loads the rule is (1 + 2 ∙ kdef) instead of (1 + kdef),
see equation (D6‐1).
The effective cross‐sectional area (index f = flange, index w = web) is thus determined
using a weighted web surface:
E mean,w 1 k def,f
A ef A f Aw (D6‐2)
E mean,f 1 k def,w
236
Glued composite members
The effective second moment of area likewise:
E mean,w 1 k def,f
I ef I f I w (D6‐3)
E mean,f 1 k def,w
The approach used here is the so‐called composite theory, for which all known equations
on stress and deformations derived via Euler‐Bernoulli beam theory apply. The only dif‐
ference here is that the moduli of elasticity are no longer constant. Given that fact, to
now take account of a full cross‐section and a single modulus of elasticity, the cross‐
sectional values have to be adapted, to take the differences in stiffness and hence the
different deformation behaviour, into consideration. As a reference value E0 for the
modulus of elasticity, the modulus of elasticity of the key material, the flanges in this
case, is often used. The individual areas can then be converted into effective areas using
the ratio of the moduli of elasticity Ei/E0: Aef,i = Ei/E0 ∙ Ai. The effective second moments of
area thus result in I ef,i,y z 2 E i E 0 dA and the internal forces and moments in
Mi,y = E0 ∙ Ief,i,y ∙ ‘, Mi,z = E0 ∙ Ief,i,z ∙ ‘, Ni = E0 ∙ Aef,i ∙ (where ‘ = derivative of the angle).
Verification of stresses in the flanges
Within a beam cross‐section under bending stress, the flanges are particularly prone to
axial compressive and tensile stresses, amid a small bending component. With a symmet‐
rical cross‐section, which is only required to absorb a bending moment, the absolute
stresses in the compression and tension flanges are equal. If a beam is subject to addi‐
tional axial stresses, the maximum stresses in the outer fibres of the compression flange
can be calculated as the sum of stresses due to the bending moment Md and axial com‐
pressive force Fd, where ymax equates to the distance from the outer fibres to the neutral
axis, tension flange analogous:
Md F
f,c,max,d y max d (D6‐4)
I ef A ef
Equation (D6‐4) clearly shows how cross‐sectional values are replaced by effective values
in accordance with composite theory, while all other details remain as to date, based on
Euler‐Bernoulli beam theory. Conversely, to calculate axial stresses in the centre of gravi‐
ty of the compression flange, equation (D6‐4) must take into account the distance yc
between the centre of gravity and the neutral axis of the beam (tension flange analogous):
Md F
f,c,d yc d (D6‐5)
I ef A ef
237
Glued composite members
When calculating the stresses, the position of the neutral axis thus always has to be de‐
termined. For symmetrical cross‐sections, this is trivial and ymax = h/2 with the beam
depth h (see equations (D6‐23) and (D6‐24) for an exemplary derivation of the position of
the neutral axis).
The calculated stresses must now be compared with the design strength values of the
compression flange:
In Equation (D6‐6) the verification of the compression flange also takes possible buckling
into consideration, which involves the use of the coefficient kc already familiar from Arti‐
cle D2 (there coefficient kc,z in equation (D2‐7)). This coefficient kc may be conservatively
determined (particularly conservative for box girders) in accordance with EC 5 with the
following slenderness ratio z (steps: z → rel → kz → kc):
c ef ef ef
z 12 (D6‐7)
b z i I A 2
b 12
ℓc is the distance between the lateral supports of the compression flange, while b is the
flange width (see Figure D6‐2). The compression flange is considered a column without
taking the lateral support from the web into account. The stresses in the tension flange
are calculated correspondingly, but without the buckling coefficient kc.
Verification of axial stresses in the web
The web must particularly absorb shear stresses generated by shear forces, although
proportions of the bending moment and axial forces generate additional normal stresses
in the web, which is why the load‐bearing capacity of the web to accommodate these
stresses must also be checked. Given the assumption of linear strain distribution over the
beam depth, the axial stresses w in the web can be calculated with composite theory in
the following general form from the stresses in the flange f:
E mean,w
w f (D6‐8)
E mean,f
238
Glued composite members
Respectively, taking into consideration the load duration and service class:
E mean,w 1 k def,f
w f (D6‐9)
E mean,f 1 k def,w
To calculate the axial stresses w in the web, the first task is to determine the flange
stresses f at the point at which w should be calculated. This is performed similarly to
the equations (D6‐4) and (D6‐5):
Md F
f,d y1 d (D6‐10)
I ef A ef
The maximum stress in the compression zone of the web w,c,max,d can then be calculated
with equations (D6‐10) and (D6‐9) as:
Md F d E mean,w 1 k def,f
w,c,max,d y w,c (D6‐11)
I ef A ef E mean,f 1 k def,w
yw,c here is the distance of the neutral axis of the beam from the point where the stress
value is calculated.
This stress must meet the following condition:
The maximum stress in the tension zone of the web is correspondingly calculated as:
Md Fd E mean,w 1 k def,f
w,t,max,d y w,t (D6‐13)
I ef A ef E mean,f 1 k def,w
yw,t is the distance of the neutral axis of the beam from the point where the stress value
is calculated.
This stress must meet the following condition:
fw,c,d and fw,t,d are the design values for compressive and tensile bending strengths of the
web. In the absence of any other values, the design values for in‐plane tensile and com‐
pressive strength of the web should be used.
239
Glued composite members
Verification of shear stresses in the web
If equation (D6‐15) is satisfied, no more accurate shear buckling analysis is required:
h w 70 b w (D6‐15)
In this case, the design shear force in each web Fw,v,Ed must maintain the following condi‐
tion (for geometric details, see Figure D6‐2):
Verification of the shear stresses in the glue line between flanges and web
The strength of a correctly glued joint between web and flanges exceeds the shear
strength of the flange and web materials and the weakest element in such a joint is nor‐
mally the rolling shear strength of the web fv,90. It is assumed that the design shear stress
mean,d in the cross‐section under consideration (namely in both glued joints between the
web and compression or tension flange, see Figure D6‐2) is uniformly distributed if flange
heights hf ≤ 4 ∙ bef, which is why the shear stress can be determined using Euler‐Bernoulli
beam theory:
Vd S f
mean,d (D6‐17)
I ef g
where
Sf First moment of area of a flange, related to the neutral axis of the beam
ℓg = 2 ∙ hf,c(t) For cross‐sections in accordance with Figure D6‐2
240
Glued composite members
The shear stress should meet the following condition:
where
fv,90,d Design rolling shear strength of the web
hf Either hf,c or hf,t, depending on whether the glued joints in the
compression or tension flange are verified
bef = bw For box girders
bef = bw/2 For I‐joists
Equation (D6‐18) must be allocated depending on flange height hf, since the shear stress‐
es in the glued joint between the flange and web are not actually uniformly distributed.
While this uneven distribution can be ignored up to a glue line depth hf of 4 ∙ bef, at
greater flange depths, either the shear stress has to be increased or, as shown in equa‐
tion (D6‐18), it must be reduced to take account of this non‐uniform distribution with
higher maximum values. However, the equations shown to date to determine normal and
shear stresses in the individual components can also be formulated more generically, as
will be shown at the end of this article.
D6.2 Glued thin‐flanged beams (stressed skin panels)
The previous section focused on glued components with thin webs and thick flanges.
However, glued components with thin flanges also exist, for which a potential stability
failure of the flanges (buckling) as well as their shear deformations, both due to their
significant slenderness, must be taken into consideration. (Glued) stressed skin panels
(Figure D6‐3) can be considered thin‐flanged beams when considering effective tributary
areas. The glued joints between the webs and flanges (sheathing) are deemed rigid in the
assessment. Consequently and also at this point, a linear strain distribution is assumed
over the depth of the composite component, Figure D6‐4 (b). The joints between flanges
and web are also frequently implemented with mechanical fasteners, whereupon the slip
between flanges and web must be taken into consideration when stresses are deter‐
mined, as in Figure B6‐4 (d). Although such mechanically jointed elements do not con‐
stitute part of the current article, they can be designed using the methods set out in
Article D7.
241
Glued composite members
H V H V H V
Effective flange width
Shear deformations in the plane of the flanges mean the normal stresses in the centre
plane of the flanges are not uniformly distributed (see Figure D6‐5). The extent to which
flanges contribute to bending stiffness or bending capacity of the composite component
thus declines with increasing lateral distance from the webs. The extent of this stress
reduction particularly depends on the ratio of the bf/ℓ and E/G values and here, bf repre‐
sents the clear distance between webs, ℓ the span, E the modulus of elasticity of the
flange in the direction of the webs and G the shear modulus of the flange, while the
mean values for in‐plane loading are used for both moduli. The effective flange width
declines with increasing E/G and bf/ℓ values.
τyx
x
y
τyx
b ef z
Figure D6‐5 Stress distribution in the flange. (STEP 1995 Article B10)
242
Glued composite members
The derivation of the effective flange width taking into consideration the shear defor‐
mations of flanges is specified in Möhler et al. (1963) and the resulting ratio obtained
between effective and actual flange widths bef/bf for simply supported beams under a
uniformly distributed load is:
b ef 1 tanh 1 2 tanh 2 2
bf
2 2
1 2 bf (D6‐19)
where
1 b f 2 b f
1 and 2
2 2
1 a a2 c and 2 a a2 c
Ey Ey
a xy and c
2G Ex
xy is the Poisson’s ratio of the flanges.
To allow the Euler‐Bernoulli beam theory to be used when designing thin‐flanged beams,
the concept of effective flange width is used. Here, the effective flange width bef is de‐
fined as the width of a fictitious flange, in which the normal stresses in the centre of
gravity of the flange calculated with the Euler‐Bernoulli beam theory correspond to the
related maximum value in accordance with the correct theory (Möhler et al., 1963). The
entire normal force in the flange as well as the section modulus of the overall cross‐
section remains identical in both cases (Beam theory and correct theory). In EC 5, the
approximation values for the effective flange width bef for internal webs are specified as
follows (for notations, see Figure D6‐6):
The corresponding approximation equation for edge webs is:
243
Glued composite members
Table D6‐1 from EC 5 provides maximum values for the effective flange widths bc,ef and
bt,ef, whereby the buckling failure is only relevant for the compression flange (determin‐
ing bc,ef). Accordingly, the effective widths bc,ef or bt,ef should not be assumed to exceed
the maximum value calculated in accordance with equation (D6‐19) which takes shear
deformation into consideration. Moreover, bc,ef should not be assumed to exceed the
maximum value in accordance with Table D6‐1 which takes buckling of the flanges into
consideration. Table D6‐1 in EC 5 deviates considerably from the table given in DIN 1052
(Table 5). Not only do the given maximum values differ slightly, but the information pro‐
vided is also far less detailed. The NA also stipulates that laminated veneer lumber (LVL)
with perpendicular layers can be treated like plywood and cross‐laminated timber like OSB.
Figure D6‐7 shows the course of effective flange width in accordance with equation (D6‐19)
and the corresponding approximation in accordance with EC 5 for shear deformations. In
the most practical cases, the ratio bf/ℓ is below 0.3.
bef bef
0,5bc,ef 0,5bc,ef
hf,c
1 1 1 1
hw
bf bw bf bw hf,t
0,5bt,ef
Figure D6‐6 Stressed skin panel (thin‐flanged beam). (STEP 1995 Article B10)
Table D6‐1 Maximum values of the effective flange width taking shear deformation (bc,ef and bt,ef) and
buckling (bc,ef) into consideration.
244
Glued composite members
1,0
(bef - bw)
0,8
a
bf
0,6
b
0,4
d c
0,2
0,0
0,0 0,2 0,4 0,6 0,8 1,0
bf /l
Figure D6‐7 Effective flange width. a particleboard equation (D6‐19), b particleboard EC 5,
c plywood equation (D6‐19), d plywood EC 5. (STEP 1995 Article B10)
Flanges subject to compression are at risk of buckling and a more accurate verification
can be conducted e.g. in accordance with the method of Halász and Cziesielski (1966). In
the absence of any more accurate buckling verification, the clear distance between webs
bf should be assumed as not exceeding double the effective flange widths bc,ef and bt,ef
determined while taking buckling into consideration (Table D6‐1). For thin‐flanged beams
with nailed or stapled joints between flanges and webs, the withdrawal capacity of the
fasteners must be sufficient to prevent any buckling of the flanges.
Verification
Thin‐flanged beams (stressed skin panels) are computed assuming rigid joints between
web and flange (sheathing). The procedure resembles that for thin‐webbed beams (Sec‐
tion D6.1), only differing in terms of the shear verification of (in this case thin sheathing)
flanges (in Section D6.1 equation (D6‐18)). Here too, a uniform distribution of shear
stress mean,d in the glued joint in question (Section 1‐1 in Figure D6‐6) is assumed and the
determined shear stress should meet the following condition:
where
fv,90,d Design rolling shear strength of flange (sheathing)
hf Either hf,c or hf,t
Also relevant is the fact that 8 ∙ hf should be replaced by 4 ∙ hf, if U‐shaped cross‐sections
are considered.
245
Glued composite members
Generalisation
b1
h1
a1+a2
z
h2
b2
Figure D6‐8 Cross‐section of a glued stressed skin panel. 1 designates the sheathing and 2 the web.
(STEP 1995 Article B10)
In the following section, using the example of a glued stressed skin panel with one‐sided
sheathing, it is shown how bending stiffnesses and stresses in the different components
can be calculated while taking a range of material stiffnesses into consideration. All geo‐
metric details are taken from Figure D6‐8. The first task is to determine the position of
the neutral axis and likewise here, applying the composite theory, the influence of differ‐
ent moduli of elasticity has to be taken into consideration:
2
z i E i Ai h 1 2 E 1 A 1 h 1 h 2 2 E 2 A 2 E 2 A 2 h1 h 2
h1
a 1 i=1 2 (D6‐23)
E 1 A1 E 2 A 2 2 2 E 1 A1 E 2 A 2
E i Ai
i=1
h1 h 2
a2 a1 (D6‐24)
2
The effective bending stiffness is:
E I ef E i I i E i A i a i2
2
(D6‐25)
i=1
The compressive stress in the centre of gravity of the flange is:
M
1,c E a (D6‐26)
I ef 1 1
E
246
Glued composite members
The compressive stress in the outer upper fibre of the flange is:
M h
1,c,max E 1 a1 1 (D6‐27)
E I ef 2
The tensile stress in the centre of gravity of the web is:
M
2,t E a (D6‐28)
E I ef 2 2
The bending stress in the outer lower fibre of the web is:
M h
2,t,max E 2 a2 2 (D6‐29)
E I ef 2
The shear stress in the joint between flange and web is (required to verify equation
(D6‐22)):
V S1 V A1 a 1
max E 1 E (D6‐30)
E I ef b 2 E I ef b 2 1
S1 is the first moment of area of the flange, which is why E1 must be selected as the mod‐
ulus of elasticity, b2 is the width of the glue line.
Equations (D6‐26) to (D6‐30) correspond to equations (D6‐4) or (D6‐17) already given
and are only formulated generally, without being based on a modulus of elasticity E0, so
recalculations such as in equation (D6‐11) are no longer needed when the stresses in the
components should be calculated by a parameter other than the reference modulus of
elasticity E0. The design stresses must not exceed the corresponding design strengths.
The stresses determined here correspond to those in the instantaneous state, while for
the stresses in the final state, the moduli of elasticity must be re‐divided by (1 + 2 ∙ kdef)
(equation (D6‐1)). Composite beams with mechanical fasteners between the web and
flange can be calculated with the ‐method presented in Article D7.
247
Glued composite members
D6.3 Literature
K.H. Solli, H.J. Blass, J.G.M. Raadschelders, original Articles B9, B10, STEP 1995.
Von Halász R. and Cziesielski E. (1966). Berechnung und Konstruktion geleimter Träger mit Stegen aus
Furnierplatten. Berichte aus der Bauforschung Heft 47, pp 75‐118.
Möhler K., Abdel‐Sayed G. and Ehlbeck J. (1963). Zur Berechnung doppelschaliger, geleimter Tafelelemente.
Holz als Roh‐ und Werkstoff 21:328‐333.
248
D7 Mechanically jointed members
and CLT elements
Original article: H. Kreuzinger
Mechanically jointed beams comprise at least two individual components joined either by
mechanical fasteners or interim layers with comparatively low shear stiffness. Shear
forces are applied and generate relative displacements between the components, due to
semi‐rigidity of the joints with mechanical fasteners or shear deformations in the interim
layers. As a result of these relative displacements, the Bernoulli hypothesis that plane
sections remain plane no longer holds and the Euler‐Bernoulli beam theory cannot be
applied to the overall cross‐section of the mechanically jointed beam. Longitudinally, the
components of mechanically jointed beams are generally without abutment. The joints
between the individual components mainly serve to accommodate shear forces, while
the mechanical fasteners used include nails, screws, bolts, connectors or punched metal
plate fasteners.
While glued joints are deemed rigid (see Article D6), interim layers with low shear stiff‐
ness, such as perpendicular layers in cross‐laminated timber subject to rolling shear, can
cause similar relative displacements between longitudinal layers to joints with mechani‐
cal fasteners. The load‐bearing capacity and stiffness of mechanically jointed beams lie
between the corresponding values for composite beams whose individual components
are not jointed and composite beams with rigid (glued) joints. Amid rising composite
stiffness between components, the load‐bearing capacity and stiffness of the overall
beam follow suit. Figure D7‐1 clarifies the effect of a semi‐rigid joint in terms of overall
deformation and the course of bending stresses over the beam depth for a full cross‐
section (A), a cross‐section comprising three loosely superimposed individual cross‐
sections (C) and a cross‐section comprising three individual cross‐sections connected via
semi‐rigid joints (B).
Assuming constant moduli of elasticity for the wood in beams A, B and C, as well as iden‐
tical loads and span, the maximum value of bending stress in the outer fibre C in beam C
in Figure D7‐1 is three times as large as the corresponding value of bending stress in the
outer fibre A for the full cross‐section of beam A. For a simply supported beam with
uniformly distributed load, this results in a ratio between the maximum deflection uC for
beam C and the maximum deflection uA for beam A of nine. If the individual components
249
Mechanically jointed members and CLT elements
are mechanically jointed, beam B, the bending stress in the outer fibre and deflection can
be lowered compared to beam C. Assuming constant moduli of elasticity and dimensions
of the beams, A ≤ B ≤ C = 3 ∙ A and uA ≤ uB ≤ uC = 9 ∙ uA applies for the system shown
in Figure D7‐1.
A VA
B VB
C VC
hi hi
h hi hi
hi hi
q
uA
uB
uC
Figure D7‐1 Deflection and bending stress distribution of a full cross‐section (A), a cross‐section comprising
three individual cross‐sections connected via semi‐rigid joints (B) and a cross‐section comprising
three loosely superimposed individual cross‐sections (C).
Figure D7‐2 Selection of possible cross‐sectional types for mechanically jointed beams.
Composite beams may comprise solely wood components or other construction materi‐
als too. Using suitable mechanical fasteners means that, in principle, any desired con‐
struction materials can be combined to form a composite beam. Examples of possible
cross‐sectional types for mechanically jointed beams are shown in Figure D7‐2 (timber‐
concrete‐composite see also Article D10). The key calculation methods used for mechan‐
ically jointed beams are the ‐method and the shear analogy method. The ‐method for
mechanically jointed beams and columns with up to three single components is optimal
for calculation by hand. Conversely, the shear analogy method is valid for any cross‐
sectional types also containing more than three individual components and allows freely
variable component cross‐sections, load or connection configurations along the beam
axis. However, the shear analogy method requires the use of computer software (plane
frame analysis programs), which can consider shear deformations in bar elements.
250
DĞĐŚĂŶŝĐĂůůLJũŽŝŶƚĞĚŵĞŵďĞƌƐĂŶĚ>dĞůĞŵĞŶƚƐ
dŚĞ͞JͲŵĞƚŚŽĚ͟;DƂŚůĞƌ͕ϭϵϱϲͿŝƐƐƵŝƚĂďůĞĨŽƌĚĞƐŝŐŶŝŶŐďĞĂŵƐĂŶĚĐŽůƵŵŶƐǁŝƚŚƵƉƚŽ
ƚŚƌĞĞ ŵĞĐŚĂŶŝĐĂůůLJ ũŽŝŶƚĞĚ ĐŽŵƉŽŶĞŶƚƐ ĂŶĚ ǁŚŝĐŚ ĂƌĞ ŶŽƚ ůŽŶŐŝƚƵĚŝŶĂůůLJ ĂďƵƚƚĞĚ͘ &Žƌ
ŵĞĐŚĂŶŝĐĂůůLJ ũŽŝŶƚĞĚ ďĞĂŵƐ ǁŝƚŚ ŵŽƌĞ ƚŚĂŶ ƚŚƌĞĞ ĐŽŵƉŽŶĞŶƚƐ͕ ^ĐŚĞůůŝŶŐ ;ϭϵϴϮͿ ƉƌŽͲ
ƉŽƐĞĚ Ă ƐŽůƵƚŝŽŶ͕ ĂůďĞŝƚ ŽŶĞ ǁŚŝĐŚ ƌĞƋƵŝƌĞƐ ƐŽůǀŝŶŐ Ă ƐLJƐƚĞŵ ŽĨ ĞƋƵĂƚŝŽŶƐ͕ ƌĞŶĚĞƌŝŶŐ ŝƚ
ŝŵƉƌĂĐƚŝĐĂůƚŽƉĞƌĨŽƌŵŵĂŶƵĂůůLJ͘dŚĞĐŽĞĨĨŝĐŝĞŶƚ JĚĞŶŽƚĞƐĂǀĂůƵĞďĞƚǁĞĞŶϬĂŶĚϭ͕ďLJ
ǁŚŝĐŚƚŚĞ^ƚĞŝŶĞƌƉĂƌƚƐ;ͼͼĂϮͿŽĨƚŚĞďĞŶĚŝŶŐƐƚŝĨĨŶĞƐƐĐŽŶƚƌŝďƵƚŝŽŶƐŽĨƚŚĞŝŶĚŝǀŝĚƵĂů
ĐŽŵƉŽŶĞŶƚƐ ǁŝƚŚŝŶ ƚŚĞ ƚŽƚĂů ĐƌŽƐƐͲƐĞĐƚŝŽŶ ĂƌĞ ƌĞĚƵĐĞĚ͘ /Ĩ Jсϭ͕ ŶŽ ƐŚĞĂƌ ĚĞĨŽƌŵĂƚŝŽŶƐ
ŽĐĐƵƌ͕ ǁŚŝůĞ ƚŚĞ ƵůĞƌͲĞƌŶŽƵůůŝ ďĞĂŵ ƚŚĞŽƌLJ ĂƉƉůŝĞƐ ĨŽƌ ƚŚĞ ŽǀĞƌĂůů ďĞĂŵ͘ &Žƌ ďĞĂŵƐ
ǁŝƚŚƵŶĐŽŶŶĞĐƚĞĚŝŶĚŝǀŝĚƵĂůĐŽŵƉŽŶĞŶƚƐ͕ JсϬĂƉƉůŝĞƐ͘dŚĞũŽŝŶƚĞĨĨŝĐŝĞŶĐLJĐŽĞĨĨŝĐŝĞŶƚ J
ŝƐƚŚĞƌĞĨŽƌĞĂŵĞĂƐƵƌĞŽĨĐŽŵƉŽƐŝƚĞƐƚŝĨĨŶĞƐƐ͘dŚĞ JͲŵĞƚŚŽĚŚĂƐůŝŵŝƚĞĚĂƉƉůŝĐĂďŝůŝƚLJƚŽ
ƐŝŵƉůLJƐƵƉƉŽƌƚĞĚďĞĂŵƐǁŝƚŚĂƉƉƌŽdžŝŵĂƚĞůLJƵŶŝĨŽƌŵůLJĚŝƐƚƌŝďƵƚĞĚůŽĂĚ͕ǁŚŝůĞĨŽƌĐŽŶƚŝŶͲ
ƵŽƵƐďĞĂŵƐ͕ŝƚĐĂŶďĞƵƐĞĚǁŝƚŚĂƉƉƌŽdžŝŵĂƚŝŽŶĂŶĚƵŶĚĞƌƐƉĞĐŝĨŝĐďŽƵŶĚĂƌLJĐŽŶĚŝƚŝŽŶƐ͘
&Žƌ ďĞĂŵƐ ǁŝƚŚ ĂŶLJ ŶƵŵďĞƌ ŽĨ ŵĞĐŚĂŶŝĐĂůůLJ ũŽŝŶƚĞĚ ĐŽŵƉŽŶĞŶƚƐ͕ ƌĂŶĚŽŵ ĂŵŽƵŶƚ ŽĨ
ƐƉĂŶƐĂŶĚĂƌĂŶĚŽŵĐŽŶĨŝŐƵƌĂƚŝŽŶŽĨĐŽŶĐĞŶƚƌĂƚĞĚŽƌĚŝƐƚƌŝďƵƚĞĚůŽĂĚƐ͕ƚŚĞƐŚĞĂƌĂŶĂůŽͲ
ŐLJ ŵĞƚŚŽĚ ;<ƌĞƵnjŝŶŐĞƌ͕ ϮϬϬϭ͖ <ƌĞƵnjŝŶŐĞƌ ĂŶĚ ^ĐŚŽůnj͕ ϮϬϬϯ͖ ^ĐŚŽůnj͕ ϮϬϬϰͿ ŝƐ ƵƐĞĚ͕ ĂůͲ
ƚŚŽƵŐŚƚŚĞƐŚĞĂƌĂŶĂůŽŐLJŵĞƚŚŽĚƌĞƋƵŝƌĞƐƚŚĞƵƐĞŽĨĐŽŵƉƵƚĞƌƉƌŽŐƌĂŵƐ͘ŽƚŚŵĞƚŚŽĚƐ
ĂƌĞ ĞdžƉůĂŝŶĞĚ ŝŶ ƚŚĞ ĨŽůůŽǁŝŶŐ ƐĞĐƚŝŽŶƐ ǁŚĞƌĞďLJ ƚŚĞ ƐŚĞĂƌ ĂŶĂůŽŐLJ ŵĞƚŚŽĚ ŝƐ ĐƌƵĐŝĂů ŝŶ
ĂƐƐĞƐƐŝŶŐĐƌŽƐƐͲůĂŵŝŶĂƚĞĚƚŝŵďĞƌĞůĞŵĞŶƚƐ͘
ϳ͘ϭ JͲŵĞƚŚŽĚ
dŚĞ JͲŵĞƚŚŽĚ ĚĞƐĐƌŝďĞĚ ŝŶ ŶŶĞdž ŽĨ ϱ ĂŶĚ ƵƐĞĚ ƚŽ ĐĂůĐƵůĂƚĞ ƐŝŵƉůĞ ŵĞĐŚĂŶŝĐĂůůLJ
ũŽŝŶƚĞĚďĞĂŵƐŝƐĂǀŝƚĂůƉĂƌƚŽĨƚŝŵďĞƌĞŶŐŝŶĞĞƌŝŶŐ͘hƐŝŶŐĐŽĞĨĨŝĐŝĞŶƚƐJ͕ĞĨĨĞĐƚŝǀĞďĞŶĚŝŶŐ
ƐƚŝĨĨŶĞƐƐĞƐĐĂŶďĞĐĂůĐƵůĂƚĞĚĨŽƌŵĞĐŚĂŶŝĐĂůůLJũŽŝŶƚĞĚďĞĂŵƐĨƌŽŵƵƉƚŽƚŚƌĞĞŝŶĚŝǀŝĚƵĂů
ĐƌŽƐƐͲƐĞĐƚŝŽŶƐ͕ǀŝĂĞƋƵĂƚŝŽŶ;ϳͲϭͿ͘dŚŝƐĞƋƵĂƚŝŽŶĐŽƌƌĞƐƉŽŶĚƐƚŽĞƋƵĂƚŝŽŶ;ϲͲϮϱͿĨƌŽŵ
ƌƚŝĐůĞϲƵƐĞĚƚŽĐĂůĐƵůĂƚĞƌŝŐŝĚůLJŐůƵĞĚĐŽŵƉŽƐŝƚĞĐŽŵƉŽŶĞŶƚƐ͕ǁŚĞƌĞďLJĐŽĞĨĨŝĐŝĞŶƚƐ J ŝ
ǁĞƌĞ ŝŶĐůƵĚĞĚ ŝŶ ĞƋƵĂƚŝŽŶ ;ϳͲϭͿ͕ ǁŚŝĐŚ ĂƌĞ ĂůǁĂLJƐ ďĞƚǁĞĞŶ Ϭ ĂŶĚ ϭ ƚŚĞƌĞďLJ ƌĞĚƵĐŝŶŐ
ƚŚĞ ^ƚĞŝŶĞƌ ƉĂƌƚƐ ;ŝͼŝͼĂŝϮͿ ŽĨ ƚŚĞ ďĞŶĚŝŶŐ ƐƚŝĨĨŶĞƐƐ ;ǁŚĞƌĞ Ăŝ ŝƐ ƚŚĞ ĚŝƐƚĂŶĐĞ ďĞƚǁĞĞŶ
ƚŚĞĐĞŶƚƌĞƐŽĨŐƌĂǀŝƚLJŽĨƚŚĞŝŶĚŝǀŝĚƵĂůĐƌŽƐƐͲƐĞĐƚŝŽŶƐƚŽƚŚĞŶĞƵƚƌĂůĂdžŝƐͿ͗
ϯ
/ ĞĨ Ϯ
¦ ŝ / ŝ J ŝ ŝ ŝ Ă ŝ
ŝсϭ
;ϳͲϭͿ
Ϯϱϭ
Mechanically jointed members and CLT elements
The joint efficiency coefficients i take into consideration shear deformations in the semi‐
rigid joint connecting two individual components, beam B in Figure D7‐1. When = 0, the
Steiner part E ∙ A ∙ a2 vanishes. In this case, the overall effective bending stiffness solely
comprises the bending stiffnesses of the individual components, beam C in Figure D7‐1.
When = 1, the effective bending stiffness peaks, corresponding to that of a composite
beam with rigid joints or with a geometrically identical full cross‐section, beam A in Fig‐
ure D7‐1. Using the effective bending stiffness (E ∙ I)ef, the ‐method facilitates defor‐
mation calculation and stress distribution of the composite cross‐section whilst using the
Euler‐Bernoulli beam theory. The ‐method is also applicable for composite columns.
The reduction coefficient was derived by Möhler (1956). By studying the deformation
relationships of composite cross‐sections with two or three individual components with‐
out butt joints within the entire beam length and that have been continuously connected
via mechanical fasteners leading to semi‐rigid joints, he presented coupled differential
equations of beam deflections and joint slips, which he then solved for a simply support‐
ed beam under uniformly distributed loads and a centred concentrated load, for a col‐
umn centrally loaded and for two‐span beams under a uniformly distributed load. Each
static system revealed its own reduction coefficients . The ‐method disregards the
shear deformation of the individual components and thus only applies to beams with a
span so large that the shear deformations of the individual components remain negligi‐
ble. The stiffnesses of the joints with continuous mechanical fasteners are considered
with a joint stiffness assumed to be constant and dispersed over the entire beam length.
Strictly speaking, the ‐value from equation (D7‐2) only applies to beams with sinusoidal
distributed loads or those centrally loaded, since these have a sinusoidal flexural buckling
curve (to derive the ‐values, see Annex 4). The mechanical reason for this is that “…the
coefficient and hence (EI)ef is only constant in the case of a sinusoidal distributed load
over the beam length” (Scholz, 2004). However, given the minimal difference between
simply supported beams with a uniformly or sinusoidally distributed load, the ‐method
also applies to simply supported beams with a uniformly distributed load. To ensure a
more comprehensive overview, the conditions to validate the ‐methods are listed here
in note form. The ‐method is valid for:
Simply supported beams with a uniformly distributed load and centrally
loaded columns
A maximum of three individual components non‐abutted over the member length
Continuous fasteners with constant stiffness over the member length
Shear deformations of the individual components are negligible (large spans)
252
Mechanically jointed members and CLT elements
According to Möhler (1956), the coefficient i for the individual components i = 1 and
i = 3 is calculated as follows (see also Annex 5):
1
i for i=1 and i=3 (D7‐2)
E i Ai s i
2
1 2
K i
For the reduction coefficient 2 of component 2, the following applies:
2 1 (D7‐3)
The last, still lacking information is the position of the neutral axis which, as shown in
Article D6, is determined as follows (for geometric definitions, see Figure D7‐3):
3
x i E i Ai h 3 2 E 3 A 3 h 2 2 E 2 A 2 h 2 h 1 2 E 1 A 1 h2
a 2 i=1 3
E 1 A1 E 2 A 2 E 3 A 3 2
E i Ai (D7‐4)
i=1
E 1 A 1 h1 h 2 E 3 A 3 h 2 h 3
2 E 1 A1 E 2 A 2 E 3 A 3
This results in the following distances between the centre of gravity of the individual
components 1 and 3 to the neutral axis:
h1 h 2 h2 h3
a1 a2 and a3 a2 (D7‐5)
2 2
where
Ei Modulus of elasticity
Ai Cross‐sectional area
I i Second moment of area
si Fastener spacing in joint between individual components
ki Slip modulus of fasteners
ℓ Member length
hi Depth of individual component
253
Mechanically jointed members and CLT elements
b1 V 1,t V 1,M
A1, E1, I1
h1 N M
S a1 W 2,max neutral axis
s1, K1, F1 y a2
z N
h2 M
x a3 h
A2, E2, I2 b2
V 2,M V 2,c
h3 N M
s3, K3, F3
V 3,MV 3,c
A3, E3, I3
b3
Figure D7‐3 Example of a composite beam with three components.
Stress analysis
Now, all parameters required to calculate normal and shear stresses in the individual
components are known, Annex 5 includes accurate information on stress analysis and the
influence of bending stiffness on stress distribution. The bending stresses in the outer
fibres of the individual components i are determined from the external moment M (cf.
equation (D6‐29)):
M h
i,m E i i (D7‐6)
E I ef 2
When calculating the normal stresses in the centre of gravity of the individual compo‐
nents, however, unlike glued composite beams, the coefficients i are taken into consid‐
eration (for an explanation, see Annex 5):
M
i,t(c) E a (D7‐7)
E I ef i i i
The overall stress of the beam is determined by superposing the normal and bending
stress in the outer fibre.
The shear stress peaks in the neutral axis (here in component 2) and amounts to (deriva‐
tion see Annex 5):
max
V max 3 E 3 A 3 a 3 0,5 E 2 b 2 h
2
(D7‐8)
E I ef b 2
254
Mechanically jointed members and CLT elements
The load on the fasteners in the joint between the individual components amounts to
(derivation see Annex 5):
Here, there is still a need to divide by the number of rows of fasteners, if multiple such
rows are installed adjacent.
Strictly speaking, the ‐method only applies to simply supported beams. EC 5 Annex B,
however, means the ‐method can also be used for the approximate calculation of con‐
tinuous and cantilevered beams. In this case, the actual length of a continuous or cantile‐
vered beam is translated into a fictitious length, which for cantilevered beams, equates
to double the length of the cantilever (ℓ = 2 ∙ ℓK). For continuous beams, the fictitious
length assumed should be 80% of the length of the span under examination (ℓ = 0,8 ∙ ℓF).
When verifying a continuous beam at the intermediate supports, the smaller length of
both adjacent spans is the key parameter. In addition, EC 5 also allows the fastener spac‐
ing in the joint to be adjusted to the course of the shear force diagram between smin and
smax (≤ 4 ∙ smin). In this case, the ‐method may be applied with an effective distance sef of
the fasteners:
Increased stresses imposed by reduced cross‐sections can also be taken into considera‐
tion. For this purpose, the normal stresses in the centre of gravity of the individual com‐
ponents in accordance with equation (D7‐7) are to be multiplied by Ai/Ai,net and the bend‐
ing stresses in the outer fibres of the individual components in accordance with equation
(D7‐6) by Ii/Ii,net.
Time‐dependent deformations of composite beams are generally only to be taken into
consideration in the ultimate limit state. However, in the event of composite beams
made from various construction materials having variable creep behaviour, creep must
also be taken into consideration for the ultimate limit state (EC 5 Section 2.2.2 (1)P).
Time‐dependent deformations are taken into consideration by reducing the moduli of
elasticity Ei of the individual components and the slip moduli Ki of the fasteners in the
joint with kdef (see also equation (D6‐1)).
255
Mechanically jointed members and CLT elements
D7.2 Shear analogy method
Unlike the ‐method, the shear analogy method by Kreuzinger (1999) can be used to
calculate any composite members under any load configuration more accurately. Details
are included in the NA to EC 5. The concept of the shear analogy centres on transforming
a real composite beam comprising multiple mechanically jointed components into a
single fictitious and homogenised beam. In this case, two components A and B of a ficti‐
tious beam are defined, as in Figure D7‐4. The fictitious component A is given the sum of
the bending stiffnesses BA along the individual neutral axis of the real components. The
fictitious component B is allocated the Steiner parts (Ei ∙ Ai ∙ zi2) of the bending stiffness BB
of the real components and a shear stiffness SB. This shear stiffness SB includes the shear
deformations of the individual components themselves and the joint slip due to the semi‐
rigidity of the joint between the individual components. Fictitious component A thus
represents the sum of the bending stiffnesses (Ei ∙ Ii) and fictitious component B the
interaction of the individual components through Steiner’s parts (Ei ∙ Ai ∙ zi2) of the bend‐
ing stiffness, shear deformation and joint slip of the overall composite beam. The follow‐
ing section focuses on the shear analogy method for mechanically jointed beams. This
approach is also particularly applicable to uniaxially loaded cross‐laminated timber ele‐
ments, since a strip of any CLT element can be considered a composite beam and thus
transformed into a plane frame model comprising fictitious beams A and B. However,
twodimensional structural systems comprising mechanically jointed layers can also be
designed by applying the shear analogy method (as described in the NA). For calculations
within the framework of second order theory, see Scholz (2004).
b1
d1 1
zs1 A
2
Transformation
z 3
B
n
Figure D7‐4 Transformation of a composite cross‐section into a fictitious cross‐section. (Scholz, 2004)
The bending stiffness BA of the fictitious component A is the sum of bending stiffnesses
(Ei ∙ Ii) of the individual components i:
n n b d 3
B E i I i E i i i
A
(D7‐11)
i1 i1 12
256
Mechanically jointed members and CLT elements
The bending stiffness BB of the fictitious component B can be calculated from the sum of
Steiner’s parts (Ei ∙ Ai ∙ zi2) of the bending stiffnesses of the individual components i to:
B n 2 n 2
B E i A i z si E i b i d i z si (D7‐12)
i=1 i=1
where zsi are the distances between the centres of gravity of the individual components i
from the centre of gravity S of the entire beam.
Since the fictitious component A does not allow for shear deformations, all the shear
deformations of the composite beam are allocated to the fictitious component B. The
shear stiffness SB hence includes all shear deformations of the individual components as
well as the relative displacement of adjacent individual components due to the semi‐
rigidity in the joint, Figure D7‐5.
The shear stiffness due to the shear deformation of individual components is recorded
via the related shear modulus Gi and the thickness di of the individual components. The
shear stiffness due to semi‐rigidity in the joint is recorded via the slip modulus ki,i+1 of the
fasteners in the joint between the individual components i and i+1. The following applies
to the relative displacement of the individual components:
t t
ui d i d i ; u i,i+1 (D7‐13)
Gi G i bi k i,i+1 k i,i+1 b i
u
t d1
1
2 d2
t t t
a x Ji di
i
z
t
n dn
ui+1
dx ui
Figure D7‐5 Deformation of a composite member (layered structure) subject to constant shear flow t.
(Scholz, 2004)
257
Mechanically jointed members and CLT elements
1 1 n1 1 d1 n1 d dn
2
i
(D7‐14)
S B
a i1 k i,i+1 2 G 1 b 1 i2 G i b i 2 G n b n
If only the semi‐rigidity of the joint is considered and the individual components them‐
selves do not allow for shear deformations, equation (D7‐14) can be simplified as follows:
1 1 n1 1
B
2
(D7‐15)
S a i1 k i,i+1
To calculate the internal forces and moments in fictitious beams A and B, these are cou‐
pled to each other via infinitely rigid web members, to ensure they undergo the same
deformation along their axis, Figure D7‐6.
Fictitious beam B
Superimposed system Rigid webs to obtain
with same deformations same deflection
Figure D7‐6 Fictitious beam with coupled fictitious components A and B.
258
Mechanically jointed members and CLT elements
While analytical solutions are available for this fictitious beam model to determine inter‐
nal forces and moments, there is also the option of using software to determine the
internal forces. The shear analogy method is designed in such a way that in practice, the
use of conventional plane frame analysis programs, which take shear deformations of
beams into consideration, is possible. Once the internal forces for the fictitious beams
have been successfully determined, this is followed by a retransformation into the real
composite system. The displacements of the fictitious beam already correspond to those
of the composite beam, with only the internal forces having to be retransformed. The
retransformation of the internal forces and moments into real internal forces and mo‐
ments is performed in proportion to the stiffnesses. Plane frame analysis programs are
hence used to determine the bending moments MA and MB and the shear forces VA and
VB for the fictitious components A and B (consideration of shear deformations must be
activated). The internal forces and moments of beam A, MA and VA, deliver the stress
components along the own neutral axis on the overall stress (Figure D7‐7 left) and the
internal forces and moments of beam B, MB and VB, add those stresses to the overall
stress which are due to the interaction of the individual components (Steiner parts, shear
deformations of individual components and slip due to semi‐rigidity of joints or perpen‐
dicular layers of CLT). Figure D7‐7 shows the individual stress components and stress
distribution determined by addition on the entire cross‐section for a layered structure
(mechanically jointed beam or CLT).
Stresses along
own neutral axis Steiner's parts Shear Resulting stresses
VA WA VB WB V A + V B W A + WB
1
2
V M
i
z
n
Figure D7‐7 Schematic presentation of the stress analysis using the example of a layered structure.
The bending stresses in the outer fibres of the real components i are calculated using the
bending moment MA of the fictitious component A:
E i I i di MA d MA d
M iA M A m,i
A
( ) i i A E i i (D7‐16)
BA 2 Ii 2 B 2
259
Mechanically jointed members and CLT elements
The bending moments of the fictitious component B, namely Steiner’s parts on the over‐
all stress, in turn, elicit the normal stresses, which are deemed constant over the individ‐
ual components (zsi are the distances between the centres of gravity of components i
from the centre of gravity S of the entire beam):
B
M
t/c,i
B
( z si ) B
E i z si (D7‐17)
B
n
z 0i E i A i
z 0s i=1 n (D7‐18)
E i Ai
i=1
From the equilibrium of forces at an infinitesimally small element come the shear stress‐
es , which can be recorded from the integration, area by area, over individual compo‐
nent cross‐sections using the shear forces of the fictitious component A:
zi
d A zi
dM A E i E z2 d 2
iA ( z i ) dz A z i dz V A Ai i i
dx dx B B 2 8 (D7‐19)
d
i
d
i
2 2
and using shear forces of the fictitious component B:
d
B B
zi zi
dM E i E d
iB (z si , z i ) dz B z si dz V B Bi z si z i i (D7‐20)
d dx
i
d dx
i
B B 2
2 2
The maximum shear stresses in the centres of gravity of the individual components gen‐
erated by the shear forces VA can be determined using the well‐known equation for
rectangular cross‐sections:
E i I i VA
iA 1.5 (D7‐21)
BA d i bi
260
Mechanically jointed members and CLT elements
The stress curves along the real composite beam are obtained by superposing individual
stresses. Figure D7‐7 clarifies the stress analysis via retransformation on the example of a
cross‐section comprising multiple layers.
The ‐method and the shear analogy method are suitable for calculating mechanically
jointed beams, while the latter shear analogy method in particular is also useful for calcu‐
lating twodimensional load‐bearing structures, i.e. solid wood panels and cross‐laminated
timber elements. Although solid wood panels and cross‐laminated timber elements with
perpendicular layers lack any semi‐rigid joints between single layers, such products can
be considered like mechanically jointed beams and calculated likewise. The individual
longitudinal layers arranged in the principal direction of the panel correspond to the
individual components of a mechanically jointed beam, while the semi‐rigid joint be‐
tween individual components is mapped by the perpendicular layers arranged perpen‐
dicular to the principal panel direction. If solid wood panels and cross‐laminated timber
elements are exposed to out‐of‐plane bending stress, the longitudinal layers shift rela‐
tively to their individual positions, which ultimately leaves the perpendicular layers dis‐
torted due to shear. This shear stress in the tangential‐radial plane is also referred to as
rolling shear (Figure D7‐8).
W
W
Tangential
Radial
Figure D7‐8 Left: Deflected three‐layered panel. Right: Rolling shear stress of the middle layer.
261
Mechanically jointed members and CLT elements
The rolling shear stiffness and strength of wood are clearly lower than the corresponding
stiffness and strength values for longitudinal shear in the tangential‐longitudinal or radial‐
longitudinal plane. The rolling shear modulus comprises around 10% of the shear modu‐
lus of the wood, while the rolling shear strength is around 40% of its (longitudinal) shear
strength. The shear analogy method can be directly applied to calculate twodimensional
structural members with cross‐wise arranged individual layers. In this case, the rolling
shear in the middle layers is taken into consideration when shear stiffness SB in accord‐
ance with equation (D7‐14), i.e. Gi = GR,i for perpendicular layers. When applying the ‐
method, the “weak” perpendicular layers loaded in rolling shear are considered to consti‐
tute a semi‐rigid joint between the longitudinal layers arranged in the principal beam
direction. This semi‐rigidity is recorded in the ‐coefficients, whereby instead of the stiff‐
ness K/s, the rolling shear stiffness GR ∙ b/d of the perpendicular layer is the basis. In this
case, b is the width and d the thickness of the perpendicular layer:
1
i for i=1 and i=3 (D7‐22)
E i Ai d i
2
1
G R,i b i 2
D7.3 Literature
H. Kreuzinger, original Article B11, STEP 1995.
Kreuzinger H. (2001). Verbundkonstruktionen. Holzbaukalender 2002, pp. 598‐621. Bruderverlag Karlsruhe.
Kreuzinger H. and Scholz A. (2003). Flächentragwerke – Berechnung und Konstruktion. Schlussbericht AIF
Forschungsvorhaben, Technische Universität München.
Möhler K. (1956). Über das Tragverhalten von Biegeträgern und Druckstäben mit zusammengesetzten
Querschnitten und nachgiebigen Verbindungsmitteln. Habilitation, Technische Universität Karlsruhe.
Schelling W. (1982). Zur Berechnung nachgiebig zusammengesetzter Biegeträger aus beliebig vielen
Einzelquerschnitten. In: „Ingenieurholzbau in Forschung und Praxis“, Herausgeber J. Ehlbeck und G. Steck;
Bruderverlag, Karlsruhe. ISBN 3‐87104‐049‐5.
Scholz A. (2004). Ein Beitrag zur Berechnung von Flächentragwerken aus Holz. Dissertation, Technische
Universität München.
262
D8 Reinforcements
Original article: H. J. Larsen
In a structural timber context, the term ‘reinforcements’ refers to measures used to
increase the ability of structural members to withstand tension or compression perpen‐
dicular to the grain or shear. The members can be boosted using internal or external
reinforcements, where cross‐sectional weakening caused by internal reinforcements
must be taken into account:
Internal reinforcements:
Glued‐in threaded rods,
Glued‐in rebars,
Fully threaded wood screws.
External reinforcements:
Glued‐on plywood,
Glued‐on LVL,
Glued‐on boards,
Punched metal plate fasteners.
In areas where notches, holes and joints loaded perpendicular to the grain are present,
as well as in double‐tapered, pitched cambered and curved beams, climate fluctuations
trigger swelling and shrinking processes, which can cause tension perpendicular to the
grain exerted by external loads to intensify significantly. This is why, in accordance with
NA, reinforcements perpendicular to the grain are required in service class 3.
This article sets out verifications for reinforcements in notches, holes and apex zones of
double‐tapered, pitched cambered and curved beams, while joints loaded perpendicular
to the grain are covered in Article E11. All verifications are given in the NA; EC 5 does not
include any design rules for reinforcements. As a general rule, any reinforcement to
increase resistance against tension perpendicular to the grain should be, conservatively,
capable of withstanding the entire tensile force perpendicular to the grain; this equates
to a state in which the wood has failed perpendicular to the grain.
Other types of reinforcements include those of supports with fully threaded screws to
enhance compression behaviour perpendicular to the grain, where the screws are insert‐
ed perpendicular to the support area. This type of reinforcement is covered in Article E5.
263
Reinforcements
D8.1 Effects of fluctuations in moisture content
Tensile stresses perpendicular to the grain are generated not only by external loads, but
also by changing levels of moisture content, where the cross‐sectional moisture gradient
is of particular significance. For members in service class 1, which have been dried to an
appropriate level of moisture content before assembly, the general assumption is that
the impact of any additional load imposed by moisture changes will be sufficiently cov‐
ered by characteristic values and partial safety coefficients. Under certain circumstances,
however, the impact of moisture changes may have to be assessed.
For example, assume the moisture content in the external area of a cross‐section (re‐
spectively 1/6) declined by 3%. This would correspond to unhindered strain of =
3 ∙ 0.002 = 0.006 (with a shrinkage coefficient of 0.2% per 1% change in moisture content
below the fibre saturation point (see Article B2)). With E90 = 300 N/mm2, it follows that
= 2/3 ∙ 300 ∙ 0.006 = 1.2 N/mm2, namely the stresses generated by hindered shrinking
are in the same order of magnitude as the short‐term strength and the risk of failure
exists. In practice, however, while such stresses are reduced by relaxation, they should
not be ignored.
(a) (b)
0,5 Δσ
u
Δε = 0,006 - Δσ
u-3% +
b/6 b/6
b
Figure D8‐1 Influence of a decrease in moisture content in the (a) external fibres of a rectangular cross‐
section, (b) unhindered strain, (c) resulting internal stresses. (STEP 1995 Article E5)
One effect of an increase in moisture content is an increase in beam depth from h to
h ∙ (1 + ), where corresponds to the strain resulting from this increase in humidity (for
a beam 600 mm deep, this would equate to an increase of 4 mm).
264
Reinforcements
The influence of moisture content parallel to the grain is generally negligible. However, in
curved beam areas, the angle d declines to d' and the radius of curvature increases
from r to r', Figure D8‐2 (a):
d
d d 1 (D8‐1)
1
r r 1 (D8‐2)
These changes indicate an increase in the distance between supports and a reduction in
curvature of (see Figure D8‐2 (b)):
a
v r 1 cos (D8‐3)
2
The reduction in curvature seen in this example results in a lower level of tensile stress
perpendicular to the grain.
h (1 - ε)ϕ/2
(c)
(d)
Figure D8‐2 Influence of an increase in moisture content on the curvature of a beam. (a) geometry after the
increase in moisture content, (b) undeformed state. e is the entire deflection at centre span.
(STEP 1995 Article E5)
265
Reinforcements
D8.2 Reinforcement of notched beams
Reinforcements of beams notched on the same side as the support are designed for a
tensile force Ft,90,d perpendicular to the grain, which corresponds to the tensile stress
perpendicular to the grain of area 1 as specified in Figure D8‐3. This tensile force can
subsequently be transferred by glued‐in rods, glued‐on reinforcement plates or fully
threaded screws, Figure D8‐4. The first step involves calculating the tensile component
perpendicular to the grain generated from an external load in the critical area 1. The
derivation for reinforcements of notches corresponds to the derivation for joints loaded
perpendicular to the grain (Ehlbeck and Görlacher, 1983) and is described in detail in
Article E11. At this point, only the required changes to the design equations compared to
those in Article E11 are shown.
hef = D ·h
h
1
a
x
Figure D8‐3 Rectangular notch on the same side as the support with critical area 1.
1 2
h =D ·h ℓad h =D ·h
h ef h ef
a ℓad a
a ℓr
2,c
tr
a2
a
2,c
a1,c
Figure D8‐4 Reinforcement of notched beams with geometric details, 1: Reinforcement with fully threaded
screw / glued‐in rod, 2: Reinforcement with glued‐on plates.
The equation specified in NA and used to determine the force component Ft,90,d generat‐
ing tension perpendicular to the grain in critical area 1 is as follows:
F t,90,d 1.3 V d 3 1 2 1
2 3
1.3 V d (D8‐4)
266
Reinforcements
The factor used in equation (D8‐4) for notches corresponds to the factor specified in
equation (E11‐15) for joints loaded perpendicular to the grain if is replaced with
a/h = hef and the term is converted, whereby a represents the distance from the loaded
edge (for notches, this is the distance between the notch corner and the upper edge, see
Figure D8‐3):
2 3
a a
3 1 2 1 1 3 2
2 3
(D8‐5)
h h
However, the stresses in the notch corners are not uniformly distributed (see also Fig‐
ure D5‐2), which results in peak stresses, which decline with increasing distance from the
notch corner. FE simulations by Henrici (1984) resulted in the factor 1.3 used in equation
(D8‐4), which involves the force component Ft,90,d increasing by 30%.
If the reinforcement is done with glued‐in rods or fully threaded screws (number 1 in
Figure D8‐4), on the one hand, the means of reinforcement used must be capable of
absorbing the tensile force Ft,90,d and on the other, the bond between the wood and
reinforcement must be guaranteed. This can be verified for fully threaded screws by
designing for stress in the direction of the screw axis (withdrawal capacity Fax,Rd and ten‐
sile capacity Ft,Rd, see Article E5; the smaller value must exceed Ft,90,d from equation ((D8‐
4)). For glued‐in rods, the bond line must be checked, in which case a uniformly distribut‐
ed bond line stress is assumed:
F t,90,d
ef,d f k1,d with ef,d (D8‐6)
n d r ad
The characteristic value for the bond line strength fk1,k is specified in the NA (see also
Table D8‐1). The effective bond line stress ef,d is calculated from the tensile force Ft,90,d in
accordance with equation (D8‐4) divided by the bond line area, which is calculated from
the effective anchorage length ℓad, the external rod diameter dr and the number n of
glued rods used. The effective anchorage length ℓad corresponds to the height of the
notched area (h ‒ hef) and is shown in Figure D8‐4. The constraint is that in longitudinal
member axis direction, only one rod can be taken into account, which must have a mini‐
mum length of 2 ∙ ℓad and which must not exceed the external diameter dr = 20 mm.
Figure D8‐4 shows, in turn, the minimum distances of the rods. The spacings a2 have to
be a minimum of 3 ∙ dr, while the minimum figure for edge distances a2,c and for end
distances a1,c is at least 2.5 ∙ dr.
267
Reinforcements
Table D8‐1 Characteristic bond line strength in reinforcements, DIN EN 1995‐1‐1/NA:2013.
Characteristic Effective anchorage length ℓad of the rod [mm]
strength [N/mm2] ≤ 250 250 ≤ ℓad ≤ 500 500 ≤ ℓad ≤ 1000
Bond line between rod and fk1,k 4.0 5.25 ‒ 0.005 ∙ ℓad 3.5 ‒ 0.0015 ∙ ℓad
borehole wall
Bond line between beam fk2,k 0.75
surface and reinforcement
plate
Bond line between beam fk3,k 1.50
surface and reinforcement
plate if shear stresses are
uniformly introduced
Laterally glued‐on reinforcement plates are accounted for correspondingly; the bond
line stress, which is assumed to be uniformly distributed, is calculated from the tensile
force Ft,90,d divided by the bond line area. However, the design value of bond line strength
fk2,d equates approximately to the design rolling shear strength of the member to be
reinforced and also takes into account a non‐uniform shear stress distribution (character‐
istic value see Table D8‐1):
F t,90,d
ef,d f k2,d with ef,d (D8‐7)
2 h h ef r
As with glued‐in rods, the effective bond line area taken into account here is the area
below critical area 1: 2 ∙ ℓr ∙ (h ‒ hef), with factor 2 due to reinforcement on both sides,
see Figure D8‐4.
268
Reinforcements
The tensile force Ft,90,d determined in accordance with equation (D8‐4) must be transmit‐
ted via the bond line to the reinforcement plates, which must, in turn, be capable of
transferring the tensile stresses formed, whereby a triangular stress distribution is taken
into consideration in this case:
F t,90,d
k k t,d f t,d with t,d (D8‐8)
2t r r
where
tr Thickness of a reinforcement plate
ℓr Width of a reinforcement plate
ft,d Design tension strength of the plate material
Coefficient kk takes the non‐uniformly distributed tensile stress into consideration, which
peaks at the notch corner and can be assumed as kk = 2.0 without further verification.
The effective area of the glued‐on reinforcement plates must include a certain ratio
between height (h ‒ hef) and width ℓr. The minimum value should prevent any crack
development in critical area 1; the maximum value should ensure that the only area of
the reinforcement plate taken into account is that located in the notch area subject to
tension perpendicular to the grain:
r
0.25 0.5 (D8‐9)
h h ef
D8.3 Reinforcement of beams with holes
The procedures used to verify reinforcements of holes are similar to those used for
notches, but the tensile force component Ft,90,d perpendicular to the grain and the effec‐
tive bond line surfaces are calculated differently. Verifications for reinforcements of
areas of rectangular and circular holes subjected to tension perpendicular to the grain in
accordance with Figure D8‐5 are shown; possible reinforcements see Figure D8‐6. In
addition, the holes to be reinforced must comply with the minimum and maximum di‐
mensions specified in Table D8‐2. If the geometric boundary conditions are fulfilled, the
tensile force to be applied may be calculated in accordance with equation (D5‐11). Simi‐
lar to the approach when designing reinforced notches, this tensile force Ft,90,d is used to
verify bond lines of glued‐in rods, bond lines and tensile stresses of glued‐on plated and
fully threaded screws.
269
Reinforcements
hro + 0.15·hd
M M
hro
hd h
V hru V
V V
ℓA a ℓz ℓz a ℓA
ℓv ℓv
Figure D8‐5 Beam with rectangular and circular hole.
Table D8‐2 Minimum and maximum dimensions for reinforced holes, DIN EN 1995‐1‐1/NA:2013.
c ℓ = distance between two adjacent holes, see also Figure D5‐7.
z
When using fully threaded screws, they are verified for a tensile force Ft,90,d in the direc‐
tion of the screw axis; in turn, with the withdrawal capacity Fax,Rd and tensile capacity Ft,Rd
from Article E5, where the smaller value must exceed Ft,90,d from equation (D5‐11).
Verification of the bond line stress for glued‐in rods:
F t,90,d
ef,d f k1,d where ef,d (D8‐10)
n d r ad
Compared to equation (D8‐6), only the determination of the effective anchorage length
ℓad changes and all other rules remain the same. In addition, however, for holes with
internal reinforcements, increased shear stresses in the hole area must be verified (see
below).
Effective anchorage length for rectangular holes (Figure D8‐5):
ad h ru or ad h ro (D8‐11)
For circular holes (Figure D8‐5):
270
Reinforcements
ℓad
ℓad
ℓad
V
ℓad
V
ℓad
hro
ℓad hd
ℓad h
V
ℓad hru
V
a2
a a
2,c
1,c
h1 hro
hd hd h
h1 hru
ar ar
a tr b
Figure D8‐6 Reinforcement of holes with geometric details.
Verification of bond line stress for glued‐on plates:
F t,90,d
ef,d f k2,d with ef,d (D8‐13)
2 a r h ad
For rectangular holes (Figure D8‐6):
h ad h 1 (D8‐14)
For circular holes (Figure D8‐6):
h ad h 1 0.15 h d (D8‐15)
271
Reinforcements
Verification of tensile stresses in glued‐on reinforcement plates, which likewise here,
correspond to verifications of those of notches:
F t,90,d
k k t,d f t,d with t,d (D8‐16)
2a r t r
where
tr Thickness of a reinforcement plate
ar Width of the reinforcement plate to be taken into consideration, see Figure D8‐6
ft,d Design tension strength of the plate material
The coefficient kk takes into consideration the non‐uniformly distributed stresses and also
in this case, can be assumed to be kk = 2.0 without further evidence required.
As with reinforced notches, a certain ratio between height and width must be maintained
for the effective area of the glued‐on reinforcement plates:
Verification of the increased shear stress in the hole area for internal reinforcements is
not specified in the NA but can be carried out using the approach by Blass and Bejtka
(2004), as well as the “Erläuterungen zur DIN 1052” (Blass et al., 2005, Section E11.4.4).
Accordingly, the key shear stresses peak at the following value:
1.5 V d
max max (D8‐18)
b h hd
where
0.2
a h
max 1.84 1 d (D8‐19)
h h
and 0.1 ≤ a/h ≤ 1.0 and 0.1 ≤ hd/h ≤ 0.4.
272
Reinforcements
D8.4 Reinforcement of double‐tapered, pitched cambered
and curved beams
As already explained in Article D4, tensile stresses perpendicular to the grain are gener‐
ated in double‐tapered, pitched cambered and curved beams, due to the action of exter‐
nal forces. Changes in moisture content trigger additional tensile stresses perpendicular
to the grain. As a general rule, verifications of the external forces acting on such beam
shapes are carried out and climate‐related tensile stresses perpendicular to the grain are
not taken into consideration. Accordingly, as specified in Article D4, the NA includes rules
to take climate‐related tensile stresses perpendicular to the grain into account. Rein‐
forcements in the apex zone subject to tensile stresses perpendicular to the grain are
frequently employed for double‐tapered, pitched cambered and curved beams, which
are exposed to changes in moisture content or where the calculated level of tensile
stresses perpendicular to the grain is excessive. The NA regulates two different measures
for reinforcement:
Reinforcements intended to absorb additional climate‐related stresses
perpendicular to the grain (→ service classes 1 and 2) and
Reinforcements to completely absorb tensile stresses perpendicular to
the grain (external forces and climate loads), (→ service classes 1 and 2,
mandatory for use in service class 3).
As with reinforcements for notches and holes, both measures involve determining the
tensile force, which is generated from the tensile stresses perpendicular to the grain. In
turn, this tensile force must then be capable of being transferred by the bond line be‐
tween the wood and glued‐in rods, by the anchorage area of screws or the bond line
between wood and glued‐on reinforcement plates. In addition, the tensile stresses in the
glued‐in rods, screws or reinforcement boards must be verified. The two reinforcement
measures outlined in the NA are discussed in the following section and understanding of
the current article presumes previous understanding of Article D4.
Figure D8‐7 shows a schematical view of the course of tensile stresses perpendicular to
the grain within the apex zone along the axis of a curved beam. In the design, a simplified
(stepped) trajectory of tensile stresses perpendicular to the grain is considered, rather
than the actual non‐linear course, which allows the values of the forces Ft,90,d which the
reinforcements have to transfe to be calculated. The tensile stresses perpendicular to the
grain decline based on stress and geometry with increasing distance from the apex and
two separate areas of tensile stresses perpendicular to the grain can be differentiated, as
shown in Figure D8‐7: an internal area, the internal half or the two internal quarters of
the apex cross‐section and an external area with lower tensile stresses perpendicular to
the grain, which encompasses both external quarters of the apex cross‐section.
273
Reinforcements
Simplification
Tension stresses
according to FE
1/2 1/2
Figure D8‐7 Tensile stresses perpendicular to the grain in the apex zone of a curved beam,
shown qualitatively.
Reinforcement to absorb additional climate‐related tensile stresses
perpendicular to the grain
For reinforced double‐tapered, pitched cambered and curved beams in service classes 1
and 2, the rule according to NA is that conditions in accordance with equations (D4‐15) and
(D4‐18) can be overlooked, if it is possible to prove that the combination of tensile stresses
perpendicular to the grain and shear stresses in the apex zone matches equation (D8‐20):
2
d t,90,d
0.3
1 (D8‐20)
f v,d h
k dis 0 f t,90,d
h ap
where
kdis = 1.3 for double‐tapered and pitched cambered beams
= 1.15 for curved beams
h0 Reference height of 600 mm
hap Beam height at apex, see Figure D4‐10 and Figure D4‐11
t,90,d Tensile stresses perpendicular to the grain at apex,
calculated with equation (D4‐13)
Equation (D8‐20) is very similar to equation (D4‐18), which was used to verify combined
tensile perpendicular to the grain and shear stresses. However, equation (D8‐20) specified
in the NA is less conservative due to the semi‐quadratic interaction between tensile and
shear stresses. In addition, the volume factor kvol from equations (D4‐16) and (D4‐17) is
replaced with (h0/hap)0.3. Although this elicits values which change somewhat, the function
274
Reinforcements
is the same as with kvol; namely to take into consideration the volume effect on the ten‐
sile strength perpendicular to the grain of wood. The coefficient kdis, that considers the
stress distribution in the apex zone, is assigned values other than those specified in EC 5
(kvol and kdis see also Article D3).
The reinforcement elements to be deployed in this case must be capable of absorbing a
portion of the additional climate‐related tensile stresses perpendicular to the grain, while
when determining the additional tensile forces perpendicular to the grain, the strain in
wood and reinforcement elements were assumed to be equal (Brünninghof et al., 1993).
Therefore, interaction between the wood and reinforcement is presumed, meaning that
in this case, the reinforcement is not configured to absorb the full extent of the tensile
forces perpendicular to the grain. Brünninghof et al. (1993) showed that the strains in the
wood and reinforcement element were only approximately equal, when around a quarter
of the tensile forces perpendicular to the grain Ft,90,d generated from external loads was
absorbed by the reinforcement elements. Moreover, most investigations centring on the
derivation of Ft,90,d assumed the use of beams b = 160 mm wide. However, given the
intensifying impact of climate as beams widen, a factor of b/160 was introduced, to take
account of the harsher situation applicable for such wider beams. Accordingly, the rein‐
forcements applied to such beams to absorb climate‐related tensile stresses perpendicu‐
lar to the grain should be designed such to withstand the tensile force Ft,90,d, which can
be calculated according to equation (D4‐13) from the tensile stresses perpendicular to
the grain. Ft,90,d is the tensile force per reinforcement (where t,90,d ∙ b ∙ a1 equates to the
entire tensile force exerted over the length a1, a ¼ of which should be absorbed by the
reinforcements):
1 1 b ba1 b
F t,90,d t,90,d b a 1 t,90,d (D8‐21)
n 4 160 n 640
where
b Beam width in mm
a1 Spacing of reinforcements in axial beam direction = tributary width
to determine the tensile force perpendicular to the grain
n Number of reinforcements within length a1, namely the number of
reinforcements per row
It must be possible to accommodate this tensile force by both glued‐in rods or glued‐on
plates as well as by the shear stresses in the joints between the reinforcement elements
and timber. The characteristic values specifying the strength of the bond line fk1,k or fk2,k
and required to verify the bond line stress are specified in Table D8‐1.
275
Reinforcements
Bond line stress in glued‐in or screwed‐in threaded rods, respectively, is verified similar‐
ly to the verification on notches and holes, as in equations (D8‐6) and (D8‐10), whereby
the tensile force Ft,90,d has already been calculated for an individual rod and factor 2 in
the numerator takes into consideration the fact that in what are now generally longer
reinforcement rods, the shear stress is no longer uniformly distributed over the rod
length. This means that the stresses involved are doubled, similar to coefficient kk in
equations (D8‐8) and (D8‐16) and a triangular distribution of shear stress is assumed over
the effective anchoring length ℓad above and underneath the beam axis:
2 F t,90,d
ef,d f k1,d with ef,d (D8‐22)
d r ad
The withdrawal capacity of threaded rods with threads similar to screws is assessed simi‐
larly to the verifications of fully threaded screws exposed to axial loads (Article E5). Here,
reference can also be made to the European Technical approvals (ETA) of these rods to
determine their characteristic withdrawal parameter.
The verification of the bond line stress of a glued‐on reinforcement plate, in turn, is
determined similarly to the verifications for notches and openings, as in equations (D8‐7)
and (D8‐13), but assuming higher values for the bond line strength (fk3,d instead of fk2,d),
since an uneven distribution of shear stresses has already been taken into consideration
in ef,d. ℓr is the length of the glued‐on reinforcement plate and ℓad is its height, above or
underneath the beam axis:
2 F t,90,d
ef,d f k3,d with ef,d (D8‐23)
ad r
The tensile stresses in a reinforcement plate of thickness tr, length ℓr and tensile strength
ft,d are verified as follows:
F t,90,d
t,d f t,d with t,d (D8‐24)
t r r
Here, compared to equations (D8‐8) and (D8‐16), coefficient kk is missing, since uneven
stress distribution was already taken into consideration when determining Ft,90,d. Moreo‐
ver, Ft,90,d was already determined per reinforcement element.
276
Reinforcements
Reinforcements to completely absorb tensile stresses perpendicular to the grain
Reinforcements are required according to NA when using double‐tapered, pitched cam‐
bered and curved beams in service class 3, but may also be used in service classes 1 and 2.
This then eliminates the need for tensile stress perpendicular to the grain verifications in
accordance with equations (D4‐15) and (D4‐18), since all tensile stresses perpendicular to
the grain are absorbed by reinforcement elements in this case. Here, the tensile forces
Ft,90,d exerted on the reinforcements must be determined more accurately, before the
aforementioned verifications can be conducted via equations (D8‐22), (D8‐23) and (D8‐24).
This is followed, in turn, by verifications of tensile stresses in the cross‐section of glued‐in
rods or screwed‐in threaded rods. In such cases, the tensile forces Ft,90,d to be applied are
determined for both internal and external areas of the beam, subject to tensile stresses
perpendicular to the grain, whereby for simplicity, the tensile forces in the external area
are assumed to be 2/3 of the values of the internal area. Likewise in this case, the deci‐
sive tensile stress perpendicular to the grain t,90,d is deemed to be the value t,90,d calcu‐
lated with equation (D4‐13).
The tensile force in a reinforcement element in the internal area (both internal quarters),
corresponds to equation (D8‐21) without factor b/640:
t,90,d b a 1
F t,90,d (D8‐25)
n
Tensile force in a reinforcement element in the external area (both external quarters):
2 ba1
F t,90,d t,90,d (D8‐26)
3 n
D8.5 Literature
H.J. Larsen, original Article E5, STEP 1995.
Blass H.J. and Bejtka I. (2004). Selbstbohrende Holzschrauben und ihre Anwendungsmöglichkeiten.
Holzbaukalender 2004, pp. 516‐541. Bruderverlag Karlsruhe.
Blass H.J., Ehlbeck J., Kreuzinger H. and Steck G. (2005). Erläuterungen zur DIN 1052:08‐2004.
Herausgeber: Deutsche Gesellschaft für Holzforschung. Bruder‐Verlag, Karlsruhe.
Brünninghof H., Schmidt K. and Wiegand T. (1993). Praxisnahe Empfehlungen zur Reduzierung von
Querzugrissen. Bauen mit Holz 11:928‐937.
Ehlbeck J. and Görlacher R. (1989). Tragverhalten von Queranschlüssen mittels Stahlformteilen,
insbesondere Balkenschuhen, im Holzbau. Forschungsbericht Universität Karlsruhe.
Henrici D. (1984). Beitrag zur Spannungsermittlung in ausgeklinkten Biegeträgern aus Holz.
Dissertation, Technische Universität München.
277
D9 Diaphragms and bracings
Original articles: T. Alsmarker, H. Brüninghoff, S. Winter
A building is not only exposed to vertical forces such as permanent loads and imposed
loads, but horizontal forces like wind or earthquake loads must also be absorbed and wind
affects a building in various ways. Its direct impact translates into pressure on one or multi‐
ple external surfaces and suction on the remaining external surfaces. Figure D9‐1 shows the
main wind load distribution in a building where the wind is blowing in a direction perpen‐
dicular to the longitudinal wall. The wind in the direction shown in Figure D9‐1 exerts pres‐
sure on the facing wall and half the roof and suction on the area away from the wind. When
the roof pitch is low, suction is also exerted on the wind‐facing half of the roof (in Fig‐
ure D9‐1: „alt“ = „or“). Notable here is the fact that the suction effect on the gable walls
acts perpendicular to the wind direction. In addition to the main wind loads, the wind can
also exert pressure or generate suction on the interior surface of the building.
+ alt
Figure D9‐1 Wind load distribution with the wind direction perpendicular to the longitudinal wall.
The arrow indicates the wind direction. (STEP 1995 Article B13)
279
Diaphragms and bracings
In addition to wind loads, horizontal bracing loads also often have to be accommodated
and the latter occur when force is applied to slender beams, deflecting them laterally
(lateral torsional buckling). This is prevented using bracing with significant stiffness per‐
pendicular to the principal load‐bearing direction of the slender beams. The bracing is
not subject to any stresses if the beams to be braced are completely straight and the
forces are transmitted centrally and vertically i.e. when deflection occurs only in the
principal plane and no stress components perpendicular to the principal load‐bearing
beam directions are developing. However, in reality, inevitable production and assembly
irregularities (imperfections) lead to deviations from the ideal position, while additional
changes in form due to wind loads or otherwise, from externally imposed horizontal
forces exacerbate any eccentricity due to imperfections. The basic load‐bearing behav‐
iour when exposed to wind perpendicular to the longitudinal wall is shown in Figure D9‐2
for a very simple building, which uses floor and wall diaphragms. In this case, the walls
are only connected to the foundation and the (horizontal) roof, which means half of the
entire wind force is directed towards the horizontal roof diaphragm, which acts as a deep
beam. The roof diaphragm rests on the side walls, which transmit the forces into the
foundations via their in‐plane shear action. The bracing system comprises multiple ele‐
ments, which must be correctly connected to ensure a complete load path of shear forces.
In addition to other measures, this includes proper fastening of the horizontal diaphragm to
the vertical diaphragms (shear walls) and hold‐down anchors of the walls to the foundation.
Figure D9‐2 In‐plane load‐bearing in a simple building, where the roof functions as a horizontal diaphragm
and the side walls function as vertical diaphragms = shear walls. (STEP 1995 Article B13)
280
Diaphragms and bracings
D9.1 Possibilities, arrangement and design of bracing
Load distribution
Lateral load resisting systems designed to distribute horizontal forces can comprise verti‐
cal elements or combined vertical and horizontal elements, where the following mini‐
mum criteria apply:
For existing floor diaphragms, at least three shear walls must be present, which
do not intersect with each other at any point and which are not all parallel.
In the absence of a floor diaphragm, at least four shear walls must be present,
of which no more than two walls respectively may intersect at any point.
There is also scope to add additional structural members to this base unit, which are
braced sufficiently to transfer all or part of the horizontal forces to which they are ex‐
posed, via the base unit into the foundation.
Bracing elements within a structural system should be arranged such as to allow a sym‐
metrical load transfer. In all other cases, the additional forces generated by the eccen‐
tricity between the centre of gravity of the introduced load and that of the bracing must
also be taken into consideration. The support forces for continuous floor diaphragms must
be determined in a manner similar to that applied to the series of simply supported beams.
a
F
F1 F2 F1 F2 F1 F2
b
I F3
F3 F3
F F
M = 0 F1 = F F1 = F2 = F/2
F ∙ b – F2 ∙ a = 0 F2 = F3 = 0 F3 = 0
F2 = F ∙ b/a = F1
F3 = F1
Figure D9‐3 Forces in shear walls, statically determinate system. (STEP 1995 Article E14)
281
Diaphragms and bracings
For each load direction, the bracing elements used should have the same stiffnesses,
since otherwise additional forces will be generated due to eccentricity. The following
approximated calculation method for statically indeterminate systems (Figure D9‐4) only
applies provided the stiffnesses of all shear walls are the same. Where a more accurate
consideration of the stiffness of the wall panels is required, as Steinmetz (1992) remarks,
the bending deformation, shear deformation and the semi‐rigidity of all fasteners must
be taken into consideration.
ex W
y
y
x,i by,i bx,i
by,i
Wx bx,i
S
ey
y,i
ys
bx,i by,i
xs x
Figure D9‐4 Shear walls, statically indeterminate system. (STEP 1995 Article E14)
The forces in the statically indeterminate system in accordance with Figure D9‐4 can be
calculated corresponding to the following equations. Eccentricities ēx and ēy are the lever
arms of the load application point to the centre of gravity of the bracing.
Centre of gravity S (xs, ys) of the bracing:
b yi x i b xi y i
xs and ys (D9‐1)
b yi b xi
The eccentricities of individual panels i from the centre of gravity S of the bracing are:
s xi x i x s and s yi y i y s (D9‐2)
The forces on wall panels i with wind direction Wx are determined as the sum (i) of the
portion of Wx per length bxi of wall panels and (ii) the force from the additional moment
Mx caused by the eccentricity ēy of the load application point (note algebraic signs!):
b xi s yi b xi b s yi b xi
H xi W x ΔM x xi W x W x e y (D9‐3)
b xi Ip b xi 2
2
b xi s yi b yi s xi
Ip is the polar moment of inertia of the wall panels about the centre of rotation = centre
of gravity S.
282
Diaphragms and bracings
The additional moment Mx also generates horizontal loads in a y‐direction with wind
direction Wx:
s xi b yi s xi b yi
H yi ΔM x Wx e y
(D9‐4)
Ip b xi s yi2 b yi s xi2
The forces on the wall panels i with wind direction Wy are similarly determined:
b yi s xi b yi b yi s xi b yi
H yi W y ΔM y W y W y e x
(D9‐5)
b yi Ip b yi 2 2
b xi s yi b yi s xi
s yi b xi s yi b xi
H xi ΔM y Wy e x
2 2
(D9‐6)
Ip b xi s yi b yi s xi
Examples for vertical bracing systems
Portal frames with moment‐resisting connections in their corners can absorb both
vertical and horizontal forces and are widely used in single‐storey hall structures in par‐
ticular. Dowel‐type fasteners may be used to produce such moment‐resisting connec‐
tions (e.g. a dowel circle) or finger joints. As a rule, for horizontal loading, portal frames
are only used to resist wind loads and when additional horizontal forces are exerted, e.g.
horizontal braking forces of crane systems, the cross‐sectional sizes would rapidly in‐
crease mainly due to the need to accommodate more fasteners. One further downside is
the limited clear height of the construction and these reasons often spawn what are
uneconomical solutions.
Cantilever columns can be used, particularly for hall structures involving crane operation
and made of reinforced concrete or steel members (e.g. reinforced concrete cantilever
column with sleeve foundations). For lower horizontal forces, meanwhile, timber poles or
glued laminated timber columns (Heimeshoff, 1983) can also be used, e.g. for agricultural
constructions or smaller hall structures. When clamping glued laminated timber columns
directly in sleeve foundations, a national German technical approval Z‐9.1‐136 applies.
However, given the high risk of fungal attack on the moulded column footing due to
moisture, the clamped end must be very carefully protected.
283
Diaphragms and bracings
A diagonal bracing may include either elements resistant to both compression and ten‐
sion, e.g. diagonals in trusses, or solely tensile braces. Normally, the set‐up includes
cross‐arrayed components, e.g. round steel bars with turnbuckles. Diagonal bracings are
either used within wall constructions or as visible elements, which often fulfil a creative
function at the same time.
Strip braces are usually used for bracing of roof structures and, in smaller buildings, also
for bracing walls or floors. They are normally nailed to their respective components. The
design load‐bearing capacity is determined either by the load‐bearing capacity of the
joint or the load‐bearing capacity of the strip brace in the net cross‐section. Strip braces
must be tensioned at the time of assembly and their advantages include fast and simple
assembly and low costs, although the assembly is often defective in practical terms.
Further downsides include the low load‐bearing capacity and the sensitivity of the mate‐
rial to temperature. If a strip brace is assembled amid very low temperatures, significant
deformation will occur during the summer months due to the steel expanding in re‐
sponse to heat. The system must withstand significant displacement before being able to
accommodate forces, which may lead to deformations of the overall structural system.
Round steel bars with turnbuckles are particularly suitable for bracing structural systems
and tend to be designed on an individual basis. Numerous options apply to the joint
design, certain examples of which are specified in Figure D9‐5. Bracings with round steel
bars can resist large forces and also have the advantage of being adjustable during as‐
sembly to align the construction. Provided corresponding detailing applies, bracings with
round steel bars can boost the overall visual impression of the construction.
Struts made of solid timber or glued laminated timber are designed taking buckling
lengths into consideration and they can resist compressive and tensile forces. The joints
can be established e.g. by nails, bolts, wood‐based panels, steel plates (Figure D9‐5 (a))
and this type of bracing may be more economical than using round steel bars. One down‐
side, however, is the more complex assembly. Once the struts have been installed, there
is no further scope to align the construction, hence the need for ultra‐precise execution.
The ability to dissipate compressive and tensile forces results in more rigid constructions
with fewer deformations.
284
Diaphragms and bracings
Vertical diaphragms (shear walls) are often executed in the form of timber frame con‐
structions, comprising solid timber members (top and bottom plate, studs) and wood‐
based panel materials as sheathing. Such shear walls, made of wood‐based panels and
timber frames, are used very frequently when constructing single‐ and multi‐storey
homes. The multiple functionality of the wall, i.e. the ability to resist to horizontal forces
as well as protect against fire, heat and sound, make this type of bracing the most eco‐
nomical option overall. When using wood‐based panels, moreover, there is great scope
for prefabrication. Suitable strength properties mean that particleboards, OSB and ply‐
wood are particularly good choices when applied as sheathing. The shear forces are
transferred via numerous slender fasteners (nails or staples) along the panel edges in the
timber frame. The joints of the wall elements must each be separately verified and the
load‐bearing capacity of the shear walls against compressive and tensile stresses must
take into consideration the combination of vertical and horizontal forces. The ends of
shear walls must be prevented from uplifting, while the compressive forces acting on the
respective opposite side must also be taken into consideration during the design (com‐
pression perpendicular to the grain on the bottom plate). Massive shear walls made from
cross‐laminated timber are an increasingly popular alternative to timber frame walls.
Examples for horizontal bracing systems
Every floor of a building and its roof must be braced with horizontal elements, while
bracing elements may include either horizontal diaphragms or diagonals respectively. For
sloping roofs, meanwhile, the bracing is normally located within the plane of the roof,
while in floors, wood‐based panels are used to accommodate vertical and horizontal
loads as well as absorbing shear forces in the horizontal plane. Edge beams and floor
beams carry the flexural forces of the horizontal diaphragm. The floor functions like a
beam, which transfers the horizontal forces into the vertical bracing. Horizontal diagonals
or roof diagonals are parts of a truss, which transfer the horizontal forces into the adja‐
cent vertical bracings. Accordingly, all joints and the compression and tension members
of this truss have to be verified. Roof‐level bracings may include round steel bars or strip
285
Diaphragms and bracings
braces in combination with purlins or small trusses. Trusses with punched metal plate
fasteners always have to be laterally braced by additional trusses at roof level. For trusses
or glued laminated timber beams, the bracing at roof level is not only to resist the hori‐
zontal forces generated, but also to protect against lateral buckling.
Procedure for conceptual and structural design
Defining the building geometry and the lateral load resisting walls and building
parts intended to accommodate horizontal forces. In areas at risk of earthquakes,
as well as defining a system to resist wind loads, a bracing system to resist earth‐
quake loads should also be determined. The ductility of the vertical bracing sys‐
tems influences the earthquake design load (Article G2). Under certain circum‐
stances, brittle bracing elements or joints can absorb higher wind loads, but when
designing for earthquake loads, this leads to uneconomical results.
Determination of horizontal forces. The current systems for accommodating wind
load, e.g. continuous columns in gable walls of halls up to roof level, must be tak‐
en into consideration, since determining forces via tributary areas alone may dis‐
tort results.
Calculation of the resulting forces in the bracing elements. For symmetrically
braced buildings, this is done in accordance with simple rules to determine the re‐
action forces in statically determinate structures. In a building in which the bracing
elements are unevenly distributed or with uneven stiffness distribution of the
bracing, e.g. due to large openings in the walls, the resulting torsional forces in the
building must be taken into consideration (cf. Figure D9‐4). The greater the dis‐
tance between the bracing elements, the greater the torsional resistance of the
structural system.
Verification: Ed ≤ Rd
Verification of joints of the bracing elements. The flow of forces from the load
application point to the foundation must be verified, as must the following joints:
‐ Introduction of the forces of immediately stressed components such as vertical
studs or rafters in horizontal elements, e.g. floor diaphragms, particularly ten‐
sile forces due to wind suction.
‐ Joints within horizontal bracing, e.g. joints of floor or wall diaphragms, joints of
horizontal trusses, joints of purlins, etc.
‐ Joints between horizontal and vertical elements, e.g. connecting floor dia‐
phragms to wall diaphragms or connecting chords of horizontal trusses to verti‐
cal diaphragms or diagonals.
286
Diaphragms and bracings
‐ Joints within vertical bracing, e.g. nailing of wood‐based panels onto timber
frames or connecting of prefabricated wall elements.
‐ Joints between vertical elements and foundations, e.g. the end anchoring of
shear walls or anchoring of steel diagonals to the bottom of columns.
Any form of joint which may lead to stress being exerted perpendicular to the
grain in the timber member should be avoided since the low strength may trigger
sudden failure of the joints in question.
Verification of deformations. The permissible scope of deformations depends on
the building type and verification is particularly important for members simulta‐
neously stressed by high vertical forces. In this case, high eccentricity of the forces
will generate significant additional forces. However, also with a view to avoiding
cracks in cladding, deformations of structural systems should be limited.
Verification of conditions during erection. It is important to guarantee the struc‐
tural stability of the building while erecting the structural system. However, when
constructing houses with prefabricated walls, a short‐term makeshift form of
bracing, e.g. via adjustable inclined steel columns is sufficient and there is general‐
ly no need for separate verification. For hall structures, however, separate verifi‐
cations may be required. In the normal case, the construction progress involves
the vertical members being erected first and subsequently connected to the hori‐
zontal bracing elements.
In the following section, floor and wall diaphragms which constitute key bracing elements
for timber structures will be explained in more detail, before discussing the design and
arrangement of bracing.
D9.2 Diaphragms
Horizontal diaphragms
Floors and roofs can be used to transmit horizontal forces to the supporting walls. In
timber frame buildings, these horizontal structures generally comprise wooden joists,
which are sheathed with a range of wood‐based panels. The double top plate of the walls
is often used as a chord of the horizontal diaphragm, while the wood‐based panels have a
staggered arrangement and are joined using nails or bolts. Alternatively, a continuous
header or trimmer joist may be used as the chord. The remainder of the article focuses
solely on horizontal diaphragms in accordance with Figure D9‐6, which are exposed to a
horizontal and uniformly distributed load.
287
Diaphragms and bracings
For this type of floor diaphragm, it is safe to assume that the load‐bearing behaviour
resembles that of a deep I‐beam with a depth of b. According to EC 5, this is applicable,
provided the span ℓ is between 2 ∙ b and 6 ∙ b. The sheathing functions as a web resisting
shear forces, while the chords act like flanges, which transmit the bending moment ex‐
erted; see Figure D9‐7, forces Fc and Ft with lever arm b.
According to NA, the simplified method may also be used for diaphragms where the span ℓ
is smaller than 2 ∙ b, if joists, which are continuous in the load direction and over the dia‐
phragm width b, transmit the loads uniformly into the diaphragm or if the diaphragm width
b is only calculated as equating to half the span of the diaphragm. (NCI for 9.2.3.2 (NA.5))
Figure D9‐6 Horizontal (floor) diaphragm. (STEP 1995 Article B13)
Fc
b Vd
Ft
l
Figure D9‐7 Principle behaviour of floor diaphragms. (STEP 1995 Article B13)
288
Diaphragms and bracings
The simplified verification presumes that the entire moment is absorbed by the chords.
Accordingly, the chords must be designed to resist tensile or compressive forces of
M max,d
F t,d F c,d (D9‐7)
b
where
Mmax,d Design value of maximum bending moment
b Diaphragm width
The sheathing must be designed for a shear flow vd (see Figure D9‐7) of
F v,d
vd (D9‐8)
b
where
Fv,d Design value of maximum shear force
b Diaphragm width
In this case, vd is assumed to be uniformly distributed over the diaphragm width b.
Finally, the spacing s of the fasteners which is used to connect the sheathing to the joists
is determined as follows:
F v,Rd
s (D9‐9)
vd
where
Fv,Rd Design load‐bearing capacity of a joint with laterally loaded fasteners
vd Shear flow from equation (D9‐8)
For simply supported diaphragms in accordance with Figure D9‐7, the shear force is
transmitted from the horizontal diaphragm to the shear walls via the edge members at
the end of the diaphragm. Here, the shear force is assumed to be uniformly distributed
along the diaphragm edge. Edge members and chords must also be correctly connected
to the top plate, to transmit the shear forces to the shear wall underneath. It is important
to ensure that a flow of forces is possible via other structural elements, if the sheathing is
not directly connected to these load‐bearing elements.
289
Diaphragms and bracings
In the proposed calculation model, the sheathing is assumed to function as one, which is
why the individual sheathing panels must be properly connected. The optimal floor dia‐
phragm design involves a staggered arrangement of sheathing panels, Figure D9‐7. How‐
ever, since the diaphragm action must be present in both directions, staggering should
therefore be oriented in the most unfavourable loading direction. When large openings
are present in the floor diaphragm, it is important to ensure that the forces in the area of
the openings are reliably transferred, which can be done for compressive and tensile
forces via trimmer beams and steel straps. Effective transmission of the shear forces is
ensured if the sheathing is appropriately nailed or screwed to the trimmer beams and
joists around the opening.
Vertical diaphragms (shear walls)
Timber frame walls generally comprise vertical studs installed at regular intervals, which
then form a frame, together with the top and bottom plates. This frame is generally
sheathed on one or both sides with different types of wood‐based panels which are
nailed or stapled onto the frame. Structurally, the wall can be considered a cantilevered
diaphragm, the top plate of which is exposed to a concentrated horizontal load. Since the
sheathing is used for bracing, this load can be effectively transferred to the foundations.
Figure D9‐8 sets out the load‐bearing behaviour schematically.
The studs are connected to the bottom and top plate with nails or other metal fasteners,
while from a structural perspective, the frame joints are deemed hinges, Figure D9‐8 (b).
The sheathing and fasteners fixing the sheathing to the frame must prevent any dis‐
placement of the timber frame and meanwhile, the fasteners are most highly stressed at
the corners, since this is where the largest slip between the frame and sheathing takes
place. The fasteners in the top left and bottom right corners will have force directions
towards the free edge of the sheathing. In Figure D9‐8, it was assumed that the studs are
anchored to the foundation. The key criterion dictating the load‐bearing capacity of a
timber frame shear wall is whether or not the uplift of the studs can be prevented. In
addition to verification of uplift, verification of the ability of the studs to withstand con‐
centrated compressive load is required. Both the load‐bearing capacity of joints as well as
the shear strength of the sheathing are additional important factors impacting on the
load‐bearing capacity of shear walls.
The load‐bearing capacity of a wall comprising multiple wall elements is calculated as the
collective load‐bearing capacities of the individual elements, even when wall elements
include a range of different sheathing material and various fasteners. The same approxi‐
mation can be used when different sheathing material and fasteners are used on both
sides of the frame, but according to EC 5, in this case, the specified load‐bearing capacity
of the weaker side should be halved. Within a wall with window openings, when calculat‐
ing the overall load‐bearing capacity for the wall, wall element lengths with these open‐
ings should be disregarded.
290
Diaphragms and bracings
y
(a) b b (b)
2 2
s
H
ϕ
y γ
t
h Tp x
r
x
Figure D9‐8 (a) Typical wall element, (b) principal structural behaviour. (STEP 1995 Article B13)
According to method A of EC 5, the design racking resistance Fi,v,Rd of each wall panel i
can be calculated as
F f,Rd b i c i
F i,v,Rd (D9‐10)
s
whereby Ff,Rd is the design joint capacity and s the fastener spacing, while coefficient ci
considers the ratio of the wall panel height to the wall panel.
The tension studs, hold‐down anchors and compression studs must resist the force
F i,v,Ed h
F i,c,Rd F i,t,Rd (D9‐11)
bi
where Fi,v,Ed equates to the design horizontal force acting on wall panel i of width bi and
height h.
291
Diaphragms and bracings
Any vertical load may be taken into consideration when determining the vertical tensile
force and must be considered when determining the vertical compressive force. To apply
method A, the end studs of the wall panel must be vertically secured, either by the verti‐
cal load or by using hold‐down anchors, while the bottom plate must also be sufficiently
anchored to the foundation, to allow transmission of the horizontal forces. However,
other design models are available in the event that the end studs are not vertically se‐
cured. In multi‐storey buildings, the wall panels must be interconnected such as to allow
transmission of the tensile forces from one floor to another.
In conclusion, the prerequisites for the validity of the simple calculation model presented
in this document should be cited (Colling, 2011):
All edges of the sheathing are supported by studs and continuously fastened
to the same.
The forces are uniformly distributed and continuously transferred via the
fasteners between the sheathing and top plate (concentrated loads cannot
be considered).
The sheathing does not buckle.
The studs and sheathing are rigid in comparison to the joints, plastic hinges
are developing in the fasteners and thus the load‐bearing capacity of the wall
panel is determined by the load‐bearing capacity of the joints.
Internal walls
The distribution of the horizontal load to the internal walls depends on the ratio of stiff‐
ness of the floor or roof diaphragm in relation to that of the wall. One assumed limit case
involves a rigid horizontal diaphragm, supported by flexible walls, while the other in‐
volves a flexible horizontal diaphragm, supported by rigid walls. The first case involves
the horizontal load being distributed to the wall panels based on the stiffness of the
walls. If there is a rigid floor diaphragm on three walls, each of equivalent stiffness, each
wall absorbs one third of the overall load. Here, it should be noted that when the walls
are arranged asymmetrically, the resulting torsional moment is to be taken into consider‐
ation. If we assume a flexible floor diaphragm supported by rigid walls, it depends on
whether the floor diaphragm is configured as a horizontal continuous beam or as simply
supported beams. A safe assumption would be that the walls at diaphragm ends are
considered supports of a simply supported beam and the internal walls as intermediate
supports of a continuous beam. When timber floor diaphragms are used on timber shear
walls, this is between both the above‐mentioned limit cases and the assumption of a rigid
floor diaphragm should be made cautiously. Rigid floor diaphragms should only be as‐
sumed, if the ratio of the diaphragm width b to the span ℓ between the internal walls is
around one.
292
Diaphragms and bracings
D9.3 Bracing
Background to the design equations in accordance with EC 5
Variables influencing actions on bracing members
Regarding actions on bracing, a distinction has to be made between compression and
bending members. Also relevant is whether the scope of consideration includes a highly
stressed individual support or a series of supports, which e.g. form a bracing structure in
the shape of a truss. The actions on the bracing members particularly depend on the
cross‐sectional and longitudinal dimensions of the structure to be braced, the support
conditions and the construction material properties, which, in turn, are determined by
the selection of a strength class, service class and the load duration class of the governing
load case. The stiffnesses of members and joints play a key role, not only as attributed of
the structure to be braced, but particularly for the bracing structure itself. Verification in
accordance with the second order theory moreover has to take geometric and structural
imperfections into consideration, e.g. in the form of initial deformations.
Individual supports of compression members
Compression members of length ℓ, which are supported (= braced) at regular intervals by
flexible (elastic) supports, generate very considerable support (= bracing) forces, assum‐
ing the deformation patterns in accordance with Figure D9‐9 (b) and (c).
Nd I = const. Nd Nd Nd
C C C C C
a) l Fd Fd Fd Fd Fd
c) Nd Nd
e
Nd Nd
Fd
a
Fd Fd=0 Fd Fd=0 Fd
Nd Nd
b) Nd Nd
e
d) l
Fd
a a
Figure D9‐9 Static system and deformations of braced compression members. (STEP 1995 Article B15)
293
Diaphragms and bracings
Möhler and Schelling (1960) showed that the minimum spring stiffness C of the supports
should have the following value:
E I
C k s
2
(D9‐12)
a3
where
k s 2 1 cos (D9‐13)
m
and where a is the length and m the number of waves, so that with ℓ = m ∙ a and with
hinged supports at both ends, a wave‐shaped deformation pattern is attained with one
wave for ks = 2 or a limitless number of waves for ks = 4 (in EC 5, ks = 4 is recommended).
If we now use the (elastic) critical load Nd of the Euler buckling case II (equation (D2‐19))
in equation (D9‐12), the following equation given in EC 5 to determine the minimum
spring stiffness follows:
k s 2 E I k s
C 2
N d (D9‐14)
a a a
The spring force Fd (see Figure D9‐10) can be calculated conservatively and in accordance
with second order theory at:
e
F d 5.2 N d (D9‐15)
2a
where e equals the initial deformation of the member axis.
2a
a a
Fd / 2 Fd / 2
Nd x f e
Nd
y EI
C
Fd=C(f-e)
Figure D9‐10 Deformation of an elastically supported beam. (STEP 1995 Article B15)
294
Diaphragms and bracings
Fd 1
200
=k
N
Fd<5,2 Nd e = d
Nd
150 2a k1
96100
58 50
0
100 300 500 700 900 2a
e
Figure D9‐11 Coefficient of bracing force as a function of initial deformation. (STEP 1995 Article B15)
Figure D9‐11 shows the emergence of support forces in the springs of Fd = Nd/58 and
Fd = Nd/96, assuming initial deformations of ℓ/300 and ℓ/500 for members made of solid
timber or glued laminated timber. These values were rounded up in EC 5 to the recom‐
mended values of Nd/50 and Nd/80.
Individual support of bending members
Burgess (1989) proposed specifying the design value of mean compressive force Nd in the
compression zone of a bending member as follows:
N crit
Nd Md (D9‐16)
M crit
whereby Ncrit and Mcrit signify the elastic buckling load and elastic buckling moment in
accordance with classical theory of stability (Article D2). EC 5 specifies the following
approximation:
Md
N d 1 k crit (D9‐17)
h
This involves calculating the coefficient kcrit for the member without bracing (lateral tor‐
sional buckling of beams, Article D2), in which the torsional stiffness of a beam is also
taken into consideration. Where kcrit = 1, no bracing is required. The method is only appli‐
cable for bending members braced along their compression zone.
Bracing of beam or truss systems
To determine the load in the bracing, it is assumed that the compression or bending
members to be braced by a bracing structure of stiffness (E ∙ I)ef have an initial defor‐
mation in the shape of a half sine wave (see Figure D9‐9 (d)).
295
Diaphragms and bracings
Nd Nd Nd Nd
(EI)ef
(a)
(b)
e
x
(c)
y
Nd Nd Nd Nd
Figure D9‐12 Deformation of braced compression members. (a) Straight line, (b) axis with initial deformation,
(c) deformed axis. (E ∙ I)ef is the stiffness of the bracing structure. (STEP 1995 Article B15)
The compressive forces Nd generate a moment (for all definitions see Figure D9‐12):
M d nN d y (D9‐18)
Disregarding the bending stiffness E ∙ Iz of the slender members to be braced, the differ‐
ential equation of the elastic curve is obtained, taking the initial deformation y0 into
consideration:
Md d 2 y y 0
y y 0 (D9‐19)
(E I) ef d2
Replacing the moment Md from equation (D9‐18) in equation (D9‐19) we have:
2
nN d
y y e sin x (D9‐20)
(E I) ef
with the initial deformation function
y 0 e sin x (D9‐21)
296
Diaphragms and bracings
The solution of the differential equation results in
2
e nN d
qd sin x (D9‐22)
nN d
1 2
(E I) ef
see also Brüninghoff (1983).
Applying equation (D9‐22) requires that the stiffness of the bracing structure (E ∙ I)ef,
including joint slip, be used as an input parameter. To simplify the design for frequently
occurring practical cases, EC 5 limits the deflection caused by qd and additional external
actions of the bracing structure to ℓ/500:
4 1
max y q d (D9‐23)
4
(E I) ef 500
If we cancel (E ∙ I)ef from equations (D9‐22) and (D9‐23) and convert the sinusoidal load
into a uniformly distributed load, the result is
nN d
qd k (D9‐24)
30
where kℓ = 1.
For spans exceeding 15 m, the expected level of execution accuracy means that the initial
deformations will no longer increase proportionally with the span, whereupon the initial
deformation can be reduced in accordance with equation (D9‐21) by a factor of
15
k (D9‐25)
where ℓ is the span in [m].
The structural engineers should verify the deflections of the bracing structure, if it is
thought likely that the above‐specified deflection limits will be exceeded. If bending
members are to be braced rather than compression members, the side exposed to com‐
pression should also be braced, to ensure equations in accordance with EC 5 are applica‐
ble. The torsional stiffness of the bending members can be taken into consideration by
calculating the compressive force in accordance with equation (D9‐17).
297
Diaphragms and bracings
Uses of bracing
Bracing is designed to resist external forces, which do not arise from the structure itself
and which thus have to be transferred to the foundations:
Wind loads,
Horizontal loads, e.g. braking forces of a crane, seismic loads.
Bracing structures are also capable of absorbing internal forces, generated by existing or
forced deviations from the planned position of the structural members. These forces can
be balanced within the correctly detailed structure itself and therefore do not generally
need to be transmitted to the foundations:
Forces due to lateral displacements of frames and columns,
Forces arising from bracing beams and compression chords of trusses;
both of which are liable to lateral torsional buckling,
Forces on intermediate supports of compression members,
Forces on the lateral supports of buckled tension chords.
Figure D9‐13 to Figure D9‐18 show examples of bracing structures.
Figure D9‐13 Transfer of wind loads from gable columns through roof and wall bracing. (STEP 1995 Article D9)
(b)
(a)
Figure D9‐14 Transfer of external loads through bracing into the foundation. (a) Braking forces of a crane,
(b) bracing to provide torsional resistance at beam ends. (STEP 1995 Article D9)
298
Diaphragms and bracings
z
x y
Figure D9‐15 Transfer of P‐ forces from inclined columns though roof and wall bracing. (STEP 1995 Article D9)
Figure D9‐16 Support reactions for inclined structures with vertical actions (left) and lateral intermediate
support to decrease column lengths (right). (STEP 1995 Article D9)
Figure D9‐17 Lateral bracing of beams at risk of buckling. (STEP 1995 Article D9)
299
Diaphragms and bracings
Figure D9‐18 Lateral support of tension chords using knee‐bracing against purlins. (STEP 1995 Article D9)
The key function of the main load‐bearing elements within a structural system is to
transmit vertical forces, e.g. self‐weight and snow on roofs, while bracing structures are
designed to absorb horizontal wind and stabilisation forces (buckling forces from the
main load‐bearing element). The structural engineer would normally handle these ac‐
tions separately and design main elements and bracing in a range of steps. In reality,
however, the system is actually a three‐dimensional structure (Figure D9‐19). In this case,
the connection of two main members, the two simple triangular trusses shown in Fig‐
ure D9‐19, with one upper and two lower purlins, adding two diagonal bracing elements
initially suffices. If the support reactions are now counted, a total of eight emerge, one of
which is required to secure the existing hinged square at eaves level. This leaves seven
support reactions with six possible equilibrium conditions in space. Thus already this
simple system remains statically indeterminate.
z y
x
Figure D9‐19 Support reactions of the three‐dimensional braced system. (STEP 1995 Article D9)
300
Diaphragms and bracings
The engineer simplifies the calculation, for example, by disregarding a support reaction in
the x‐direction, see Figure D9‐19, which means the horizontal forces generated from the
bracing are thus transferred to the two supports in the y‐direction. This only holds, how‐
ever, if the actual displacements of the supports deemed to be fixed in the y‐direction
are nearly the same; otherwise the “omitted” support will have to resist forces in the x‐
direction resulting from the rotation of the structure in plan. Therefore, this assumption
is only possible, because the deformations of the chords are negligible and result only in
the bracing being displaced in the y‐direction. If the length of the posts and diagonals
change and a connection slip in the truss joints occurs, this results in shear deformation
as shown in Figure D9‐20. Accordingly, the vertical element at the support will remain
vertical without causing any horizontal reaction in the “omitted” support.
y
x
Figure D9‐20 Deformation of a bracing structure. (STEP 1995 Article D9)
In sloping roofs, the span of the bracing is assumed to equal to the length of the roof
area and is assumed to be a planar structure. The spatial load‐bearing behaviour gener‐
ates deviation forces in the apex zone, which should also be taken into consideration
(Figure D9‐21).
301
Diaphragms and bracings
(a)
F1 F1 F1 F1
(a)
Figure D9‐21 Vertical load of beams caused by horizontal actions if the beams are not straight.
(STEP 1995 Article D9)
For the calculation model, it is both practical and advisable for roofs with a significant
incline to disregard any transfer of shear forces in the apex zone, leaving two cantilevers
(Figure D9‐22). If the system and loading are symmetrical and the deformation of the
chords is disregarded, the horizontal displacement of the apex is identical on both brac‐
ings. No deviation forces resulting from the forces in the chords of continuous bracings
are exerted at the end of the cantilevers in this case. What do emerge, however, are two
sets of horizontal support forces in the x‐direction, which must be monitored and offset
within the structure (internal forces).
Figure D9‐22 Support reactions of system, considered as cantilever. (STEP 1995 Article D9)
It may also be advisable to provide additional supports to the bracing system at the apex
and/or at intermediate locations, which transfer the lateral forces towards the eaves.
Given a sufficiently rigid support, this results in fewer deflections due to the shortened
span of the bracing and thus reduced lateral forces.
302
Diaphragms and bracings
This measure makes particular sense, for example, in constructions with trusses with
punched metal plate fasteners as the main structural elements, which are usually nar‐
rowly spaced. This reduced spacing between the trusses does not allow for any bracing
elements with sufficient beam depth (stiffness) between the compression chords of the
trusses, if the span is assumed to be that of the main structural elements (Figure D9‐23).
Figure D9‐23 Support of the apex (additional support to shorten the span of the bracing).
(STEP 1995 Article D9)
Details of bracing
Types of bracing
Bracing systems are mainly formed as trusses by adding diagonal members to the main
structural elements, e.g. beams and purlins carrying vertical actions. Prefabricated truss‐
es can also be used and installed between the members to be braced. Although this
article only covers trusses, beams, diaphragms or individual supports can also be used for
bracing purposes. In most cases, the chords of the bracing are formed by the main struc‐
tural elements and if the latter are trusses, the compression chord to be supported
should be part of the bracing. If the main structural elements are beams, the bracing
system should be placed in their compression zone, which is then supported against
lateral deviations. Existing purlins may be bracing elements and diagonals are additionally
installed. At this point, there is also scope to select a range of geometries, which are
linked to various characteristics.
Crossed diagonals resisting only tensile forces (Figure D9‐24)
Generally made of steel, e.g. as round steel bars with turnbuckles,
Simple assembly,
Purlins will be subject to additional loads, when, as often happens,
they are used as posts in bracing trusses.
303
Diaphragms and bracings
(b) (a)
Figure D9‐24 Crossed diagonals resisting only tensile forces. (a) Braced beam, (b) purlin. (STEP 1995 Article D9)
W‐trusses (Figure D9‐25)
No additional forces in the purlins,
Due to generally fluctuating stress directions, all diagonals must resist to both
tension and compression, which is why timber members are recommended,
Securing the diagonals subject to compression to the purlins can help reduce
their buckling lengths.
(a)
Figure D9‐25 W‐truss. (a) Braced beam. (STEP 1995 Article D9)
N‐trusses (Figure D9‐26)
The application is recommended for places where one loading direction generates
larger bracing forces than another,
The vertical posts subject to compression here have lower buckling lengths than
the diagonals (here, subject to tension).
304
Diaphragms and bracings
(b) (a)
Figure D9‐26 N‐truss. (a) Braced beam, (b) purlin. (STEP 1995 Article D9)
K‐trusses (Figure D9‐27)
The forces in the struts are halved, with unchanged strut inclination,
Many connections are required,
Buckling lengths are comparatively low,
They offer larger openings if used as vertical bracing systems (Figure D9‐28),
The diagonals support the purlins at midpoint, reducing the buckling length
in the plane of bracing.
(a)
(b)
Figure D9‐27 K‐truss. (a) Braced beam, (b) purlin. (STEP 1995 Article D9)
(a)
Figure D9‐28 Wall bracing. (STEP 1995 Article D9)
305
Diaphragms and bracings
Trussed beam (Figure D9‐29)
Simple assembly, e.g. flat steel sections can be screwed or nailed onto the purlins,
The trussed beam can be optimally exploited by selecting a parabolic shape, since
the tensile force remains constant with a uniformly distributed load exerted,
Only effective in one loading direction,
The ability to accommodate non‐uniformly distributed loads must be ensured
using specific measures.
(b) (a)
D D
Z D D D Z
Z Z
Z Z
Figure D9‐29 Trussed beam. (a) Braced beam, (b) purlin. (STEP 1995 Article D9)
Connections
Connections can be detailed in a number of ways and examples shown in Figure D9‐30 to
Figure D9‐32 have worked well in practice. Timber diagonals can be connected with steel
plates using nails or dowels as appropriate fastener types, as in Figure D9‐30. For slotted‐in
steel plates, timber member and steel plate should be collectively predrilled beforehand
and here, since more closely spaced nails are possible as compared to non‐predrilled nail
holes, this paves the way for compact, rigid and high load‐bearing connections. When the
forces in the diagonals are lower, unilaterally applied nailing plates can also be used.
However, since predrilling is normally not used, the connection area is around three
times as large as for slotted‐in plates with predrilled nailing. Moreover, the diagonals are
subject to bending moments due to the eccentric connection, which should be taken into
consideration at the time of verification. Steel diagonals can be connected via cold‐
formed steel connectors like angle brackets or flat bars, as in Figure D9‐31. The example
in Figure D9‐32 is particularly easy to assemble. The round steel diagonals can be guided
through the chords and connected on the rear side. For this purpose, steel parts allowing
a wide‐ranging angle of inclination of the diagonal are available.
306
Diaphragms and bracings
(a) (b)
(c)
(d)
(c)
Figure D9‐30 Timber diagonals connected via slotted‐in steel plate. (a) Braced beam, (b) slotted‐in steel plate,
(c) timber diagonal, (d) predrilled nailed connection. (STEP 1995 Article D9)
(a)
(b)
(d)
(c)
(e)
Figure D9‐31 Connection of steel diagonals with angle brackets and flat bars. (a) Angle bracket,
(b) braced beam, (c) fastener, (d) purlin, (e) turnbuckle. (STEP 1995 Article D9)
(d)
(a) (e)
(l)
(c)
Figure D9‐32 Connection of steel diagonals on the rear side of the chords. (a) Braced beam,
(l) nailed steel fitting, (c) steel rod, (d) purlin, (e) connection with timber block and fasteners.
(STEP 1995 Article D9)
307
Diaphragms and bracings
D9.4 Literature
T. Alsmarker, H. Brüninghoff, S. Winter, original Articles B13, B15, D9, E14, STEP 1995.
Brüninghoff H. (1983). Determination of Bracing Structures for Compression Members and Beams.
Paper 16‐15‐1, CIB‐W18 Meeting 16, Lillehammer.
Brüninghoff H. et al. (1989). Holzbauwerke: Eine ausführliche Erläuterung zu DIN 1052, Teil 1 bis 3. Deutsches
Institut für Normung e.V., Deutsche Gesellschaft für Holzforschung e.V., 1. Auflage. Beuth Verlag, Berlin.
Burgess H.J. (1989). Suggested Changes in Code Bracing Recommendations for Beams and Columns.
Paper 22‐15‐1, CIB‐W18 Meeting 22, Berlin.
Heimeshoff B. (1983). Einspannung von Stützen aus Brettschichtholz durch Verguß in Betonfundamenten.
Holzbau Statik Aktuell, Ausgabe Juli 1983/7.
Möhler K. and Schelling W. (1968). Zur Bemessung von Knickverbänden und Knickaussteifungen im Holzbau.
Der Bauingenieur 43(2).
Steinmetz (1992). Die Aussteifung von Holzhäusern am Beispiel des Holzrahmenbaues. Holzbau Statik Aktuell,
Ausgabe Juli 1992/1.
308
ϭϬ dŝŵďĞƌͲĐŽŶĐƌĞƚĞ
ĐŽŵƉŽƐŝƚĞ ƐƚƌƵĐƚƵƌĞƐ
KƌŝŐŝŶĂůĂƌƚŝĐůĞ͗͘ĞĐĐŽƚƚŝ
dŝŵďĞƌͲĐŽŶĐƌĞƚĞĐŽŵƉŽƐŝƚĞƐƚƌƵĐƚƵƌĞƐƌĞƉƌĞƐĞŶƚĂĨŽƌŵŽĨŵŝdžĞĚĐŽŶƐƚƌƵĐƚŝŽŶ͕ŝŶǁŚŝĐŚ
ƚŚĞďĞĂŵƐŽƌĐŽůƵŵŶƐĂƌĞŵĂĚĞŽĨƚŝŵďĞƌĂŶĚĐŽŶĐƌĞƚĞĐŽŶŶĞĐƚĞĚ ǀŝĂƐĞŵŝͲƌŝŐŝĚũŽŝŶƚƐ
ƵƐŝŶŐ ŵĞĐŚĂŶŝĐĂů ĨĂƐƚĞŶĞƌƐ͘ dŚĞ ŵĂũŽƌŝƚLJ ŽĨ ƚŝŵďĞƌͲĐŽŶĐƌĞƚĞ ĐŽŵƉŽƐŝƚĞ ƐƚƌƵĐƚƵƌĞƐ ĨĞĂͲ
ƚƵƌĞĐŽŶĐƌĞƚĞŝŶƚŚĞĐŽŵƉƌĞƐƐŝŽŶnjŽŶĞĂŶĚƚŝŵďĞƌŝŶƚŚĞƚĞŶƐŝŽŶnjŽŶĞŽĨĂďĞĂŵĂůůŽǁŝŶŐ
ƚŚĞ ŵŽƐƚ ĂĚǀĂŶƚĂŐĞŽƵƐ ĨĞĂƚƵƌĞƐ ŽĨ ĞĂĐŚ ŽĨ ƚŚĞƐĞ ĐŽŶƐƚƌƵĐƚŝŽŶ ŵĂƚĞƌŝĂůƐ ƚŽ ďĞ ƵƚŝůŝnjĞĚ͘
dŚĞ ƐƚƌĞŶŐƚŚ ĂŶĚ ƐƚŝĨĨŶĞƐƐ ŽĨ ƚŚĞ ĐŽŶĐƌĞƚĞ ƉĞĂŬ ǁŚĞŶ ĞdžƉŽƐĞĚ ƚŽ ĐŽŵƉƌĞƐƐŝǀĞ ƐƚƌĞƐƐ͘
DĞĂŶǁŚŝůĞ͕ŝĨƚŝŵďĞƌŝƐƵƐĞĚŝŶƚŚĞƚĞŶƐŝůĞnjŽŶĞ͕ƚŚŝƐĞůŝŵŝŶĂƚĞƐƚŚĞŶĞĞĚĨŽƌĐŽŶĐƌĞƚĞŝŶ
ƚŚĞ ƐĂŵĞ͕ ĂƐ ŝƐ ŶŽƌŵĂůůLJ ŝŶĐůƵĚĞĚ ŝŶ ƌĞŝŶĨŽƌĐĞĚ ĐŽŶĐƌĞƚĞ ĐƌŽƐƐͲƐĞĐƚŝŽŶƐ͕ ǁŚŝĐŚ ĐĂŶŶŽƚ
ďĞĂƌ ƐƚƌĞƐƐ ǁŚĞŶ ĐƌĂĐŬĞĚ ĂŶĚ ƚŚĞƌĞĨŽƌĞ ƌĞƉƌĞƐĞŶƚƐ ŽŶůLJ ĂĚĚŝƚŝŽŶĂů ůŽĂĚ͘ dŚŝƐ ĞŶĂďůĞƐ
ĐƌŽƐƐͲƐĞĐƚŝŽŶƐǁŚŝĐŚƌĞƚĂŝŶŚŝŐŚƐƚŝĨĨŶĞƐƐďƵƚƌĞůĂƚŝǀĞůLJůŽǁƐĞůĨͲǁĞŝŐŚƚ͘dŚĞůŽĂĚͲďĞĂƌŝŶŐ
ĐĂƉĂĐŝƚLJ ŝƐ ǀŝƌƚƵĂůůLJ ĚŽƵďůĞ ƚŚĂƚ ŽĨ ĂŶ ĂůůͲƚŝŵďĞƌ ĨůŽŽƌ ĂŶĚ ƚŚĞƌĞ ŝƐ ƉŽƚĞŶƚŝĂů ƚŽ ŽďƚĂŝŶ
ƚŚƌĞĞƚŽĨŽƵƌƚŝŵĞƐƚŚĞďĞŶĚŝŶŐƐƚŝĨĨŶĞƐƐ͘&ŝŐƵƌĞϭϬͲϭƐŚŽǁƐĂĐŽŵƉĂƌŝƐŽŶŽĨƚŚĞĚĞĂĚ
ǁĞŝŐŚƚƐ ŽĨ ƚŚƌĞĞ ĚŝĨĨĞƌĞŶƚ ĨůŽŽƌ ĐŽŶƐƚƌƵĐƚŝŽŶƐ͖ ĚĞƉĞŶĚŝŶŐ ŽŶ ƚŚĞ ƐƉĂŶ ĨŽƌ ĂŶ ŝŵƉŽƐĞĚ
ůŽĂĚŽĨƋсϮ͘ϱŬEͬŵϮ͘
(a)
10
g,kN/m2
yyyy
9
8
(c)
yyyy
7
(b)
6
5
yyyy
4
(b)
3 (q)
yyyy
(c)
2
(a)
1
2 4 6 8 10
l,m
&ŝŐƵƌĞϭϬͲϭ ^ĞůĨͲǁĞŝŐŚƚĚĞƉĞŶĚŝŶŐŽŶƚŚĞƐƉĂŶĨŽƌǀĂƌŝĂďůĞĂĐƚŝŽŶŽĨƋсϮ͘ϱŬEͬŵϮ͘;ĂͿĂůůͲƚŝŵďĞƌĨůŽŽƌ͕
;ďͿƚŝŵďĞƌͲĐŽŶĐƌĞƚĞĨůŽŽƌ͕;ĐͿƌĞŝŶĨŽƌĐĞĚĐŽŶĐƌĞƚĞĨůŽŽƌ͘;ĂĐĐŽƌĚŝŶŐƚŽEĂƚƚĞƌĞƌ͕ϭϵϵϯ͕^dWϭϵϵϱ
ƌƚŝĐůĞϭϯͿ
ϯϬϵ
Timber‐concrete composite structures
The use of timber‐concrete composite floors can effectively reduce floor deflection, a
common problem affecting all‐timber floors. The fact that it is also far easier to dampen
vibrations in timber‐concrete composite floors than in all‐timber equivalents also facili‐
tates compliance with serviceability limit state conditions during verifications.
The in‐plane rigidity of the concrete compression zone of a timber‐concrete composite
floor is often considered indefinite compared to a timber floor, which is a key benefit
when the building is exposed to earthquake loads in particular, since rigid floor dia‐
phragms help the building retain its shape. For timber floors, of course, this cannot be
generally guaranteed. The prerequisite for this favourable behaviour in response to
earthquake load is effective and force‐locking anchoring of the timber beams and the
concrete slab in the surrounding masonry (see Figure D10‐2).
Applying a concrete slab to existing timber floors is an approach which can also improve
the sound insulation properties of floors. Firstly, the significant increase in mass reduces
the airborne sound transmission, while the greater damping values of the composite
floor also mean lower structural sound transmission. Moreover, the concrete slab repre‐
sents an effective barrier to limit the spread of fire and means such constructions have
greater fire resistance than floors made of timber beams alone. The timber beams them‐
selves outperform corresponding steel or reinforced concrete precast beams in fire resis‐
tance terms.
Figure D10‐2 Example of a timber‐concrete composite floor in a masonry construction subject to earthquake
load. (a) Main beam, (b) secondary beam, (c) brick tiles, (d) concrete slab, (e) reinforcement,
(f) glued‐in dowels using epoxy resin, (g) steel stirrups connecting concrete slab with masonry,
(h) ring beam made of reinforced concrete. (STEP 1995 Article E13)
310
Timber‐concrete composite structures
For this reason, timber‐concrete floors represent a cost‐effective alternative to those
made of reinforced concrete, particularly when renovating older buildings, since this
approach eliminates the need to replace existing timber floor beams. Generally speaking,
reinforcing timber beam floors which no longer meet present‐day requirements by apply‐
ing an additional concrete slab will suffice. The scope of any economic comparison in this
case has to go beyond the square‐metre costs of any alternative, since potential savings
can emerge elsewhere, e.g. in foundations, lower earthquake loads exerted due to the
lighter construction or less formwork and equipment required. Accordingly, timber‐
concrete composite structures are becoming a particularly popular option in some Euro‐
pean countries, e.g. to reinforce existing timber beam floors in masonry structures, floors
in new constructions or when making bridges. In Scandinavia, meanwhile, prefabricated
timber‐concrete elements are also used as wall panels, with the concrete slab giving the
necessary racking strength and the timber studs installed for the required out‐of‐plane
bending stiffness and necessary stiffening against buckling.
Experiences of the past 20 years show that despite variation in how both construction
materials respond, timber‐concrete composite structures exhibit outstanding load‐
bearing and deformation behaviour. During design and construction, there is a need to
focus on how the timber and concrete respond, particularly in terms of creep and shrink‐
ing. Subsequent sections will set out the main areas to scrutinise when designing such
constructions.
D10.1 Connection types
Figure D10‐3 shows an overview of the most frequently used connections between con‐
crete slab and timber beams and in this case, the joints have been classified in terms of
stiffness. Nailed, screwed and dowelled connections, employing fasteners perpendicular
to the joint line and as shown in Figure D10‐3 (a) 1 to 3, show the lowest levels of stiff‐
ness. Screws arranged at an angle to the joint line (Figure D10‐3 (a) 4) or surface con‐
nectors such as split ring, shear plate, toothed‐plate connectors or punched metal plate
fasteners, which do not penetrate the wood as deeply, are stiffer than connections with
dowel‐type fasteners loaded perpendicular to the fastener axis (Figure D10‐3 (b)). Even
greater stiffness is possible with shear keys where pre‐bored notches in the wood are
filled with concrete and which function as dowels after curing accordingly (Figure D10‐
3 (c)). Basically rigid connections are formed with glued joints in the manner shown in
Figure D10‐3 (d).
The effective bending stiffness E ∙ Ief of the composite cross‐section for the various con‐
nections of Figure D10‐3 amounts to around 50% (Figure D10‐3 (a)) and up to 100%
(Figure D10‐3 (d)) respectively of the bending stiffness of a cross‐section with a rigid
connection. In the event of unlimited connection stiffness, linear strain distribution over
311
Timber‐concrete composite structures
the cross‐sectional height can be assumed. The various moduli of elasticity of timber and
concrete can thus be taken into consideration, as is done when calculating composite
beams made of solid timber and wood‐based materials (see Articles D6 and D7).
Where the semi‐rigidity of the timber‐concrete connection can no longer be ignored, the
slip in the joint line must be taken into account as part of the stress analysis. In this case,
the composite cross‐section no longer remains plane when shear loads are exerted (see
Figure D10‐4), whereupon verification of the composite structure can be performed via
the ‐method shown in Article D7. This is generally required for the types of connection
shown in Figure D10‐3 (a) to (c). For semi‐rigid connections, it is also important to note
that concrete in the tension zone is prone to cracking, which will render it no longer able
to exhibit bending or axial stiffness or cross‐sectional load‐bearing behaviour. This can be
taken into consideration using an iterative calculation, in which a very low modulus of
elasticity is assigned to the concrete tensile area.
Figure D10‐3 Examples of timber‐concrete connections. (a1) Nails, (a2) dowels made of reinforcing steel bars,
(a3, a4) screws, (b1) split ring connector, (b2) toothed‐plate connector, (b3) steel tubes,
(b4) punched metal plate fasteners, (c1) pre‐bored notches with fasteners preventing uplift,
(c2) dovetail‐shaped notches with fasteners, (c3) notches with pre‐tensioned steel rods,
(c4) dowel‐laminated timber (Brettstapel) with steel shear plates, (d1) truss made of steel rods
glued to timber beam, (d2) steel plate glued to timber beam. (STEP 1995 Article E13)
312
Timber‐concrete composite structures
D10.2 Mechanical performance
The material properties of the concrete required for the calculation are the characteristic
strength value fck, the mean modulus of elasticity Ecm and the creep coefficient ; numeric
details of which are included in EC 2. For timber members, meanwhile, the characteristic
bending strength fm,k, modulus of elasticity E0,mean and the deformation factor kdef have to
be known. However, the key property for the load‐bearing behaviour is the stiffness of the
timber‐concrete connection, as identified by the slip modulus Kser or Ku. The size of the slip
modulus determines the stress distribution over the height of the composite section.
(a) bc
q
Ac ,Ec ,Ic
hc
ht y ho
δ
s s At ,Et ,It z
bt
l
(b) δ
Ac ,Ec ,Ic Q
Mc Nc Q
M
Nt Q = Kδ
Mt At ,Et ,It
σc σm,c
(c)
- -
+
-
= +
+
+
ε σ σt σm,t
Figure D10‐4 Load‐bearing behaviour of a timber‐concrete composite beam with semi‐rigid connections.
Above: composite cross‐section does not remain plane; middle: concrete slab under compressive
and bending stress, timber beams under tensile and bending stress and joint under shear load;
below: constant increase in strain distribution, normal stress distribution comprising longitudinal
and bending stresses. (STEP 1995 Article E13)
313
Timber‐concrete composite structures
The load‐bearing capacity and stiffness of the timber‐concrete connection are to be
determined by testing in accordance with EN 26891. In short‐term tests, the load‐defor‐
mation behaviour of connections can be determined, also taking into consideration any
interim layers (see ho in Figure D10‐4) between the concrete and timber. An example test
set‐up is shown in Figure D10‐5. The short‐term tests conducted in accordance with
EN 26891 can be used to determine the slip modulus Kser (in EN 26891 namely ks).
Even when plastic deformations may actually occur when reaching the ultimate limit
state of the concrete and joints, linear‐elastic behaviour can still be assumed when de‐
termining the internal forces and moments. To determine plastic deformations as accu‐
rately as possible, a nominal secant modulus is used for the modulus of elasticity of the
concrete (see EC 2) and an equivalent secant modulus for the slip modulus of the joints.
Meanwhile, the stiffness values of the concrete to be applied when determining the
internal forces and moments are calculated while assuming an uncracked cross‐section,
although when verifying stresses of the concrete cross‐section, the tensile strength of the
concrete is not taken into account. On the compression side, the compressive strength is
assumed to equate to the limit value for compressive stress which applies in the event of
plastic deformations of the concrete (see EC 2). If not all the concrete cross‐section is
under compressive stresses, there is a need to use either a corresponding reinforcement
or omit the cracked cross‐sectional part. The value used for the slip modulus of the joint
is Ku, whereby Ku = 2/3 ∙ Kser. In this case, the value Ku represents an average and not a 5%
quantile value.
F F
Figure D10‐5 Possible test arrangement to determine the load‐deformation behaviour of a timber‐concrete
connection. (STEP 1995 Article E13)
314
Timber‐concrete composite structures
Accordingly and also when verifying the ultimate limit state, in the process of determin‐
ing internal forces and moments, average values for the stiffness properties of the con‐
struction materials and connections should be used. EC 2 specifies only nominal values
for the modulus of elasticity, which are considered to constitute average values. If aver‐
age values were used for the slip moduli of the connections and the modulus of elasticity
of the concrete collectively with 5% quantile values for the modulus of elasticity of tim‐
ber, the stresses determined for the timber cross‐sections would be underestimated –
namely on the unsafe side. For verification of stresses, of course, the characteristic
strength values have to be used.
Since both concrete and timber are prone to creep and concrete in particular is subject
to longitudinal shrinkage, this must be taken into consideration when verifying composite
sections, e.g. by reducing the stiffness properties of the concrete, timber and connection
as applicable. For timber members with typical semi‐rigid connections, the impact of
creep on the distribution of internal forces and moments in the composite section can be
disregarded, since similar creep behaviour is assumed for both timber and connections.
For timber‐concrete composite structures conversely, the individual cross‐sections respond
differently to changes in the prevailing climate. The reduction in stiffness properties repre‐
sents an approximation, which equates to an overestimation of the actual deformation
occurring. Accordingly, it is conservative from a structural engineer’s perspective.
Shrinkage of the concrete in the longitudinal direction reduces the overall deformation
and thus the load exerted on the fasteners, although this exacerbates the deflection of
the composite beam at the same time. This additional deflection can be offset by in‐
stalling a precamber on the construction before applying the concrete. Since most of the
shrinkage deformation has already occurred during the process of curing the concrete
and the composite construction is supported during this period, the impact of said
shrinkage is considerably less. Still further reduction is possible via the cracks in the con‐
crete, which arise from tensile stresses that developed when shrinking is prevented. The
way in which concrete and timber behave varies respectively in the event of prevailing
climatic changes – concrete is particularly sensitive to changes in temperature, while
timber is more significantly affected by changes in moisture content – but this is only
problematic if the connections between the timber and concrete are very rigid and the
members concerned are particularly long.
315
Timber‐concrete composite structures
D10.3 Recommendations for design and execution
The following list includes certain basic recommendations for design and execution,
which should also help avoid errors:
Dry wood should be used in principle. In addition, propping should be provided for
longer after the concreting than would apply for all‐concrete slabs.
Fasteners must be protected against corrosion depending on the service class.
This can be done by hot‐dip galvanising, passivating, coating or otherwise using
stainless steel fasteners.
When thick concrete slabs are used, reinforcement is required to avoid wider
cracks developing on the tension side of the concrete cross‐section. In addition,
the cracked zone must be taken into consideration when determining internal
forces and moments.
When applying the concrete, the wood must be protected against humidity. This
can be done, for example, by using protective films or mixing additives into the
concrete, which enable the water content in the concrete to be minimised. This
also, moreover, reduces the shrinkage of the concrete. Some wood species, e.g.
larch, contain extractives which slow down the curing process or may prevent it
on a localised basis.
When a large span is involved, softer semi‐rigid joints are preferable, since this
facilitates efforts to dissipate the constraint actions caused by the shrinkage of the
concrete.
D10.4 Literature
A. Ceccotti, original Article E13, STEP 1995.
RILEM TC 111 CST (1992). Behaviour of timber‐concrete load‐bearing structures. Proc. of ACMAR‐Ravenna
International Symposium, Dept. of Civil Engineering, University of Florence.
Natterer J. (1993). Constructions en bois II. Notes du cours IBOIS, EPFL Lausanne.
316
D11 System strength
Original article: H. J. Blass
For structural systems comprising multiple similar members, positioned equidistantly and
connected via a continuous load distribution system, EC 5 allows the design load‐bearing
capacity values to be increased by a load distribution factor. One example of such a struc‐
tural system is a timber beam floor, in which the beams are linked by panel sheathing.
The load distribution factor takes into consideration two different impacts, which boost
the strength of members within a system:
The low likelihood that a member with low strength will be installed at the specific
point where stresses tend to peak,
The positive correlation between the strength and stiffness of timber members.
This positive correlation explains why more rigid members, which, by nature, are exposed
to above‐average load, can also accommodate this load. In contrast, less rigid members,
most of which are also weaker, have to carry less load. Accordingly, the load distribution
acts to offset the impact of the comparatively large natural scatter of timber properties.
Both for concentrated as well as uniformly distributed loads, the load distribution leads
to higher characteristic values for the load‐bearing capacity of members within a struc‐
tural system. For concentrated loads, the load distribution system dissipates part of the
load onto neighbouring members and thus eases the forces exerted on the member
under most strain. Figure D11‐1 shows the cross‐section of a timber beam floor under
the impact of a concentrated load.
Figure D11‐1 Floor construction under the impact of a concentrated load. (STEP 1995 Article B16)
317
System strength
When the load is uniformly distributed, the effect of the load distribution is less clear.
When members have identical stiffness, they would tend to deform to the same extent,
even in the absence of any load distribution system. However, since stiffness values tend
to vary in reality, less stiff members are prone to more extensive deformations than
stiffer equivalents. Line (b) in Figure D11‐2 shows the various deflections of timber beams
in a floor without a load distribution system. If a load distribution system is present, there
will be less overall variation of deflection (see line (a) in Figure D11‐2). In this case, the
load distribution system dissipates part of the stress from the softer to the more rigid
members.
This also applies when the behaviour of individual members is no longer linear. If, due to
microscopic cracks or plastic deformations, member stiffness declines, forces are redis‐
tributed to the members remaining intact or those located in the elastic area. The dam‐
aged member can thus still – albeit to a more limited extent – contribute to the overall
load‐bearing capacity of the system.
F F F F
W S A A
(a)
(b)
Figure D11‐2 Effect of a load distribution system on deflection of timber beams. (a) with and (b) without load
distribution system; W: member with lower stiffness; S: member with higher stiffness; A: member
with medium stiffness. (STEP 1995 Article B16)
D11.1 Application in EC 5
The system strength of such similar members, which are connected by a continuous load
distribution system and at equidistant intervals, is taken into consideration in EC 5 using a
coefficient ksys, with which the relevant strength parameters can be increased. The basis
assumption is ksys = 1.1, provided the continuous load distribution system is in a position
to accommodate the forces exerted by one member on another. Verification of compli‐
ance with the ultimate limit state of the load distribution system should be ensured for
short load duration. For laminated timber decks, meanwhile, with prestressed or glued
laminations, a higher value may be used for ksys, Figure D11‐3.
318
System strength
1.2
Laminations pre‐stressed
or glued together
1.1
ksys
Nailed or screwed
laminations
1
1 2 3 4 5 6 7 8 9 10
Number of loaded laminations
Figure D11‐3 System strength factor ksys for laminated deck plates of solid timber or glued laminated members.
D11.2 Load distribution in various structural systems
Floors and flat roofs
Foschi, Folz and Yao (1989) derived load distribution (system strength) factors for use in
designing beams in floors and flat roofs. In the process, they assumed stress was uni‐
formly distributed and the ratio between permanent and variable loads was varied. The
behaviour of the members and joints was also assumed to be linear‐elastic.
The system strength factor was derived using a reliability analysis applied to an individual
beam within a construction, which involved examining the change in the probability of
failure of an individual beam, when the beam functioned as part of a load‐bearing sys‐
tem. Parameter studies could illustrate the influence of the various variables.
The support conditions, size, number and interval distance of the beams and the ratio
between permanent and variable loads did not impact significantly on the system
strength factor ksys, while the following parameters increased the system strength factor
and enhanced the load‐bearing capacity of the system in the process:
Larger stiffness of the load distribution system in relation to beam stiffness,
Larger variance in the modulus of elasticity of the beams,
A closer correlation between modulus of elasticity and bending stiffness.
The variance in bending strength of the beams also impacts significantly on the size of
the system strength factor. Very small or large values for the coefficient of variation
(COV) respectively result in only marginal values for the system strength factor, the value
of which peaks for COV values between 0.20 and 0.30. For a typical floor structure,
ksys = 1.15 was determined, which correlates well to the value of 1.1 in EC 5.
319
System strength
Trusses with punched metal plate fasteners
Wolfe and McCarthy (1989) and Wolfe and LaBissoniere (1991) examined load distribu‐
tion in roof trusses with punched metal plate fasteners. Tests were performed with three
different truss configurations; initially with the individual trusses, followed by the entire
roof structure, where a roof sheathing made of 12 mm‐thick plywood acted as the load
distribution system. The interaction of the individual trusses within the roof structure was
estimated, which involved comparing the load‐deformation behaviour of the truss as an
individual beam subjected to linear load with the corresponding behaviour of the same
truss as part of the roof structure subject to a uniformly distributed load.
As expected, the interaction of the trusses in the roof structure increased the load‐
bearing capacity, while variation in deflection declined. Once individual trusses had
reached their load‐bearing capacity and been partially damaged, they were still capable
of accommodating a portion of their original load, since the latter was redistributed to
the still undamaged trusses. Where only one of the individual trusses in the system was
exposed to a load along its top chord, 40 to 70% of the load was transferred via the
sheathing to the neighbouring trusses. The load distribution meant the ratio of the load‐
bearing capacity of the roof structure relative to that of the weakest truss went from 1.09
to 1.47. These values, which depend on the effectiveness of the load distribution system
and the truss positioning within the roof structure, show that the design values of the
load‐bearing capacity of the entire truss, namely the members and connections, could be
increased by 10% thanks to the load distribution effect. For most structures, an accepta‐
bly conservative estimate would be ksys = 1.1.
Sheet piling
If the planks of a retaining wall are connected, e.g. by a tongue and groove joint, this will
increase the load‐bearing capacity of the retaining wall compared to that of the individual
planks. In this case, the connection between the individual planks functions as a load
distribution system. This connection also ensures virtually uniform deflection of the
planks, however large the scatter in bending stiffness.
A numerical analysis (Van der Linden et al., 1994) was performed with properties of ekki
(Lophira alata), including moisture content exceeding 30% (fm,mean = 103 N/mm2 and
E0,mean = 17600 N/mm2). The coefficient of correlation between the bending strength and
modulus of elasticity was assumed to be 0.73, while the coefficient of variation for bend‐
ing strength and modulus of elasticity was assumed at 15%. The plastic deformation
capacity of the timber was taken into consideration when exposed to high bending com‐
pressive stresses and a related reduction in stiffness before reaching the bending
strength limit. Using a non‐linear finite‐element program, the load‐bearing capacities of
simulated walls were calculated assuming wide‐ranging plank properties, while the sys‐
tem strength factor was determined as the ratio of the characteristic load‐bearing capacity
320
System strength
of a sheet piling system with ten planks to the characteristic load‐bearing capacity of the
single planks. This resulted in ksys = 1.15, although this is only applicable for bending
strength, since the load distribution system for normal forces in the planks is ineffective.
D11.3 Literature
H.J. Blass, original Article B16, STEP 1995.
Foschi R.O., Folz B.R. and Yao F.Z. (1989). Reliability‐based design of wood structures.
Structural Research Series, Report No. 34, University of British Columbia.
Van der Linden M.L.R., Van de Kuilen J.‐W.G. and Blass H.J. (1994). Application of the Hoffman yield criterion
for load sharing in timber sheet piling. Pacific Timber Engineering Conference, Gold Coast.
Wolfe R.W. and LaBissoniere T. (1991). Structural performance of light‐frame roof assemblies. II.
Conventional truss assemblies. Research Paper FPL‐RP‐499, Forest Products Laboratory, USA.
Wolfe R.W. and McCarthy M. (1989). Structural performance of light‐frame roof assemblies. I. Truss assemblies
with high truss stiffness variability. Research Paper FPL‐RP‐492, Forest Products Laboratory, USA.
321
E
Joints
E1 Joints in timber structures
Original article: P. Racher
Basic tasks of any structural design include the conceptual design of the structure, the
choice of the bracing system, structural design of the individual members and the ease
with which the structure itself can be formed. However, the joints are particularly im‐
portant, since those used in timber construction tend to be weaker than the members
being joined. The serviceability limit state is also very largely dictated by the mostly semi‐
rigid joints. Selecting a joint design within a timber construction requires consideration of
more than just the load and the load‐bearing capacity. Other aspects to be taken into
account include the visual appeal, cost efficiency and production process, as well as the
erection method and the preferences of the structural engineer or architects (Natterer et
al., 1991). This precludes a one‐size‐fits‐all approach to decide on the optimal joint for a
timber construction. Instead, key is to use the simplest possible joints and minimise the
fastener types used for a better final construction result.
The main joints differentiated are glued joints, carpentry joints and joints using various
metal fasteners. Rigid glued joints include, for example, glued‐in steel rods, which are
focused on in more detail in Article E10 and also include the finger joints used in glulam
production. Carpentry joints or contact connections meanwhile, such as step joints or
mortise and tenon joints are introduced in Article E9. This article focuses on joints using
metal fasteners. The key fasteners are introduced, while important design‐relevant
points are summarised. Since all featured joints are explained in more detail in subse‐
quent articles, the current scope merely features an overview and generally applicable
design rules, which should facilitate understanding of all following articles.
E1.1 Types of metal fasteners used in timber joints
Traditional metal fasteners can be divided into two main groups, based on the type of
force transmission between the members. The first features dowel‐type fasteners,
whereby forces exerted generate bending and tensile stresses in the fasteners as well as
embedment and shear stresses in the wood along the shank. This category includes nails,
staples, bolts, screws, dowels and threaded rods and these are the most frequently used
fasteners. The second group includes “surface‐type” fasteners such as split ring and
toothed‐plate connectors and punched metal plate fasteners, in which most of the force
transmitted is focused on the surface area of the member.
325
Joints in timber structures
Dowel‐type fasteners
Nails
Nails, a popular choice for construction parts such as horizontal and vertical diaphragms
and trusses, are the most frequently employed fasteners in timber construction and
come in wide‐ranging sizes, shapes and materials (see Figure E1‐1). The most common
types used in timber construction are round smooth shank nails, although nails with a
square cross‐section or profiled shank are also available. The dimensions of nails corre‐
spond to a range of standardised measurements in European countries, but are normally
2 to 8 mm in diameter and 40 to 200 mm long. When producing nailed joints, predrilling
of nail holes may be required to prevent the timber splitting or make it easier for the
nails to penetrate timber with a greater density. In timber constructions, meanwhile,
nails tend to be used in single‐shear joints with timber, steel or wood‐based materials.
For more on nailed joints, please refer to Article E3.
(a) (c)
(b) (d)
Figure E1‐1 Nail shapes. (a) Round smooth shank nail, (b) spiral nail, (c) ringed shank nail,
(d) machine driven nails. (STEP 1995 Article C1)
Staples
Staples as shown in Figure E1‐2 are the classic fastener choice for timber frame buildings
(at least in German‐speaking countries), given the rapidity with which such constructions
can be completed. Staples are made of high‐tensile and ductile steel, since the produc‐
tion process imposes a reshaping angle of 90° on the staple material, which it must with‐
stand undamaged. Staples are generally resinated and can only be processed using spe‐
cial fastener tools, while their very slender nature precludes any hammering in. Stapled
joints are covered in more detail in Article E3.
326
Joints in timber structures
Figure E1‐2 Staples.
Bolts, dowels and threaded rods
Bolts (Figure E1‐3 on the left) are generally made of steel, including hexagonal or square
heads and nuts and are between 12 and 30 mm in diameter. To make it easier to drive
home the bolts, as specified in EC 5, predrilled holes 1 mm larger than the bolt diameter
can be used. This hole tolerance reduces the stiffness of bolted joints. Accordingly and with
the visual appearance in mind, dowels or fitted bolts without clearance are often preferred
to bolts. Dowels are steel rods with a round cross‐section and driven into suitably predrilled
holes, see Figure E1‐3 on the right. Bolted and dowelled joints are covered in E4. Joints with
threaded rods are designed as bolted joints (with the external thread diameter).
Figure E1‐3 Left: bolt, right: dowel.
Screws
The fastest evolving fastener is the self‐tapping wood screw. Some examples are illus‐
trated in Figure E1‐4, which also differentiates the two basic types of such screws, name‐
ly “partially threaded” (with a partially smooth shank) and “fully threaded”, the latter of
which come in sizes up to an external thread diameter of 14 mm and up to 1.5 m long.
Depending on the technical approval or ETA and the timber type used, it may even be
possible to use self‐tapping screws of 14 mm in diameter without predrilling. However,
such predrilling is possible, and can be done, for example, using the core (= inner thread)
diameter. Other examples of screws include those with a thread shape in accordance
with DIN 7998, which, from a diameter of 8 mm onwards, can only be screwed into
327
Joints in timber structures
predrilled holes. However, the advent of self‐tapping screws has seen these traditional
screws used less and less frequently, so they will not be covered in any more detail at this
point (design of screwed joints with threads in accordance with DIN 7998 resembles to
that of self‐tapping screws). Screwed joints are covered in Article E5.
Figure E1‐4 Examples of self‐tapping screws.
Surface‐type fasteners
Connectors
Increasing the load‐bearing area allows connectors to accommodate considerable loads.
In truss joints, for example, this paves the way for virtually perfectly hinged joints, if a
single connector is employed for each joint, instead of multiple dowel‐type fasteners.
Figure E1‐5 shows typical forms of split ring, shear plate and toothed‐plate connectors,
whereby the installation and load capacity depend largely on the accuracy of grooving
and boring. Split ring and shear plate connectors are made of cast aluminium, cast steel
or steel and are between 60 and 260 mm in diameter. The second connector type is the
toothed‐plate connector, made of cast steel or hot dip galvanised steel sheet and be‐
tween 38 and 165 mm in diameter. Larger connectors are used when connecting glulam
members and connector diameters of up to 75 mm for solid timber. To absorb the mo‐
ments generated in the joints, which press the timber members apart, connector joints
must be secured using fasteners such as bolts or screws, which are capable of accommo‐
dating such tensile forces. The diameter of the bolt washers should correspond to around
half the diameter of the connector. Double‐sided connectors are used for timber‐to‐
timber joints. They absorb the force exerted directly between the surfaces of the timber
members to be joined and such joints are generally established at the construction site.
Single‐sided connectors can be used for both steel‐to‐timber joints as well as timber‐to‐
timber joints. They allow joints to be prefabricated, meaning only the bolts need be in‐
stalled on‐site. When these single‐sided connectors are used, the bolts loaded perpendicu‐
lar to their axis accommodate the force transmitted by the base plate of the connector.
Connectors are covered in Article E6.
328
Joints in timber structures
Figure E1‐5 Common connectors: (a) double‐sided split ring connector, (b) single‐sided shear plate connector,
(c) double‐ and single‐sided toothed‐plate connector. (STEP 1995 Article C1)
Punched metal plate fasteners
Punched metal plate fasteners (Figure E1‐6) allow joints to be established between two
pieces of wood in the same plane and are made from galvanised steel sheets or stainless
steel, 0.9 to 3.0 mm thick. Producing joints with punched metal plate fasteners requires
the use of special factory tools. They are mainly used for lightweight timber trusses, with
individual timber members each at least 35 mm thick. Since these trusses tend to have a
very low stiffness perpendicular to the truss axis, careful transport and assembly are
required to avoid damaging the joints. A range of punched metal plate fasteners from
different manufacturers is available and here, the structural engineer should refer to the
corresponding manufacturer details, which are recognised by a certification body. The
relevant standard in this case is EN 14250 “Timber structures ‒ Product requirements for
prefabricated structural members assembled with punched metal plate fasteners”.
Punched metal plate fasteners are covered in Article E7.
Figure E1‐6 Typical punched metal plate fastener. (STEP 1995 Article C1)
329
Joints in timber structures
E1.2 Load‐deformation behaviour of joints
For design purposes, overall timber structural calculations have to be combined with
localised joint analysis and the key consideration here is joint behaviour, which impacts
on both the diagrams of internal forces and moments and the deformation of the struc‐
ture. The behaviour of a joint can be determined by tests carried out in accordance with
EN 26891 “Joints made with mechanical fasteners. General principles for the determina‐
tion of strength and deformation characteristics.” Figure E1‐7 shows experimental load‐
deformation curves of joints with various fasteners loaded perpendicular to their axis, in
which the load F is defined as the load per shear plane.
40 (a)
F(kN)
(b)
30
(f)
(c)
20
(d)
(e)
10
(g)
0 2 4 6 8 10 12
u(mm)
Figure E1‐7 Experimental load‐deformation curves for joints exposed to tensile load parallel to the grain.
(a) glued joint (12.5 ∙ 103 mm2), (b) split ring connector (100 mm), (c) double‐sided toothed‐plate
connector (62 mm) (Hirashima, 1990), (d) dowel (14 mm), (e) bolt (14 mm), (f) punched metal
plate fastener (104 mm2), (g) nail (4,4 mm). (STEP 1995 Article C1)
Undoubtedly, the key feature, as is clearly shown in Figure E1‐7, is that laterally loaded
joints with mechanical fasteners, unlike rigid glued joints, undergo significant defor‐
mations, which structural engineers must take into consideration.
330
Joints in timber structures
However, the increased calculation effort due to the semi‐rigid joints can be deemed
positive overall:
Safety aspect: a semi‐rigid joint can withstand a higher degree of deformations
before failure → failure is signalled.
Load distribution aspect: plastic deformation or creep may result in load
redistribution, which can help relieve highly stressed areas.
The elastic‐plastic, ductile behaviour of laterally loaded timber joints with mechanical
fasteners, particularly dowel‐type fasteners, is due to the interaction between the plastic
deformation of the fasteners and the crushing (so‐called embedment) of the wood un‐
derneath the fasteners, if specific edge and end distances and spacings are maintained.
The required minimum distances guarantee that almost all brittle failure modes like
splitting can be prevented, which paves the way for ductile behaviour.
However, the ductile behaviour of laterally loaded joints with dowel‐type fasteners de‐
pends not only on the specific embedment strength of the timber, the yield moment of
the fastener and the geometric boundary conditions (edge, end distances, spacing), but
also on how slender the fastener used actually is. Figure E1‐8 shows one example of a
dowelled joint, where the slenderness of the timber‐to‐timber joint with one dowel and
two shear planes can be defined as the ratio of the thickness t2 of the middle timber
member and the diameter d of the dowel.
17,5
F (kN)
t2 /d = 12
14,0
t2 /d = 6
10,5
7,0
t2 /d = 2
3,5
0
0 3 6 9 12 15
u (mm)
Figure E1‐8 Influence of slenderness on the load‐deformation behaviour of a timber‐to‐timber joint
loaded in tension parallel to the grain. (STEP 1995 Article C1)
331
Joints in timber structures
The direction of the forces transmitted also impacts on the joint behaviour. If a single
fastener is used, this influence is dictated by the fastener diameter and the grain direc‐
tion of the timber. Tests (Smith and Whale, 1986) have shown that the load‐bearing
capacity of joints with fasteners of diameter up to around 8 mm is independent of the
angle between the force and grain directions. Depending on this angle, however, stresses
perpendicular to the grain may reduce the ductility and load‐bearing capacity of the joint
by causing the timber to split prematurely. To prevent brittle fracture and splitting how‐
ever, it is possible to enhance the ductility of a joint by reinforcing the timber member in
the area of the joint. Effective reinforcements may include punched metal plate fasteners
or wood‐based materials, which can be pressed or glued onto both internal sides of the
timber members, or the insertion of fully threaded screws (see Article E12).
E1.3 Design of joints with dowel‐type fasteners
The key metal fasteners from an engineer’s perspective are those in the dowel‐type
category, such as nails, staples, screws, bolts and dowels, to which the above statements
concerning load‐bearing capacity and deformation behaviour apply. In this section, we
now reconcile these statements in terms of general design principles and equations,
which help clarify the individual articles on fasteners.
Minimum distances
For all dowel‐type fasteners, minimum values have been defined for edge and end dis‐
tances and spacings, most of which are based on experience or comprehensive experi‐
mental investigations. The minimum distances defined in Figure E1‐9 vary according to
the fasteners involved and constitute a prerequisite for the design equations specified in
EC 5. Certain fundamental and significant aspects of fastener distances are as follows:
The spacing parallel to the grain a1 influences the effective number of fasteners
and hence the overall load‐bearing capacity of the joint, since an insufficient
spacing a1 may trigger premature splitting of the joint along the set of fasteners
in the grain direction.
The required spacing perpendicular to the grain a2 does not exceed a1, where
however no splitting will take place perpendicular to the grain.
The required distance to the unloaded edge/end does not exceed that to the
loaded edge/ end, where insufficient distances to the loaded edge/end lead to
premature splitting.
The spacings a1 and a2 (and the end distance a3) influence the block shear
behaviour of timber members in the area of the joint; the smaller these distances,
the greater the likelihood of a block shear failure (Article E13).
332
Joints in timber structures
a2 a2
a2 a2
a1 a1 a1 a1
a4,c
F a
4,t
F
a a
3,t 3,c
Figure E1‐9 Spacing parallel to the grain a1, spacing perpendicular to the grain a2, loaded end distance a3,t,
unloaded end distance a3,c, loaded edge distance a4,t, unloaded edge distance a4,c.
Types of loading and failure modes
Dowel‐type fasteners are subject to two kinds of loading; firstly in shear, when a force
perpendicular to the fastener axis is exerted and secondly withdrawal/pushing in, when
force is exerted in the direction of the fastener axis. Combined stresses are also possible.
Joints with laterally loaded fasteners
The behaviour of joints with laterally loaded fasteners was discussed in Section E1.2 and,
assuming compliance with minimum distances, is mainly dictated by the embedment
strength, the yield moment of the fastener and the joint geometry. For example, possible
failure modes of a timber‐to‐timber joint with two shear planes are shown in Figure E1‐10.
For these modes, the load‐bearing capacity Fv,R can be calculated by force and moment
equilibrium conditions. Johansen (1949) was first to publish this method for timber‐to‐
timber joints, which is why the equations used to determine the load‐bearing capacity bear
his name.
333
Joints in timber structures
t1 t2 t1
F F
F
g h j k
Figure E1‐10 Possible failure modes of a timber‐to‐timber joint with two shear planes.
Failure modes j and k are characterised by an inclination of the fastener due to the emer‐
gence of one or more plastic hinges, from where the rope effect emerges. This generates
(tensile) forces along the shank of the fastener, the intensity of which depend on how
deeply the fastener is anchored in the timber and which press the joint members against
each other. A smooth shank nail will generate less rope effect than a self‐tapping screw.
The tensile forces Fax generated by the rope effect in the shank direction can be taken
into account during design. The contribution of the rope effect (see expression Fax/4 in
equations (E1‐3) and (E1‐4)) to the load‐bearing capacity is limited in EC 5, depending on
the fastener type used, to the following percentages of the Johansen part:
Round, smooth shank nails: 15%
Square, smooth shank nails: 25%
Ringed shank or spiral nails: 50%
Screws: 100%
Bolts: 25%
Dowels: 0%
This means that for screws, ¼ of the withdrawal capacity Fax can be taken into account,
but a maximum of 100% of Fv,R. Accordingly, if the Johansen load‐bearing capacity of a
screwed joint is Fv,R = 2 kN and the withdrawal capacity Fax of the screw is 6 kN, then
6 kN/4 = 1.5 kN < Fv,R = 2 kN and the total load‐bearing capacity of the joint is
2 kN + 1.5 kN = 3.5 kN. Conversely, with withdrawal capacity of 10 kN, this results in
Fax/4 = 2.5 kN > Fv,R = 2 kN. In this case, only 2 kN can be considered as rope effect, result‐
ing in a total load‐bearing capacity of 4 kN. However, for round, smooth shank nails, if the
Johansen part amounts to e.g. Fv,R = 1 kN, the rope effect part Fax of a round, smooth
334
Joints in timber structures
Failure mode h: Embedment strength of the middle member is reached:
Failure mode j: Embedment strength in all members is reached and a plastic hinge is
formed:
f h,1,k t 1 d 4 2 M y,Rk F
F v, Rk 1.05 2 1 ax,Rk (E1‐3)
2 f t 2
d 4
h,1,k 1
Failure mode k: Embedment strength in all members is reached and three plastic hinges
are formed:
2 F
F v, Rk 1.15 2 M y,Rk f h,1,k d ax,Rk (E1‐4)
1 4
where
Fv,Rk Characteristic capacity per fastener and shear plane
ti Member thickness, i = 1 or 2
fh,i,k Characteristic embedment strength in timber member i
d Fastener diameter
My,Rk Characteristic yield moment of fastener
Fax,Rk Characteristic withdrawal capacity of fastener
Ratio of embedment strength values, = fh,2,k/fh,1,k
335
Joints in timber structures
The pre‐factors 1.05 or 1.15 in equations (E1‐3) and (E1‐4) counterbalance the different
partial safety factors M for steel and wood and the modification factor kmod, which is
applied to determine the design load‐bearing capacity of the joint, although the moisture
content and load duration only influence the embedment strength of the timber, not the
yield moment of the fastener. Since Fv,Rd is determined with kmod and a partial safety
factor of M = 1.3 (in Germany), this adversely affects the steel components (M = 1.00) in
joints, offset by an increase of 5 or 15% respectively. Figure E1‐11 shows the load‐bearing
capacities in accordance with Johansen (namely without the rope effect) for a double‐
shear timber‐to‐timber joint, where the respective failure mechanisms in accordance
with Figure E1‐10 are specified depending on the timber thicknesses. A clear interaction
between timber thicknesses and failure mechanisms is seen, whereby the thicker the
timber members, the more plastic hinges emerge. A ductile failure mode, involving up to
two plastic hinges per fastener and shear plane, can only be guaranteed with sufficient
timber thicknesses.
10000
k
8000
j
6000
Fv,Rk
4000
2000
g, h
0
0 20 40 60 80 100
t1
Figure E1‐11 Load‐bearing capacities according to Johansen in [N] depending on the timber thickness in [mm],
equations (E1‐1) to (E1‐4), with dowel d = 12 mm and Fax,Rk = 0, t2 = 2 ∙ t1, fu = 600 MPa,
k = 380 kg/m3. Failure modes g to ka see Figure E1‐10.
Joints with axially loaded fasteners
Dowel‐type fasteners may also be exposed to systematic load in the direction of their
axis; namely tensile load when pulled out and compressive load when pushed in, the
latter of which can be exploited using self‐tapping screws e.g. to increase compression
perpendicular to the grain resistance (Article E5). In this respect, the load‐bearing capacity
in the shank direction depends on the surface geometry of the fasteners and the with‐
drawal resistance of a self‐tapping screw far exceeds that of a ringed shank nail, while
dowels have no appreciable withdrawal resistance. The shank direction in comparison to
the grain direction of the timber is also a key variable, since the withdrawal resistance of a
fastener arranged in parallel to the grain is far lower than when perpendicular to the same.
336
Joints in timber structures
As a general rule, axial load may generate the following failure mechanisms, which should
be verified:
Withdrawal failure (fastener pulled out of the timber),
Head pull‐through failure (fastener head is pulled through the timber),
Tensile failure of the fastener (rupture).
There is also a need to examine whether groups of fasteners exposed to withdrawal may
be at risk of block shear failure (Article E13). When loaded in compression parallal to the
shank axis, i.e. when the fastener is pushed in, meanwhile, the fasteners may fail due to
buckling.
Influence of loading direction
The influence of the angle between the force and grain directions is very high, due to
anisotropy of the wood. The greater the angle between the force and grain directions,
the greater the tensile stress perpendicular to the grain and, in turn, the higher the risk of
brittle failure of the timber member. This must be taken into consideration by the de‐
signer and is outlined in more detail in Article E11. Brittle failure triggered by a tensile
force perpendicular to the grain is unrelated to other brittle failure mechanisms, which
may, for example, occur in groups of fasteners or joints with low timber thicknesses
loaded parallel to the grain, although tensile stresses perpendicular to the grain and
shear stresses peaking are responsible for the brittle behaviour in the latter case.
Embedment strength declines alongside the rise in tensile stress perpendicular to the
grain with increasing angle between force and grain directions. However, since the em‐
bedment strength is a system rather than material parameter and the fastener diameter
impacts on the embedment strength, this effect can be practically ruled out for fasteners
with smaller diameters (d < 6 mm). Accordingly, for dowelled and bolted joints, plus
screwed joints with diameters d > 6 mm, the embedment strength is determined as a
factor of the angle between force and grain directions, while for nails, staples and
screws where d ≤ 6 mm, the embedment strength is determined independently of . In
all cases, however, a distinction is made between the embedment strengths with or
without predrilling.
Moreover, depending on the loading situation of the fasteners, lateral and axial loads
may be exerted simultaneously, whereupon both load components are taken into con‐
sideration via a linear or quadratic interaction.
337
Joints in timber structures
Simplified design
The NA sets out simplified design rules, which can be used as an alternative to Johansen
equations, subject to compliance with specific required timber thicknesses. The reasons
behind these regulations are as follows:
The targeted failure mechanism results in two plastic hinges per shear plane.
Plastic behaviour is contingent on greater penetration depths of the fasteners.
Slender fasteners result in more ductile joints than sturdy fasteners.
Premature splitting determines the load‐bearing capacity, particularly in joints
with many fasteners.
Simplified equations elicit more positive results for slender fasteners.
If we take the example of a double‐shear timber‐to‐timber joint (Figure E1‐10 and equa‐
tions (E1‐1) to (E1‐4)), this helps clarify a simplified design according to NA. NA states that
the load‐bearing capacity per fastener and shear plane can be determined as follows:
2
F v, Rk 2 M y,Rk f h,1,k d (E1‐5)
1
Contingent on compliance with the following required minimum timber thicknesses:
M y,Rk
t 1,req 1.15 2 2 Side member (E1‐6)
1 f h,1,k d
4 M y,Rk
t 2,req 1.15 Middle member (E1‐7)
1 f h,2,k d
It is very easy to derive equations (E1‐5) to (E1‐7) based on Figure E1‐11. The required
timber thickness corresponds to the specific thickness, at which failure mode j moves to
failure mode k. This means that equations (E1‐3) and (E1‐4) are equated and solved in
accordance with t. A simplified design means that only failure mode k is taken into consid‐
eration in accordance with Figure E1‐10 with two plastic hinges per shear plane or, in other
words, failure mode k is forced by the required timber thicknesses. Provided the required
timber thicknesses in accordance with equations (E1‐6) and (E1‐7) are complied with, this
eliminates any need to take failure modes g, h and j into consideration, allowing the load‐
bearing capacity to be directly determined from the equation (E1‐5). However, the rope
effect part Fax,Rk can also be taken into account. For nailed joints, these NA regulations have
338
Joints in timber structures
been even further simplified. The embedment strength is specified independently of the
angle between the force and grain directions and the minimum timber thicknesses are
only specified as a function of the nail diameter. Simplified design is no longer addressed
in Article E2 and the individual articles on the various types of fasteners, but even so,
according to NA, it can be applied for all dowel‐type fasteners.
Joints with multiple shear planes
Joints with more than two shear planes can be calculated using Johansen equations,
whereby each shear plane is deemed to constitute a part of a series of double‐shear
timber‐to‐timber or steel‐to‐timber joints. Here, it should be noted that not all failure
modes are prone to occur within a joint with multiple shear planes or rather that the
failure modes in the respective shear planes must be reciprocally compatible. According‐
ly, in a timber‐to‐timber joint, a mode j failure in accordance with Figure E1‐10 is only
possible in both external shear planes. Article E15 takes a closer look at joints with multi‐
ple shear planes.
Stiffness
The structural engineer must also ensure consideration of the semi‐rigidity of timber joints
with mechanical fasteners as required and this semi‐rigidity is determined via the slip mod‐
ulus Kser. The slip modulus of a joint in the ultimate limit state Ku is determined as follows:
2
K u K ser (E1‐8)
3
Ku is required, for example, to determine the effective bending stiffnesses of mechanical‐
ly jointed beams (e.g. nailed I‐beams). The values for Kser depend on the fastener type
and diameter and on the density. The equations specified in Table E1‐1 for Kser apply per
shear plane and fastener of timber‐to‐timber joints and are based on comprehensive test
results. For steel‐to‐timber joints meanwhile, the values have to be multiplied by the
factor 2.0. If the mean densities m of the connected timber members differ, Kser should
be calculated with the geometric mean of both densities.
339
Joints in timber structures
Table E1‐1 Values for Kser in N/mm for some fastener types. See also EN 1995‐1‐1:2010 Table 7.1.
Fastener type Kser (d in mm, m in kg/m3)
Dowels, screws, nails (predrilled), d m1.5
Bolts with or without clearance* K ser
23
Nails (non‐predrilled) d 0.8 m1.5
K ser
30
Staples d 0.8 m1.5
K ser
80
* The clearance should be added separately to the deformation.
E1.4 Additional general rules for dowel‐type fasteners
For joints with dowel‐type fasteners, certain additional and generally valid regulations
and statements apply, covering various aspects of a joint. These aspects, the most im‐
portant of which will be summarised in the following section, concern both influences on
the load‐bearing capacity through the shape of the joint, which can generate additional
stresses in the members, as well as execution rules, although the allocation may not
always be clear in the latter case.
Execution rules
Predrilling
Dowel‐type fasteners can be driven into members made of timber or wood‐based materials
with or without predrilling of holes. While bolted and dowelled joints always require
predrilling of timber members, nails and screws may be usable without such predrilling,
depending on the wood species and fastener diameter. It follows that joints with hardwood
and densities exceeding 500 kg/m3 must always be predrilled4. Such predrilling is also advis‐
able for nails of diameter exceeding 6 mm. In principle, staples are driven in without
predrilling. For nails, the drilling diameter is around 0.8 ∙ d and for self‐tapping screws, the
core diameter (= inner thread diameter) is the basis for predrilling. Steel plates in steel‐to‐
timber joints can be predrilled for bolted and dowelled joints with up to d + 1 mm. Mean‐
while, self‐tapping dowels are also available and can be driven in without predrilling of steel
4
Recent research projects deal with a new generation of self‐tapping screws for hardwoods that can be
inserted without predrilling.
340
Joints in timber structures
plates. For joints with bolts and threaded rods, there is also scope to drill holes in the tim‐
ber members with d + 1 mm, although this clearance has to be taken into consideration
when calculating the deformations. The timber members of joints with dowels and fitted
bolts, conversely, are predrilled with the nominal diameter of the fasteners.
Minimum penetration depth
Minimum penetration depths help ensure ductile failure behaviour and must also be
taken into account when using nails and screws in members with the tip of the nail or
screw. This means that e.g. for smooth shank nails in timber‐to‐timber joints, a minimum
penetration depth of 8 ∙ d applies, while for non‐smooth nails, the figure is 6 ∙ d. Since
these minimum penetration depths represent robustness requirements, sufficient load‐
bearing capacity can still be guaranteed, even in the event of unwanted gaps present
between the members e.g. due to climate‐related deformations.
Minimum number of fasteners
For all dowel‐type fasteners, robustness requirements dictate that at least two fasteners
be used for load‐bearing joints (with a few exceptions, such as fixing casings or battens).
For dowelled joints meanwhile, the joint must include at least four shear planes as well as
two dowels. Another advisable step, particularly for a dowelled joint, involves replacing a
certain number of dowels with fitted bolts, to hold the joint together.
Fire resistance and corrosion
Another challenge imposed on the structural engineer is compliance with fire resistance
requirements, one of the solutions to which involves “concealing” fasteners in the tim‐
ber. This is also what underpins visually appealing solutions. Further information on the
fire resistance of joints is included in Article G1.
When designing joints subject to harsh environmental conditions, the risk of corrosion
must also be borne in mind. Initially, the design focus should be on ensuring that no
water can ingress on the connection face. For unprotected constructions, applying a
cover may help protect against sun and water effectively. Under sub‐optimal conditions,
corrosion can be prevented by coatings or using stainless metals, while the planner
should also consider the compatibility of the metal with wood preservatives. For this
reason, e.g. contact between aluminium or steel and wood that has been treated with
wood preservatives containing copper can be problematic.
341
Joints in timber structures
Influences on load‐bearing capacity
Groups of fasteners
Timber joints with mechanical fasteners tend to incorporate more than just one fastener,
which means an uneven load distribution between them, even when the load withstood
by the joint is exerted at the centre of gravity of the joint. The load‐bearing capacity of a
joint corresponds to the total collective forces accommodated by the individual fasteners
in the failure state. If the loads on the individual fasteners now differ significantly when
the joint fails and the load exerted on some of the fasteners is far below their capacity, as
is possible with the brittle behaviour of a joint, the load‐bearing capacity of the joint is
deemed lower than the sum of capacities of the individual fasteners. This explains the
reduction in load‐bearing capacity per fastener in joints with multiple fasteners for specif‐
ic fastener types. In principle, the various influences on the load distribution in joints
impact on both joints with multiple identical fasteners as well as those with multiple
different fasteners. Accordingly, the effective characteristic capacity of a joint with a row
of fasteners Fv,ef,Rk, featuring fasteners consecutively arranged in force and grain direc‐
tion, is determined with an effective number of fasteners nef:
The effective number nef was determined using regression analyses on numerous exper‐
imental test results for each fastener type and is explained in more detail in the individual
articles and in Article E13.
Another failure mode affecting groups of fasteners is block shear. If many fasteners are
closely aligned or arranged linearly, the load‐bearing capacity of the joint may be limited
due to part of the member rupturing (Figure E1‐12). The block shear failure of a fastener
group is attributable to reaching the shear strength at member planes parallel to the
force and grain directions and the tensile strength at a member plane perpendicular to
the same. The failure is initiated in the form of a rupture in one of the resistance areas S,
followed by failure in the area perpendicular to the first rupture area. Given that timber
is a brittle material, the two strengths cannot be collectively pooled, which means the
member strength has to be determined separately for the net cross‐section St for tensile
stress and Sc for shear respectively. The design block shear strength is the greater of the
two values (see Article E13).
342
Joints in timber structures
(a) (b) F
F
Sc
St St
Sc
Figure E1‐12 Blockshear failure in a joint: (a) Tension failure in the net area St,
(b) shear failure in the net area Sc. (STEP 1995 Article C1)
Swelling and shrinkage
Particular attention should be paid to cross‐sectional, dimensional changes of the timber
caused by varying moisture levels. Areas in which fasteners prevent swelling and shrink‐
age may be subject to tensile stresses perpendicular to the grain, which lead to members
rupturing (Figure E1‐13 (a)). To prevent or limit such rupture, the area in which swelling
and shrinkage and thus strains are restrained must be limited, with fasteners clustered
on one side of the connected members (Figure E1‐13 (b)). Fasteners solely intended to
hold the joint together should be driven into elongated holes. In other cases, e.g. mo‐
ment‐resisting frame corners, the area where strains are restrained should be limited to
around a metre and/or reinforcements should be applied.
(a) (b)
Figure E1‐13 Joint details: (a) Splitting due to shrinkage, (b) correctly executed joint with oval holes.
(STEP 1995 Article C1)
343
Joints in timber structures
Eccentricities
Joints and members within structures should be arranged as symmetrically and concen‐
trically as possible, particularly if subject to considerable stresses. Even so, eccentricities
may occur for the following reasons (Figure E1‐14):
Type of fastener used,
Layout of the joint,
Layout of the structural system.
In eccentrically arranged fasteners, such as connectors, the impact of bending moment
on unilaterally stressed members must be limited through fasteners under tensile load,
such as bolts with washers. However, eccentricities can often be eliminated by arranging
fasteners and members correctly, as shown in Figure E1‐14 (b) and Figure E1‐14 (d). In
other cases, secondary forces arising from eccentricities (moment, shear, tension) which
act on the fasteners and members must also be taken into account during the design.
(a) (b)
(c) (d)
Figure E1‐14 Eccentricities in structures generated by arrangements of fasteners (a) or members (c).
Modified arrangements, to avoid eccentricities (b), (d). (STEP 1995 Article C1)
Combination of different types of fasteners
At times, particularly in additionally reinforced joints, different types of fasteners may be
used in a joint to transfer forces. In such cases, the load‐bearing capacity should be de‐
termined taking the stiffness of the various fasteners into consideration. To prevent
overloading due to excessive differences in stiffness, however, gluing and mechanical
fasteners must not be combined.
344
Joints in timber structures
F/2
F
F/2
nA
nB
Figure E1‐15 Joint with a combination of fasteners: nA dowels and nB nails. (STEP 1995 Article C1)
Conservatively, a joint is designed assuming elastic behaviour of fasteners and the design
load Fd is allocated via the slip moduli of the fastener. If a force is exerted on two types of
fasteners (see Figure E1‐15), the equilibrium condition and compatibility of the defor‐
mation can be determined as follows:
n FA F
F F i n A F A n B FB and u B (E1‐10)
i=1 K u,A K u,B
where Ku,A and Ku,B are the slip moduli in the ultimate limit state.
Consistency of units
Finally, a statement valid for all design situations, but particularly relevant here, given the
numerous empirical equations involved in joint design. One of the key control mecha‐
nisms during the design process is checking of units, which is imperative. This approach
also allows errors to be detected very quickly when e.g. converting area loads into line‐
and individual loads.
However, the many empirical equations used in timber construction compound the diffi‐
culty of unit checking. Many equations have been derived from regression analyses based
on numerous test results with a large scatter, which require parameters to be deployed
in specific units and where the unit of the result is already defined. One example would
be equation (E1‐11), which is used to determine Kser for a staple:
d 0.8 m
1.5
K ser (E1‐11)
80
The exponents clearly show that equation (E1‐11) is empirical and based on regression
analyses. Diameter d in [mm] and the mean density m in [kg/m3] have to be implement‐
ed, to determine Kser in [N/mm]; no derogation from these units is possible.
345
Joints in timber structures
E1.5 Literature
P. Racher, original Article C1, STEP 1995.
Hirashima Y. (1990). Lateral resistance of timber connector joints parallel to grain direction.
International Engineering Conference, Vol. 1:254‐261, Tokyo.
Johansen K.W. (1949).Theory of timber connections. International Association of Bridge and Structural
Engineering (IABSE), Publication 9, Basel.
Natterer J., Herzog T. and Volz M. (1991). Holzbau Atlas Zwei. Edition française, Presses polytechniques
et universitaires romandes, Le Mont‐sur‐Lausanne.
Smith I. and Whale L.R.J. (1986). Mechanical timber joints. TRADA Research Report 18/86,
Hughenden Valley, England.
346
E2 Johansen model
Original article: B. O. Hilson
Figure E2‐1 shows joints with laterally loaded dowel‐type fasteners. Examples of dowel‐
type fasteners include nails, staples, screws, dowels or bolts.
F F F F F
F 2F 2F
(a) (b)
Figure E2‐1 Laterally loaded joints with dowel‐type fasteners: (a) single‐shear (one shear plane per fastener);
(b) double‐shear (two shear planes per fastener).
Allowable stresses within such joint types were previously determined based on short‐
term tests with comparatively few tested joints. A common method used to determine
allowable values consisted of e.g. specifying the mean value of the load‐bearing capacity
of multiple similar test specimens and then dividing this by an overall safety factor, to
record the influences of variations in loads and resistances, on the degree of craftman‐
ship as well as the load duration. The results obtainable from such tests, however, do not
generally suffice to enable reliable estimates of the characteristic load‐bearing capacities,
as is required for EC 5. Similarly, tests with representatively selected joints to determine
characteristic values are infeasible, due to the great variety of joints and correspondingly
huge number of tests, which would be required.
347
Johansen model
Accordingly, mechanical models are used, so that characteristic load‐bearing capacities
can be calculated based on specific material properties and joint dimensions. The equa‐
tions contained in EC 5 to calculate the capacity of joints are based on the work of Johan‐
sen (1949) and dictate that capacity is limited to the point where the embedment
strength is reached in at least one of the connected timber members and in specific cases
by the simultaneous emergence of plastic hinges in the fastener. The failure mechanism
depends on joint geometry as well as material properties, particularly the plastic bending
moment of the dowel or the embedment strength of the timber or wood‐based material.
Numerous tests were performed, to verify the validity of the Johansen equations. All
these tests, e.g. conducted by Möller (1951), Aune and Patton‐Mallory (1986) or Hilson et
al. (1990) revealed effective conformity between test and theory, provided that frictional
influences and tensile forces in the fasteners excluded from the theoretical scope could
be limited to a negligible level.
E2.1 Material properties
As already shown in Article E1 (equations (E1‐1) to (E1‐4)), to determine the capacity of a
laterally loaded joint, the properties required are “embedment strength fh” and “yield
moment My”. All other required parameters are geometric details such as timber thick‐
nesses t1/2 or fastener diameter d. The property still required for equations (E1‐3) and
(E1‐4), namely “withdrawal capacity Fax,R” is a parameter added at a later stage, which is
independent of the Johansen model and which takes the increased capacity due to the
rope effect into consideration. The rope effect can only occur in fasteners, which end up
inclined when failing, as is explained in more detail in Section E2.6.
The embedment strength fh of the timber or wood‐based material is defined as the peak
stress calculated in an embedment test and a property related to a system rather than a
material property, dictated by how the fastener and timber interact. A typical test ar‐
rangement is shown in Figure E2‐2. To ensure uniform impression (embedment) in the
timber or wood‐based material, it is important to avoid any deformation of the fastener
as far as possible. This is generally done by clamping the ends of the fastener into the
lateral steel plates as well as limiting the specimen thickness t, generally to twice the
fastener diameter.
348
Johansen model
B B
t
F
Figure E2‐2 Typical embedment test arrangement. Clamping of the fastener into lateral steel plates (B),
obtaining uniform embedment of the fastener in the test specimen (A). (STEP 1995 Article C3)
F
Fmax
0 u
Figure E2‐3 Typical load‐embedment relationship. (STEP 1995 Beitrag C3)
A typical load‐embedment relationship is as shown in Figure E2‐3. The embedment
strength is defined as the load F reached before a displacement of 5 mm in the test di‐
vided by the projected area of the fastener in the test specimen:
F
fh (E2‐1)
d t
In the above, t is the specimen thickness and d the fastener diameter.
The embedment u is the relative displacement between the fastener and specimen,
namely between BB and A in Figure E2‐2. Additional indications to define the embed‐
ment strength are included in EN 383 “Timber structures – test methods – determination
of embedment strength and foundation values for dowel‐type fasteners”, while experi‐
mental devices to define the embedment strength are described in Rodd et al. (1987).
Furthermore, the yield moment My of the fastener is a key parameter for the Johansen
equations. The procedure for determining the yield moment is described in EN 409 “Timber
structures – test methods – determination of the yield moment of dowel‐type fasteners”.
349
Johansen model
E2.2 Johansen equations for single‐shear
timber‐to‐timber joints
One prerequisite for deriving the Johansen equations is the rigid‐plastic material behav‐
iour assumed for the fastener when exposed to bending stress and for the timber when
exposed to embedment stress. The assumed load‐embedment relationship of the fasten‐
er in the timber is shown in Figure E2‐4. This approximation simplifies the calculation but
has virtually no impact on the final result. To derive design equations, the first task is to
identify the possible failure modes, so that a design model can then be derived. In tim‐
ber‐to‐timber joints with fasteners in single‐shear, analogously to Figure E1‐10, the fail‐
ure modes shown in Figure E2‐5 can occur in accordance with Johansen.
F
Fmax
0 u
Figure E2‐4 Simplified load‐embedment relationship. (STEP 1995 Article C3)
t1 t2
a b c d e f
Figure E2‐5 Possible failure modes of timber‐to‐timber joints with a laterally loaded fastener in single‐shear.
350
Johansen model
The following section shows examples of how two failure mechanisms can be derived;
cases a and d in accordance with Figure E2‐5. The derivation of further failure mechanims
is shown in Annex 6. The following designations are used:
t1 and t2 are the timber thicknesses or penetration depths of the
fastener in member 1 or 2
fh,1,k is the characteristic embedment strength of member 1
fh,2,k is the characteristic embedment strength of member 2
d is the fastener diameter
My,Rk is the characteristic yield moment of the fastener
Fv,Rk is the characteristic load‐bearing capacity per shear plane
is the ratio of the embedment strengths: = fh,2,k/fh,1,k
Failure mechanism a – embedment failure in member 1:
Fv,Rk
f h,1,k
d
t1 t2
Fv,Rk
Figure E2‐6 Failure mechanism a.
In the event of a mode a failure as shown in Figure E2‐6, the embedment strength in
member 1 is the key parameter. The equilibrium of forces V = 0 in member 1 results in:
Equation (E2‐2) corresponds precisely to the equation for failure mechanism a of a tim‐
ber‐to‐timber joint in EC 5.
351
Johansen model
Failure mechanism d – embedment failure in both members and
one plastic hinge in member 2:
Fv,Rk
a1
b2
f h,1,k
f h,1,k f h,1,k
f h,2,k
My,k
My,k
A
a1a1
f h,1,k
b1 f h,2,k a1a1
t1 t2 b1
Fv,Rk t1
Figure E2‐7 Failure mechanism d. Right: free‐body diagram of fastener.
In the event of a mode d failure as shown in Figure E2‐6, both members reach the em‐
bedment strength and a plastic hinge forms. The following applies:
f h,2,k
f h,1,k
The shear force in the fastener is equal to zero at the point where the bending moment
peaks:
f h,2,k d b 2 f h,1,k d b 1 f h,1,k d b 2 f h,1,k d b 1 0
Consequently:
b 2 b1
The equilibrium of moments at the fastener MA = 0 is:
b2 b a1
M y,k f h,2,k d b 2 f h,1,k d b 1 a 1 b 2 1
2 2
a
f h,1,k d a 1 b 2 b 1 a 1 1
2
352
Johansen model
Now fh,2,k is replaced by ∙ fh,1,k, b2 by b1/ and a1 = (t1‐b1)/2 and the equation is trans‐
formed:
2 2 t 1 4 M y,k
2
b1 t 1 b1 0
2 2 2 f h,1,k d
Solving the quadratic equation reveals:
t1 4 2 M y,k
b1 2 1 (E2‐3)
2 2
f h,1,k d t 1
In addition, the equilibrium of forces V = 0 of the fastener in member 1 equates to:
F v,Rk f h,1,k d b 1
Inserting equation (E2‐3) reveals:
f h,1,k d t 1 4 2 M y,k
F v,Rk 2 1 (E2‐4)
2 f d t 2
h,1,k 1
Equation (E2‐4) (E2‐4) is reflected in the equation for failure mechanism d of a timber‐to‐
timber joint in EC 5. Equation (E2‐4) corresponds to the Johansen part in EC 5, the contri‐
bution of the rope effect as well as the pre‐exponential factor 1.05 are missing (explana‐
tion of factor 1.05 see Article E1).
E2.3 Johansen equations for double‐shear
timber‐to‐timber joints
The same approach employed for single‐shear joints can also be used to derive Johansen
equations for double‐shear joints and possible failure modes are shown in Figure E2‐8
(see also Figure E1‐10). When deriving the equations, it is important to note that the
Johansen equations are determined per shear plane, whereby the joints are conceptually
separated in the middle (namely at t2/2) in accordance with Figure E2‐8. Accordingly, the
Johansen equations for double‐shear joints are merely expansions of the single‐shear
joint equations.
In EC 5, the capacities according to equations (E2‐7) and (E2‐8) are further expanded, to
take account of (i) the rope effect Fax,Rk/4 (see Section E2.6) and (ii) the pre‐factors 1.05
or 1.15 are applied, which take into account the varying partial safety factors M and kmod
for steel (My,k) and timber (fh,k). Equations (E2‐5) and (E2‐6) conversely do not reveal any
353
Johansen model
changes in comparison to EC 5, since (i) both failure modes described do not lead to any
inclination of the fastener and hence no rope effect takes place and (ii), in terms of mate‐
rial parameters, only reveal embedment failure (with M and kmod for timber) and no
plastic hinge (with M for steel).
t1 t2 t1
F F
F
g h j k
Figure E2‐8 Possible failure modes of timber‐to‐timber joints with a laterally loaded fastener in double‐shear.
The resulting equations are as follows:
Failure mechanism g (see Figure E2‐6 and equation (E2‐2)):
Failure mechanism h (see Annex 6, failure mechanism b, equation (20), here with half
width t2/2):
t2
F v,Rk f h,2,k d (E2‐6)
2
Failure mechanism j (see Figure E2‐7 and equation (E2‐4) or Annex 6, failure mechanism e):
f h,1,k d t 1 4 2 M y,k
F v,Rk 2 1 (E2‐7)
2 2
f h,1,k d t 1
354
Johansen model
Failure mechanism k (see Annex 6, failure mechanism f, equation (25)):
2
F v,Rk 2 M f h ,1,k d (E2‐8)
1
E2.4 Johansen equations for steel‐to‐timber joints
with thick steel plates
In steel‐to‐timber joints with sufficiently thick steel plates, the steel plate acts as a
clamped restraint for the fastener. The embedment strength of the steel plate consider‐
ably exceeds the equivalent figure for timber or wood‐based materials and is assumed to
be indefinite when deriving Johansen equations. This explains why plastic hinges occur in
the shear plane between steel plate and timber in the case of failure modes developing
plastic hinges. For a steel plate thickness at least equivalent to the fastener diameter, the
formation of plastic hinges in the shear plane between steel plate and timber can be
assumed. The derivations for steel‐to‐timber joints resemble those for timber‐to‐timber
joints, while possible failure mechanisms in accordance with Johansen for steel‐to‐timber
joints are shown in Figure E2‐9.
t1 t2
a b c d e f g h j/l k m
Figure E2‐9 Possible failure modes of steel‐to‐timber joints.
The derivation is shown as an example for failure modes d and e of a single‐shear steel‐
to‐timber joint.
Failure mechanism d
In the event of failure in accordance with mode d, as in Figure E2‐9 and Figure E2‐10,
embedment strength is reached in the timber member and a plastic hinge forms in the
shear plane.
355
Johansen model
The equilibrium of moments at the fastener MA = 0 is:
a b a1
M y,k f h,1,k d a 1 b 1 a 1 1 f h,1,k d b 1 a 1 1 0
2 2
Inserting a1 = (t1‐b1)/2 and transforming reveals:
4 M y,k
b 12 2 t 1 b 1 t 12 0
f h,1,k d
Consequently:
4 M y,k
b1 t 1 2 1
f h,1,k d t 12
In addition, the equilibrium of forces V = 0 of the fastener in the timber member
equates to:
F v,Rk f h,1,k d b 1
Inserting of b1 results in:
4 M y,k
F v,Rk f h,1,k d t 1 2 1 (E2‐9)
f d t 2
h,1,k 1
Equation (E2‐9) is included in the equation for the failure mechanism d of a steel‐to‐
timber joint in EC 5 and expanded there with the rope effect, but incorrectly without the
pre‐factor, which takes into consideration the differing partial safety factors and kmod for
steel and timber.
Fv,Rk
a1 b1
f h,1,k f h,1,k
f h,1,k My,k
A
My,k
a1 a1 b1
f h,1,k
t1 a1
Fv,Rk
Figure E2‐10 Failure mechanism d. Right: free‐body diagram of fastener.
356
Johansen model
Failure mechanism e
In the event of failure in accordance with mode e, as in Figure E2‐9 and Figure E2‐11, the
embedment strength has been reached in the timber member, and two plastic hinges
form. The equilibrium of moments at the fastener MA = 0 is:
b1
2 M y,k f h,1,k d b 1 0
2
Solving for b1:
M y,k
b1 2
f h,1,k d
Fv,Rk
f h,1,k f h,1,k
My,k
My,k A
My,k
My,k
b1 b1
Fv,Rk
Figure E2‐11 Failure mechanism e. Right: free‐body diagram of fastener.
In addition, the equilibrium of forces V = 0 of the fastener in the timber member
equates to:
F v,Rk f h,1,k d b 1
Inserting b1:
M y,k
F v,Rk f h,1,k d 2 2 M y,k f h,1,k d (E2‐10)
f h,1,k d
Equation (E2‐10) is reflected in the equation for the failure mechanism e of a steel‐to‐
timber joint in EC 5 and expanded there with the rope effect and the pre‐factor 1.15
(2 ∙ 1.15 = 2.3).
357
Johansen model
E2.5 Johansen equations for steel‐to‐timber joints
with thin steel plates
External thin steel plates cannot provide a clamped restraint of the fastener in the steel
plate, hence the emergence of plastic hinges in the shear planes between external steel
plates and the middle timber member. Accordingly, the corresponding equations were
derived assuming a pinned restraint of the fasteners in the steel plates. A steel plate is
deemed thin, if the thickness of the steel plate does not exceed half the fastener diame‐
ter d. For double‐shear joints, if the middle section comprises a thin steel plate, the
same equations apply as are used in the case of thick steel plates. The connection geom‐
etry allows the fastener to form a plastic hinge in the steel plate area, although this pre‐
supposes sufficient load‐bearing capacity of the steel plate. If the thickness of the steel
plate is between 0.5 ∙ d and d, the load‐bearing capacity of the joint is determined from a
linear interpolation between the load‐bearing capacity values for thin and thick steel
plates. All possible failure modes of a steel‐to‐timber joint with thin steel plates are
shown in Figure E2‐9. For failure mode a, as in Figure E2‐9 and Figure E2‐12, embedment
failure of the timber without a plastic hinge forming in the fastener, the load‐bearing
capacity in accordance with Johansen is derived here as an example.
Fv,Rk
a1 a1 b1
t1
Fv,Rk
Figure E2‐12 Failure mechanism a of a single‐shear steel‐to‐timber joint with thin steel plate.
In the shear plane, the moment = 0:
a b1 a 1
f h,1,k d a 1 b 1 a 1 1 f h,1,k d b 1 a 1 2 0
2
Inserting a1 = (t1‐b1)/2 and transforming:
2 2
b1 2 t 1 b1 t 1 0
358
Johansen model
Consequently:
b1 t 1
2 1 0,4 t 1
In addition, the equilibrium of forces V = 0 of the fastener in the timber member
equates to:
F v,Rk f h,1,k d b 1
Inserting b1:
Equation (E2‐11) corresponds to the equation for the failure mode a of a steel‐to‐timber
joint in EC 5.
E2.6 Rope effect
The equations in accordance with Johansen apply for joints with laterally loaded dowel‐
type fasteners, in which failure occurs due to embedment failure in the timber or wood‐
based material or due to a combination of embedment failure and plastic hinges in the
fasteners. When deriving the equations, it is assumed that no normal forces in the direc‐
tion of the axis occur in the fasteners. For fasteners capable of transmitting normal forc‐
es, however, this is incorrect. In all failure mechanisms, which lead to the fasteners inclin‐
ing, normal forces are exerted and, at times, can lead to considerable load increases in
comparison to pure Johansen cases. Unlike fasteners with threads or nuts and washers
(bolts), smooth shank fasteners such as dowels or smooth shank nails can only withstand
low normal forces. Figure E2‐13 shows an example failure of a nailed joint, in which the
tip of the nail being pulled out is obvious. Figure E2‐14 shows the same failure mecha‐
nism in a fully threaded screw, in which no pulling out of the screw tip is visible.
Figure E2‐13 The nail is pulled out of the timber.
359
Johansen model
Friction force
Tensile force in the direction of the screw axis
Force component parallel
to the shear plane Compression component perpendicular
to the shear plane
Figure E2‐14 The screw tip is not pulled out of the timber.
The anchoring generates a tensile force in the fasteners in the direction of its axis. In
Figure E2‐14 meanwhile, the triangle of forces emerging at the current deformation is
marked. The emerging tensile force can be divided into two components parallel and
perpendicular to the shear plane respectively, while the compressive force component
on the shear plane generates frictional force between both timber members, which has a
positive effect on the load‐bearing capacity of the joint. As explained in the following
section, the force component parallel to the shear plane is disregarded.
Even without any plastic deformation of the fasteners, contact between the timber sur‐
faces generates compressive forces, which lead to friction and thus a rope effect. Bejtka
and Blass (2002) determined the load‐bearing capacity taking into consideration the rope
effect and originating from a screw inserted at an angle . In the derivation of the rope
effect shown here, an undeformed state is actually expected. In the deformed state,
namely where the fastener is inclined due to significant joint deformations, the actual
load‐bearing capacities are higher, since the friction between the shank of the fastener
and the surrounding timber may generate additional normal forces along the fastener
axis, through which, as shown in Figure E2‐14, a force component parallel to the shear
plane can emerge. However, the additional force component parallel to the shear plane
can only develop with the fastener considerably inclined, meaning significant joint defor‐
mation. Since such significant joint deformation cannot be practically guaranteed, these
additional force components parallel to the shear plane are not taken into consideration.
An additional important note reveals that the frictional forces can only form when com‐
pressive stress is exerted perpendicular to the joint (see force components H1 and H2 in
Figure E2‐15). With this in mind, the rope effect does not apply to laterally loaded joints,
which are also subject to tensile forces perpendicular to the shear plane. One example of
a joint subject to tensile and shear loads is the connection of a tension diagonal to the
chord of a truss.
360
Johansen model
F2,k
s1 s2
P H1
H1 H2
P H2 My,k D
Q2
N1
My,k N2
Q1
x2
D x1
F1,k f 1,k,2
f1,k,1 My,k
f h,2,k
f h,1,k
My,k
Figure E2‐15 Equilibrium of forces for a timber‐to‐timber joint with fully threaded screws inserted at
an angle and for the failure mechanism with two plastic hinges per shear plane.
f1,k,1 and f1,k,2 are the characteristic values of the withdrawal parameter in members 1 and 2.
The moment in the shear plane is:
f h,1,k d x 12 f h,2,k d x 22
M 1 M y,k and M 2 M y,k
2 cos 2 2 cos 2
With M1 = M2 and fh,2,k = ∙ fh,1,k:
4 M y,k cos 2
x 12 x 22 (E2‐12)
d f h,1,k
Equilibrium of forces V = H = 0 in the shear plane:
361
Johansen model
where:
f h,1,k d x 1
Q1 (E2‐17)
cos
f h,2,k d x 2
Q2 (E2‐18)
cos
f 1,k,1 d s 1
N1 (E2‐19)
cos
f 1,k,2 d s 2
N2 (E2‐20)
cos
Inserting equations (E2‐15), (E2‐17) and (E2‐19) in equation (E2‐13), the following is
obtained:
Likewise, inserting equations (E2‐16), (E2‐18) and (E2‐20) in equation (E2‐14):
The equilibria of forces F1,k = F2,k and fh,2,k = b ∙ fh,1,k elicit:
Equations (E2‐12) and (E2‐23) elicit:
x1
f 1,k,2 s 2 f 1,k,1 s 1 tan
f h,1,k 1 tan 1
(E2‐24)
2 M y,k cos 2 tan f 1,k,1 s 1 f 1,k,2 s 2
2 2
2
1 f h,1,k d 2
2 f h,1,k 1 1 tan
2
362
Johansen model
f 1,k,1 s 1 f 1,k,2 s 2
x1 x2 (E2‐25)
f h,1,k tan
Equating equations (E2‐23) and (E2‐25):
f 1,k,1 s 1 f 1,k,2 s 2 f 1,k,1 s 1 f 1,k,2 s 2 tan (E2‐26)
f h,1,k tan f h,1,k 1 tan
Equation (E2‐26) only has a solution, if f1,k,1 ∙ s1 =f1,k,2 ∙ s2, since otherwise, the solution
would be ‐1 = tan2. This means that the screw tensile force in members 1 and 2 must be
identical. Accordingly, the lower withdrawal capacity is the decisive parameter. For
f1,k,1 ∙ s1 =f1,k,2 ∙ s2, equation (E2‐24) can be adapted:
2 2 M y,k cos 2
x1 (E2‐27)
1 f h,1,k d
Now, x1 from the simplified equation (E2‐27) is inserted in equation (E2‐21):
2
F1,k f 1,k,1 d s 1 tan 1 tan 2 M y,k f h,1,k d cos (E2‐28)
2
1
The load‐bearing capacity Fv,Rk for timber‐to‐timber joints with fully threaded screws
inserted at an angle (or dowel‐type fasteners capable of withstanding axial tensile
forces and inserted at an angle ) for the failure mode with two plastic hinges per shear
plane amounts to:
2
F v,Rk F ax,Rk tan 1 tan 2 M y,k f h,1,k d cos 2 (E2‐29)
1
1
cos f 1,k,1 d s 1
F ax,Rk min (E2‐30)
1 f d s2
cos 1,k,2
363
Johansen model
Fasteners are often inserted perpendicular to the grain direction with = 0° whereupon
equation (E2‐29) is simplified as follows:
2
F v,Rk F ax,Rk 2 M y,k f h,1,k d (E2‐31)
1
Equation (E2‐31) now comprises the Johansen part of the load‐bearing capacity, see also
equation (E2‐8), and the part of the rope effect Fax,Rk. The load‐bearing capacity in
accordance with Johansen accordingly increases with the rope effect by ∙ Fax,Rk. In EC 5,
the friction coefficient is assumed to be = 0.25 = ¼.
E2.7 Literature
B.O. Hilson, original Article C3, STEP 1995.
Aune P. and Patton‐Mallory M. (1986). Lateral load‐bearing capacity of nailed joints based on the yield theory ‒
theoretical development and experimental verification. Forest Products Laboratory,
Research Papers FPL 469 & 470.
Bejtka I. and Blass H.J. (2002). Joints with inclined screws. Paper 35‐7‐4, CIB‐W18 Meeting 35, Kyoto.
Hilson B.O., Whale L.R.J. and Smith I. (1990). Characteristic properties of nailed and bolted joints under
short‐term lateral load. Part 5 ‒ Appraisal of current design data in BS5268:Part 2:1984 Structural Use
of Timber. Journal of the Institute of Wood Science 11(6):208‐212.
Johansen K.W. (1949). Theory of timber connections. International Association of Bridge and Structural
Engineering. Publication No. 9:249‐262. Bern.
Möller T. (1951). En ny metod för beräkning av spikförband. Report No 117, Chalmers University of Technology.
Rodd P.D., Anderson C., Whale L.R.J. and Smith I. (1987). Characteristic properties of nailed and bolted joints
under short term lateral load. Part 2 ‒ Embedment test apparatus for wood and wood‐based sheet materials.
Journal of the Institute of Wood Science 11(2):60‐64.
364
E3 Joints with nails and staples
Original articles: B. O. Hilson
Nails are the most frequently used fasteners in timber constructions and come in a wide
range of lengths, diameters and surface finishes. Nailed joints are easy to manufacture
and generally exhibit ductile behaviour, which facilitates a balanced load distribution in
nail groups. In addition, the ductile behaviour helps dissipate energy when cyclic loads
such as earthquakes are imposed, which may improve the load‐bearing behaviour of the
construction. The most commonly used type is the smooth shank nail with a circular
cross‐section, which is produced from a wire with a minimum tensile strength of
600 N/mm2 and with standard diameters ranging between 2 and 8 mm. The nails can be
blank or protected against corrosion e.g. by galvanising. The nail head is generally flat and
circular with a diameter of around double the nail shank diameter. However, certain
types of nails have smaller heads, so they can be driven in flush to the timber surface.
Certain countries also manufacture nails with a square cross‐section. Nails can be driven
in manually or with pneumatic nail guns.
The load‐bearing behaviour of a nail can be improved, in terms of withstanding both
forces perpendicular to the nail axis as well as in a shank direction (withdrawal) by modi‐
fying the nail surface. One approach would be to deform the surface of smooth shank
nails by rolling on grooves (rings) or deforming to a helical thread (spirals). Another would
involve helically twisting nails with a square cross‐section. This process not only alters the
nail area but also hardens the steel and thus boosts the yield strength. Hot‐dip galvaniz‐
ing, etching or coating with resin or plastics are further means of improving the load‐
bearing behaviour.
Since staples are very similar fasteners to nails, there is very little difference in their load‐
bearing behaviour. However, there are two main things to bear in mind. One is the fact
that the steel grade used for staples tends to be far higher than that for nails, as reflected
in the yield moment. The other is the need to monitor the angle between the staple
crown and the grain direction of the wood. Nails and staples are shown in Figure E3‐1,
while the backgrounds, design equations and derivations thereof can be taken from
Articles E1 and E2. A nailed joint must comprise at least two nails.
365
Joints with nails and staples
(a) (c)
(b) (d)
Figure E3‐1 Left: Nail shapes. (a) Round smooth shank nail, (b) spiral nail, (c) ringed shank nail,
(d) machine driven nails. (STEP 1995 Article C1) Right: Staples.
E3.1 Predrilling
If nails are driven into wood with a higher density, there is a risk of the wood splitting or
the nail buckling. Accordingly, timber members must be predrilled with around 80% of
the nail diameter, either when the timber has a characteristic density from 500 kg/m3 or
the nail diameter exceeds 6 mm. There are three key advantages to predrilling:
The embedment strength and hence the capacity when the nails are laterally
loaded are increased,
The nail spacings and edge distances can be reduced, which allows more compact
joints to be achieved,
Joints are subject to less displacement movement when loaded.
One further important factor is the thickness of the timber members to be connected,
given that the splitting risk rises with decreasing thickness. This is why both EC 5 and the
NA (NDP to 8.3.1.2) stipulate minimum timber thicknesses depending on density.
E3.2 Nail distances and spacings
Nails must be arranged at sufficient distances from the end grain and edges as well as
each other, to prevent splitting. Figure E1‐9 shows how a range of distances are defined.
The end is deemed loaded, when the force transmitted by the nail has a component
pushing in the direction of the end. Otherwise, the end is deemed unloaded. The in‐
creased splitting risk means greater end distance requirements are imposed on the load‐
ed end than on the unloaded end. Similarly, the edges may be loaded or unloaded. Ac‐
cording to EC 5, the minimum nail distances and spacing are based on many years of
experience and specified in Table E3‐1 for timber‐to‐timber joints. The values for mini‐
mum nail distances differ from one wood species to the other and are mainly a factor of
the splitting risk, shear strength, wood density and nail diameter. Predrilling considerably
reduces the tendency of the wood to split and thus allows denser clustering of such nails.
366
Joints with nails and staples
Table E3‐1 Minimum nail distances and spacings, d = nail diameter in mm, 0° ≤ ≤90° = angle between
force and grain directions. See also EN 1995‐1‐1:2010 Table 8.2.
Distances and spacings Minimum distances and spacings
Non‐predrilled Predrilled
k ≤ 420 kg/m3 420 kg/m3 < k ≤ 500 kg/m3
367
Joints with nails and staples
E3.3 Embedment strength
EC 5 recommends the following characteristic embedment strength values for timber‐to‐
timber nailed joints with nail diameter up to 8 mm, regardless of the angle between the
force and grain directions.
Timber (solid timber, glued laminated timber) and laminated veneer lumber (LVL):
Non‐predrilled:
0.3
f h,k 0.082 k d (E3‐1)
Predrilled:
where k is the characteristic density in kg/m3 and d is the nail diameter in mm.
For joints between timber and wood‐based materials, additional equations are provided
in EC 5 and in the NA to determine the embedment strength of wood‐based panels (ply‐
wood, hardboards, particleboards, OSB, gypsum plasterboards and cement‐bonded par‐
ticleboards, for equations, see also Section E12.1).
The equations for the characteristic values of embedment strength are based on regres‐
sion analyses and were derived from numerous embedment tests, which encompassed
wide‐ranging timber densities, wood species and nail diameters. An illustration of the
origin of equations (E3‐1) and (E3‐2) is shown in Whale et al. (1989).
E3.4 Yield moment
EC 5 includes equations for the characteristic yield moment My,k (in Nmm) for smooth
shank nails, which have been manufactured from wire with a minimum tensile strength
of 600 N/mm2.
368
Joints with nails and staples
For smooth shank nails with a round cross‐section, the following applies:
And for nails with a square cross‐section:
2.6
M y,k 0.45 f u d (E3‐4)
where fu is the wire tensile strength in N/mm2 and d is the nail diameter in mm (side
length for square nails).
Like equations (E3‐1) and (E3‐2), which are used to define embedment strengths, equa‐
tions (E3‐3) and (E3‐4) are also based on the regression analyses of numerous test re‐
sults. Exponent 2.6 is based on the observations of Werner and Siebert (1991), who no‐
ted how the nail yield stress declines with increasing diameter.
Other nail types must be tested in accordance with EN 409 “Timber structures ‒ Test
methods ‒ Determination of the yield moment of dowel‐type fasteners”, to determine
applicable values for My,k. These values for My,k can be obtained from the declarations of
performance (DOPs, see Article A1) for the individual nail types.
E3.5 Deformations in nailed joints
Nailed joints deform under loads, just like joints established with other mechanical fas‐
teners. This is shown in Figure E3‐2 using a typical load‐deformation diagram showing a
test on a laterally loaded nailed joint with two shear planes.
F
Fmax
Fser
0 uinst u
Figure E3‐2 Typical load‐deformation diagram of a nailed joint. Fmax is the maximum load, Fser is the load at
serviceability limit state and uinst is the elastic initial deformation. (STEP 1995 Article C4)
369
Joints with nails and staples
The elastic initial deformation, which occurs when exposed to serviceability load Fser, can
be estimated using the slip modulus Kser in accordance with EC 5. The following applies:
Non‐predrilled:
m
0.8 1.5
d
K ser (E3‐5)
30
Predrilled:
dm
1.5
K ser (E3‐6)
23
In this respect, m is the mean density in kg/m3, d the diameter in mm and Kser in N/mm.
Consequently:
F ser
u inst (E3‐7)
K ser
Due to creep, the final deformation in nailed joints may exceed the level of initial elastic
deformation. The final deformation can be estimated from (see also Article F1):
Equations (E3‐5) and (E3‐6) apply for timber‐to‐timber joints, while for steel‐to‐timber
joints, the determined value should be doubled, leading to deformations which amount
to only half those of timber‐to‐timber joints. When the mean densities among the mem‐
bers to be connected vary, the geometric mean for m is used to determine Kser.
Deformations in joints often represent a significant proportion of the overall defor‐
mations in a nailed timber construction under a serviceability load and should thus be
taken into account during calculations. Within a nailed truss, the portion of deflection
due to semi‐rigidity of the joints often exceeds that due to elastic deformation of the
truss members. In nailed composite beams, meanwhile, simple beam theory cannot be
applied. For stresses and deflections, displacements must also be taken into considera‐
tion, e.g. between the flange and web of a nailed I‐joist, which can be done using the
calculation methods described in Article D7. The semi‐rigidity also impacts on the mo‐
ment‐rotation behaviour of joints with nailed‐on plywood panels in portal frames.
370
Joints with nails and staples
E3.6 Axially loaded nails ‒ withdrawal
The load‐bearing capacity of axially loaded smooth shank nails (Figure E3‐3) is low and
the long‐term behaviour unfavourable. Accordingly, instead of smooth shank nails only
ringed shank nails may be used for permanent or long‐term loading in an axial direction.
The point at which the withdrawal resistance peaks is when nails are driven in perpendic‐
ular or at an angle to the grain. Nails arranged in parallel to the grain direction, mean‐
while, generally offer negligible withdrawal resistance and are unsuitable for transferring
forces in axial direction. In addition, varying the wood moisture content reduces the
withdrawal resistance of smooth shank nails.
l
tpen
th
Figure E3‐3 Axially loaded nail arranged perpendicular to the grain. (STEP 1995 Article C5)
In addition to a tensile failure of the nail, the following possible failure modes for the
nailed joint shown in Figure E3‐3 must be taken into consideration:
Withdrawal of the nail from the timber member containing the nail tip
(withdrawal parameter fax,k),
Pull‐through of the nail head through the plate material
(head pull‐through parameter fhead,k).
The NA regulates the use of smooth shank nails in more detail and states that smooth
shank nails in predrilled holes should not be subject to axial loads. The NA also defines
load‐bearing capacity classes for nails that determine the parameters fax,k (withdrawal
parameter) and fhead,k (head pull‐through parameter) required to determine the with‐
drawal resistance. It can be assumed that smooth shank and ringed shank nails of load‐
bearing capacity class 1 can be exposed to 60% of the withdrawal resistance, if perma‐
nently exposed to axial loads in connections in coupled purlins in roofs with a maximum
30% incline. The experimental determination of the parameters and limit values for classify‐
ing the nail types in load‐bearing capacity classes is defined in E DIN 20000‐6:2012 (draft).
371
Joints with nails and staples
Other factors which impact on the withdrawal resistance of nails include the density of
the wood into which the nail is driven and the nail surface quality. Cement‐coated or hot‐
dip galvanised nails, ringed shank and spiral nails and nails with square cross‐sections
perform more favourably in response to withdrawal loads than uncoated or galvanised
smooth shank nails. Another advantage of ringed shank and spiral nails lies in the fact
that their withdrawal resistance is hardly affected by changes in wood moisture content.
Empirical equations to determine withdrawal resistance were determined for a number
of combinations. For nails, the characteristic withdrawal resistance Fax,Rk should be as‐
sumed as the smallest of the following values:
For nails without smooth shank:
f ax,k d t pen
F ax,Rk min 2 (E3‐9)
f head,k d h
For smooth shank nails:
f ax,k d t pen
F ax,Rk min 2 (E3‐10)
f ax,k d t f head,k d h
Here, d = diameter, dh = head diameter of the nail, t = thickness of the timber member on
the nail head side and tpen = penetration depth on the nail tip side. The characteristic
values of the withdrawal parameter fax,k and the head pull‐through parameter fhead,k are
functions of the characteristic density of the wood used and are specified in the NA or
the declarations of performance of the nail manufacturer.
The EC 5 also includes minimum values for the penetration depth tpen. Also noteworthy is
the fact that when considering the rope effect for nails loaded laterally, not the entire
withdrawal resistance may be taken into account, but only specific portions of the load‐
bearing capacity calculated in accordance with Johansen theory: a maximum of 15% for
smooth shank, round nails and 25% for smooth shank, square nails. For non‐smooth
shank nails, the limit is at 50%.
372
Joints with nails and staples
E3.7 Effective number of nails
The load‐bearing capacity of a series of n nails arranged in the grain direction must be
reduced by an effective number nef, if the nails are not arranged offset with respect to
each other by at least 1 ∙ d perpendicular to the grain direction (see also Article E13 on
nef). This kind of offset arrangement is often practically applied, which is why nef is usually
irrelevant for nailed joints. In addition, the need for nef for nailed and above all stapled
joints is disputed; in DIN 1052 nailed and stapled joints need not be subject to any reduc‐
tion in the actual number n.
In EC 5, the following equation applies with kef from Table E3‐2:
k ef
n ef n (E3‐11)
Table E3‐2 Values for kef, interim values may be linearly interpolated.
Nail spacing kef
Non‐predrilled Predrilled
a1 ≥ 14 ∙ d 1.0 1.0
a1 = 10 ∙ d 0.85 0.85
a1 = 7 ∙ d 0.7 0.7
a1 = 4 ∙ d ‐ 0.5
E3.8 Combined loads
When a nail is subject to simultaneous lateral load Fv,Ed and withdrawal load Fax,Ed, both
load components overlap. For smooth shank nails, a linear interaction applies. However,
according to NA, for smooth shank nails in connections in coupled purlins, the following
less conservative equation may be applied:
1.5 1.5
F v,Ed F
ax,Ed 1 (E3‐12)
F
F v,Rd ax,Rd
For non‐smooth shank nails, a quadratic interaction of both load components can be used.
373
Joints with nails and staples
E3.9 Staples
Staples are very similar fasteners to nails, which means correspondingly little variation in
their load‐bearing behaviour. However, two things must be taken into consideration
above all. The first is the fact that the steel quality used for staples is usually far higher
than that for nails (minimum tensile strength of steels for staples of 800 N/mm2), which is
reflected in the yield moment. Secondly, the angle between the staple crown and the
grain direction of the wood must be taken into account. The basic rule is that the load‐
bearing capacity of stapled joints can be assumed as equivalent to that of two nails of
equivalent diameter, provided the angle between the staple crown and grain direction is
at least 30°. For an angle below 30°, the load‐bearing capacity of laterally loaded stapled
joints must be reduced by 70%. A stapled joint must comprise at least two staples and
the penetration depth must be at least 14 ∙ d.
The characteristic yield moment is (EC 5/A2:2014):
The applicable minimum distances differ from those for nails, details are specified in
EC 5. There is no effective number of fasteners: nef = n.
E3.10 Literature
B.O. Hilson, original Articles C4, C5, STEP 1995.
Werner H. and Siebert W. (1991). Neue Untersuchungen mit Nägeln für den Holzbau.
Holz als Roh‐ und Werkstoff 49:191‐198.
Whale L.R.J., Smith I. and Hilson B.O. (1989). Characteristic properties of nailed and bolted joints under short
term lateral load. Part 4 ‒ The influence of testing mode and fastener diameter upon embedment test data.
Journal of the Institute of Wood Science 11(5):156‐161.
374
E4 Joints with bolts and dowels
Original articles: J. Ehlbeck, H. Werner
Dowels are slender, cylindrical fasteners made of steel with smooth or even slightly
grooved surfaces. Based on this definition, each dowel has a diameter of between 6 and
30 mm, with diametric tolerance defined as ‐0.1/+0.5 mm. Holes for dowels must be
predrilled with the nominal diameter. The holes in steel plates for steel‐to‐timber joints
may be drilled 1 mm larger than the dowel diameter; resulting additional deformations
have to be taken into consideration. Similarly, bolts are dowel‐type fasteners made of
steel, which include a head and nut. They are inserted into predrilled holes and then
tightened such as to ensure that the wood or steel parts are secured in close proximity.
When necessary, the bolts must also be retightened, if the wood has reached its equilib‐
rium moisture content. The predrilled holes may be up to 1 mm larger than the diameter
of the bolt. For joints with fitted bolts, the bolts are driven into a hole, which is of equiva‐
lent dimensions to the diameter of the bolt. Each of the bolt also has washers on both
sides, the side lengths or diameter of which should be at least 3 ∙ d and they should be at
least 0.3 ∙ d thick (d is the bolt diameter). The washers should have a full contact surface
with the wood.
Dowelled joints have excelled when it comes to transferring large forces. Not only is this
joint type economic, it is also easy to set up. However, dowelled joints require at least
one dowel to be replaced with a fitted bolt, to guarantee the joint cohesion or help pre‐
vent a gap between the individual joint members. In comparison to normal bolted joints,
dowelled joints are more rigid and accordingly, bolted joints should be avoided at any
point where good dimensional stability and high construction rigidity is required. Joints
with fitted bolts, conversely, like dowelled joints, do not feature any predrilled holes of
excessive size and can thus be as rigid and dimensionally stable as dowelled joints. Bolted
and dowelled joints can be executed as purely timber‐to‐timber joints, but also in the
form of joints connecting timber and wood‐based materials or steel‐to‐timber joints.
Figure E4‐1 Left: bolt. Right: dowel.
375
Joints with bolts and dowels
A dowel and a bolt are shown in Figure E4‐1. Background details and design equations for
laterally loaded joints and derivations thereof can be taken from Articles E1 and E2. The
NA stipulates that a load‐bearing bolted or dowelled joint should have at least four shear
planes and at least two fasteners. If only one fastener is used, only half the load‐bearing
capacity may be assumed. Another execution rule is that for external steel plates, fitted
bolts must be used instead of dowels.
E4.1 Fastener distances and spacings
Similarly, as for nails, dowels and bolts must be arranged at sufficient distances from the
end grain, edges and each other, to avoid the risk of the wood splitting. Limited distances
between fasteners or excessive proximity to the edges will lead to premature failure of
wood due to brittle failure, e.g. splitting. Figure E1‐9 shows how the different distances
are defined. The minimum spacings of bolts and dowels differ, due primarily to the size of
the washers. However, regulations governing the minimum distances for bolts and dow‐
els are complex, since the end distance a3,c to the unloaded end depends on the size of
the angle between the force and grain directions. The end distance can be reduced if
is smaller. The minimum distances for bolts are specified in Table E4‐1 and those for
dowels in Table E4‐2.
Spacing a1 (parallel to the grain) ‐ (4 + cos) ∙ d
Spacing a2 (perpendicular to the grain) ‐ 4 ∙ d
Distance a3,t (loaded end) ‐ max(7 ∙ d; 80 mm)
Distance a4,c (unloaded edge) ‐ 3 ∙ d
376
Joints with bolts and dowels
Spacing a1 (parallel to the grain) ‐ (3 + 2 ∙ cos) ∙ d
Spacing a2 (perpendicular to the grain) ‐ 3 ∙ d
Distance a3,t (loaded end) ‐ max(7 ∙ d; 80 mm)
Distance a4,c (unloaded edge) ‐ 3 ∙ d
E4.2 Embedment strength
For bolts and dowels up to a diameter of 30 mm, the characteristic embedment strength
fh,0,k when loaded parallel to the grain is (in N/mm2):
where k the characteristic density in kg/m3 and d the fastener diameter in mm.
Equation (E4‐1) corresponds to equation (E3‐2) for predrilled nailed joints. Unlike nailed
joints, however, for bolted and dowelled joints, the influence of the angle between the
force and grain directions on embedment strength has to be taken into consideration,
since for large fastener diameters, the embedment strength declines with increasing
angle between the force and grain directions:
f h,0,k
f h,α,k (E4‐2)
k 90 sin 2 cos 2
where = angle between the force and grain directions.
In equation (E4‐2), the Hankinson equation (3) in Annex 2 re‐emerges, see also Figure D1‐
5 and equation (D1‐1), which defines the design format for compressive stress at an
angle to the grain direction. The influence of the angle between the force and grain
directions on the embedment strength is shown in Figure E4‐2. The reduction in embed‐
ment strength with increasing angle is more prominent in softwoods than hardwoods.
377
Joints with bolts and dowels
fh,α 1,00 d = 6 mm
fh,0 (a)
0,80 d = 30 mm
d = 6 mm
0,60 (b)
d = 30 mm
0,40
0,20
0,00
0 15 30 45 60 75 90
α ( ˚)
Figure E4‐2 Ratio fh,a / fh,0 depending on the angle between the force and grain directions;
(a): Hardwood; (b): Softwood. (STEP 1995 Article C6)
The coefficient kc,90, which resembles to the factor k90 from equation (E4‐2), is already
known from Article D1 (cf. equation (D1‐4)). There, the verifications for compressive
stress at an angle to the grain direction were introduced as well as the factor kc,90, to take
the type of loading, splitting risk and degree of compressive deformation into considera‐
tion. Compressive forces also appear under the fasteners in joints subject to lateral loads.
Factor k90 is defined in EC 5 as follows:
E4.3 Yield moment of the fasteners
The following relationship is applied in EC 5 to determine the characteristic value of the
yield moment of round steel bolts and steel dowels:
where fu is the tensile strength of the fastener.
Equation (E4‐4) corresponds to equation (E3‐3) for smooth shank nails, although the
background of both equations is different (Blass et al., 2000). Similarly, for bolts and
dowels, the yield moment is generally determined in accordance with EN 409.
378
Joints with bolts and dowels
E4.4 Verification of serviceability
The load‐deformation behaviour of joints with dowel‐type fasteners can be shown using
load‐slip curves. Figure E4‐3 specifies idealised load‐slip curves for bolted and dowelled
joints with around the same load‐bearing capacities, where Fmax,est indicates the estimated
maximum load.
1,00
Fmax, est
F
(a)
0,75
(b)
0,50
0,40
0,25
0,00
0 1 2 3 4 5 6 7 8 9 10
uinst uinst u (mm)
(a) (b)
Figure E4‐3 Idealised load‐slip curves of a (a) dowelled and a (b) bolted joint. (STEP 1995 Article C7)
379
Joints with bolts and dowels
The slip modulus Kser derived from the curve is a characteristic parameter of the overall
joint. Based on numerous test results, EC 5 gives a slip modulus Kser for bolted and dow‐
elled joints per shear plane and fastener under a serviceability load as follows:
d m1.5
K ser (E4‐5)
23
where Kser in Nmm, m in kg/m³ and d in mm.
For bolted joints with clearance, the clearance must be added to the displacements (see
Figure E4‐3 curve (b)). If the two timber members to be connected have differing mean
densities, the geometric mean for m should be applied:
Equation (E4‐5) applies for timber‐to‐timber joints. For steel‐to‐timber joints, the value
for Kser determined using equation (E4‐5) must be multiplied by factor 2.0.
Accordingly, the following applies, similar to nailed joints:
F ser
u inst (E4‐7)
K ser
The final deformation can be estimated as follows:
E4.5 Axially loaded fasteners
Dowels cannot be exposed to withdrawal loads, while for bolted joints, for the rope effect
portion, up to 25% of the portion in accordance with Johansen theory may be applied. The
withdrawal resistance is calculated from the minimum value of the tensile strength of the
bolt and the load‐bearing capacity of the washer, which corresponds to the compression
strength perpendicular to the grain between the washer and timber surface.
380
Joints with bolts and dowels
E4.6 Interaction between multiple fasteners in a row
For joints exposed to load in the grain direction, the force exerted is distributed unevenly
on the fasteners in the direction of force (especially if brittle failure occurs). This means
that the load‐bearing capacity of a joint with multiple fasteners in a row does not equate
to the sum of the load‐bearing capacities of the individual fasteners. The load‐bearing
capacity of such a joint with n dowels or bolts is calculated with an effective number nef:
n
n ef min 0.9 a1 (E4‐9)
n 4
13 d
Equation (E4‐9) is based on comprehensive experimental results. Provided there is no
premature splitting of the wood and plastic deformations are possible, the forces can be
distributed between the individual fasteners, as applies, for example, for joints exposed
to compression stress perpendicular to the grain, in wich nef = n applies. For dowelled
joints, according to NA, a further rule applies to determine the effective number of fas‐
teners. For moment‐resisting joints with multiple dowel circles, nef = 0.85 ∙ n applies.
Additional information and background details to the effective number of fasteners are
included in Article E13.
E4.7 Literature
J. Ehlbeck, H. Werner, original Article C6, C7, STEP 1995.
Blass HJ, Bienhaus A and Krämer V (2000) Effective bending capacity of dowel‐type fasteners.
CIB‐W18 Meeting 33, Paper 33‐7‐5. Delft, the Netherlands.
381
E5 Joints with self‐tapping screws
Original article: J. Ehlbeck, W. Ehrhardt
Great progress has been made in recent decades in the development of wood screws.
Meanwhile, timber construction mainly features the use of self‐tapping screws, which, as
circumstances allow and depending on the declaration of performance and density, can
be deployed up to a diameter of 14 mm without predrilling. If predrilling takes place, the
drilling diameter is roughly equivalent to the inner thread diameter. Self‐tapping screws
have differing diameters, but tend to have a shank diameter that is smaller than the
external thread diameter, but exceeds that of the inner thread (= core diameter). The
nominal diameter of a self‐tapping screw is generally the external thread diameter. Such
screws come in numerous varieties (Figure E5‐1): partially threaded self‐tapping screws
with a smooth shank area, fully threaded screws, various screw heads and threads or
steel varieties.
Screwed joints are particularly ideal for connecting steel plates, steel members and
wood‐based materials to timber, but can also be used for purely timber‐to‐timber joints.
In this case, most of the joints feature one shear plane only. One additional and im‐
portant area of use is that of reinforcements (Articles D8 and E12), in which self‐tapping
screws are used as reinforcements in joints loaded perpendicular to the grain, notched
beams, openings and curved beams or as reinforcements in beam supports (see Section
E5.3). Reinforcements almost exclusively feature the use of fully threaded screws, the
use of which is explicitly regulated in the NA.
The European product standard EN 14592 “Timber structures ‒ Dowel‐type fasteners ‒
Requirements” also applies to self‐tapping screws. Due to the variety of screws and the
associated variation in steel properties, diameters and threads, there is a need to have
screws checked by a notified body, whereupon the results will be declared by the manu‐
facturer as characteristic values of the yield moment, tensile capacity or withdrawal
parameter (declaration of performance). Alternatively, the characteristic values used to
calculate the load‐bearing capacity of screwed joints can also be specified in a European
Technical Assessment (ETA) (see also Article A1).
383
Joints with self‐tapping screws
Figure E5‐1 Typical self‐tapping screws.
This means that the parameters required for joint design, such as yield moment My,k,
withdrawal parameter fax,k and diameter d required to determine the embedment
strength for modern self‐tapping screws are stipulated in European Technical Assess‐
ments (ETA) or declared by the manufacturer in declarations of performance (DOP).
Accordingly, the engineer must consult the technical documents for the screw types used
(ETA or DOP).
One additional key detail in the ETAs concerns the torque moment. Self‐tapping screws
come with special drive systems (usually Torx), to initiate the required torque levels in
the screws to insert them. However, these torques must not be at a level which leads to
the screws being damaged due to the torque level initiated. The ratio of torque failure
moment to insertion moment must be at least 1.5 for the longest screws (see
EN 14592:2009).
Background details, design equations for laterally loaded joints and derivations thereof
can be taken from Articles E1 and E2. The NA sets out that a screwed joint must comprise
at least two screws, except when used to secure casings, battens, wind braces, rafters,
purlins and similar.
E5.1 Load‐bearing behaviour
Self‐tapping screws in timber joints can transmit high forces in the direction of the screw
axis thanks to their thread, which ensures an effective bond between fasteners and the
surrounding wood, as well as the high steel quality involved and this is a quality structural
engineers can exploit. Self‐tapping screws, which are subject to axial load when pulled
out, often allow far higher forces to be transmitted than screws subject to lateral load.
When subject to lateral load, screws may have lower load‐bearing capacities than nails or
dowels of the same nominal diameter (and the same tensile strength), since the bending
resistance of the threaded part is far lower, due to the lower core area. For fully threaded
screws however, when exposed to lateral loads, the rope effect can result in a load‐bearing
capacity for the screw, which exceeds that of a corresponding dowel of equivalent nominal
384
Joints with self‐tapping screws
diameter. For small fastener diameters (e.g. nails), the angle between the force and grain
directions of the wood has hardly any influence on embedment strength and hence load‐
bearing capacity. For larger fastener diameters meanwhile (e.g. bolts or dowels), a no‐
ticeable impact on the embedment strength of the members to be connected is apparent.
This means that screwed joints where d < 6 mm can be designed like nailed joints, while
for d ≥ 6 mm, the decrease in embedment strength with increasing angle between the
force and grain directions has to be considered. Even so, most ETAs for self‐tapping
screws allow any reduction in embedment strength where > 0° to be ignored for
screws where d ≥ 6 mm.
E5.2 Design of joints with self‐tapping screws
The regulations of EC 5 on joints with screws are confusing and sometimes outdated. The
basic regulations for screws subject to lateral load still originate from the period before
the introduction of modern, self‐tapping screws and only apply to partially threaded
screws, namely those with a partially smooth shank and an external thread diameter
equivalent to that of the smooth shank. As withdrawal resistance has become an increas‐
ingly important parameter with the emergence of modern screws, the regulations ap‐
plied in EC 5 for this type of loading are unclear and often contradict the respective dec‐
larations of performance or ETAs of the manufacturer. The latter often also include more
favourable values for the respective screws than EC 5, while it is preferable to reference
regulations on minimum distances from the individual ETAs rather than EC 5. Since both
the Johansen model and the rope effect are relevant for modern self‐tapping screws with
partial or full threads (Article E2), the simplest approach involves extracting the parame‐
ters required to calculate the load‐bearing capacity, such as yield moment, embedment
strength or withdrawal resistance, directly from the ETAs or the declarations of perfor‐
mance of the used screws rather than using the equations of the EC 5 concerning with‐
drawal or head pull‐through resistances. The very high load‐bearing capacity in the screw
axis means the portion of the rope effect to be taken into account for screws may be up
to 100% of the total according to Johansen theory. The wide variety of available screws
(differing threads, head shapes, steel qualities, inner thread diameter, external diameter,
etc.) also mean that in the foreseeable future, the resistances of the individual screw
types, e.g. the withdrawal resistance, must be defined in tests and consequently taken
from the technical documents.
Laterally loaded screws
Depending on the joint type (e.g. a single‐ or double‐shear timber‐to‐timber joint), the
load‐bearing capacities are to be determined using the Johansen model with rope effect
as shown in Article E2. The required values such as yield moment, withdrawal parameter
or minimum distances are to be taken from the ETAs or the declarations of performance.
385
Joints with self‐tapping screws
As already discussed in Section E5.1, the embedment strength of screws of nominal di‐
ameter up to 6 mm is determined with the equations for (non‐predrilled) nails (equation
(E3‐1)) and that for screws where d > 6 mm is determined with the equations for bolts
(equations (E4‐1) and (E4‐2)). When multiple screws are arranged in a row in the grain
direction, the risk of splitting means an effective number nef < n must be applied. This
effective number is, in turn, calculated with the equation for nails (screws with a nominal
diameter of up to 6 mm, equation (E3‐11)) or those for bolts (d > 6 mm, equation (E4‐9)).
Axially loaded screws
Wood screws are particularly ideal when it comes to absorbing greater withdrawal loads
and are thus often deployed in joints where the fasteners are arranged in an inclined or
cross‐wise set‐up. An inclined arrangement means the screws are subject to biaxial
stresses; in the screw axis (when pulled out) and perpendicular to the screw axis (Johan‐
sen theory), whereby, given the high load‐bearing capacity in the screw axis, high‐load
bearing joints can be formed. The vast majority of the load is accommodated in the longi‐
tudinal screw direction in this case, given the clearly higher stiffness than when laterally
loaded. The lateral load‐bearing capacity is not considered for screws arranged at an
angle, since the equations used to define the embedment strength in EC 5 only apply to
fasteners arranged in a direction perpendicular to the grain. However, the influence of
friction on the shear plane can be taken into consideration for design purposes. For axial‐
ly loaded screws, only the penetration depth of the threaded part can be taken into
account. The following failure modes must be checked (the head pull‐off resistance must
exceed the tensile strength of the screw):
Withdrawal failure of the threaded part in the member with the screw tip
(parameter fax),
Pull‐through failure of the screw head or withdrawal failure of the
threaded part in the member with the screw head (parameter fhead or fax),
Tensile failure of the screw (tensile strength ftens),
Buckling failure of the screw when exposed to axial compression
(see Section E5.3),
Block shear failure (see Article E13).
The required parameters plus the minimum spacings and end and edge distances are
specified in the technical documents. The minimum penetration depth of the thread on
the screw tip side should usually be at least 6 ∙ d and the screwing angle between the
screw axis (force) and grain direction must generally be at least 30°. There are also ETAs
for which the angle between the screw axis and grain direction can be between 0 and 90°.
386
Joints with self‐tapping screws
An effective number of screws nef must also be taken into consideration. This reflects the
fact that due to the risk of brittle failure, the load‐bearing capacity of a group of n screws
is lower than the total of the individual load‐bearing capacities of the screws (N. B.: nef is
applicable to groups of screws, which are loaded by a systematic tensile force acting in
the axial direction, but not for the rope effect):
0.9
n ef n (E5‐1)
E5.3 Buckling under compressive load
The verification of a screwed joint when exposed to compression is determined by the
minimum resistance value against pushing‐in and buckling. This verification is carried out
e.g. when using fully threaded screws as reinforcement in beam supports, if the load‐
bearing capacity of a support perpendicular to the grain is supposed to be increased, see
Figure E5‐2 on the left (Bejtka and Blass, 2006). For cross‐wise arranged screws, buckling
analysis must be performed for the compressed screw. The resistance against pushing‐in
can be determined with the withdrawal parameter fax,k, while the buckling verification is
performed in accordance with EC 3. The relative slenderness ratio is determined by tak‐
ing the elastic foundation of the screw in the timber into consideration. Figure E5‐2 on
the right shows test specimens with buckled fully threaded screws.
Reinforcement
Figure E5‐2 Left: Example of a reinforcement in a beam support. Right: Test specimens with buckled screws
(Bejtka, 2005).
387
Joints with self‐tapping screws
The design load‐bearing capacity Fax,Rd in the direction of the screw axis is determined as
follows:
F ax,Rd min f ax,d d ef ; c N pl,d (E5‐2)
The first term in equation (E5‐2) shows that the resistance of a screw against being
pushed in corresponds to the withdrawal resistance, since the withdrawal parameter fax,d
is also used to determine the push‐in resistance.
The resistance against buckling in accordance with EC 3 is determined as follows (partial
safety factor here is M1 = 1.1 (Germany) for steel in the buckling analysis):
1 for k 0.2
c 1
for k 0.2
(E5‐3)
k k k
2 2
where
k 0.5 1 0.49 k 0.2 k2
(E5‐4)
and with the relative slenderness ratio
N pl,k
k (E5‐5)
N ki,k
Here:
d 12 Characteristic, plastic normal force,
N pl,k f y,k
4 related to the inner thread diameter
Characteristic yield strength of the
fy,k
screw in the core area
Characteristic elastic buckling load, depending on
N ki,k c h E S I S the inner thread diameter, characteristic density
and elastic foundation of the screw in the timber
90
c h 0.19 0.012 d k Elastic foundation modulus
180
k Characteristic density of the timber
210 000 d 14
E S I S Bending stiffness of the inner thread diameter
64
d1 Inner thread diameter of the fully threaded screw
Angle between screw axis and grain directions
388
Joints with self‐tapping screws
The design load‐bearing capacity F90,Rd of a reinforced support amounts to:
kc,90 is the coefficient for compression perpendicular to the grain, cf. equation (D1‐4). The
effective contact length ℓef, which is applied during compressive stress exposure perpen‐
dicular to the grain to determine the effective contact surface, can also be taken from
Article D1.
However, failure due to compression perpendicular to the grain may occur not only on
the lower edge of the timber beam, but also on the level of the screw tips. Compression
perpendicular to the grain at the level of the screw tip, meanwhile, is verified with the
lower term of equation (E5‐6), where Aef = ℓef,2 ∙ b represents the effective area at the
level of the screw tips, which is subject to compression perpendicular to the grain. Aef
depends on the type of load (direct or indirect), the type of load introduction (single‐ or
double‐sided) and the resulting load distribution and is explained in Figure E5‐3 and
Figure E5‐4. Comprehensive simulations have shown that the stress distribution is ap‐
proximately linear when a direct load is introduced, which means ℓef,2 can be calculated
directly from the thread length ℓs (Figure E5‐3). When an indirect load is applied, con‐
versely, simulations show the emergence of exponential stress distribution (Figure E5‐4).
Accordingly, for indirect loading, the following equations derived from regressions apply
(Bejtka and Blass, 2006):
3.3 s
Single ‐ sided load introduction: ef 0.25 s e h
(E5‐7)
3.6 s
Double ‐ sided load introduction: ef 0.58 s e h
389
Joints with self‐tapping screws
F F
F F
ℓ ℓ
h ef,2 h ef,2
ℓS ℓS
≈ 45° ≈ 45°
ℓ ℓ
Figure E5‐3 Above: Load distribution with direct load introduction (Bejtka, 2005). Below: Effective area
at level of screw tips with direct load introduction (Bejtka and Blass, 2006).
F/2 F/2
ℓ ℓ
h ef,2 h ef,2
ℓS ℓS
ℓ ℓ
Figure E5‐4 Above: Load distribution with indirect load introduction (Bejtka, 2005). Below: Effective area
at level of screw tips with indirect load introduction (Bejtka and Blass, 2006).
390
Joints with self‐tapping screws
E5.4 Joint stiffness
For laterally loaded screws, equations to determine Kser can be taken from the provisions
for nails (for screws of up to 6 mm (external thread diameter)) or for bolts (for screws
from 6 mm in diameter). For axially loaded screws, empirical equations to determine
Kser = Kax are specified in the technical documents of the individual screws.
One example shown here is how the stiffness parallel to the shear plane can be calculat‐
ed for a joint with cross‐wise arranged screws, as shown in Figure E5‐5. For joints with
cross‐wise arranged screws, one screw is pushed in and the other pulled out.
D
F F
Shear plane
Figure E5‐5 Cross‐wise arranged screws, inclined by angle .
The stiffness of the joint in the direction of the shear plane Kser can be determinined from
the rigidities Kax in the direction of the screw axes using the principle of virtual work,
while waiving the portions due to lateral load and friction.
Normal force N in screw axis due to force F (F/2 per screw):
F
N (E5‐8)
2 cos
Normal force in screw axis due to virtual force "1":
1
N (E5‐9)
2 cos
391
Joints with self‐tapping screws
Accordingly, it is now possible to calculate the deformation w in the shear plane direction
for one side of the joint (for n = 2 screws and assuming the normal force is evenly distrib‐
uted over the thread length → superimposition of two rectangles):
2 N N N1 N1 N 2 N 2 1 F 1 F
w 2 (E5‐10)
n=1 K ax K ax K ax K ax 2 cos 2 cos 2 K ax cos 2
The stiffness Kser of the joint is as follows for each side of the shear plane:
F
K ser,SF 2 K ax cos 2 (E5‐11)
w
Consequently, the overall stiffness with equivalent thread lengths on both sides of the
joint is as follows:
K ser K ax cos
2
(E5‐12)
E5.5 Combined loading
When screws are simultaneously subject to lateral loads Fv,Ed and axial loads Fax,Ed, the
following equation must be satisfied:
2 2
F v,Ed F ax,Ed
1 (E5‐13)
F v,Rd F ax,Rd
E5.6 Literature
J. Ehlbeck, W. Ehrhardt, original Article C8, STEP 1995.
Bejtka I. (2005). Verstärkung von Bauteilen aus Holz mit Vollgewindeschrauben. Dissertation.
Universität Karlsruhe.
Bejtka I. and Blass H.J. (2006). Self‐tapping screws as reinforcement in beam supports. Paper 39‐7‐2,
CIB‐W18 Meeting 39, Florence.
392
E6 Joints with connectors
Original articles: H. J. Blass
The term 'connectors' is applicable to so‐called split ring, shear plate and toothed‐plate
connectors. Split ring and shear plate connectors, as shown in Figure E6‐1, are used in
both timber‐to‐timber as well as steel‐to‐timber joints; usually together with bolts. While
split ring connectors can only be used in timber‐to‐timber joints, shear plate connectors
are also suitable for joining wood with steel. Shear plate connectors are used in timber‐
to‐timber joints, if e.g. the connectors are to be pre‐installed prior to assembly or if the
construction has to be assembled and disassembled multiple times. Split ring and shear
plate connectors come in wide‐ranging shapes and sizes and with diameters ranging
between 60 and 260 mm. They are always circular, since they are inserted into pre‐milled
depressions in the timber and can be made of cast aluminium alloys, steel, cast steel or
solid oak. Most split ring and shear plate connectors used in Europe are described in
EN 912 “Timber fasteners ‒ Specifications for connectors for timbers”. EN 912 sets out
descriptions of split ring connectors made of metal as type A, shear plate connectors as
type B and split ring connectors made of solid oak as type D.
Figure E6‐1 Joint with a split ring (left) and a shear plate connector (right). (STEP 1995 Article C9)
Multiple procedures are required when it comes to fabricating a joint using split ring or
shear plate connectors. The first involves milling the depression for the connector, which
is usually done at the same time as drilling the bolt hole (see Figure E6‐2 on the left).
Next, the connector is placed in the corresponding depressions and the timber members
to be connected are assembled. Finally, the bolts are slid into the holes and the nuts
tightened (see Figure E6‐2 on the right). If no bolts are usable, e.g. for connectors on the
upper side of deep glulam beams, screws can also be used instead of bolts.
393
Joints with connectors
Figure E6‐2 Left: Drilling the bolt hole and milling the depression for the connector. Right: Assembly of a joint
with split ring connectors. (STEP 1995 Article C9)
Toothed‐plate connectors as shown in Figure E6‐3 are used similarly to the split ring and
shear plate connectors in timber‐to‐timber and steel‐to‐timber joints, normally together
with bolts. While split ring and shear plate connectors are inserted into pre‐milled de‐
pressions in the timber, toothed‐plate connectors are pressed into the pieces of timber
to be connected. To avoid any gap due to the base plate (up to 3 mm) of the toothed‐
plate connectors between the timber members, the base plates can also be placed in
pre‐milled depressions in the wood. While double‐sided toothed‐plate connectors are
only usable in timber‐to‐timber joints, single‐sided toothed‐plate connectors can also be
used when connecting timber and steel. Like shear plate connectors, single‐sided
toothed‐plate connectors are also used in timber‐to‐timber joints, e.g. if the connector is
to be installed already prior to assembly or if the construction has to be assembled and
disassembled multiple times. Given the need to press the “teeth” into the pieces of
wood, which becomes increasingly difficult with rising density, toothed‐plate connectors
can only be installed in timber members with a characteristic density of less than around
500 kg/m3.
Figure E6‐3 Joint with double‐sided (left) and single‐sided toothed‐plated connector (right).
(STEP 1995 Article C10)
394
Joints with connectors
Toothed‐plate connectors come in many shapes and sizes, with diameters between 38
and 165 mm. Although they tend to be circular, square and oval shapes are also possible.
Toothed‐plate connectors comprise cold‐formed steel strip, hot‐dip galvanised steel or
malleable cast iron. Most of the toothed‐plate connectors used in Europe are also de‐
scribed in EN 912 “Timber fasteners ‒ Specifications for connectors for timbers”. EN 912
describes toothed‐plate connectors as type C.
Joints with toothed‐plate connectors are manufactured similarly to bolted joints. First of
all, the holes for the bolts are drilled. The connectors are then placed between the mem‐
bers to be connected and pressed together, during which the teeth of the toothed‐plate
connectors engage with the wood. Since the pressing‐in process requires considerable
force, either a special high‐strength bolt or a hydraulic press is used for this process. The
bolts required for the joint can only be used for the task of pressing in for small connect‐
or diameters of up to around 65 mm. To avoid excessive indentations perpendicular to
the grain, when pressing in with bolts, oversized washers should be used to distribute the
load. After being pressed in, the final bolts are assembled and tightened. As an alterna‐
tive to bolts, screws can also be used, as is done for joints with split ring and shear plate
connectors.
For both connector types, cross‐sectional weaknesses must be taken into consideration
when verifying the timber members. Here, the NA specifies so‐called shortfall areas for
all common connector types used in Germany.
E6.1 Load‐bearing behaviour
For double‐sided connectors (split ring and double‐sided toothed‐plate connector), the
force is initially transmitted from one member via embedment stresses into the connect‐
or and from there, via the shear resistance of the connector to the second member. The
bolts accommodate the resulting eccentricity and keep the joint together. For single‐
sided connectors (shear plate and single‐sided toothed‐plate connector), meanwhile,
forces are distributed slightly differently: once the force has been transmitted to the
connector, the bolt is exposed to embedment stress between the connector and bolt.
Subsequently, the shear resistance of the bolt means that for timber‐to‐timber joints, the
force is transmitted from the bolt to the second connector, while for steel‐to‐timber
joints, it is transmitted directly to the steel member. The hole diameter in single‐sided
connectors must accordingly correspond to the bolt diameter plus a small additional
tolerance. This tolerance means initial slip for joints with single‐sided connectors.
395
Joints with connectors
Split ring and shear plate connectors
When performing tensile tests on joints with split ring or shear plate connectors, the
most common cause of failure observed is a shear block failure of the timber in front of
the connector. The calculation model used to determine the load‐bearing capacity of
such connector joints thus assumes the occurrence of such failure. Although unevenly
distributed in reality, embedment stresses are assumed to be uniformly distributed and
to act in parallel to the direction of the force applied. These are then, in turn, transferred
to the member via shear stresses (Figure E6‐4 centre). The sudden shear failure means
such joints may fail, even if only very minor deformations are involved, which also means
interaction of the bolt and connector cannot be guaranteed. The load‐bearing capacity of
the bolt is disregarded, since the bolt only begins to carry load following an initial slip,
due to clearance. Figure E6‐5 shows a tensile connection after shear failure in both the
middle and one side member.
he a1
a3,t
dc
Figure E6‐4 Stresses in a joint with a split ring or shear plate connector with associated shear areas.
(STEP 1995 Article C9)
Figure E6‐5 Shear failure in middle and side member of a joint with split rings subject to tension.
(STEP 1995 Article C9)
396
Joints with connectors
Assuming that the shear failure of the wood is the key cause of failure for tensile joints,
the load‐bearing capacity depends on the shear area As in front of the connector and the
shear strength of the wood itself. The shear area within the connector is not taken into
consideration, since most tests showed that the wood core within the connector sheared
off before the maximum load was reached, meaning it did not help boost the total load‐
bearing capacity in any way. The shear failure of the wood in front of the connector,
however, only occurs when the wood has sufficient embedment strength. Accordingly,
for larger distances to the loaded end a3,t, embedment failure is the main factor dictating
the load‐bearing capacity of the joint (Figure E6‐4 on the right).
If joints with connectors are loaded at an angle exceeding 30° or so to the grain direction
or if loaded in compression, other failure mechanisms are triggered. Joints with angles
between the force and grain directions of between around 30 and 150° tend to fail due
to the wood splitting (Figure E6‐6 on the left). Joints loaded in compression, meanwhile,
exhibit combined embedment and splitting failure (Figure E6‐6 on the right). In this case,
however, the splitting only occurs following considerable embedment deformations
under the connector and bolt. In a state of failure therefore, both connector and bolt
transmit some of the force, although this interaction between connector and bolt only
emerges in joints subject to compression. Compared to joints subject to tension or loads
acting at an angle of less than 150°, which show brittle failure, greater plastic defor‐
mations occur in this case before the load peaks. Since the wood core within the con‐
nector of compressed joints is prone to shear failure before this point, the effective em‐
bedment area of the bolt is reduced by the area within the connector.
Figure E6‐6 Left: Splitting failure in joints with a split ring loaded at an angle of 60° to the grain. (STEP 1995
Article C9). Right: Combined embedment‐splitting failure in a joint with a split ring loaded in
compression. (Photo: TU Delft, Vermeijden and Kuipers)
397
Joints with connectors
The load‐bearing capacity Fv,0,R of a joint with connectors subject to tensile load parallel
to the grain is hence the smaller of the two values for resistance to shear and embed‐
ment respectively (definitions see also Figure E6‐4 on the right):
f A (shear)
F v,0,R min v s (E6‐1)
f h d c he (embedment)
where
Fv,0,R Load‐bearing capacity of a split ring or shear plate connector,
fv Apparent or mean shear strength of the wood,
As Shear area per connector,
fh Embedment strength of the wood,
d c Connector diameter and
he Insertion depth of the connector in the wood.
As numerous tests conducted by Kuipers and Vermeyden (1964) have shown, the shear
strength declines with increasing shear area, which explains the reference to ‘apparent
shear strength’, as determined by many tests. However, as applies to all joints, certain
minimum wood thicknesses are required, given the risk of splitting failure for smaller
wood thicknesses, rather than shear or embedment failure, resulting in a lower load‐
bearing capacity (Scholten, 1944).
Although shear or embedment failure actually only affects joints loaded at an angle be‐
tween the force and grain directions of up to around 30° and splitting failure, a different
failure mechanism, occurs in joints loaded up to an angle of 150°, many tests showed
that even for those joints loaded at an angle between the force and grain directions
exceeding 30°, a calculation employing the approach of equation (E6‐1), namely without
taking splitting into consideration, would suffice. This can be explained by the fact that
spacings and edge distances affect the load‐bearing capacity in similar ways when various
failure mechanisms occur. If splitting failure does occur when load is applied at an angle
to the grain direction, increasing the distance to the loaded end results in the area of
wood exposed to tension perpendicular to the grain expanding. The same applies when
shear failure occurs in joints loaded parallel to the grain. Here, likewise, increasing the
distance to the loaded end increases the size of the stressed areas in the wood, although
shear stress in this case. Tests could show that the “higher load‐bearing capacity due to
greater end distance” was similar for both failure modes “tension perpendicular to the
grain” and “shear”. Accordingly, equation (E6‐1) suffices as a means of verifying both
failure modes. Only when the distance to the end is excessive and the maximum load is
reached without the wood splitting or shearing off is no further increase in load‐bearing
capacity expected, even with increasing edge distance.
398
Joints with connectors
Toothed‐plate connectors
In many tests with joints with toothed‐plate connectors, the observed failure mode was
embedment failure, under both connector teeth and bolts. The teeth were also bent in
some types of toothed‐plate connectors. Conversely, for tensile joints featuring short end
distances, the failure modes involved splitting and shear failure of the wood. Since joints
with toothed‐plate connectors tend to undergo prominent plastic deformations before
the maximum load is reached, interaction between connectors and bolts can be as‐
sumed. Figure E6‐7 shows a joint after the maximum load has been reached, with both
plastic embedment deformations of the wood under the connector teeth and bolt and
plastic deformation of the bolt and connector teeth clearly apparent.
The load‐bearing capacity of joints with toothed‐plate connectors thus takes into consid‐
eration the interaction of connector and bolt:
where
Fv,Rk Characteristic load‐bearing capacity of the joint with connector and bolt
FC,Rk Characteristic load‐bearing capacity of the toothed‐plate connector (type C)
FB,Rk Characteristic load‐bearing capacity of the bolt in accordance with EC 5 with
the characteristic values of embedment strength and yield moment of the bolt
(Article E4)
Figure E6‐7 Embedment failure under the connector teeth and bolt. The teeth and bolt are deformed
plastically. (STEP 1995 Article C10)
399
Joints with connectors
E6.2 Load‐bearing capacity of split ring and
shear plate connectors
The calculation model of equation (E6‐1) was transposed via numerous tests to a design
model, where, ensuring compliance with minimum distances and recommended mini‐
mum thicknesses, many different kinds of split ring and shear plate connectors, joint
geometries, densities and angles between the force and grain directions can be calculat‐
ed. To transpose equation (E6‐1) to a design model, tests were used to determine regres‐
sion functions for the apparent shear strength fv and the embedment strength fh:
As previously explained, shear strength decreases with increasing shear area (Kuipers and
Vermeyden, 1964), and the embedment strength is determined as a function of density,
since density is the key influencing parameter. The next step involves using Figure E6‐4 to
determine the shear area As, assuming a3,t = 2 ∙ dc and he = hc/2 = dc/8 (hc = connector
height):
2 2
1 d d d
A s 2 a 3,t h e a 3,t d c c 2 2 d c c 2 d c d c c 2.11 d c2 (E6‐4)
2 4 8 8
The reference density was k = 350 kg/m3; which, together with equation (E6‐4), results in:
35 d c
1.5
F v,0,Rk min (E6‐5)
31.5 d c h e
Equation (E6‐5) shows the design model, although it only applies for the checked configu‐
ration. Correction coefficients are also required, which take the influence of differing
densities into consideration (other than the reference density 350 kg/m3), varying wood
thicknesses and the influence of the distance to the loaded end a3,t. The final design
equation meanwhile, regulated in EC 5 for split ring and shear plate connectors, reads as
follows:
k k k k 35 d c1.5
F v,0,Rk min 1 2 3 4 (E6‐6)
k 1 k 3 31.5 d c h e
400
Joints with connectors
k1 is the correction coefficient for varying wood thicknesses t1 and t2 and takes into con‐
sideration the unfavourable effect of thinner side or middle members on load‐bearing
capacity:
t t
k 1 min 1; 1 ; 2 (E6‐7)
3 he 5 he
k2 is the correction coefficient for the distance from the loaded end (where the above‐
mentioned assumption of a3,t = 2 ∙ dc can be found back):
a
k 2 min k a ; 3,t (E6‐8)
2d c
ka = 1.25 for joints with one connector per shear plane and ka = 1.0 for multiple connect‐
ors per shear plane. The angle between the force and grain directions in this case peaks
at 30°; for other values of , k2 = 1.0.
k3 takes into consideration the increased load‐bearing capacity for densities exceeding
350 kg/m3, where characteristic densities over 610 kg/m3 are no longer taken into con‐
sideration:
k 3 min 1.75; k (E6‐9)
350
Finally, we have:
When a load is exerted at an angle to the grain direction, the characteristic load‐bearing
capacity Fv,0,Rk determined in accordance with equation (E6‐6) must be reduced as follows
(similar to determining the embedment strength of a bolted joint loaded at an angle to
the grain, equation (E4‐2)):
F v,0,Rk
F v,α,Rk (E6‐11)
k 90 sin 2 cos 2
where k90 = 1.3 + 0.001 ∙ dc.
401
Joints with connectors
E6.3 Load‐bearing capacity of toothed‐plate connectors
Just as for those with split ring and shear plate connectors, the load‐bearing capacity of
joints with toothed‐plate connectors was determined from numerous test results, which
covered wide‐ranging connector types, joint geometries, densities and angles between
the force and grain directions, while complying with minimum distances and recom‐
mended minimum thicknesses. As equation (E6‐2) shows, the load‐bearing capacity of a
joint with toothed‐plate connectors is determined by adding the individual load‐bearing
capacities for connectors and bolts, since connectors and bolts work together in toothed‐
plate connectors, unlike split ring and shear plate connectors. The approach taken to
determining the load‐bearing capacity of the bolt was already explained in Article E4.
Accordingly, reference will only be made at this point to determining FC,Rk of the toothed‐
plate connector as a portion of the total load‐bearing capacity. When determining the
load‐bearing capacity of the connector, differences are established between two differ‐
ent types (note: connector types C6 to C9 are not normally used in Germany):
Types C1 to C9 comprise plates, with edges, which are bent up into teeth.
Types C10 and C11 comprise a plate with added conical teeth.
One difference emerging in comparison to split ring and shear plate connectors comes in
how the connector diameter is defined. For toothed‐plate connectors, the following
applies (all dimensions in mm):
dc: Diameter of connectors with teeth of types C1, C2, C6, C7, C10, C11
dc: Side length of connectors with teeth of types C5, C8, C9
dc: Root of the product of the side lengths of connectors with teeth of types C3, C4
Analogous to equation (E6‐1), the calculation model of equation (E6‐2) is transformed
into a design model:
Coefficients k1 to k3 are very similar to those of the design equations for split ring and
shear plate connectors and have the same significance. Coefficient k1 is identical:
t t
k 1 min 1; 1 ; 2 (E6‐13)
3 he 5 he
402
Joints with connectors
Coefficient k2 differs for the two connector types, C1 to C9 or C10/C11. For connectors of
types C1 to C9, meanwhile, the following applies:
a 3,t
k 2 min 1; (E6‐14)
1.5 d c
where for a3,t, the following value is taken with d = bolt diameter in mm:
For toothed‐plate connectors of types C10 and C11, the following applies for k2 and a3,t:
a
k 2 min 1; 3,t with a 3,t max 1.5 d c ; 7 d ; 80 mm (E6‐16)
2d c
For coefficient k3, only the upper limit of density changes in comparison to the split ring
and shear plate connectors (610 kg/m3 reduced to 525 kg/m3):
k 3 min 1.5; k (E6‐17)
350
E6.4 Minimum distances and spacings
As with all joint types, minimum distances must also be ensured for joints with connect‐
ors, which means the design equations are valid and no unwanted premature splitting
takes place. The minimum spacings and distances to be maintained for split ring and
shear plate connectors are specified in Table E6‐1 (for definitions of the minimum dis‐
tances, see Figure E1‐9). Minimum distances for (toothed‐plate) connectors of types C1
to C9 are specified in Table E6‐2 and those for types C10 and C11 in Table E6‐3. Notwith‐
standing the above, the corresponding minimum distances for the bolts must be ob‐
served in accordance with EC 5 (Table E4‐1). EC 5 Section 10.4.3 also sets out require‐
ments for bolt diameters in combination with connectors.
403
Joints with connectors
Table E6‐1 Minimum distances of split ring and shear plate connectors, dc = connector diameter in mm,
0° ≤ ≤ 90° = angle between force and grain directions. See also EC 1995‐1‐1/A2:2014 Table 8.7.
Spacing a1 (parallel to the grain) ‐ (1.2 + 0.8 ∙ cos) ∙ dc
Spacing a2 (perpendicular to the grain) ‐ 1.2 ∙ dc
Distance a3,t (loaded end) ‐ 2.0 ∙ dc
< 30° 1.2 ∙ dc
Distance a3,c (unloaded end)
30°≤ ≤ 90° (0.4 + 1.6 ∙ sin) ∙ dc
Distance a4,t (loaded edge) ‐ (0.6 + 0.2 ∙ sin) ∙ dc
Distance a4,c (unloaded edge) ‐ 0.6 ∙ dc
Table E6‐2 Minimum distances of toothed‐plate connectors of types C1 to C9, dc = connector diameter in
mm, 0° ≤ ≤ 90° = angle between force and grain directions. EC 1995‐1‐1/A2:2014 Table 8.8.
Spacing a1 (parallel to the grain) ‐ (1.2 + 0.3 ∙ cos) ∙ dc
Spacing a2 (perpendicular to the grain) ‐ 1.2 ∙ dc
Distance a3,t (loaded end) ‐ 1.5 ∙ dc
< 30° 1.2 ∙ dc
Distance a3,c (unloaded end)
30°≤ ≤ 90° (0.9 + 0.6 ∙ sin) ∙ dc
Distance a4,t (loaded edge) ‐ (0.6 + 0.2 ∙ sin) ∙ dc
Distance a4,c (unloaded edge) ‐ 0.6 ∙ dc
Table E6‐3 Minimum distances of toothed‐plate connectors of types C10 and C11, dc = connector diameter
in mm, 0° ≤ ≤ 90° = angle between force and grain directions. See EC 1995‐1‐1:2010 Table 8.9.
Spacing a1 (parallel to the grain) ‐ (1.2 + 0.8 ∙ cos) ∙ dc
Spacing a2 (perpendicular to the grain) ‐ 1.2 ∙ dc
Distance a3,t (loaded end) ‐ 2.0 ∙ dc
< 30° 1.2 ∙ dc
Distance a3,c (unloaded end)
30°≤ ≤ 90° (0.4 + 1.6 ∙ sin) ∙ dc
Distance a4,t (loaded edge) ‐ (0.6 + 0.2 ∙ sin) ∙ dc
Distance a4,c (unloaded edge) ‐ 0.6 ∙ dc
404
Joints with connectors
E6.5 Effective number of connectors
Similar to joints with dowel‐type fasteners, for joints with connectors, an effective num‐
ber nef also applies when more than two connectors are arranged in a row parallel to the
grain and force:
n
n ef 2 1 n 2 (E6‐18)
20
in which n is the number of connectors arranged in a row parallel to the grain and force.
Joints with toothed‐plate connectors are subject to ductile failure when complying with
the minimum distances, which is why calculations for this connector type should take the
actual number of fasteners (nef = n) into account.
E6.6 Joint stiffness
Stiffness values for mechanical joints are required to verify serviceability, as well as to
calculate the load‐bearing capacity of mechanically jointed beams. The ultimate slip
modulus Ku used to verify load‐bearing capacity includes two‐thirds of the corresponding
value of Kser. Given that the stiffness values of tested joints cover wide‐ranging limits, it is
almost impossible to reliably estimate the influence of different variables on the defor‐
mation behaviour. Accordingly, a simple relationship was selected instead, showing the
slip modulus as a function of the connector diameter and the mean density of the timber.
The influence of the angle between the force and grain directions, the wood moisture
content, wood thickness and the number of connectors per shear plane was disregarded.
The following equations apply for determining Kser in N/mm. Split ring and shear plate
connectors and toothed‐plate connectors of types C10 and C11:
dc m
K ser (E6‐19)
2
Toothed‐plate connectors of types C1 to C9:
dc m
K ser 1.5 (E6‐20)
4
where dc is the connector diameter in mm and m is the mean density of the wood in
kg/m3.
405
Joints with connectors
E6.7 Literature
H.J. Blass, original Articles C9, C10, STEP 1995.
Kuipers J. and Vermeyden P. (1964). Research on timber joints in the Netherlands. Rapport 4‐64‐15,
Onderzoek v‐7, Stevin‐Laboratorium, Technische Hogeschool Delft.
Scholten J.A. (1944). Timber‐connector joints ‒ their strength and design. Technical Bulletin No. 865,
USDA Forest Service, Washington, D.C..
406
E7 Punched metal plate fasteners
Original article: L. R. J. Whale
Punched metal plate fasteners are defined in EN 1075 “Timber structures ‒ Test methods ‒
Joints made with punched metal plate fasteners” as a fastener comprising a metal plate
with teeth (“nails”) punched in one direction, which stick out from the plate at right
angles. With punched metal plate fasteners, two or more timber members of equivalent
thickness can be connected, whereby at least one punched metal plate fastener must be
attached to each member side (symmetrical joint). A typical punched metal plate fasten‐
er is shown in Figure E7‐1. Punched metal plate fasteners are generally produced from
galvanised or stainless steel plates, between 0.9 and 3.0 mm thick. Plates with punched‐
out integral “nails” were originally used in the USA in the late 50s, after refining conven‐
tional hand‐nailed steel or plywood gusset plates. Both systems proved suitable for man‐
ufacturing joints in the same plane, but the punched metal plate fasteners were better
suited when it came to prefabricating trusses in the factory. They could transfer forces
via smaller joint areas, which also helped save on material.
Figure E7‐1 Typical punched metal plate fastener. (STEP 1995 Article C11)
Nowadays, punched metal plate fasteners are in widespread use for truss joints and
joints in many other plane timber structures (Figure E7‐2 (a)), which means that many
different punched metal plate fasteners encompassing wide‐ranging “nail” arrange‐
ments, lengths and shapes have been developed. However, the factors governing the
load‐bearing capacity of these different punched metal plate fasteners are all the same
and this enables uniform design, as was accommodated in EC 5 for these fastener types.
407
Punched metal plate fasteners
Aef
(e)
(d)
Aef
(a) (b)
Figure E7‐2 Typical joint: (a) view and (b) effective joint areas. (STEP 1995 Article C11)
E7.1 Parameters influencing the load‐bearing capacity of
connections with punched metal plate fasteners
In a connection with punched metal plate fasteners, the force is initially transferred from
the timber member to the nails, then from the nails to the steel plate, via the joint line
and finally, in turn, via the nails to the member on the other side of the joint. The load‐
bearing capacity of a connection with punched metal plate fasteners is thus reached
when one of the following two criteria apply: either the load‐bearing capacity of the nails
(anchorage) in one of the connected members is attained, or the load‐bearing capacity of
the net cross‐section of the steel plate in one of the joint lines between the members is
crucial. The parameters influencing both these load‐bearing capacity criteria can be
summarised as follows (definitions see Figure E7‐3).
y
FM M
anet
Aef l FM
F
α
bnet x
γ
β
e
Figure E7‐3 Connection with punched metal plate fastener ‒ geometry and loading. (STEP 1995 Article C11)
408
Punched metal plate fasteners
Parameters influencing the load‐bearing capacity of the nails (anchorage)
is the angle between the force direction and the main direction of the punched
metal plate fastener (the main direction is that where the tensile capacity of the
plate peaks, often parallel to the direction of the punched‐out nails) and is gener‐
ally the angle at which the individual nails are generally loaded. On the one hand,
this influences the embedment area in the timber, which is stressed by a nail, sec‐
ondly, the bending resistance of the nail itself, which often shows a rectangular
cross‐section.
is the angle between the force and grain directions; namely the angle to the
wood grain direction, under which the nails exert stress on the wood.
The wood species or strength class of the wood in the joint, namely its resistance
against embedment or withdrawal stress as exerted via the punched‐out nails.
Aef is the effective area of the punched metal plate fastener in each member,
namely the contact area between the punched metal plate fastener and wood,
taking tolerances for the positioning of the punched metal plate fastener into con‐
sideration. The effective joint area is defined as the smallest area that can be as‐
sumed, given that the plate is shifted arbitrarily from its correct position by a spe‐
cific tolerance (± 5 mm to the timber edges and 6 x nominal plate thickness to the
end grain) (see also Figure E7‐2 (b)).
rmax is the distance between the centre of gravity of the punched metal plate
fastener and the point of Aef situated furthest from the centre of gravity. Ip is the
polar moment of inertia of the effective area Aef about its centre of gravity. What
must be noted is that rmax and Ip only have an influence, when moments are to be
transferred over the punched metal plate fastener.
Parameters influencing the load‐bearing capacity of the steel plate (steel capacity)
is the angle between the main plate direction and the direction of the joint line
between the timber members (this angle influences the net cross‐section of the
steel along the joint line of the connection, see Figure E7‐3).
ℓ is the length of the joint line covered by the plate. The corresponding projected
lengths, parallel and perpendicular to the main plate direction, transfer the
stresses in both orthogonal plate directions in the joint line between the timber
members.
The steel grade, from which the punched metal plate fastener was manufactured,
namely the strength properties of the steel.
These parameters are included in design equations, with which the load‐bearing capacity
of joints can be determined depending on the specific strength properties of the
punched metal plate fastener.
409
Punched metal plate fasteners
E7.2 Determining characteristic values for punched metal
plate fasteners based on test results
According to EC 5, the following characteristic strength values are required for the design
of connections with punched metal plate fasteners:
fa,,,k is the characteristic nail load‐bearing capacity per unit of area under angles
and ; requiring fa,0,0,k = nail load‐bearing capacity for = 0° and = 0° and
fa,90,90,k = nail load‐bearing capacity for = 90° and = 90°.
ft,0,k is the characteristic plate tensile capacity per length unit in the main plate
direction ( = 0°).
ft,90,k is the characteristic plate tensile capacity per length unit perpendicular to
the main plate direction ( = 90°).
fc,0,k is the characteristic plate compressive capacity per length unit in the main
plate direction ( = 0°).
fc,90,k is the characteristic plate compressive capacity per length unit perpendicular
to the main plate direction ( = 90°).
fv,0,k is the characteristic plate shear capacity per length unit in the main plate
direction ( = 0°).
fv,90,k is the characteristic plate shear capacity per length unit perpendicular to the
main plate direction ( = 90°).
Each of these properties is to be determined from tests described in EN 1075. The 5%‐
quantiles are converted into design values by multiplying by the modification factor kmod
and dividing by the partial safety factor of the material M. The corresponding modifica‐
tion factors for timber are to be used to determine the nail load‐bearing capacity, while
for the plate load‐bearing capacities, the coefficients kmod = 1.0 and M = 1.25 can be
assumed. The nail load‐bearing capacity fa,,,k is to be determined over the entire range
of ‐ and ‐values. If a sufficient number of values in the fa,,,k‐plane have been deter‐
mined, interpolation between them is possible.
The characteristic nail load‐bearing capacity per plate in the grain direction ( = 0°)
fa,,,k is obtained from tests on joints (Figure E7‐4) with plate angles = 0°, 30°, 60° and
90°. A bilinear function can be adapted to these data (Figure E7‐5), eliciting the coeffi‐
cients k1, k2 and 0, which are then used in the following equations:
f a,0,0,k k 1 for 0
f a,α,0,k (E7‐1)
f a,0,0,k k 1 k 2 0 for 0 90
410
Punched metal plate fasteners
a,
0˚ < α < 90˚; β = 0˚
F α F
Figure E7‐4 Standard test specimen = 0°. (STEP 1995 Article C11)
fa,0,0 1
-k1
1
-k2
F
fa,90,0
α0
30 45 60 90 α(˚)
Figure E7‐5 Derivation of coefficients k1, k2, 0. (STEP 1995 Article C11)
The characteristic nail load‐bearing capacity per plate in the main plate direction ( = 0°)
fa,,,k is obtained from tests with T joints (Figure E7‐6 on the left) with grain angles = 45°
and 90° ( = 0° was already checked with the test setup of Figure E7‐4). The test results
reveal the interaction between the nail load‐bearing capacity and , Figure E7‐6 on the
right. For this diagram, test specimens with additional angles were checked.
fa,0,0
β
fa,0,30
fa,0,45
fa,0,60
fa,90,90 fa,0,90
F 30 45 60 90 β(˚)
Figure E7‐6 Standard test specimen = 0°. (STEP 1995 Article C11)
411
Punched metal plate fasteners
Once curves have been fitted, which represent the lower limits of the relationship be‐
tween fa,0,0 and fa,90,0 (bilinear curve, Figure E7‐5) for straight joints and between fa,0,0 and
fa,0,90 or fa,90,90 (sinusoidal curve, Figure E7‐6 on the right) for T‐joints, EC 5 specifies an
interpolation method for arbitrary values fa,,,k between these extreme values:
f a,α,0,k f a,α,0,k f a,90,90,k
max 45 for 45
a,0,0,k a,0,0,k a,90,90,k
f a,α,0,k f f f sin max , (E7‐2)
a,0,0,k f a,0,0,k f a,90,90,k sin
f
max , for 45 90
The characteristic course of the fitted area fa,,,d in accordance with equation (E7‐2) is
compared with test data of a typical punched metal plate fastener in Figure E7‐7.
E7.3 Design of connections with punched metal plate fasteners
Nail load‐bearing capacity (anchorage)
Induced stresses F,d or M,d from forces FEd and moments MEd, which act on the punched
metal plate area, can be calculated as follows:
F A,Ed
F,d (E7‐3)
A ef
M A,Ed
M,d (E7‐4)
Wp
412
Punched metal plate fasteners
where
FA,Ed Design value of the force acting on a punched metal plate fastener
(corresponds to half the force in the member) in the centre of gravity
of the effective joint area Aef (Aef defined in Section E7.1).
MA,Ed Design value of the moment acting on an effective joint area.
Wp Section modulus with r = distance from the centre of gravity of the
punched metal plate fastener:
W p r dA
A ef
2 2
F,d M,d
1 (E7‐5)
f a,α,β,d f a,0,0,d
413
Punched metal plate fasteners
Plate load‐bearing capacity (steel capacity)
To verify the plate load‐bearing capacity in a joint line of a connection, in accordance
with EC 5, all forces and plate load‐bearing capacities are resolved into both orthogonal
main directions x and y (Figure E7‐3):
where
FEd Resulting force of an individual punched metal plate fastener in the shear plane
(compression = negative).
FM,Ed Force caused by the moment in the shear plane where FM,Ed = 2 ∙ MA,Ed/ℓ
(Figure E7‐3).
The following ultimate limit state condition of the plate load‐bearing capacity must be
complied with in each joint area / shear plane:
2 2
F x,Ed F y,Ed
1 (E7‐8)
F x,Rd F y,Rd
where
Fx,Ed, Fy,Ed Design values of the forces in x‐ and y‐directions,
Equations (E7‐6) and (E7‐7)
Fx,Rd, Fy,Rd Design values of the plate load‐bearing capacities
The characteristic values of the plate load‐bearing capacities in both these directions are
determined as the maximum tensile and shear strengths of the plate as follows (here,
only the equations for tensile forces Fx,Ed > 0 and Fy,Ed > 0 are given, whereas for compres‐
sive forces, the design values are determined in a similar manner):
f t,90,k cos
F y,Rk max (E7‐10)
k f v,90,k sin
414
Punched metal plate fasteners
Coefficient k in equation (E7‐10) takes into consideration the influence of tensile or com‐
pressive force in the x‐direction of the plate on the shear strength of the plate in the y‐
direction:
The constants 0 (equation (E1‐9)) and kv are to be determined from tests (similar to the
coefficients k1, k2 and 0) and can be taken from the technical documents for the plate
types used.
E7.4 Additional rules
In addition to the abovementioned design rules, Section 10 of EC 5 also includes rules for
trusses made with punched metal plate fasteners, concerning limit values for lateral
deflections, distortions, curvatures or oblique positions during the assembly. The NA also
requires verification of the transport and assembly conditions.
In addition, there is also a need to meet the requirements of EN 14250 “Timber struc‐
tures ‒ Product requirements for prefabricated structural members assembled with
punched metal plate fasteners”. As well as formulating requirements for materials and
members (for example fire behaviour), information on the CE mark is also provided. In
addition to this European regulation, there is also a national draft, E DIN 20000‐6 “Applica‐
tion of building products in construction works – Part 6: Dowel‐type and non dowel‐type
fasteners”, describing further requirements, which are not regulated in the European
standards. Among other things, for punched metal plate fasteners deployed in members
with an overall length of > 12 m, the partial safety factors for the plate characteristics must
be increased by a factor of 1.5. The reason for this regulation is that the new European
standard includes less stringent requirements, which do not suffice to comply with the
national standard to date.
Values for the slip modulus Kser per mm² of the joint area can be taken from the technical
documents for the punched metal plate fasteners.
E7.5 Literature
L.R.J. Whale, original Article C11, STEP 1995.
415
E8 Cold‐formed steel connectors
Original article: E. Gehri
Cold‐formed steel connectors such as joist hangers or joist anchors have largely replaced
carpentry joints. This is primarily due to the ease of assembly of the steel connectors and
the scope to eliminate complex machining of the timber elements to be connected.
Figure E8‐1 shows examples of timber joints with cold‐formed steel connectors. The steel
is generally between 1 and 4 mm thick and either galvanised or made of stainless steel.
The joint between connector and timber is usually via ringed shank nails or self‐tapping
screws without any predrilling of the wood. The fact that the fastener holes have been
pre‐punched into the connectors guarantees swift and easy assembly at the construction
site. For concealed beam hangers however, namely in which joists are connected via
dowels, the joists must be split and drilled beforehand.
417
Cold‐formed steel connectors
E8.1 Load‐bearing behaviour
The load‐bearing capacity of the joints using cold‐formed steel connectors is not only
influenced by the capacity of the steel‐to‐timber joint, but also by the capacity of the
timber members and steel. In particular, tensile stresses perpendicular to the grain
emerging in the timber beams may trigger failures, even before the load‐bearing capacity
of the nailed or screwed joint has been reached. Examples of joints in which tension
failures perpendicular to the grain may occur include joist anchors and joist hangers as
shown in Figure E8‐1. Joints subject to tension perpendicular to the grain can be de‐
signed using one of the approaches outlined in Article E11.
For most cold‐formed steel connectors, steel failure in the net cross‐section is not the
decisive criterion. This is due to limiting the number of pre‐punched nail holes in the steel
and hence limiting the force introduced by the fasteners. However, most joints including
cold‐formed steel connectors are subject to plastic deformations of the steel before the
load‐bearing capacity is actually attained.
Since many cold‐formed steel connectors include at least two steel‐to‐timber joints in
different planes, these joints tend to be eccentrically loaded, which means the fasteners
concerned are often subject to combined lateral and axial loads.
Load‐bearing capacity of the nailed steel‐to‐timber joint
In accordance with EC 5, the design value of the load‐bearing capacity Fv,Rk per nail for
single‐shear joints with thin steel plates (namely for t ≤ 0.5 ∙ d where t is the steel plate
thickness) is the lower of the values obtained from the following equation (see also Arti‐
cle E2):
0.4 f h,k t 1 d
F v,Rk min F ax,Rk (E8‐1)
1.15 2 M y,Rk f h,k d
4
For joints with thick steel plates (namely for t ≥ d), meanwhile, the design value of the
load‐bearing capacity can be obtained from the following equation:
f h,k t 1 d
4 M y,Rk F
F v,Rk min f h,k t 1 d 2 1 ax,Rk (E8‐2)
2
f h,k d t 1 4
F
2.3 M y,Rk f h,k d ax,Rk
4
418
Cold‐formed steel connectors
For steel plate thicknesses of between 0.5 ∙ d and d, linear interpolation is permissible and
any variation in load‐bearing capacity according to equations (E8‐1) or (E8‐2) is based on
the clamping effect of the nail in the steel plate. Tests with nailed steel‐to‐timber joints
(Ehlbeck and Görlacher, 1982) showed, however, that also for steel plates of thickness
t = 2.0 mm and ringed shank nails of diameter d = 4.0 mm, the steel plate acted as a
clamped restraint for the nail, if the nails were conically shaped under the head (Figure E8‐
2) and driven into tight fitting holes in the steel plates. In these cases, the load‐bearing
capacities in accordance with EC 5 for thick steel plates were consistently attained.
t
D dk d1 d2
35º
s k L1
Ln
Figure E8‐2 Ringed shank nails for steel‐to‐timber joints with a conical shank under the head.
(STEP 1995 Article C13)
The characteristic value of the embedment strength fh,k depends on the nail diameter d in
mm and the characteristic timber density k in kg/m3 and is for non‐predrilled timber
(equation (E8‐3) corresponds to equation (E3‐1)):
0.3
f h,k 0.082 k d (E8‐3)
Given the varying cross‐section in the non‐smooth section of the nail shank and the work
hardening during manufacture, the characteristic value of the nail yield moment must be
determined by tests. Calculating the yield moment from the plastic section modulus and
the tensile strength of the nail wire is actually not applicable for ringed shank nails. Euro‐
pean technical assessments (ETA) or declarations of performance (DOP) for ringed shank
nails or screws reveal information about the yield moments to be applied or the charac‐
teristic values of load‐bearing capacities Fv,Rk (laterally loaded fasteners) and Fax,Rk (axially
loaded fasteners) are already specified per fastener and shear plane. The values specified
there are normally those derived for typical characteristic densities and must be reduced
when using wood with a density lower than the reference density. The characteristic
values of the load‐bearing capacity of nailed joints in accordance with EC 5 are defined as
for all fasteners under the prerequisite specific minimum nail distances. Since the holes in
the steel are pre‐punched, fastener distances are given. When designing joints with cold‐
formed steel connectors, however, the minimum distances must be verified.
419
Cold‐formed steel connectors
E8.2 Joist hangers
Joist hangers are frequently used as supports for solid timber and glulam timber beams and
are manufactured in wide‐ranging shapes and dimensions. Figure E8‐3 shows an example
of a joist hanger for a timber‐to‐timber connection.
40 B 40 < 3mm BH 42
20 13 20 7 2 33 7
e2
7
27,5 17 20
20 20
60 20 20 HN
H 20 20
HH 20 15
20
e1 80
> 4mm
Figure E8‐3 Joist hanger for a header‐joist connection. (STEP 1995 Article C13)
When the joist hanger is exposed to a uniaxial load, the load acts on its plane of sym‐
metry. The joist transmits the force via contact pressure into the base plate of the joist
hanger and simultaneously via the steel‐to‐timber connection into the joist hanger and
then, in turn, via the second steel‐to‐timber connection into the header. For joist hangers
with relatively few fasteners in the joist connection, most of the load is transmitted via
contact with the base plate. When designing a joist hanger, however, it is safe to assume
that the force will act at the centre of gravity of the joist connection. However, if the
transmission of force is mainly by contact with the base plate, the resulting force shifts
away from the header. The joint between the joist hanger and header is accordingly
stressed by an eccentric force, which exerts both lateral and axial loads on the fasteners.
Joist hangers with straps fixed onto the upper side of the header, tend to have fewer
fasteners at the header and joist level. In this type of joist hanger, the load is normally
transmitted via contact into the header. The load‐bearing capacity of such joist hangers is
almost impossible to calculate, hence the need for tests to determine the same. Unlike
the joist hanger shown in Figure E8‐3, tensile failure of the steel may also be a key crite‐
rion dictating load‐bearing capacity, particularly for steel less than 2 mm thick.
420
Cold‐formed steel connectors
Under normal circumstances, the nail distances to the end of the joist do not reach the
minimum values defined in EC 5. For this reason, a lower load‐bearing capacity would
have to be assumed for these nails. As soon as the nails deform under load, however, an
increasing portion of the load is transmitted to the base plate of the joist hanger via
contact pressure. Riberholt (1975) introduced a mechanical model, with which the por‐
tion of the load in the base plate can be estimated. Since the load‐bearing behaviour of
the nailed joint and of the support on the base plate can be seen as plastic, the steel‐to‐
timber joint capacity and of the contact pressure can be added.
E8.3 Joist anchors
Joist anchors are used to connect timber members arranged at right angles to each other,
e.g. to transfer wind suction forces or suspend beams. Most connections feature two diag‐
onally arranged joist anchors, although joist anchors with one and two rows of holes to
accommodate nails both exist. Joist anchors are generally only exposed to tensile forces.
When designing joist anchors, three different components must be considered:
The load‐bearing capacity of the nailed joint. Assuming that the force is exerted
on the corner of the joist anchor, the nailed joint is subject to eccentric load
(Figure E8‐4).
The load‐bearing capacity of the joist anchors in the net cross‐section, which is
also eccentrically loaded.
The load‐bearing capacity of the timber elements. Any stresses perpendicular
to the grain can be verified using the design methods shown in Article E11.
e
Fd
Figure E8‐4 Tension load on a joist anchor. (STEP 1995 Article C13)
421
Cold‐formed steel connectors
E8.4 Literature
E. Gehri, original Article C13, STEP 1995.
Ehlbeck J. and Görlacher R. (1982). Mindestnagelabstände bei Stahlblech‐Holznagelung. Forschungsbericht,
Versuchsanstalt für Stahl, Holz und Steine, Universität Karlsruhe.
Riberholt H. (1975). Berechnung von Stahlblech‐Holz Verbindungsteilen in Dänemark. Bauen mit Holz
77:534‐536.
422
E9 Contact joints
(Carpentry joints)
Original article: J. Ehlbeck, M. Kromer
A timber construction that usually comprises bar‐shaped members obtained from the
tree stem is only considered effective, if the individual elements can be assembled to‐
gether into a whole appropriately. Joints transmit internal forces generated by the exter‐
nal loads from one element to the other and two or more individual elements are united
at the junctions of the construction. These joints are frequently contact joints (carpentry
joints) working in compression, within which the forces to be absorbed are only transmit‐
ted by contact and, where applicable, friction. Certain joints also feature additional wooden
or metal fasteners, intended to secure items into position, but may also contribute to the
process of force transmission. Despite the numerous variations of traditional contact joints,
they can all be traced back to specific basic forms. These forms, namely half‐lap joint, step
joint, mortise and tenon joint and cogged joint, are shown in Figure E9‐1.
The specified basic forms for carpentry joints and variations thereof are used as longitu‐
dinal joints to extend timber elements in a direction parallel to the grain or as transverse
joints; featuring bars at right angles or in a criss‐cross pattern. Although carpentry joints
are not regulated in EC 5, the NA to EC 5 (NCI NA.12) includes design equations and min‐
imum dimensions for step, tenon and peg joints.
(a) (c)
)
(d)
(b)
Figure E9‐1 Basic forms: (a) half‐lap joint, (b) step joint, (c) mortise and tenon joint,
(d) cogged joint. (STEP 1995 Article C12)
423
Contact joints (Carpentry joints)
E9.1 Step joints
A step joint is a connection designed to transmit compressive forces of inclined struts in
timber chords. The compressive force of the strut is transmitted to the front surface of
the joint by contact and then via shear stress in the loaded end to the chord. Previously,
an additional tenon was often used to secure things in place, but today this is normally
done using a rafter nail, wood screws or laterally nailed‐on straps. The main timber engi‐
neering applications are the simple step joints shown in Figure E9‐2, particularly single
step and heel joints as well as both forms combined in the form of a double step joint.
β β β
tv tv tv tv
1 2
lv lv lv
1
lv
2
(a) (b) (c)
Figure E9‐2 Implementation options for typical step joints: (a) single step ‐ notch in front, (b) notch in heel,
(c) double step ‐ combination of both. (STEP 1995 Article C12)
When dimensioning a step joint, evidence is initially required to show that the load‐
bearing capacity of the existing contact surfaces suffices. For normal strut inclinations of
30 to 60°, only the front face is taken into consideration for the transmission of compres‐
sive forces and not the rear. The size of the front surface can be determined from the
member width b and the notch depth tv. The design stresses c,,d, generated by the
compressive force Fc,,Ed, act in the front surface A of the joint at an angle to the grain
direction and must meet the following condition:
c,α,d
1 (E9‐1)
f c,α,d
where
F c,α,Ed
c,α,d (E9‐2)
A
424
Contact joints (Carpentry joints)
and
f c,0,d
f c,α,d (E9‐3)
2 2
f c,0,c f c,0,d
sin cos sin cos
4 2 2 4
2 f c,90,d 2 f v,d
Equation (E9‐3) shows that the resistance depends on compressive strengths parallel and
perpendicular to the grain as well as the shear strength. The joint, which normally fea‐
tures the front surface at an angle to the chord (and strut), generates a combination of all
three stress components (origin of equation (E9‐3) see Annex 2, equation (5)). To maxim‐
ise the compressive force transmitted via a single step joint, the inclination of the front
surface should be selected such as to ensure that it corresponds to half the angle be‐
tween the strut and timber chord. If this is done, it is possible to ensure that the angle
between the force and grain directions of the wood is minimised; in both strut and chord
alike ( = /2), see Figure E9‐3.
E /2 E
E /2
Figure E9‐3 Definition of angles in a single step joint.
It is also advisable to ensure compliance with the following condition when selecting the
notch depths tv:
h 4 for 50
tv (E9‐4)
h 6 for 60
The shear stresses in the chord can be assumed to be uniformly distributed, but loaded
end lengths > 8 ∙ tv must not be taken into consideration when determining the shear
area. For larger loaded end lengths, the distribution of shear stresses becomes increas‐
ingly uneven (peaking at the notch end), and expanding the shear area to a greater ex‐
tent would lead to unrealistically high load‐bearing capacities. In addition, the need for
an effective width bef in accordance with equation (D1‐17) must be taken into account,
given the potential impact of cracks on shear strength. Step joints must also be secured
against unintentional movement.
425
Contact joints (Carpentry joints)
It should also be noted that the transmission of eccentric forces in the strut may gener‐
ate additional bending moments, which must be taken into account during design. Like‐
wise, the load‐bearing capacity of the chord, which is subject to tensile and bending
moment stresses, must be verified in the net cross‐section.
Multiple step joint
It seems advisable to modify the existing joint geometry (single step and heel joint), to
ensure the timber cross‐sections can be used as efficiently as possible. Using a multiple
step joint means the complete depth of the strut can be used to accommodate all the
forces exerted. The interlocking between the strut and chord is achieved using multiple
heel joints. A multiple step joint is illustrated in Figure E9‐4 and is characterised by a low
notch depth and as many heels as possible. The maximum heel number depends on the
strut depth, notch depth and the angle between strut and chord, since the multiple step
joint also requires an appropriate incline of the front surface, which should correspond to
half the angle between the strut and timber chord.
Figure E9‐4 Multiple step joint. (Blass and Enders‐Comberg, 2012)
Comprehensive experimental tests showed that while the multiple step joint has similar
load‐bearing capacities to the double step joint combining a single step and heel joint,
only a third of the notch depth is required, which also helps mitigate any cross‐sectional
weakening of the chord (Blass and Enders‐Comberg, 2012). Using hardwood (e.g. ash or
beech) for the edge lamellae of the chord can significantly boost the load‐bearing capacity
and stiffness of the joint.
426
Contact joints (Carpentry joints)
Design proposal
Using the multiple step joint focuses the resulting compressive force into the centre of
the strut, meaning no additional moments are generated. When multiple step joints with
the maximum number of heels were tested, this usually resulted in compression failure
perpendicular to the grain in the chord (Blass and Enders‐Comberg, 2012). Accordingly,
for angles ≥ 45°, in the ultimate limit state design, compression failure perpendicular to
the grain is assumed. With this in mind, the contact area calculated for the chord is ex‐
panded by 30 mm to the left and right. Since hybrid‐glulam chords are prone to compres‐
sion failure perpendicular to the grain in the softwood lamella underneath that made of
beech, the compression strength perpendicular to the grain is determined in the area of
transition from beech to softwood. Here, a load distribution angle of 45° is assumed,
allowing for additional expansion of the cross‐sectional area. The bending resistance of
the beech lamella is not taken into consideration, but contingent on a sufficient loaded
end length, so that compression failure perpendicular to the grain in the softwood is
always decisive. Otherwise, the loaded end should be checked for possible shear failure.
The following design proposal is included in Enders‐Comberg and Blass (2014) and the
corresponding geometric details can be taken from Figure E9‐5.
b
hs
tv
Lc / Lv
Figure E9‐5 Multiple step joint, geometric details (Blass and Enders‐Comberg, 2014).
Verification of compression capacity perpendicular to the grain Fc,90,R:
A c f c,90 k c,90
F c,90,R (E9‐5)
sin
where the length of the compression area is Lc
hs
Lc (E9‐6)
sin
and the compression area is Ac
A c b L c 2 30 mm (E9‐7)
427
Contact joints (Carpentry joints)
Verification of shear capacity Fv,R:
Av f v
F v,R (E9‐8)
cos
where the length of the shear area is Lv
hs
Lv (E9‐9)
sin
and the shear area is Av
Av bLv (E9‐10)
E9.2 Mortise and tenon joints
Mortise and tenon joints are used to connect individual members in floor, wall and roof
constructions where the joint has to transfer shear forces. Nowadays, economic reasons
mean it is only feasible to keep producing tenon joints using CNC‐controlled machining
equipment, which normally means straight tenons with central tenon position, low posi‐
tioned tenons where the underside of the member runs through or an intermediate form
with the tenon position varied. In this respect, central tenons are used to connect timber
members of the same sectional height, while variable and low positioned tenons are used
for mortises in trimmer joists or joists with higher cross‐sections. Figure E9‐6 shows a
typical tenon with its geometric details. For manually processed joints, the tenon height
hz amounts to at least a third of the member depth h, while for newer constructions,
although the tenon height tends to be in line with the latest machining equipment, it
must be at least h/6. The minimum and maximum dimensions for tenon lengths ℓz are 15
and 60 mm according to NA and the entire length ℓz of the tenon must bear on the mor‐
tise. Mortises should only be arranged centrally or in the compressive bending stress area
of the member. The cross‐sectional weaknesses caused by mortises must be taken into
consideration at the time of design, while the rule for the members with the tenon
means that ho must be at least as large as hu. Further requirements include the lower
notch hu not exceeding h/3 and the members with the tenon requiring a height‐width
ratio of 1.5 ≤ h/b ≤ 2.5. Figure E9‐7 sets out the complex geometric boundary conditions
in graphical form. The background to these geometric boundary conditions is the fact
that the calculation model was validated using test specimens, with values within these
limits. Any extrapolation may lead to non‐conservative results.
428
Contact joints (Carpentry joints)
hz/h ≥ 1/6
ho ℓz hu/h ≤ 1/3
he ho ≥ hu
h ho + hz ≤ h
hz 15 mm ≤ ℓz ≤ 60 mm
1.5 ≤ h/b ≤ 2.5
hu V
x ≤ 0.4 · h
b
Figure E9‐6 Tenon joint.
1.0
0.8 hz /h ≥ 1/6
0.4
hu /h ≤ 1/3
0.2 ho ≥ hu
0.0
0.0 0.2 0.4 0.6 0.8 1.0
hz/h
Figure E9‐7 Geometric boundary conditions for tenons.
The design of tenon joints can be performed similarly to that for notched beams (Equa‐
tion (D5‐9) in Article D5), although the compression strength perpendicular to the grain
fc,90 must also be verified. The characteristic load‐bearing capacity FRk of the tenon with a
tenon width b is therefore determined as follows (dimensions in Figure E9‐6), where
h ≤ 300 mm:
2
b h k k f
FRk min 3 ef e z v v,k (E9‐11)
1.7 b z,ef f c,90,k
The upper term in equation (E9‐11) verifies tension perpendicular to the grain and has
been reduced to shear stress verification. This verification equates to the failure mode
“crack propagation in the grain direction” (for more background details, see Article D5).
The effective width bef in accordance with equation (D1‐17) takes into consideration the
influence of cracks on shear strength, while the factor kz is based on test results and
takes influences from the tenon joint geometry into consideration:
hz h
2
h
kz 1 2 1 z 2 e (E9‐12)
he
h e h
429
Contact joints (Carpentry joints)
The coefficient kv takes into consideration the risk of a crack developing and reduces the
shear strength. kv must be less than 1 and corresponds to the coefficient kv for notched
beams from equation (D5‐10).
The lower term, conversely, represents a verification of compression perpendicular to
the grain and is derived from test results, the pre‐exponential factor 1.7 results from the
adaptation to the test results. The pre‐exponential factor corresponds to the coefficient
kc,90, which is applied when verifying compression perpendicular to the grain (see Article
D1, equation (D1‐4)) and takes into consideration, among other things, the fact that the
compression perpendicular to the grain is distributed over areas outside the directly
loaded partial area, whereby the compression strength perpendicular to the grain rises
(see also Figure D1‐3). The load distribution beyond the tenon length ℓz is ℓz also deter‐
mined via the effective tenon length ℓz,ef (here too, similar to the verification of compres‐
sion perpendicular to the grain, effective area Aef in equation (D1‐5)):
30 mm
z,ef min z (E9‐13)
2 z
In mortise and tenon joints, the members with the tenon may fail due to tension perpen‐
dicular to the grain, similarly to a notched beam, and due to compression perpendicular
to the grain in the supported surface of the tenon and must accordingly be verified
(equation (E9‐11)). In mortise and tenon joints in floor constructions, however, the
member with the mortise is also subject to tension perpendicular to the grain, since the
tenon joint works like a joint loaded perpendicular to the grain in this case, meaning the
member with the mortise is loaded perpendicular to the grain direction. The verifications
required in this case are explained in Article E11.
430
Contact joints (Carpentry joints)
E9.3 Dovetail joint
The emergence of computer‐controlled systems (CNC machines) for use in fabricating
engineered timber components means that even what were originally traditional car‐
pentry joints are coming back into vogue. The joint referred to here, namely a dovetail
joint, is regulated in a general national technical approval for a joint between headers
and joists in service classes 1 and 2 (Z‐9.1‐649), Figure E9‐8. The joist failure in the same,
caused by a combination of tension failure perpendicular to the grain and shear failure in
the area underneath the dovetail tenon, resembles the failure of a notched beam;
whereby the load‐bearing capacity of the header is mainly limited by the tension strength
perpendicular to the grain, like the behaviour of a joint loaded perpendicular to the grain.
Test observations showed that the joist actually failed in a similar way to a notched sup‐
port, the header like a member with a joint loaded perpendicular to the grain. The posi‐
tion of the cracks due to tension perpendicular to the grain in the joist differs from that
for notched beams, the cracks are not at the tip of the notch, but rather at the transition
between the straight tenon and lower fillet. This applies similarly to the header, for which
cracks due to tension perpendicular to the grain also occur at the height of the transition
between the straight tenon and lower fillet. Accordingly, the design method for notched
beams or joints loaded perpendicular to the grain as included in the EC 5 and NA can be
used to underpin the design of dovetail joints connecting headers and joists, while for
equations, see also Z‐9.1‐649.
bN ℓz bH
bZ bZ tef tz
hZ
hN hZ hH
r r
E
E
J J
(a) (b) (c)
Figure E9‐8 Single‐sided dovetail joint, (a) 3D overview, (b) joist, (c) header.
431
Contact joints (Carpentry joints)
The design load‐bearing capacity F90,Rd of a dovetail joint subject to shear force acting in
insertion direction amounts to the following per joint:
18 h H h Z r
2
k ab h Z 6.5 t ef h H 0.8 f t,90,d
hZ r 2
hH
F 90,Rd min (E9‐14)
k v bN h Z r
f v,d
1.5
where (see also Figure E9‐8):
h H Header depth in mm, hH ≥ 140 mm
b H Header width in mm, bH ≥ 60 mm
hZ Tenon height parallel to the lateral surface of the header in mm
r Mortise radius parallel to the lateral surface of the header in mm,
15 mm ≤ r ≤ 60 mm
tef Effective joint depth in mm, tef = min (bH , 100 mm)
ft,90,d Design tension strength perpendicular to the grain (in Z‐9.1‐649: ft,90,k = 0.5 MPa)
kab Coefficient to consider single‐ or double‐sided joints:
kab = 1 for single‐sided joints
1
k ab min for double‐sided joints
b H 200
h N Joist depth in mm, 140 mm ≤ hN ≤ 280 mm
b N Joist width in mm, bN ≥ 60 mm
ℓZ Tenon length in mm, 25 mm ≤ ℓZ ≤ 30 mm
1
kn
kv min
h 1 0.4 z 1 2
N hN
Notch ratio, ≥ 0.4, = cos ∙ (hZ ‐ r)/hN,
Joist inclination in °
5 for solid softwood
kn
6.5 for glulam
fv,d Design shear strength (in Z‐9.1‐649: fv,k = 2.5 MPa)
bZ Tenon width where bZ ≥ 0.8 ∙ bN
Tenon cone angle where 4° ≤ ≤ 12°
Dovetail milling angle where 10° ≤ ≤ 18°
432
Contact joints (Carpentry joints)
The upper expression in equation (E9‐14) outlines the verification of the header and its
source resembles that shown in Article E11 for joints loaded perpendicular to the grain
(equation (E11‐35). The lower expression in equation (E9‐14) includes verification of the
joist and is along the lines of the verification outlined in Article D5 for non‐reinforced
notched beams (equation D5‐9).
E9.4 Peg joints
EC 5 does not include any details of peg joints, although the NA does contain design rules
for joints with solid oak pegs, which were derived from test results. The characteristic
load‐bearing capacity FRk in N of a peg loaded laterally and with a constant cross‐section
(normally round or octagonal with diameter d in mm) is determined as follows, where
20 mm ≤ d ≤ 30 mm (the only items checked were solid oak pegs between 20 and 30 mm
in diameter, which means equation (E9‐15) is only valid within these limits.):
2
FRk 9.5 d (E9‐15)
The density of the timber members to be connected must be at least 350 kg/m3, and the
angle between the force and grain directions exerts no impact on FRk. The minimum
wood thickness is 2 ∙ d, while the minimum distances (edge and end distance and spac‐
ing) are also 2 ∙ d.
E9.5 Literature
J. Ehlbeck, M. Kromer, original Article C12, STEP 1995.
Blass H.J. and Enders‐Comberg M. (2012). Fachwerkträger für den industriellen Holzbau. Karlsruher Berichte
zum Ingenieurholzbau, Band 22, KIT Scientific Publishing Karlsruhe, 161 S.
Enders‐Comberg M. and Blass H.J (2014). Treppenversatz – Leistungsfähiger Kontaktanschluss für Druckstäbe.
Der Bauingenieur 89(4):162‐171.
433
E10 Glued‐in steel rods
Original article: C. J. Johansson
Joints with glued‐in rods are mainly used in structures using glued‐laminated timber or
laminated veneer lumber, some examples of which are shown in Figure E10‐1 and Fig‐
ure E10‐2. The rods are used to prevent cracks in curved and notched beams due to
tensile stresses perpendicular to the grain or transmit forces in a construction or portion
thereof, e.g. when connecting a column to a foundation or a corner of a portal frame.
Glued‐in steel rods are also used for rehabilitation and repair, during which they may be
subject to lateral or axial load or both combined. The manufacturer must also possess a
proof of suitability for gluing load‐bearing timber members (DIN 1052‐10). Joints with
glued‐in steel rods are not included in EC 5, although the design and execution are regu‐
lated in the NA (NCI NA.11). The advantages of using glued‐in rods are:
Scope to transmit significant concentrated loads,
Scope to arrange rods in parallel with the grain,
Very rigid joint when exposed to axial load,
Effective fire resistance, since the surrounding wood protects the steel.
Figure E10‐1 Glued‐in rods to avoid cracks (pitched cambered beam, opening, notch). (STEP 1995 Article C14)
Figure E10‐2 Glued‐in rods in a clamped column end joint and in a moment‐resisting joint; a filled with mortar,
b steel fitting. (STEP 1995 Article C14)
435
Glued‐in steel rods
Instead of glued‐in rods, steel rods with a thread in accordance with DIN 7998 are an
increasingly popular choice for reinforcements perpendicular to the grain direction;
driven into predrilled holes like screws and hence mechanically connected with the wood
via the thread. Glued‐in steel rods, however, are best‐suited for arrangements parallel to
the grain. Reinforced joints are introduced in Article E12.
E10.1 Materials and manufacture
Ensuring reliable adhesion between a smooth steel surface and the adhesive is far from
easy. Accordingly, to ensure mechanical interlocking between the rod and adhesive,
threaded rods or RC rebars are preferred. All statements of the NA thus only apply for
joints with threaded rods or rebars. Generally speaking, rods of diameter between 12 and
24 mm are used. The rods are glued‐in, either by injecting adhesive (Figure E10‐3) or
pouring the adhesive into the hole and inserting the rod. The drilled holes are around 1
to 4 mm larger than the external diameter of the thread, to ensure sufficient room in
which to inject the adhesive.
a
d b c
Figure E10‐3 Injection of a joint with a glued‐in rod; a sealing, b adhesive bleed, c adhesive ingress, d a block
prevents the rod from being pushed out by the adhesive. (STEP 1995 Article C14)
Different adhesives are used, such as phenol resorcinol formaldehyde resin, dual‐
component polyurethanes and dual‐component epoxy resins. The adhesive selection
depends on the manufacturing method and the type of load to which the rods are sub‐
ject. In Germany the adhesive requires a general national technical approval for use with
glued‐in steel rods. For hole diameters that exceed the rod diameter, the key factors
determining the withdrawal strength of the rod are the strength and durability of the
bond line.
Phenol resorcinol formaldehyde resins are a tried and tested option when it comes to
structural timber, particularly e.g. glulam. Riberholt (1988), however, advocates that
phenol resorcinol resins should not be used for joints manufactured using the injection
method and holes of excess size, since the shrinking of the adhesive when it hardens
reduces its strength. Dual‐component polyurethane adhesives, conversely, frequently
react sensitively to higher temperatures (Aicher, 1992) and should therefore only be used
436
Glued‐in steel rods
in this respect if a general technical approval has been issued. This is particularly true for
joints which were established using the injection method and which are permanently
subject to high loads.
For laterally loaded glued‐in rods, the choice of adhesive is less important, since the loads
are mainly transmitted via compressive – and at times tensile – stresses in the bond line
and the mechanical behaviour resembles that of laterally loaded dowel‐type fasteners.
E10.2 Laterally loaded glued‐in rods
The lateral load‐bearing capacity of rods can be determined using the Johansen equa‐
tions for dowel‐type fasteners (expanded to take the rope effect into consideration), cf.
Articles E1 and E2. The provisions for bolts apply, including the steel rods with threads in
accordance with DIN 7998 (not glued‐in), Article E4.
Gluing‐in the rods enables virtually infinite friction coefficients between steel and timber
surfaces. Meanwhile, for rods glued‐in perpendicular to the grain, Rodd et al. (1989)
successfully demonstrated a considerable increase in load‐bearing capacity and stiffness
in comparison to non‐glued‐in rods. This also means that the embedment strength fh,,k
determined with Equation (E4‐2) can be increased by a factor of 1.25, if the steel rods are
glued‐in perpendicular to the grain. For rods glued‐in in parallel to the grain, the embed‐
ment strength is very low and only 10% of the value for rods glued‐in perpendicular to
the grain. For angles between the force and grain directions of between 0 and 90°, the
embedment strength can be determined by linear interpolation. The minimum distances
for glued‐in steel rods are specified in Table E10‐1.
Table E10‐1 Minimum distances for laterally loaded glued‐in steel rods in accordance with NCI NA.11 (2013).
Minimum distances
Steel rods glued‐in in parallel to the grain:
a2 = spacing, a2 = 5 ∙ d
a2,c/t = loaded/unloaded end distance a2,c = 2.5 ∙ d
a2,t = 4 ∙ d
Steel rods glued‐in perpendicular to the grain See dowels, Table E4‐2
437
Glued‐in steel rods
E10.3 Axially loaded glued‐in rods
Axially loaded glued‐in rods may fail in three different modes:
Tensile failure of the steel rod,
Failure of the bond line or the wood along the wall of the hole,
Failure of the timber member.
If multiple glued‐in steel rods of a joint are not uniformly stressed, the failure mode “ten‐
sile failure of the steel rod” has to become the key criterion. The plastic deformation of
the most heavily loaded steel rod then makes it possible to redistribute load among the
individual steel rods in the joint. Non‐uniform tensile stresses in a group of steel rods can
occur, e.g. due to inconsistent tightening of nuts or deformations in connecting steel
parts. The latter failure type, of the timber member, can be checked by verifying the
tensile stress of the member in the net cross‐section or block shear. Both the other fail‐
ure types are involved when determining the design withdrawal resistance Fax,Rd for a
glued‐in rod:
A ef f y,d
F ax,Rd min (E10‐1)
d ad f k1,d
The load‐bearing capacity for the tensile failure of the steel rod is determined via its
effective cross‐section Aef and its yield strength fy,d. The design of the bond line, mean‐
while, resembles that of reinforcements with strength values for bond lines, see Article
E12. ℓad is the glued‐in length of the steel rod, d the diameter and fk1,d the design strength
for the bond line, the characteristic value of which is given in Table D8‐1 (Article D8).
The required minimum distances for axially loaded rods are given in Table E10‐2. Under
specific boundary conditions, there is also scope to use lower minimum distances for
steel rods glued‐in in parallel to the grain in glulam (see general national technical ap‐
proval No. Z‐9.1‐791).
The minimum glued‐in length ℓad,min for axially loaded rods is:
0.5 d
2
ad,min max (E10‐2)
10 d
438
Glued‐in steel rods
The minimum glued‐in lengths should always be complied with, since tests have shown
that shorter glued‐in lengths are prone to brittle failure. Moreover, for multiple axially
loaded glued‐in steel rods, group effects may emerge, leading to a so‐called “group tear‐
out” resembling a block shear failure and which lead to the group of fasteners being
completely pulled out. There are no relevant indications contained in the NA. For groups
of glued‐in steel rods, therefore, there is a need to identify an effective number nef = n0.9
as is done for axially loaded screw groups, if plastic steel failure is not the key criterion
(for more details, see Article E13). Glued‐in steel rods can also be used for joints loaded
perpendicular to the grain, in which case the tensile stresses perpendicular to the grain
must also be verified, as is explained in Article E11. Also important for axially loaded rods
is preventing corrosion of the rods. Riberholt (1988) highlights the risk of seeing the bond
line between the steel and wood destroyed due to the spread of rust, which is why glued‐
in rods must be protected by applying e.g. a zinc coating against corrosion. Riberholt
(1986) stated that certain adhesives, such as epoxy resins, already provide effective cor‐
rosion protection.
Table E10‐2 Minimum distances of axially loaded glued‐in steel rods in accordance with NCI NA.11 (2013).
Minimum distances
Steel rods glued‐in parallel to the grain:
Spacing a2 = 5 ∙ d
Edge distance a2,c = 2.5 ∙ d
Steel rods glued‐in perpendicular to the grain:
Spacing in beam width a1 = 4 ∙ d
Edge distance in beam width a1,c = 2.5 ∙ d
Spacing in beam length a2 = 4 ∙ d
End distance in beam length a2,c = 2.5 ∙ d
E10.4 Literature
C.J. Johansson, original Article C14, STEP 1995.
Aicher S. (1992). Testing of adhesives for bonded wood‐steel joints. Meeting IUFRO S 5.02, Bordeaux.
Riberholt H. (1986). Glued bolts in glulam. Report No. R 210, Technical University of Denmark.
Riberholt H. (1988). Glued bolts in glulam: Proposal for CIB Code. Paper 21‐7‐2, CIB‐W18 Meeting 21, Parksville.
Rodd P.D., Hilson B.O. and Spriggs R.A. (1989). Resin injected mechanically fastened timber joints.
Second Pacific Timber Engineering Conference, Auckland.
439
E11 Joints loaded perpendicular
to the grain
Original article: J. Ehlbeck, R. Görlacher
The term “joints loaded perpendicular to the grain” is used to refer to load introductions
into a member with a force component at right angles to the member’s grain direction.
The load‐bearing capacity of a joint loaded perpendicular to the grain is firstly limited by
the load‐bearing capacity of the joint itself (with various fasteners such as nails, screws,
bolts, glued‐in rods, punched metal plate fasteners or connectors), secondly when the
tensile strength perpendicular to the grain of the timber member has been attained.
Regardless of the fastener types used, however, joints loaded perpendicular to the grain
generate tensile stresses perpendicular to the grain when introducing shear forces into
the member. In timber construction, tensile stresses perpendicular to the grain are gen‐
erally considered particularly serious, given the comparatively low strength and stiffness
of the wood perpendicular to the grain direction. This means, for example, while the
characteristic tensile strength parallel to the grain is 14 N/mm² for normal coniferous
timber, the characteristic tensile strength perpendicular to the grain for the same is just
0.4 N/mm², namely less than 3% of the tensile strength parallel to the grain. In addition,
wood is excessively brittle when tensile stress is applied at right angles to its grain direc‐
tion. If the tensile strength perpendicular to the grain has been attained for a joint loaded
perpendicular to the grain, this will lead to the wood splitting in the grain direction and
ultimately to the failure of the joint. Despite all these issues, which can occur when load
introductions are exerted perpendicular to the grain, joints loaded perpendicular to the
grain are still frequently used in timber construction, whereby the tensile stresses gener‐
ated perpendicular to the grain direction may be the key factor dictating the load‐bearing
capacity of the joint, regardless of the fastener. Some examples of typical joints loaded
perpendicular to the grain with possible crack paths are shown in Figure E11‐1.
441
Joints loaded perpendicular to the grain
The key tensile stresses perpendicular to the grain, which generally overlap with shear
and longitudinal stresses, can be determined using the finite element method. Similar to
the approach used to calculate notched beams, this results in stresses not directly com‐
parable with the characteristic tension strength perpendicular to the grain, which is de‐
termined using prismatic test bodies. Accordingly, this is also an area in which fracture
mechanics may prove useful, as was already described in relation to notched beams in
EC 5 (Article D5).
Additional possibilities for handling issues with tensile stress perpendicular to the grain in
joints include conventional stress criteria, based on the concept of an effective width.
This approach is contained in the NA for joints loaded perpendicular to the grain with
multiple fasteners and as with notches and openings (Article D5), the design of joints
loaded perpendicular to the grain has seen fracture mechanics approaches become
standard practice in addition to the more traditional stress criteria.
E11.1 Notes on reducing the risk of failure due to tension
perpendicular to the grain
The following section outlines the key parameters influencing the load‐bearing capacity
of joints loaded perpendicular to the grain, which also includes structural detailing hints
on how to reduce the risk of cracks due to tension perpendicular to the grain. Infor‐
mation on reinforced joints, meanwhile, is included in Article E12, while a joint loaded
perpendicular to the grain with mechanical fasteners is schematically illustrated in Fig‐
ure E11‐2. The force F90 generating tension perpendicular to the grain is introduced via
fasteners into a beam.
442
Joints loaded perpendicular to the grain
alrr l l
h1 h
2 hi
hn
h
hbee
bt
F90 F90
F90
2 2
Figure E11‐2 Joint perpendicular to the grain (designations). (STEP 1995 Article C2)
The following parameters influence the load‐bearing capacity:
The ratio of the distance he of the row of fasteners, which is furthest away from
the loaded edge and the beam depth h. To reduce the risk of cracks due to ten‐
sion perpendicular to the grain, the fasteners should be as close as possible to the
unloaded edge, namely he/h should be maximised.
Multiple fasteners in a row distribute the load over a larger joint area and reduce
the maximum value of the tensile stresses perpendicular to the grain generated as
a result. This influence is particularly helpful for larger joint widths or fasteners
spaced relatively far apart.
Expanding the beam width b or beam depth h (if he is increased) boosts the load‐
bearing capacity. Here, it is important to note that depending on the fastener type
used, it is frequently only possible to employ part of the beam width to accom‐
modate the tensile stresses perpendicular to the grain generated.
Additional rows of fasteners reduce tensile stresses perpendicular to the grain in
the uppermost row of fasteners, if ar or he becomes larger.
The incidence of brittle failure means the tensile strength perpendicular to the
grain of the wood depends on the stressed volume, which means the load‐bearing
capacity of joints loaded perpendicular to the grain is also affected by beam size.
443
Joints loaded perpendicular to the grain
E11.2 Verification procedure based on fracture mechanics
The verification for joints loaded perpendicular to the grain is conducted in a manner
resembling that for notched beams (Article D5), which is based on fracture mechanics.
Based on linear‐elastic fracture mechanics and the assessment of test results, the follow‐
ing verification was recorded for softwood subject to loads as shown in Figure E11‐3 in
EC 5 (Van der Put and Leijten, 2000; Leijten and Jorissen, 2001):
he
F 90,Rk 14 b w (E11‐1)
he
1 h
The following applies:
F90,Rk Characteristic tensile capacity perpendicular to the grain in N
b Beam width in mm
he Distance of the fasteners arranged furthest away from the loaded edge in mm
h Beam depth in mm
w 0.35
pl
max 100 for punched metal plate fasteners
w
1
1 for all other joints
where wpl = width of the punched metal plate fastener parallel to the grain in mm
FFEd FFEd
hbee
FEd,1V FVEd,2 h
1 2
b
t
Figure E11‐3 Load exerted at an angle to the grain direction. (STEP 1995 Article C2)
444
Joints loaded perpendicular to the grain
The basic approach used to derive equation (E11‐1) can be taken from Article D5; only
certain general notes will be made at this point. The (already simplified) original form of
equation (E11‐1) reads as follows (Van der Put and Leijten, 2000):
G G c he
F 90 b (E11‐2)
0.6 he
1
h
The root terms in equations (E11‐1) and (E11‐2) are typical of linear‐elastic fracture me‐
chanics approaches and equation (E11‐1) already includes key linear‐elastic fracture
mechanics parameters, such as the value of critical fracture energy Gc. This is why equa‐
tion (E11‐1) only applies for softwoods, since the only results obtained came from tests
with softwood. Material properties, such as tensile strength perpendicular to the grain
ft,90 for example, are thus excluded from equation (E11‐1) and only geometric infor‐
mation is present.
One issue with the abovementioned verification is the fact that the applicable scope is
limited to the joint configuration shown in Figure E11‐3. For joints in practice, however,
most configurations include multiple rows of fasteners. Configurations other than that
shown in Figure E11‐3 are used, since one advantage of distributing fasteners over a longer
length in the grain direction is reducing the peak tensile stresses perpendicular to the grain
and hence increasing the load‐bearing capacity. In such cases, the verification shown in
equation (E11‐1) does not apply. To remedy this loophole, the NA uses the verification
already specified in DIN 1052:2008 for the configuration of choice and various wood spe‐
cies and types of wood‐based materials, which is derived in the following section. This
verification, as has already been noted, is based on the known concept of a stress criterion.
E11.3 Verification procedure in accordance with NCI NA.8.1.4
The stress verification in accordance with NCI NA.8.1.4 is based on theoretical and experi‐
mental examinations conducted by Ehlbeck and Görlacher (1983). A joint loaded perpen‐
dicular to the grain is dimensioned by comparing the calculated value of the tensile stress
perpendicular to the grain with the characteristic tensile strength perpendicular to the
grain; taking into consideration all impacts on load‐bearing capacity. The failure behaviour
of the wood when tensile load is applied perpendicular to the grain is particularly brittle,
which means expanding the volume subject to tensile stress adversely affects strength. The
tensile strength perpendicular to the grain and its characteristic value are determined by
tests on comparatively small samples, which means recording the volume effect is not
possible. This is a factor that must be taken into consideration, alongside the joint type and
geometry when designing joints loaded perpendicular to the grain.
445
Joints loaded perpendicular to the grain
The load generating tension perpendicular to the grain as a portion of the total load
The design of a joint loaded perpendicular to the grain with respect to the tensile
strength perpendicular to the grain of the member depends on the extent of the load
generating tension perpendicular to the grain in the member. Since the load from such
joint is introduced in the member by both compression as well as tension perpendicular
to the grain, the first task involves defining the load component generating tension per‐
pendicular to the grain. Accordingly, an equation to define this load as a portion of the
total load from a joint is derived in the following section using the example of a single‐
span girder. The girder includes the span ℓ, depth h and width b and is subject to stress
from the concentrated load F at mid‐span. Now, the position of the load application point
should be observed. If the load F is exerted on the top side of the beam, this will gener‐
ate exclusively compression perpendicular to the grain in the beam. Conversely, if the
load is applied to the bottom of the beam, only tension perpendicular to the grain is
generated, as in Figure E11‐4.
We cite the general case, in which the load F is applied at a point with an arbitrary dis‐
tance a to the underside of the beam, as in Figure E11‐5. The load F now generates both
compression and tension perpendicular to the grain. The load portion generating tension
perpendicular to the grain Ft is designated with . The factor is 1 if the load is exerted
on the underside of the beam.
If the load is applied to the upper edge of the beam, = 0 applies:
Ft F (E11‐3)
F c 1 F (E11‐4)
h F
ℓ/2 ℓ/2
Figure E11‐4 Single‐span girder with load applied to the upper or lower edge.
446
Joints loaded perpendicular to the grain
h1 F
h
a
ℓ/2 ℓ/2
Figure E11‐5 Load application at an arbitrary point at mid‐span.
The factor for load introduction between the upper and lower edges of the beam is not
initially known and has to be determined. For this purpose, the beam is cut at the height
of the load application point, namely at the distance h1 from the upper edge of the beam,
see free‐body diagram in Figure E11‐6. Subsequently, the deflection of the upper beam
section shown in Figure E11‐6 is determined, taking into consideration the internal forces,
tensile force Ft and shear stress , which impact on the cut surface. Given the condition
that the deflection of the cut beam portion must be equivalent to the deflection of the
entire beam, an equation to define the load portion generating tension perpendicular
to the grain can be derived.
Ft
W
ℓ/2 ℓ/2
First of all, certain base equations from Euler‐Bernoulli beam theory are required, so that
the internal stresses and moments , MFt and M and resulting individual deflection com‐
ponents can be defined. Shear stresses are calculated as:
V S
I b
The shear force amounts to:
F
V
2
447
Joints loaded perpendicular to the grain
The first moment of area of the cut cross‐sectional part with cross‐sectional area a ∙ b
amounts to:
h/2
h1
h a h a a h
S a be a b ab 1 ab 1 e h/2
2 2 2 2 2 a
The second moment of area for a rectangular cross‐section b x h is:
bh3
I
12
The above equations can be used to define the shear stress of the entire beam illustrated
in Figure E11‐6 at the height of the cut:
V S F h 12 3 F a h1
ab 1 (E11‐5)
I b 2 3
2 bh b bh3
The cut beam part of height h1 is subject to moments caused by the load component Ft
and the shear stress , Figure E11‐6. The moment MFt caused by the force Ft amounts to:
Ft
M Ft (E11‐6)
4
The moment Mcaused by shear stress results in:
h1 h
Mτ dA 1 b (E11‐7)
2 A 2 2
Inserting equation (E11‐5) into equation (E11‐7) allows the moment M to be expressed
in dependence of the force F:
h1 3 F a h1 3 F a h 12
Mτ b 3
3
(E11‐8)
2 2 bh 4h
448
Joints loaded perpendicular to the grain
With that, all the values required to determine the deflection of the beam are in place.
Both the load component Ft and the shear stress result in a deflection of the beam part
shown in Figure E11‐6. The deflection caused by the load component Ft results from
inserting Ft in the equation known from the Euler‐Bernoulli beam theory to determine
the deflection of a beam exposed to a concentrated load at mid‐span:
Ft 3 F 3
f Ft (E11‐9)
48 E I 1 48 E I 1
Inserting equation (E11‐6) into equation (E11‐9) means that the deflection may also be
be expressed in dependence of the moment MFt:
Ft 3 M 4 3 M 2
f Ft Ft Ft (E11‐10)
48 E I 1 48 E I 1 12 E I 1
The shape of the moment diagram caused by a concentrated load applied at mid‐span
corresponds to that of the shear stresses on the cut beam shown in Figure E11‐6. The
deflection f of the cut beam of the height h1 caused by these shear stresses can accord‐
ingly be determined by inserting (E11‐8):
2 2 2 3
M τ 3 F a h1 2 3 F a h1
fτ (E11‐11)
12 E I 1 4h 3
12 E I 1 48 h 3 E I 1
Adding deflection components fFt + f allows the following equation:
F 3 3 F a h 12 3 F 3 3 a h 12
f 1 f Ft f τ (E11‐12)
48 E I 1 3
48 h E I 1 48 E I 1 h3
The deflection of the beam at mid‐span caused by the concentrated load F is independ‐
ent of the distance a from the point at which the force is applied to the underside of the
beam. The deflection f1 for the cut beam under observation, in accordance with equation
(E11‐12), must therefore be equivalent to the total deflection f of the entire beam,
whereby the total deflection is determined without taking shear deformations into con‐
sideration:
F 3 F 3 3 a h 12
f f1 (E11‐13)
48 E I 48 E I 1 h3
449
Joints loaded perpendicular to the grain
After transforming:
I 3 a h 12 h3 3 a h 12
1 3 (E11‐14)
I 1 h3 h
1
h3
By solving for and replacing h – a with h1, equation (E11‐14) can be transformed as
follows:
h 13 3 a h 12 h a 3 3 a h a
2 3 2
h 3a h 2a
3
3
3
3
3
3
h h h h h
2 3
a a
1 3 2 (E11‐15)
h
h
The portion of load F generating tension perpendicular to the grain was determined
based on the distance a of the load application point from the underside of the beam and
the beam depth h. Equation (E11‐15) also applies for loads, which are not applied at mid‐
span as during derivation, only the coefficients change, but are cancelled again. Superim‐
posing concentrated loads allows any load distribution of choice to be attained. Equation
(E11‐15) is thus generally applicable, e.g. also for distributed loads.
Equation (E11‐15) is also used when there is a need to determine the governing tensile
stress components perpendicular to the grain of reinforced notches, Equation (D8‐4) in
Article D8.
Design based on a stress criterion
The stress state in a member is explained using a joint loaded perpendicular to the grain
and with multiple fasteners as an example. The volume subject to tension perpendicular
to the grain influences the stress distribution, which means the joint geometry must be
taken into consideration when determining the stress. Figure E11‐2 includes a definition
of the geometric impact variables for a double‐sided joint loaded perpendicular to the
grain. For simplicity, it is assumed that the load F90 to be transmitted is uniformly distrib‐
uted to all fasteners, although the actual load distribution is uneven. It varies depending
on the local strength and stiffness properties of the wood as well as deformation of the
members in the area of the joint. We now examine the stress distribution perpendicular
to the grain direction in one row of fasteners and in the member subject to tension per‐
pendicular to the grain (Figure E11‐7).
450
Joints loaded perpendicular to the grain
h1
Figure E11‐7 Distribution of tensile stresses perpendicular to the grain in multiple rows of fasteners.
The uppermost fastener in row (1) generates tension perpendicular to the grain 1(1)
after force F(1) is introduced into the member. The tensile stress perpendicular to the
grain peaks at the point at which the load is introduced and declines with increasing
distance from the fastener in a parabolic shape. The tensile stress perpendicular to the
grain is zero at the upper edge of the member, while the stresses perpendicular to the
grain introduced by the fasteners in rows (2) and (3) lead to the stress distributions
shown with the maxima 2(2) and 3(3). The respective stress distributions due to the
different rows of fasteners must now be superimposed. To the stress exerted at the
fastener position in row (1) caused by load F(1), the respective stresses at this position
caused by F(2) and F(3) from fasteners in rows (2) and (3) have to be added. For the rows
of fasteners underneath row (1), the stress components are also superimposed. Subse‐
quently, however, comes the need to take compression perpendicular to the grain into
consideration; that develop similar to tension perpendicular to the grain. The tensile and
compressive stresses perpendicular to the grain superpose, the resulting tensile stresses
perpendicular to the grain are therefore lower than the same in row (1). The key criteri‐
on for tensile stresses perpendicular to the grain of a joint loaded perpendicular to the
grain is thus always the fastener row furthest from the loaded member edge. The sum
of tensile stresses perpendicular to the grain of a joint loaded perpendicular to the grain
with n rows of fasteners in the area of a fastener in row (1) amounts to:
n
(1) i(1) 1(1) 2(1) 3(1) (E11‐16)
i1
451
Joints loaded perpendicular to the grain
The additional stress components from series (2) and (3) which are added to the tensile
stress perpendicular to the grain 1(1) are initially unknown and depend on the respective
stress distributions. The effective tensile stress components i(1) in the uppermost fas‐
tener row from row (i) are also smaller by factor (h1/hi)2 than the component 1(1) from
the first row of fasteners:
2
h1
i(1) 1(1) (E11‐17)
hi
Using this relationship, which is determined by numerical investigations, the sum of ten‐
sile stresses perpendicular to the grain in row (1) can also be determined:
2
n h1
ges
(1)
1
(1)
(E11‐18)
i1 hi
To calculate the overall stress in row (1) in accordance with equation (E11‐18), the size of
the load Ft(1), that is generating tension perpendicular to the grain and that is a portion of
the load F(1), is required, as well as the area subject to tension perpendicular to the grain.
The load portion Ft(1) introduced by the fasteners in row (1) can be calculated using equa‐
tions (E11‐3) and (E11‐15):
Ft F
(1) (1)
(E11‐19)
For joints loaded perpendicular to the grain with n rows of fasteners and assuming the
force F is uniformly distributed over the rows of fasteners (F(1) = F/n), the force Ft(1) of the
fasteners in row (1) can be calculated as:
2
(1) F n h
Ft 1 (E11‐20)
n i1 h i
452
Joints loaded perpendicular to the grain
ℓef
V (1)
he
ar
F90
Figure E11‐8 Distribution of tensile stresses perpendicular to the grain in longitudinal beam direction and
with effective width ℓef.
Now, the area exposed to tensile stress perpendicular to the grain is still missing. In the
longitudinal beam direction, i.e. parallel to the grain direction of the member subject to
tension perpendicular to the grain, however, the stresses are non‐linearly distributed and
decline with increasing distance from the joint (Figure E11‐8). Accordingly, the effective
area Aef subject to tension perpendicular to the grain must be known to determine the
tensile stresses perpendicular to the grain:
A ef ef t ef (E11‐21)
The integral of the stress distribution corresponds to the area of a fictional stress distri‐
bution with constant tensile stresses perpendicular to the grain (1) and effective length
ℓef, rectangular area in Figure E11‐8. The effective joint width ℓef depends on geometric
boundary conditions:
ef a r2 c h
2
(E11‐22)
The factor c was empirically determined using test results:
3
4 he h
c 1 e (E11‐23)
3 h h
The effective joint (penetration) depth tef depends on the respective fastener type.
Similarly to when the fasteners are arranged in multiple rows, distributing the fasteners
in the grain direction over an extended length ar serves to improve the load‐bearing
behaviour. However, this approach also increases the volume subject to tensile stresses,
which, in turn, adversely affects the load‐bearing capacity. Overall however, distributing
the fasteners over a wide area is considered beneficial.
453
Joints loaded perpendicular to the grain
According to equation (E11‐20), the key tensile stress perpendicular to the grain t,90 in
the fastener row placed furthest away from the loaded member edge is therefore (with
the abbreviation k1):
2
F (1) F 90 n h 1 F 90
t,90 k1 (E11‐24)
A ef n A ef i1 h i A ef
2
1 n h
k1 1 (E11‐25)
n i1 h i
The factor k1 considers the fact that, when using multiple rows of fasteners, the tensile
stresses perpendicular to the grain are lower in the row of fasteners furthest away from
the loaded member edge.
To verify the load‐bearing capacity of a joint loaded perpendicular to the grain, the ten‐
sile stress must be compared with the tension strength perpendicular to the grain. Since
tensile failure is classed as a brittle form of failure, the joint capacity cannot be readily
verified using a stress criterion. The influence of volume on the tension strength perpen‐
dicular to the grain is taken into consideration, by multiplying the tension strength,
benchmarked to a reference volume of 0.1 m³ and with a coefficient kvol:
F 90
t,90 k 1 f t,90 k vol (E11‐26)
A ef
The coefficient kvol depends on the size of the effective area Aef subject to tension per‐
pendicular to the grain in accordance with equation (E11‐21):
‐0.2 ‐0.2 2
k vol 13 A ef with A ef in mm (E11‐27)
Since joints loaded perpendicular to the grain always involve concentrated loads being
introduced in members, it is advisable to convert equation (E11‐26) to a load F90, which
then allows the load which the joint can withstand to be calculated directly.
0.8
f t,90 k vol A ef f t,90 A ef
F 90 13 (E11‐28)
k 1 k 1
454
Joints loaded perpendicular to the grain
The verification can also be expressed by inserting equation (E11‐22):
c 0.8 k2
F 90 13 t ef h
0.8
f t,90 (E11‐29)
k1
The factor k2 allows the impact of the number of adjacent fasteners on stress distribution
to be taken into consideration:
0.4
a 2
k 2 1 r (E11‐30)
c h
An approximation function allows the factor k2 to be expressed in simplified form:
ar
k 2 0.7 1.4 1 (E11‐31)
h
The expression c0.8/ in equation (E11‐29) depends solely on the quotient he/h and can
also be simplified by an approximation function:
2
c 0.8 h
6.5 18 e (E11‐32)
h
Equation (E11‐29) is transformed accordingly:
k2 h
2
6.5 18 e t ef h f t,90
0.8
F 90,max 13 (E11‐33)
k1
h
Equation (E11‐33) for the maximum shear load F90,max allows the load‐bearing capacity of
a joint loaded perpendicular to the grain to be calculated assuming the failure of the
member subject to tension perpendicular to the grain.
455
Joints loaded perpendicular to the grain
The design equation specified in the NA (NCI to 8.1.4) for joints loaded perpendicular to
the grain is based on equation (E11‐33) and reads as follows (for geometric details see
Figure E11‐2):
where
h
2
F 90,Rd k s k r 6.5 18 e t ef h f t,90
0.8
(E11‐35)
h
a
k s max 1; 0.7 1.4 r (E11‐36)
h
n
kr 2
(E11‐37)
n h
1
i1 h i
The following applies:
ks Coefficient, which takes into account an increase in the tension strength perpen‐
dicular to the grain due to multiple adjacently arranged fasteners within a fastener
group (corresponds to factor k2 in equation (E11‐33), see also equation (E11‐31)).
kr Coefficient kr considers the fact that for joints with multiple (n) rows of fasteners,
arranged on top of each other, lower tensile stresses perpendicular to the grain
are exerted in the fastener row furthest away from the loaded member edge (cor‐
responds to factor k1 in equation (E11‐33), see also equation (E11‐25)).
tef Effective joint (penetration) depth in mm; different values apply for single‐ or dou‐
ble‐sided cross joints as well as for different joint types, see NA.
The specified design of joints loaded perpendicular to the grain in equations (E11‐35) to
(E11‐37) must only be implemented for ratios he/h ≤ 0.7. In the NA, no verifications are
required for ratios he/h > 0.7, while the key criterion for determining the load‐bearing
capacity of joints with dowel‐type fasteners are the Johansen cases (with rope effect).
Joints with he/h < 0.2 may only be stressed by a short‐term load. Additional regulations
apply for joints with ar / h > 1 and Fv,Ed > 0.5 ∙ F90,Rd; which always require reinforcement.
In addition, rules apply for multiple groups of fasteners.
456
Joints loaded perpendicular to the grain
E11.4 Literature
J. Ehlbeck, R. Görlacher, original Article C2, STEP 1995.
Ehlbeck J. and Görlacher R. (1983). Tragverhalten von Queranschlüssen mittels Stahlformteilen, insbesondere
Balkenschuhen, im Holzbau. Forschungsbericht der Versuchsanstalt für Stahl, Holz und Steine,
Universität Karlsruhe.
Van der Put T.A.C.M. and Leijten A.J.M. (2000). Evaluation of perpendicular to grain failure of beams caused
by concentrated loads of joints. Paper 33‐7‐7, CIB‐W18 Meeting 33, Delft.
Leijten A.J.M. and Jorissen A.J.M. (2001). Splitting strength of beams loaded by connections perpendicular
to grain, model validation. Paper 34‐7‐1, CIB‐W18 Meeting 34, Venice.
457
E12 Reinforced joints
Original article: H. Werner
The load‐bearing capacity of the joint area of timber members is the primary weak point
when considering the load‐bearing behaviour of the overall timber construction. Dowel‐
type fasteners such as nails, dowels or screws are often used to connect members, in
configurations which could include timber‐to‐timber, timber‐to‐wood‐based products or
steel‐to‐timber joints. Taking the load‐bearing capacity of the cross‐section of the connect‐
ed timber members as the benchmark, the effectiveness of such joints generally ranges
from around 40 to 60%, which means the structural joint detailing has a crucial impact on
the performance and thus efficiency of a structure. If the key influence parameters on the
load‐bearing and deformation behaviour of the joint as well as the different failure types
are known, targeted changes made to the structural detailing may improve the load‐
bearing capacity of the joint, reducing the extent of the member cross‐section required.
As was already shown in previous articles (primarily E1), joints with dowel‐type fasteners
may have a range of failure mechanisms. Embedment failure alone occurs if the timber or
wood‐based material under the fastener is completely plasticised and the fastener re‐
mains straight. The key parameters for this failure mode are the embedment strength of
the timber or wood‐based material plus the joint geometry (thickness of the members to
be connected, spacings and end and edge distances of the fasteners). Another failure
mode affecting joints with dowel‐type fasteners is a bending failure of the fastener and
simultaneous embedment failure of the timber or wood‐based material. When such
combined failure occurs, one or two plastic hinges emerge per shear plane and the fas‐
tener becomes inclined in the area of the shear plane leading to plastic embedment
deformations. Here too, the embedment strength, joint geometry and yield moment of
the fastener are the key parameters impacting on the load‐bearing capacity. The third –
brittle ‒ failure mode occurs when the timber splits or in the event of block shear failure
and mainly affects joints with multiple fasteners. Complying with the minimum fastener
distances helps avoid premature splitting; the abovementioned ductile failure modes
described (embedment or combined failures involving bending deformation of the fas‐
tener and embedment) are only seen when complying with minimum distances.
459
Reinforced joints
Figure E12‐1 Double‐shear timber‐to‐timber joint with glued‐on plywood (left) and double‐shear
steel‐to‐timber joint with slotted‐in steel plate and glued‐on wood‐based panels (right).
While using fasteners made of high‐strength, ductile steel and timber of higher density
can boost the load‐bearing capacity of even non‐reinforced joints, while reinforcement
applied in the form of glued‐on wood‐based panels or fully threaded screws is intended
to avoid premature splitting or local shear failure in the area of the joint. This approach
can meet two aims: firstly, avoiding premature failure and secondly, in comparison to unre‐
inforced joints, allowing lower fastener spacing. The group effect, which has a particularly
adverse effect on joints with multiple fasteners arranged in a row in the force and grain
direction (see also Article E13), can also be minimised. In addition, the embedment strength
and hence load‐bearing capacity of the individual fastener can be increased.
The first choice for glued‐on reinforcements are wood‐based panels with an embedment
strength far exceeding that of the timber to be reinforced. Reinforcing joints with this
approach not only allows embedment strengths close to shear planes of connected
members to be increased, but also offers additional reinforcement against tension per‐
pendicular to the grain in the area of the fasteners which, in turn, reduces susceptibility
to splitting. The results obtained by Blass and Werner (1988) with glued‐on beech ply‐
wood in the shear plane area of softwood joints showed that gluing‐on panels significant‐
ly reduces the susceptibility of softwood to splitting and exploits the higher embedment
strength of the plywood in the area of the joint where the loading stresses peak, thus
increasing the overall load‐bearing capacity of the joint. Two examples of such rein‐
forcement are shown in Figure E12‐1.
Other means of increasing the load‐bearing capacity can be achieved with nail‐plate
connectors (Kevarinmäki et al., 1995; Blass and Schmid, 2001) or glued‐on glass‐grain
mats (Haller et al., 1998). A nail plate is pressed in, in the area of the joint, where the
central nail plate area is without punch‐outs, whereupon the individual reinforced mem‐
bers are connected with a dowel‐type fastener. If the joint is loaded laterally, both the
nail plate and timber are exposed to embedment stresses. This method considerably
increases the load‐bearing capacity of the joint and helps prevent any splitting of the
460
Reinforced joints
timber. The nail‐plate connector can be considered a refined version of the single‐sided
toothed‐plate connector (Article E6). In this case, however, the load‐bearing behaviour
differs due to the fact that for toothed‐plate connectors, the critical parameter is the
embedment strength of the timber in the area of the teeth, whereas for nail‐plate con‐
nectors, the embedment strength of the nail plate itself is the key variable. A joint rein‐
forced with nail‐plate connectors as well as a close‐up view of the cross‐section through
the joint, including bending deformation of the fastener and embedment deformation of
the timber and nail plate, is shown in Figure E12‐2. Since both the reinforcement meth‐
ods mentioned, nail‐plate connectors or glued‐on glass grain mats, are not customary in
Germany, no further detail will be provided in this article.
The wood in the joint area of connections with dowel‐type fasteners can also be rein‐
forced with self‐tapping screws, arranged perpendicular to the grain and at right angles
to the actual dowel‐type fasteners in two possible configurations, as shown in Figure E12‐
3. Fully threaded screws can be arranged at a distance to the dowel‐type fasteners in the
grain direction (Figure E12‐3 left), which ensures that any crack propagation will not
extend beyond the fully threaded screws, although any increase in load‐bearing capacity
(beyond the Johansen case) will not be possible. Installing fully threaded screws directly
underneath the fasteners (Figure E12‐3 right), conversely, not only prevents splitting but
also boosts load‐bearing capacity due to the contribution of the screw acting as a “beam”
supporting the fasteners. This kind of reinforcement therefore additionally increases the
load‐bearing capacity of a joint not at risk of splitting.
The following section explains reinforcement measures with glued‐on wood‐based panels
and fully threaded screws in more detail.
Figure E12‐2 A joint with tube fasteners that has been reinforced with nail‐plate connectors (left),
Close‐up view of cross‐section with reinforced joint and embedment deformations of the
reinforcement (right).
461
Reinforced joints
Figure E12‐3 Joints reinforced with fully threaded screws. Left: reinforcement with a distance to the dowels.
(Schmid, 2002) Right: opened specimen after test with reinforcement directly underneath the
dowel. (Bejtka, 2005a)
E12.1 Joints reinforced with glued‐on wood‐based panels
Article E2 derived the Johansen equations for single‐ and double‐shear timber joints and
the load‐bearing capacity of reinforced joints can be derived similarly, as will subsequently
be shown in detail citing the example of a steel‐to‐timber joint.
Derivation of the load‐bearing capacity Fv,s,R of a reinforced joint
Taking the example of a laterally loaded steel‐to‐timber joint, featuring thick external
steel plates and two shear planes (Figure E12‐4), equations for the failure mode with two
plastic hinges are derived for calculating reinforced joints. The input parameters for the
calculation are the joint geometry (thickness of the timber member t2, thickness of the
reinforcement s, fastener diameter d), the embedment strengths of the timber fh and the
reinforcement fh,s as well as the yield moment My of the fastener. In all the specified
equations below, if the thickness s of the reinforcement is set to zero, the Johansen
equations for unreinforced joints from Article E2 emerge.
462
Reinforced joints
Fv,s,R Fv,s,R
t1 s t2 s t1
b2 s
b2
f h,s
fh
A
My
My
Fv,s,R
2 Fv,s,R
Figure E12‐4 Left: double‐shear steel‐to‐timber joint with thick, external steel plates, embedment strength
in timber and reinforcement is exceeded and two plastic hinges have formed. Right: free‐body
diagram of the fastener.
The equilibrium of forces V = 0 results in:
The equilibrium of moments at the fastener MA = 0 is:
b s
2 M y f h,s f h d s f h d b 2 s 2
s
0
2 2
Solving for b2:
4 M y f
b2 s 2 h,s 1 s (E12‐2)
fh d fh
Inserting equation (E12‐2) in equation (E12‐1) reveals the load‐bearing capacity Fv,s,R:
4 M
2 f
F v,s,R f h,s d s f h d s h,s 1 s
y
(E12‐3)
fh d fh
463
Reinforced joints
The load‐bearing capacity Fv,s,R of a reinforced joint can be derived similarly for all other
failure modes and joint typologies (see also Article E2 and Annex 6). In the following
section, equations to calculate the load‐bearing capacity for all failure mechanisms are
specified for single‐ and double‐shear timber‐to‐timber and steel‐to‐timber joints with
reinforced joint areas. To simplify matters, equal thickness and embedment strength of
the reinforcement panels arranged at the shear planes of a joint are assumed. The fol‐
lowing designations apply:
t1 , t2 Thickness of the timber members or penetration depth
s Thickness of the reinforcement
fh,1 , fh,2 Embedment strength of the members
fh,s Embedment strength of the reinforcement
My Yield moment of the fastener
d Diameter of the fastener
Ratio of the embedment strengths of the members: = fh,2/fh,1
Ratio of the embedment strength of the reinforcement and member 1:
= fh,s/fh,1
Reinforced timber‐to‐timber joints
Reinforced, single‐shear joints made of timber or wood‐based materials
The load‐bearing capacity Fv,s,R per fastener is determined as the lowest value from the
following equations (E12‐4) to (E12‐9):
f h,1 d t 1
F v,s,R
1
s
2
t2 2
t2 s s
2
st 2 s
2
3 t2
2
1 4
2 1 t
2
2
2
4 8
2
4
2
2
2 t 2 (E12‐6)
t1 1 t1 t1 t1 t1 t1 1
f h,1 d t 1
t2 s
1
f h,s d s 4
1 t1 t1
(embedment failure in both members (case c in Figure E2‐5))
464
Reinforced joints
f h,1 d 2 4 My
t 1 4 s t 1 4 s
2
F v,s,R t 12 4 s 2
2 f h,1 d
(E12‐7)
f h,s d s
(embedment failure in both members, plastic hinge in member 2
(case d in Figure E2‐5))
f h,1 d
F v,s,R f h,s d s
1 2
4 M y (E12‐8)
t 2 4 s 1 2 t 22 4 s 2 2
2
t 4 s
f h,1 d
(embedment failure in both members, plastic hinge in member 1
(case e in Figure E2‐5))
2 f h,1 d 2 1 2M y
F v,s,R s s 2 s f h,s d s (E12‐9)
1 2 f h,1 d
(embedment failure and plastic hinges in both members (case f in Figure E2‐5))
Reinforced, double‐shear joints made of timber or wood‐based materials
The load‐bearing capacity Fv,s,R per fastener and shear plane is determined as the lowest
value from the following equations (E12‐10) to (E12‐13) (member 2 is at the centre):
f h,1 d 2 2 4 M y
t 1 4 s t 1 4 s
2
t 1 4 s
2
F v,s,R
2 f h,1 d
(E12‐12)
f h,s d s
(embedment failure in both members, plastic hinge in member 2
(case j in Figure E2‐8))
465
Reinforced joints
2 f h,1 d 2 1 2M y
F v,s,R s s 2 s f h,s d s (E12‐13)
1 2 f h,1 d
(embedment failure and plastic hinges in both members
(case k in Figure E2‐8))
Reinforced steel‐to‐timber joints
Reinforced, single‐shear joints with thin steel plates
The load‐bearing capacity Fv,s,R per fastener for single‐shear joints with thin steel plates
(namely for a steel plate of thickness t ≤ 0.5 ∙ d) is determined as the lowest value from
the following equations (E12‐14) and (E12‐15):
(embedment failure (case a in Figure E2‐9))
2M y
F v,s,R f h,1 d 1 s
2
s f h,s d s (E12‐15)
f h,1 d
(embedment failure and plastic hinge (case b in Figure E2‐9))
Reinforced, single‐shear joints with thick steel plates
The load‐bearing capacity Fv,s,R per fastener for single‐shear joints with thick steel plates
(namely for a steel plate thickness t ≥ d) is determined as the lowest value from the fol‐
lowing equations (E12‐16) to (E12‐18):
4 M y
F v,s,R f h,1 d 2 t 1 2 2 s 4 s t 1 t 1 2 s
2 2
f h,1 d (E12‐17)
f h,s d s
(embedment failure and plastic hinge (case d in Figure E2‐9))
466
Reinforced joints
4 M y
F v,s,R f h,1 d 1 s
2
s f h,s d s (E12‐18)
f h,1 d
(embedment failure and 2 plastic hinges (case e in Figure E2‐9))
Reinforced, double‐shear joints with steel plate as central member
The load‐bearing capacity Fv,s,R per fastener and per shear plane is determined as the
lowest value from the following equations (E12‐19) to (E12‐21):
4 M y
F v,s,R f h,1 d 2 t 12 2 2 s 2 4 s t 1 t 1 2 s
f h,1 d (E12‐20)
f h,s d s
(embedment failure and plastic hinge (case g in Figure E2‐9))
4 M y
F v,s,R f h,1 d 1 s
2
s f h,s d s (E12‐21)
f h,1 d
(embedment failure and 2 plastic hinges (case h in Figure E2‐9))
Reinforced, double‐shear joints with external, thin steel plates
The load‐bearing capacity Fv,s,R per fastener and per shear plane is determined as the
lowest value from the following equations (E12‐22) and (E12‐23):
2 2M y
F v,s,R f h,2 d 1 s s f h,s d s (E12‐23)
f h,2 d
(embedment failure and plastic hinge (case k in Figure E2‐9))
467
Reinforced joints
Reinforced, double‐shear joints with external, thick steel plates
The load‐bearing capacity Fv,s,R per fastener and per shear plane is determined as the
lowest value from the following equations (E12‐24) and (E12‐25) (whereby equation
(E12‐25) corresponds to the above‐derived equation (E12‐3)):
4 M
2 f
F v,s,R f h,s d s f h d s h,s 1 s
y
(E12‐25)
fh d fh
(embedment failure and 2 plastic hinges (case m in Figure E2‐9))
Verification of the bond line for glued‐on wood‐based panels
When reinforcing a joint with wood‐based panels, the bonding surface AL must also be
verified and meet the following condition:
f h,s
AL n sd (E12‐26)
fv
where
fh,s Embedment strength of the wood‐based panel
fv Effective shear strength of the wood‐based panel (out‐of‐plane, rolling shear)
or timber, whichever is smaller
n Number of fasteners
s Panel thickness
d Diameter of the fastener
The gluing‐on and the rigid joining it provides between timber and the reinforcement
mean glued‐on wood‐based panels are the only type of reinforcement that can be relied
on to also reduce the risk of block shear in joints with multiple fasteners (Article E13).
This is not possible for mechanical reinforcements (fully threaded screws), which provide
far greater semi‐rigidity in the joint.
468
Reinforced joints
Embedment strength of reinforcing wood‐based panels
Embedment strength of various wood‐based panels
The embedment strength of wood‐based panels is dictated by the type of panel and its
build‐up. For plywood, particleboards, OSB and hardboards, EC 5 specifies the following
equations to calculate the characteristic embedment strength in N/mm2 (where k is
expressed in kg/m3 and d and t in mm):
Nails and screws in non‐predrilled plywood: fh,k =0.11 ∙ k ∙ d‐0.3
Dowels, bolts, nails and screws in predrilled plywood: fh,k = 0.11 ∙ (1 – 0.01 ∙ d) ∙ k
Nails and screws in non‐predrilled particleboards or OSB: fh,k = 65 ∙ d‐0.7 ∙ t0.1
Dowels, bolts, nails and screws in predrilled particleboards or OSB:
fh,k = 50 ∙ d‐0.6 ∙ t0.2
Nails in hardboards: fh,k = 30 ∙ d‐0.3 ∙ t0.6
The best choice when selecting a reinforcement measure for joints is wood‐based panels
with an embedment strength that considerably exceeds that of the wood to be reinforced.
E12.2 Joints reinforced with fully threaded screws
Joints reinforced with fully threaded screws tend to be dowelled joints, in which timber
members are subject to loading in the grain direction. The reinforcing elements (= fully
threaded screws) are always arranged perpendicularly to the grain direction and the
dowel axis, to prevent premature splitting (for joints with multiple fasteners), see Fig‐
ure E12‐3. If the reinforcement is arranged directly underneath the dowels rather than at
a distance from them, however, the load‐bearing capacity of the reinforced joint can be
increased even further (increased embedment capacity).
Fully threaded screws, arranged at a distance from the dowels
When inserting fully threaded screws at a distance from the dowels (Figure E12‐3 on the
left), the sole function of the screws is to reinforce against tension perpendicular to the
grain and prevent premature splitting and therefore premature failure at a lower load‐
bearing capacity than the respective Johansen case. The reinforcement prevents cracks
extending any further than the fully threaded screws, although it is not possible to in‐
crease load‐bearing capacity further (beyond the Johansen case). This means the effec‐
tive number of fasteners can be reconciled with the actual number:
n ef n (E12‐27)
469
Reinforced joints
This strengthening approach is comparable to installing fully threaded screws to reinforce
notched beams or joints loaded perpendicular to the grain (Article D8). The screws are
subject to axial load and have to absorb the tensile force perpendicular to the grain,
which would otherwise trigger splitting in a non‐reinforced joint. The calculation model
derived by Schmid (2002) to determine the axial force component (tensile force) in the
reinforcing screw depending on the external load component parallel to the grain applies
only to joints with one fastener row in the grain direction and reinforcing screws installed
perpendicular to the grain and fastener axis. Although this kind of reinforcement can
reliably prevent splitting, it cannot eliminate block shear failure. It is advisable to dimen‐
sion fully threaded screws exposed to withdrawal load and which act as a reinforcement
against splitting for 30% of the load per dowel and per shear plane:
whereby Fax,90,Rk corresponds to the characteristic withdrawal resistance of a fully thread‐
ed screw of penetration depth ef = a4,c (see Article E5) and Fv,Rk is the characteristic load‐
bearing capacity of a dowel in accordance with Johansen (see Article E2) of a unrein‐
forced joint.
Fully threaded screws, arranged directly underneath the dowels
As well as preventing the risk of splitting, installing fully threaded screws directly under‐
neath the fasteners (Figure E12‐3 on the right) also helps increase the load‐bearing ca‐
pacity, since the screw functions as a “beam”, which supports the fasteners. This kind of
reinforcement measure also means the load‐bearing capacity of a joint not at risk of
splitting can be increased beyond the Johansen load‐bearing capacity. Similar to the
process of extending Johansen equations for joints with glued‐on wood‐based panels or
nail connectors, the Johansen model can also be extended for these reinforcement types
(see also Figure E12‐6). The variable required in addition to the Johansen parameters
(embedment strength fh,i, yield moment My, member thickness ti, fastener diameter d) is
the load‐bearing capacity RVE of the reinforcement element, whereby RVE may be identi‐
cal or greater than the force component FVE from the dowel, Figure E12‐5. Accordingly, a
distinction is established between “soft” or “hard” reinforcement. In a “soft” reinforce‐
ment, the screw functions like a semi‐rigid support for the dowel, whereas its role in a
“hard” reinforcement is that of a rigid support. This is an important distinction, since
depending on whether the support in question is semi‐rigid or rigid (whether or not the
reinforcement element has scope to bend), differing failure mechanisms may occur with‐
in the classic Johansen cases (involving no, one or two plastic hinges per shear plane).
Figure E12‐6 shows the range of different failure mechanisms, citing the example of a
single‐shear timber‐to‐timber joint for the Johansen case with two plastic hinges per
shear plane. In the "hard" failure mechanism, a possible plastic hinge of the dowel will
develop at the same position as the screw.
470
Reinforced joints
RVE
FVE
FVE
G VE
15 mm
RVE
FVE
FVE
15 mm G VE
Figure E12‐5 Above: „soft“ reinforcement, below: „hard“ reinforcement. (Blass et al., 2006)
Figure E12‐6 also shows the internal forces and moments at the fastener belonging to the
failure mechanisms and with which the Johansen equations can now be extended. The
load‐bearing capacity R3 resulting from the equilibrium of forces and moments is derived
as an example in the following case, featuring the “hard reinforcement” of a single‐shear
timber‐to‐timber joint with two plastic hinges per shear plane. Additional cases are con‐
tained in Bejtka (2005a) and Blass et al. (2006). The load‐bearing capacity R3 is deter‐
mined from the equilibrium of forces and moments in the joint (for internal forces and
moments, see the second case from above in Figure E12‐6). The equilibrium of forces
V = 0 reveals:
The equilibrium of moments at the fastener MA = 0 reveals:
p p
2 M y f h,2 d p f h,1 d p p F1,VE,3 2 p 0
2 2
Solving for F1,VE,3 reveals the following, with = fh,2/fh,1:
My f h,1 d p
F1,VE,3 3 (E12‐30)
p 4
Inserting equation (E12‐30) in equation (E12‐29) results in the load‐bearing capacity R3:
My f h,1 d p
R3 1 (E12‐31)
p 4
471
Reinforced joints
However, equation (E12‐31) only applies if the failure case is deemed “hard”, namely,
when the lateral load‐bearing capacities of the reinforcement element R1,VE and R2,VE
exceed the corresponding force components F1,VE,3 and F2,VE,3. The force components
F1,VE,3 and F2,VE,3 are also derived from the equilibrium of forces and moments in the shear
plane. For F1,VE,3, equation (E12‐30) applies, meaning:
My f h,1 d p
R 1,VE F1,VE,3 3 (E12‐32)
p 4
F2,VE,3 is determined similarly. The equilibrium of moments at the fastener MB = 0 reveals:
p p
2 M y f h,1 d p f h,2 d p p F 2,VE,3 2 p 0
2 2
Solving for F2,VE,3 reveals the following, with = fh,2/fh,1:
My f h,1 d p
F 2,VE,3 3 1 (E12‐33)
p 4
Consequently:
My f h,1 d p
R 2,VE F 2,VE,3 3 1 (E12‐34)
p 4
Or with R1,VE and = R2,VE/R1,VE:
F 2,VE,3 My f h,1 d p
R 1,VE 3 1 (E12‐35)
p 4
472
Reinforced joints
Figure E12‐6 Failure mechanisms for the Johansen case with two plastic hinges per shear plane and
internal forces and moments at the fastener, top down: „soft“, „hard“, „soft‐hard”, „hard‐soft“.
(Blass et al., 2006)
The load‐bearing capacity of a single‐shear timber‐to‐timber joint is therefore deter‐
mined for one of the possible failure cases (those involving a “hard” reinforcement) and
for the Johansen case with two plastic hinges per shear plane with equations (E12‐31),
(E12‐32) and (E12‐35). Accordingly, the number of possible failure cases in a reinforce‐
ment featuring fully threaded screws arranged directly underneath the dowels clearly
increases, since per Johansen case, different failure mechanisms of the reinforcement
elements must be taken into consideration (see Figure E12‐6, cases „soft“, „hard“, „soft‐
hard“ and „hard‐soft“).
473
Reinforced joints
Load‐bearing capacity of the reinforcement elements
To design a reinforced dowelled joint, with fully threaded screws arranged directly under‐
neath the dowels, the only thing now missing is the value for the load‐bearing capacity RVE
of the reinforcement elements and here, two cases can be distinguished, see Figure E12‐7.
ℓS ℓS
The load‐bearing capacity RVE is calculated in the same way as the load‐bearing capacity
of a Johansen joint with slotted‐in steel plate (= screw). The corresponding equations for
the case of one dowel reinforced with a screw (Figure E12‐7 on the left) are given in
equation (E12‐36); the equations for the second case can be referenced from literature
(Bejtka, 2005b):
f h,S d S S
16 M
min f h,S d S S
y,S
R VE 2 1 (E12‐36)
f h,S d S 2S
4 M y,S f h,S d S
474
Reinforced joints
E12.3 Literature
H. Werner, original Article 9 (Band 3), STEP 1995.
Bejtka I. (2005a). Verstärkung von Bauteilen aus Holz mit Vollgewindeschrauben.
Dissertation Universität Karlsruhe.
Bejtka I. (2005b). Self‐tapping screws as reinforcements in connections with dowel‐type fasteners.
Paper 38‐7‐4, CIB‐W18 Meeting 38, Karlsruhe.
Blass H.J., Bejtka I. and Uibel T. (2006). Tragfähigkeit von Verbindungen mit selbstbohrenden Holzschrauben
mit Vollgewinde. Karlsruher Berichte zum Ingenieurholzbau Band 4. Universität Karlsruhe.
Blass H.J. and Schmid M. (2001). Verstärkung von Verbindungen. Bauen mit Holz 103:40‐48.
Blass H.J. and Werner H. (1988). Stabdübelverbindungen mit verstärkten Anschlussbereichen. Bauen mit Holz
90:601‐607.
Haller P., Wehsener J. and Chen C.J. (1998). Development of joints by compressed wood and glassgrain
reinforcement. COST C1 ‒ Control of the semi‐rigid behaviour of civil engineering structural connections,
International Conference, Liège.
Kevarinmäki A., Kangas J., Nokelainen T. and Kanerva P. (1995). Nail‐plate reinforced bolt joints of Kerto‐FSH
structures. Publication 51, Helsinki University of Technology.
Schmid M. (2002). Anwendung der Bruchmechanik auf Verbindungen mit Holz.
Dissertation Universität Karlsruhe.
475
E13 Joints with multiple fasteners
Original article: H. J. Blass
Timber joints with mechanical fasteners usually contain more than just a single fastener.
However, this means the load distribution between multiple fasteners is always uneven,
even if the load transmitted by the joint is applied to the centre of gravity of the joint.
The load‐bearing capacity of a joint corresponds to the sum of the loads withstood by the
individual fasteners in the ultimate limit state. If now, in the ultimate limit state of the
joint, the loads exerted on the fasteners vary and some fasteners are loaded with a load
lower than their ultimate capacity, the load‐bearing capacity of the joint is lower than the
collective load‐bearing capacities of the individual fasteners. This explains the reduction
of the load‐bearing capacity per fastener in joints with multiple fasteners for specific
fastener types. As a general rule, the various influences on load distribution in joints act
both on joints with multiple identical fasteners as well as on those with different fasten‐
ers. For joints with multiple fasteners, brittle failure mechanisms may also emerge, to
which the Johansen model (Article E2) is inapplicable and which fail at lower loads than
those calculated with the Johansen equations. In response, either joints can be rein‐
forced (Article E12), an effective number nef < n can be used or the brittle failure mecha‐
nisms can be separately verified. This article starts by examining the load distribution
between individual fasteners in timber joints and the concept of the effective number of
fasteners nef as well as illustrating the failure mode “wood failure in joints with multiple
glued‐in rods subject to axial stresses”. Finally, the various failure modes occurring in
joints with laterally loaded fasteners are shown, before explaining the verification ap‐
proach for block shear failure.
E13.1 Load distribution in accordance with the theory of elasticity
Lantos (1969) developed a model to calculate the load distribution in timber joints, based
on a linear‐elastic load‐deformation relationship of the individual fasteners without initial
slip and assuming uniformly distributed normal stresses in the members to be connected.
The validity of his model for laterally loaded fasteners is limited to the lower load range,
in which the joint behaviour can be considered elastic, and to joints loaded parallel to the
grain. Cramer (1968) adopts a similar approach when taking the non‐uniform normal
stress distribution in the cross‐sections into consideration and its influence on the axial
stiffness of the connected members. However, the elastic solution according to Lantos
can be applied to calculate the load‐bearing capacity of joints with linear‐elastic
477
Joints with multiple fasteners
load‐deformation behaviour and multiple fasteners arranged in a row and in the direction
of the load. Examples include timber‐to‐timber or steel‐to‐timber joints with screws
arranged at an angle. The solution for the general case of non‐linear load‐deformation
relationships is specified in Wilkinson (1986).
Figure E13‐1 shows a single‐shear timber‐to‐timber joint with nails in which the load is
transferred between members 1 (M1) and 2 (M2) in discrete steps at the fastener loca‐
tions. Here, each step represents the load transferred by the respective fasteners. In the
deformed state, the load in member 1 between fasteners i and i + 1 corresponds to the
total load in the joint reduced by the loads transmitted by fasteners 1 to i. The original
length s between both fasteners is extended by u1,i in this case. Similarly, the correspond‐
ing section of member 2 is extended by s to s + u2,i, while the forces transferred by fas‐
teners i and i + 1 result in displacements of uf,i or uf,i+1.
s s s
1 2 i i+1 n-1 n
M2
F
M1 uf,i s+u2,i
s+u1,i uf,i+1
Figure E13‐1 View of the undeformed (upper) and section of the deformed (lower) joint area.
M1: member 1; M2: member 2. (STEP 1995 Article C15)
A comparison of the lengths and deformations in both members (see Figure E13‐1) re‐
veals that:
The deformation in the individual fastener can be replaced by:
Ff
uf (E13‐2)
K
where Ff is the load transferred by the fastener and K is the slip modulus.
478
Joints with multiple fasteners
The elongation of the members can be expressed by:
Fm s
u (E13‐3)
EA
where Fm is the load present in the member between two fasteners.
If equations (E13‐2) and (E13‐3) are inserted into equation (E13‐1), after some transfor‐
mations, the resulting load on the most stressed fastener at the start or end of the fas‐
tener row is as follows (where the connected members have the same axial stiffness E ∙ A):
m 1n 1
F 1 F 1 m 1 1 m 1 m 2 (E13‐4)
n
m1 m 2
n
m 1n 1
F n F m 1n‐1 1 m 1n‐1 m 2n‐1
m 1n m 2n
(E13‐5)
where
1
(E13‐6)
E 1 A1
1
E 2 A2
2 4
m1 (E13‐7)
2
2 4
m2 (E13‐8)
2
1 1
2 K s (E13‐9)
E
1 1A E 2 A2
479
Joints with multiple fasteners
Lantos demanded to design the most stressed fastener. Since the fasteners at the start or
end of the fastener row are most stressed, they determine the load‐bearing capacity of
the joint. Based on this solution, the differences in individual fastener loads are dictated
by the axial stiffnesses of the connected members, the number of fasteners in a row, the
spacing among fasteners and the slip modulus. The linear‐elastic approach by Lantos
presented here remains relevant for joints, which respond in a linear‐elastic manner up
to failure. For joints with multiple fully threaded screws, arranged at an angle and in a
row, the Lantos method can be used to identify and design the most stressed screw.
E13.2 Parameters influencing the load distribution in joints
In addition to the influence of varying elongation in the connected members, other pa‐
rameters can also significantly influence the load distribution between the individual
fasteners of a timber joint.
Plastic deformations and creep
Isyumov (1967) selected a more general approach to calculate the load distribution be‐
tween individual fasteners arranged in a row and parallel to the load. His solution takes
into consideration the non‐linear load‐deformation behaviour of the fasteners, which
triggers load redistribution in the joint with increasing total load. As soon as the fastener
subject to the greatest stress exhibits plastic deformations, its stiffness declines in com‐
parison to other fasteners. Since more rigid elements in a parallel system are subject to
comparatively higher stresses, this redistributes the load to the less heavily loaded fas‐
teners in the middle of the row. These load redistributions offset the load peaks de‐
scribed by Lantos (1969) and increase load‐bearing capacities in comparison to completely
elastic joints. The same applies to the influence of time‐dependent deformations in the
joint. Creep also results in a relative loss of stiffness, which, in turn, leads to load redistri‐
butions. Given that the scale of the creep deformations rises with increasing loading,
greater creep deformations are likely to occur at the start and end of a fastener row,
which leads to a more uniform load distribution in the joint.
Plastic deformations of joints and the load redistributions they cause are actually the
basis for design in accordance with EC 5. The calculation of the load‐bearing capacity of
joints with dowel‐type fasteners in accordance with EC 5 is attributable to Johansen
(1949). Johansen presumed ideal rigid‐plastic behaviour for both fasteners and wood
under embedment stress (see Article E2). Subject to sufficient edge distances and spac‐
ings and if the joint fails in accordance with one of the mechanisms described by Johansen,
the joint shows plastic load‐deformation behaviour. However, if splitting of the timber
along the row of fasteners or block shear occurs at load levels well below the potential
plastic capacity, a full redistribution of the load within the joint is prevented, which is why
480
Joints with multiple fasteners
splitting or block shear significantly reduce the load‐bearing capacity of joints with multi‐
ple fasteners. Sufficient edge distances and spacings lead to reduced splitting of the
timber in the joint area, since the shear and tensile stresses perpendicular to the grain
generated by the wedge effect of the fasteners decline with increasing distance. Accord‐
ingly, the greater the distances in the grain direction, the more favourable the plastic
deformation behaviour when reaching the ultimate load, which, contrary to elastic theo‐
ry, ends up increasing the load‐bearing capacity of joints with multiple fasteners. Another
means of preventing splitting and guaranteeing plastic failure instead is to reinforce the
joint area e.g. by glued‐on plywood or fully threaded screws (Article E12). Joints rein‐
forced accordingly reach a load‐bearing capacity which corresponds to the collective
load‐bearing capacities of the individual fasteners. No reduction need be taken into ac‐
count due to the group effect in this case.
Considerable plastic deformations occur in the following timber joints with mechanical
fasteners: joints with toothed‐plate connectors, nailed joints and joints with other slen‐
der, dowel‐type fasteners. Plastic deformations, which generally precede any failure in
joints with toothed‐plate connectors, explain why an interaction of the toothed‐plate
connector and bolt can be assumed for these joints. Conversely, for joints with split ring
or shear plate connectors, which are often prone to brittle failure, the bolt is not taken
into consideration when determining the load‐bearing capacity (Article E6). Creep de‐
formations, however, affect all types of timber joints with mechanical fasteners.
Fabrication tolerances
In joints with predrilled holes or pre‐milled depressions such as dowelled joints or joints
with connectors, differences in load‐bearing behaviour between individual fasteners of a
joint are exacerbated by manufacturing inaccuracies. Such manufacturing inaccuracies
may include, for example, misalignment of the bolt holes, deviating boreholes in the
wood, variations in the hole diameter or in the initial position of the bolts in the holes.
Dannenberg and Sexsmith (1976) as well as Isyumov (1967) stress the key impact of such
manufacturing inaccuracies on load distribution within joints with connectors (split ring,
shear plate or toothed‐plate connectors), which, in turn, affects their load‐bearing capac‐
ity. According to Wilkinson (1986), manufacturing inaccuracies and differences in the
load‐deformation behaviour of individual fasteners are the main cause of non‐uniform
load distribution in timber joints with mechanical fasteners, while the influence of varying
degrees of elongation of the connected members can be disregarded. Manufacturing inac‐
curacies such as misalignments of holes result in an initial slip for some fasteners within
the joint. If the joint is then loaded, these fasteners only start to accommodate loads
when the joint deformation exceeds the initial slip (see bolt No. 4 in Figure E13‐2). If the
joint fails before the plastic deformation stage due to splitting, the fasteners affected by
initial slip do not carry any load at all. Tests conducted by Massé et al. (1989) show e.g. for
bolted joints with Douglas fir glulam that the load‐bearing capacity per bolt declined by
481
Joints with multiple fasteners
more than 50%, when the number of bolts in the joint was increased from one to four.
These results underline the need to ensure sufficient edge distances and spacings for
joints with fabrication tolerances to guarantee plastic behaviour and hence allow the
loads to be redistributed within the joint. Manufacturing inaccuracies can largely be
avoided by ensuring precise manufacturing with CNC machinery.
4
5
F1
1
3
2
4
0
10 20
Ftot
Figure E13‐2 Example load distribution in a bolted joint according to Wilkinson (1986). F1 is the load
of the individual bolt and Ftot is the total load of the joint. (STEP 1995 Article C15)
Differences in the load‐deformation behaviour of the individual fasteners
As well as fabrication tolerances, variation in the characteristics of the wood itself in the
joint area impact on the load distribution. Knots, cracks, resin pockets, slope of grain or
density variation all alter the load‐deformation behaviour of the individual fasteners.
E13.3 Influence of the number of fasteners
In principle, when determining the load‐bearing capacity, ideal‐plastic behaviour is initial‐
ly assumed, whereupon the influence of the number of fasteners is taken into considera‐
tion by reducing the load‐bearing capacity per fastener. The so‐called effective number
nef of fasteners is then specified in EC 5 per fastener type. For nailed joints, meanwhile,
the number n of fasteners arranged in a row is reduced by an exponent kef, which de‐
pends on nail spacing. For bolted and dowelled joints, however, the effective number nef
depends on the diameter of the fastener as well as the spacing within a row of fasteners.
482
Joints with multiple fasteners
E13.4 Glued‐in rods – group tear‐out
One exceptional case for joints with multiple fasteners is that of joints with multiple
glued‐in rods subject to axial tensile stress. In such joints, a possible failure mode is the
so‐called group tear‐out, which involves the entire group of fasteners, including the
wooden block surrounding the rods, being torn out, as shown in Figure E13‐3. This failure
mode resembles the brittle failure modes mentioned above for joints with laterally load‐
ed multiple fasteners arranged in a row along the grain direction. Group tear‐out is also
due to premature failure of the wood in response to tensile or shear stresses; exacerbat‐
ed by varying loads in individual rods at the moment of failure. The total load‐bearing
capacity thus tends to be lower than the collective load‐bearing capacities of the individ‐
ual rods. As was already explained in Article E10, minimum bond lengths are required to
ensure ductile failure, so that the brittle failure mode in question can be avoided.
Since this failure mode is not explicitly taken into account in EC 5, there is a need to con‐
sider an effective number of fasteners nef = n0.9 as is done for joints with groups of screws
subject to tensile stresses (EC 5, Equation 8.41), if plastic steel failure is not to prevail.
The NA to EC 5, for example, stipulates that “the key criterion dictating the load‐bearing
capacity of the joint should be the load‐bearing capacity of the rod and not the strength
of the wood or bond line, if non‐uniform loading cannot be ruled out.” (NCI NA.11.2.3
(NA.2)); which means steel failure of the rods should be decisive. If this is not the case,
the group tear‐out should be verified, following the procedure used for block shear fail‐
ure in Section E13.5.
Figure E13‐3 Failure mode “group tear‐out” in a group of glued‐in rods (Tlustochowicz et al., 2011).
483
Joints with multiple fasteners
E13.5 Block shear failure
Before explaining how block shear failure of timber joints is verified, as is set out in Annex
A of EC 5, all possible failure modes of timber joints with laterally loaded fasteners will be
covered. The method given in Annex A is limited to steel‐to‐timber joints, but block shear
failure may also affect timber‐to‐timber joints.
Failure modes of joints with multiple fasteners and with timber members
loaded parallel to the grain
The possible failure modes of joints with multiple fasteners and with timber members
loaded parallel to the grain vary very widely. In addition to the ductile failure modes
covered in the Johansen model and depending on the dimensions of the timber mem‐
bers, brittle modes such as splitting or shear failures may also occur. Joints with multiple
dowels or bolts in a row are particularly prone to such brittle failure modes, even if the
required minimum distances are complied with, reference to cases (i), (ii), and (iv) in
Figure E13‐4.
Splitting, as shown in case (iv), is considered by applying an effective number of fasteners
nef, allowing the Johansen equations to be reused and the joint with a load‐bearing ca‐
pacity reduced by nef/n to be designed. One way to prevent such splitting and achieve the
full potential load‐bearing capacity in accordance with Johansen is to reinforce the joint
(depending on the type of reinforcement, the load‐bearing capacity may be increased
even further, see Article E12). Failure modes (i) and (ii) in accordance with Figure E13‐4,
however, are only partially considered with the effective number nef and must still be
separately verified. Row shear‐out, case (i), is not explicitly regulated in EC 5, although
block shear failure of a group of fasteners, namely multiple rows featuring multiple fas‐
teners in a row, can be verified using Annex A of the EC 5. Shear failures (i) and (ii) can
only, based on the current state of the art, be effectively prevented by gluing on wood‐
based panels, since only this ensures a sufficiently rigid joint between the timber member
and reinforcing element.
484
Joints with multiple fasteners
Verification block shear failure
The “block shear” failure mode, an example of which is shown in Figure E13‐5, primarily
affects steel‐to‐timber joints with compact fastener arrangements. One of the first mod‐
els used to verify this phenomenon was developed by Foschi and Longworth (1975) for
nailed steel‐to‐glulam joints. The basic concept is that either the shear strength of the
wood along the lateral and lower shear areas or the tensile strength of the wood in the
external fastener row, furthest away from the end grain, is decisive, see failure areas in
Figure E13‐5 on the right. A third possible failure mechanism is ductile failure in accord‐
ance with Johansen. Since both shear and tensile failure parallel to the grain are brittle
failure mechanisms, it is impossible to completely superpose both stress components.
According to Foschi and Longworth (1975), the load‐bearing capacity of a fastener group
loaded parallel to the grain is the minimum value of the Johansen load‐bearing capacity
and the combined load‐bearing capacity comprising shear capacity along the shear areas
and tensile capacity parallel to the grain (with a failure area perpendicular to the grain)
between both external rows of fasteners.
ℓv,i
(a) ℓt,i
t1
Cross-sectional area
perpendicular to the
Shear area Anet,v grain Anet,t
(b)
tef
Figure E13‐5 Block shear failure, (a) entire block sheared out, (b) block partially sheared out over the member
thickness. The areas shown are either subject to shear out (Anet,v) or rupture (Anet,t).
485
Joints with multiple fasteners
While the EC 5 verification currently used for steel‐to‐timber joints is similar, in this case,
the larger value of the shear or tensile capacity is decisive for block shear failure:
where
Fbs,Rk Characteristic block shear load‐bearing capacity
Anet,t Net cross‐section perpendicular to the grain
ft,0,k Characteristic tension strength parallel to the grain
Anet,v Net shear area parallel to the grain
fv,k Characteristic shear strength
The required net cross‐sections Anet,t to determine tensile capacity and Anet,v to determine
shear capacity are calculated in simplified form in EC 5.
To calculate the net cross‐section Anet,t perpendicular to the grain, as shown in Fig‐
ure E13‐5, the sum of the fastener spacings ℓt,i perpendicular to the grain are multiplied
by the thickness of the timber block having sheared out. Depending on the failure mode
of the joint, a timber block may shear out over the entire height of the timber member t1
or only over a partial block of height tef:
t,i t 1 failure modes c, f, j / l
A net,t i (E13‐11)
i t,i t ef all other failure modes
where
t1 Thickness of the timber member or penetration depth of fasteners
tef Effective thickness, depending on the failure mode
ℓt,i Sum of fastener spacings perpendicular to the grain, see Figure E13‐5 (a)
The effective thickness tef corresponds to the thickness over which the embedment
strength is reached. Depending on the failure mode of the joint, a partial block as shown
in Figure E13‐5 (b) with an effective thickness smaller than the member thickness, tef < t1,
or a block over the entire timber member thickness t1 may rupture or shear out, Fig‐
ure E13‐5 (a). If no plastic hinges occur and the fasteners do not rotate, the embedment
strength is reached over the entire member thickness, which means that in the event of
block shear failure, a block ruptures or shears out over the entire member thickness or
penetration depth. The effective thickness corresponds, in turn, to the thickness t1 of the
timber member or penetration depth of the fastener and is specified in the upper term of
equation (E13‐11). For clarity, reference can be made to Figure E13‐6, which shows all
possible failure modes of steel‐to‐timber joints. A block shear failure over the entire
486
Joints with multiple fasteners
thickness of the timber member t1 can only occur in failure modes c, f and j/l, since this is
the only configuration where the embedment strength is reached over the entire mem‐
ber thickness (in Figure E13‐6, thicknesses t1 or t2). For all other failure modes shown in
Figure E13‐6, an effective thickness tef, smaller than that of the timber member thickness,
must be used, as in the lower term of equation (E13‐11).
t1 t2
a b c d e f g h j/l k m
Figure E13‐6 Failure modes according to Johansen for steel‐to‐timber joints.
The effective thickness tef is determined via the Johansen equations (Article E2). Via equi‐
librium considerations, the position of the plastic hinge can be determined, which, in
turn, reveals the effective thickness tef (see schematic illustration in Figure E13‐5 (b)).
As an example, tef shall be determined for a single‐shear steel‐to‐timber joint with a thick
steel plate if this joint fails with two plastic hinges per shear plane according to Johansen;
reference is made to equation (E2‐11) and failure mode e in accordance with Figure E13‐6
or Figure E2‐11:
M y,k
F v,Rk f h,1,k d b 1 f h,1,k d 2 2 M y,k f h,1,k d (E13‐12)
f h,1,k d
where b1 is the distance between both plastic hinges or the width of the area where
embedment strength has been reached (Figure E2‐11).
To now determine the block shear load‐bearing capacity of such a joint with two plastic
hinges per shear plane, b1 = tef applies:
M y,k
t ef b 1 2 (E13‐13)
f h,1,k d
All other cases and the corresponding effective thicknesses tef can be taken from EC 5
Annex A.
487
Joints with multiple fasteners
The net shear area Anet,v, Figure E13‐5, also depends on the failure mode of the fastener
group, which impacts on load distribution within the group and is thus determined simi‐
larly to the net cross‐section Anet,t with t1 or the effective thickness tef:
v,i t 1 failure modes c, f, j / l
i
A net,v 1 (E13‐14)
v,i t,i 2 t ef all other failure modes
2 i i
where
t1 Thickness of timber member
tef Effective thickness or penetration depth of fasteners,
depending on the failure mode
ℓv,i Sum of fastener spacings and end distances parallel to the grain, see Figure E13‐5
(a), both external rows have to be summed up: in Figure E13‐5 (a), i = 8
ℓt,i Sum of fastener spacings perpendicular to the grain, see Figure E13‐5 (a)
The lower term of equation (E13‐14) means that in the event of one or more plastic
hinge emerging or for a short penetration depth, the net shear area is assumed to be the
sum of the lateral areas along the external rows of fasteners and the shear area under‐
neath the fastener group, since such a case involves only a partial block shearing out over
the entire thickness (Figure E13‐5 (b)), through which the lower shear area is activated as
well as the lateral shear areas. If no plastic hinges form and the fasteners remain straight,
only the lateral shearing areas along the external rows of fasteners and over the member
thickness or penetration depth of the fasteners are taken into consideration, see the
upper term of equation (E13‐14), since this involves a block to shear out over the entire
member thickness or penetration depth t1, meaning the lower shear area is not activated.
The fact that EC 5 determines the block shear load‐bearing capacity as the maximum
value of the shear and tensile capacity of the net area means that if the weaker com‐
ponent fails, the total load is absorbed by the other component. If the upper term of
equation (E13‐10), the net tensile capacity, is less than the lower term of equation (E13‐10),
the net shear capacity, this means that the joint may fail in the net cross‐section perpen‐
dicular to the grain, in which case the entire load will be absorbed by the shear area.
488
Joints with multiple fasteners
Coefficients 1.5 and 0.7 in equation (E13‐10) result from considerations of characteristic
strengths. The characteristic tensile strength parallel to the grain ft,0,k of the timber
member is increased by 50%, given the very low likelihood of large knots being present in
the very restricted, localised area prone to failure, which would mean low tensile
strength values. The increase thus corresponds to a volume effect with respect to tensile
stress (Article D3). The shear strength, in turn, is reduced by 30%, because non‐uniform
shear stress distributions with peak stresses need to be taken into consideration.
E13.6 Literature
H.J. Blass, original Article C15, STEP 1995.
Bejtka I. (2005). Verstärkung von Bauteilen aus Holz mit Vollgewindeschrauben. Dissertation,
Universität Karlsruhe. Band 2 der Karlsruher Berichte zum Ingenieurholzbau.
Cramer C.O. (1968). Load distribution in multiple‐bolt tension joints. Journal of the Structural Division,
ASCE 94(ST5):1101‐1117.
Dannenberg L.J. and Sexsmith R.G. (1976). Shear‐plate load distribution in laminated timber joints.
Report No. 361, Department of Structural Engineering, Cornell University, Ithaca, New York.
Foschi R.O. and Longworth J. (1975). Analysis and design of Griplam nailed connections.
Journal of the Structural Division 101(12):2537‐2555.
Isyumov N. (1967). Load distribution in multiple shear‐plate joints in timber. Forestry Branch Departmental
Publication No. 1203, Department of Forestry and Rural Development, Ottawa.
Johansen K.W. (1949). Theory of timber connections. International Association of Bridge and Structural
Engineering, Publication 9:249‐262.
Lantos G. (1969). Load distribution in a row of fasteners subjected to lateral load. Wood Science 1(3):129‐136.
Massé D.I., Salinas J.J. and Turnbull J.E. (1989). Lateral strength and stiffness of single and multiple bolts in
glued‐laminated timber loaded parallel to grain. Contribution No. C 029, Engineering and Statistical Research
Centre, Agriculture Canada, Ottawa.
Quenneville J.H.P. and Mohammad M. (2000). On the failure modes and strength of steel‐wood‐steel bolted
connections loaded parallel‐to‐grain. Canadian Journal of Civil Engineering 27:761‐773.
Tlustochowicz G., Serrano E. and Steiger R. (2011). State‐of‐the‐art review on timber connections with
glued‐in steel rods. Materials and Structures 44(5):997‐1020.
Wilkinson T.L. (1986). Load distribution among bolts parallel to load. Journal of Structural Engineering
112(4):835‐852.
489
E14 Moment‐resisting joints
Original article: P. Racher
To create a timber structure, timber members have to be connected together and to
other members. In many traditional structures (e.g. trusses, see also Article E15), joints
are mainly subject to normal forces or also shear forces. In design, these joints are as‐
sumed to be a hinge, given that the fasteners are arranged within a limited space and the
transmission of moments is limited. Depending on which static system is selected, the
joints may also be capable of transferring bending moments in addition to normal and
shear forces. The arrangement of joints in a configuration designed to transfer such bend‐
ing moments paves the way for economic structures. For example, moment‐resisting
corners can be arranged in portal frame structures, allowing the use of three‐hinged
frames among others, for larger spaces with larger clearances and smaller foundations.
Two configurations are possible when manufacturing moment‐resisting joints. Glued
joints can be established, e.g. in the form of large finger joints (Figure E14‐1), whereby
the members have finger joints over the entire member cross‐section and are glued
together under compressive force. Glued joints are considered rigid. The other option
would be using mechanical fasteners, although joints with mechanical fasteners will
always be semi‐rigid.
Figure E14‐1 Examples for glued joints, frame corners with large finger joints and an interim piece.
(photo on the left: H. Damm, photo on the right: H. Brüninghoff)
491
Moment‐resisting joints
The manufacture and load‐bearing capacity of large finger joints is regulated in
EN 14080. As well as setting out the requirements for adhesives, finger joints, devices
and the manufacture of finger joints, the standard also encompasses requirements for
the climatic boundary conditions to be complied with during the processing and curing of
the adhesive. The expense also increases, when the size of the members or transport
requirements dictate that the joints have to be manufactured on the construction site. In
particular, when additional measures are needed to ensure compliance with the required
climate constraints, such as a minimum temperature of 20°C, greater costs may be in‐
curred. In Germany, glued joints can only be established by manufacturers which have a
special “proof of suitability for gluing load‐bearing wood members” (DIN 1052‐10). The
need to comply with such requirements explains why such glued joints usually cost far
more to manufacture than joints with mechanical fasteners. Figure E14‐2 shows exam‐
ples of moment‐resisting joints with mechanical fasteners and the corresponding load
distribution, the latter of which depends on the joint geometry as well as the fastener
arrangement. The axial and/or shear forces transferred by the fasteners are dictated by
the position of the centre of rotation C, so that the resulting internal moments balance out
the external moments. At the same time, the equilibrium of forces also has to be respected.
Figure E14‐2 (a) shows how a guardrail post is connected to the cross beam of a pedes‐
trian bridge. The individual fasteners of the group are axially loaded by the force F and
the resulting moment M. When nails and screws are used, the axial load‐bearing capacity
depends on the withdrawal capacity of the fasteners, their tensile capacity and, where
applicable, the head pull‐through capacity. Figure E14‐2 (b) shows a possible configura‐
tion for a column with a moment‐resisting support at the base, which is subject to normal
and shear loads as well as the moment itself. In the layout shown, the internal forces and
moments are transferred to the steel parts in the foundation through normal and shear
forces, while the fasteners are subject to lateral loads. The timber, meanwhile, is subject
to forces acting both in parallel and perpendicular to the grain direction and limiting
stresses on the timber exerted perpendicular to the grain in such joint configurations is a
priority. One common solution used for frame corner joints with mechanical fasteners is
shown in Figure E14‐2 (c). Here, dowels are the first choice of fasteners and usually ar‐
ranged in a circular form. In a moment‐resisting frame corner, normal and shear forces
must also be accommodated alongside the moment. The following sections focus on this
type of moment‐resisting joint in particular.
The prerequisite for the configuration shown in Figure E14‐2 (c) is that the members to
be connected must not be in the same plane, but either beams or columns are imple‐
mented comprising two components. In this case, filling elements are also often arranged
between both load‐bearing components. If joints are designed where the members to be
connected have different angles to the grain, differences in the degree of shrinking and
swelling of wood perpendicular and parallel to the grain must be considered. This applies in
the configuration shown when connecting beams and columns. Varying moisture content
492
Moment‐resisting joints
can generate greater stresses perpendicular to the grain in the joint area which may
result in splitting. Therefore, the magnitude of stresses perpendicular to the grain must
be limited. This can be done, e.g. by limiting the height of members or gluing on wood‐
based panel materials. There is also scope to deploy self‐tapping fully threaded screws to
accommodate tensile stresses perpendicular to the grain and secure the corners or to
use cross‐laminated timber or crossbanded LVL members with reinforcing cross layers.
F
C M
Figure E14‐2 Examples of moment‐resisting joints with mechanical fasteners (top) and load distribution
on fasteners resulting from the moment only (bottom). (STEP 1995 Article C16)
493
Moment‐resisting joints
this purpose, the tensile force is not parallel to the grain but results in compression per‐
pendicular to grain stresses on the timber at the end of the nailed steel‐to‐timber con‐
nection. In addition, the upper edge of the beam in the joint area is notched to such an
extent that restraints can be ruled out.
In the frame corner configuration shown in Figure E14‐4, the moment is transferred in
the form of compressive and tensile force into the column and tension strut respectively.
Here, it is worth noting that the rafter is stressed by relatively high shear forces in the
relatively confined space between column and strut.
Tensile force
a
Moment
Compressive force
Figure E14‐3 Left: Division of the corner moment in tensile and compressive force. Right: The angle between
the grain direction of the beam and the system line of the hinged joint allows rotation of the steel
plate when the beam depth changes (= shrinkage/swelling).
Figure E14‐4 Frame corner in a portal frame structure. Division of the corner moment in a tensile and a
compressive force. (photo: Finnforest)
494
Moment‐resisting joints
Compared to glued or curved frame corners, in corner joints with mechanical fasteners,
the transport dimensions may be far more compact, since any assembly can take place
on the construction site itself. Here, the requirements in terms of assembly conditions
and the assembly effort required are both significantly lower than for glued joints. It is
also relatively easy to adapt beams and columns to the course of the internal moments
and forces using tapered cross‐sections. In comparison to curved frame corners, for
example, this approach allows for larger clearances.
Moment‐resisting joints, particularly with dowels arranged in a circular form which is
important for practice, are not regulated in EC 5, although the NA provides some addi‐
tional details, primarily concerning the effective number of fasteners. The design ap‐
proach presented in this article and the design of reinforcements is based on Heimeshoff
(1977) and not part of the NA.
E14.1 Influence of moment‐resisting joints on the
load‐bearing behaviour
When calculating the internal forces and moments of a structure, joints with mechanical
fasteners are often considered pinned or fixed. Figure E14‐5 (a) and Figure E14‐5 (b) show
the course of the bending moment for the horizontal member for both assumptions.
In reality, joints with mechanical fasteners are semi‐rigid with a stiffness Kr, see Fig‐
ure E14‐5 (c), which depends, among other things, on the arrangement, number and fas‐
tener types as well as the properties of the connected members. The stiffness of the joints
influences the deformation behaviour of the structure and, for statically indeterminate
systems, the distribution of internal forces and moments (Leijten, 1988; Komatsu, 1992).
q q q
q·ℓ2 Mj
0 12
q·ℓ2 Kr Mf Kr
q·ℓ2 24
8 (E·I)→ ∞
ℓ ℓ ℓ
495
Moment‐resisting joints
To take the influence of the rotation stiffness Kr of the joint on the distribution of internal
forces and moments into consideration, a coefficient r is defined and functions as a
benchmark for the ratio between rotation stiffness Kr of the joint and the bending stiff‐
ness (E∙I)Ri of the connected member:
E I Ri
Kr r r Kr (E14‐1)
I Ri
E
where
Kr Rotation stiffness of joint
(E∙I)Ri Bending stiffness of connected member (here: beam, index Ri)
ℓ Span of the connected member (here: beam)
The corner moment Mj shown in Figure E14‐5 (c) of a frame with a semi‐rigid corner joint
will assume a value which, depending on the coefficient r, varies between the limit cases
shown in Figure E14‐5 (a) and (b). This corner moment can be determined using equation
(E14‐2) and disregarding longitudinal and shear deformations:
q 2 1 q 2 1
Mj (E14‐2)
h
8 1.5 3 8 h 3 E I Ri
r 1.5
K r
where
h Frame height
ℓ Span or length of the beam
Ratio of the bending stiffnesses of beam and columns, see also Figure E14‐6 top
q Uniformly distributed load on beam
For the limit case “pinned“ (Kr = 0 → r = 0 ), equation (E14‐2) assumes the value Mj = 0
(Figure E14‐5 (a)), for the limit case „fixed“ (Kr → ∞ → r → ∞ ) and with = 0
((E ∙ I)St → ∞), the corner moment is Mj = (q ∙ ℓ2)/12 (Figure E14‐5 (b)). Using the coeffi‐
cient M, it is now possible to specify the influence of the joint stiffness on the corner
moment Mj in relation to the corner moment with moment‐resisting joint (Kr → ∞
→ r → ∞):
M j r
M (E14‐3)
M j r
496
Moment‐resisting joints
While Figure E14‐5 shows the values for the beam moments Mj and Mf for the lower and
upper limits of Kr and r, Figure E14‐6 shows an evaluation taking the stiffness Kr of the
frame corner into consideration. The diagrams reveal the influence of rotational spring
stiffness on the ratio M for various geometry and stiffness ratios.
When the values of the coefficient r are below 6, the corner moment is decreasing signi‐
ficantly. A joint is considered pinned if the ratio M ≤ 0.20, which corresponds to a value
for the coefficient r of 0.5. A fixed joint can be approximated at M ≥ 0.85, whereby the
coefficient r must reach values of at least 8 to 12. In all other cases (0.5 < r ≤ 8 ‐ 12),
the semi‐rigidity of the joint must be taken into consideration when determining the
internal forces and moments.
Kr (E·I)Ri Kr
(E·I)Ri
h D=
(E·I)St
(E·I)St
ℓ
1 1
0.8 0.8
0.6 0.6
= 0,5 = 2
M
M
0.4 0.4
ℓ/h = 2 ℓ/h = 2
0.2 ℓ/h = 10 0.2 ℓ/h = 10
ℓ/h = 20 ℓ/h = 20
0 0
0 5 10 15 20 25 0 5 10 15 20 25
r r
Figure E14‐6 Above: frame geometry. Below: Influence of the joint stiffness r on the course of M
for different ℓ/h‐ratios, below left: = 0.5; below right: = 2.
In Figure E14‐7, the ratio between the corner moment Mj and the field moment Mf of the
beam, depending on ℓ/h and r, is specified (where = 1.0). An economical design of
frame members is possible if the moments in the field and corners are approximately
equivalent. The structural engineer can dictate the moment distribution accordingly
when detailing the frame corner by selecting the joint stiffness appropriately. The rota‐
tional spring stiffness of the corner joint also dictates the deformation of the structure.
The lower the joint stiffness, the larger the deflection in the centre of the beam span.
497
Moment‐resisting joints
1.7 r → ∞
1.5
r = 20
1.3
r = 12
Mj /Mf
1.1 r = 8
0.9
r = 4
0.7
0.5
2 4 6 8 10
ℓ/h
Figure E14‐7 Influence of joint stiffness on the ratio between corner moment (Mj) and field moment (Mf)
for = 1 and different values of r.
E14.2 Design of moment‐resisting joints
When designing moment‐resisting joints, the force distribution in the joint and/or fas‐
teners must be taken into account and depends on the fastener configuration. The fol‐
lowing section focuses on the mechanical behaviour of a widespread configuration of
moment‐resisting joints, in which dowel‐type fasteners are arranged in circular form. The
moment‐resisting joint shown in Figure E14‐8 is taken as an example. In the joint layout
shown, the fasteners are subject to loads perpendicular to the fastener axis and in this
case, the angle i between the force and grain directions varies for each of the individual
fasteners. The behaviour of the wood in response to embedment stress, however, de‐
pends on the angle i to the grain direction, which results in the slip moduli for individual
fasteners varying within the joint (Ohashi and Sakamoto, 1989). However, this dependency
of slip moduli on the angle i between the force and grain directions is not taken into
account in the design.
(a) (c) xi
FM,i
ri
ω yi
ω
αi
M rj FM,k
(b) C rk
ω
FM,j dsk
Figure E14‐8 Moment‐resisting joint: (a) joint in parallel members, (b) frame corner with dowels arranged
in a circular form, (c) geometry and forces on fasteners. (STEP 1995 Article C16)
498
Moment‐resisting joints
To determine the rotational spring stiffness Kr, the mean values for the initial slip moduli
Kser of timber‐to‐timber joints with mechanical fasteners specified in EC 5 are used and
are independent of the angle between the force and grain directions (see also Table E1‐1).
For steel‐to‐timber joints, meanwhile, these values are doubled according to EC 5. The
mean value of the slip modulus in the ultimate limit state Ku,mean is 2/3 ∙ Kser (equation
(E1‐8)). This value is then to be divided by the partial safety factor M = 1.3. To determine
the load distribution, the members in the joint area are assumed to be infinitely rigid,
since they have far greater stiffness compared to fasteners. Considering the fastener
displacement allows the rotation of the joint to be calculated, which, in turn, depends on
the distance between the fasteners and the centre of rotation. This centre is generally
the centre of gravity of the fasteners. The equilibrium of moments based on the centre of
rotation C in accordance with Figure E14‐8 (c) is:
n
M FM,j r j (E14‐4)
j=1
where
FM,j Load on the jth fastener
rj Distance of the jth fastener to the centre of rotation C
n Number of fasteners
Assuming linear behaviour of the fasteners, the following relationships can be derived
(small angle: tan ≈ ), Figure E14‐8 (c):
FM,k ds k F FM,j F
ds k and tan M,k M,i (E14‐5)
Kk rk K k rk K j r j K i ri
where Kk/j/i = slip modulus of kth, jth, ith fastener in the force direction.
The load FM,i of the ith fastener exerted by the moment can now be calculated using equa‐
tions (E14‐4) and (E14‐5):
K i ri
FM,i M (E14‐6)
Kr
where Kr is the rotational spring stiffness of the joint:
n 2
K r K j rj (E14‐7)
j=1
499
Moment‐resisting joints
Considering the stiffness of the fasteners, the rotational spring stiffness of the joint can
be calculated for different fastener arrangements and depending on the joint geometry.
The rotational spring stiffnesses of the joints shown in Figure E14‐9, each with equal
fasteners, are specified in equations (E14‐8) and (E14‐9) while assuming linear‐elastic
load‐deformation behaviour.
For a circular fastener arrangement, Figure E14‐9 (a) applies:
K r K n 1 r 12 n 2 r 22 (E14‐8)
The rotational spring stiffness of the joint configuration in accordance with Figure E14‐9 (b)
is calculated as:
n n
K r K x i2 y i2 (E14‐9)
i=1 i=1
n1
≥ 6 ·d
r1
C C
r2 yi x
ymax ni
n2
y
xi
xmax
(a) (b)
Figure E14‐9 Geometry of different fastener arrangements in moment‐resisting joints:
(a) circular arrangement with two circles, (b) rectangular arrangement.
M r1 K r M K r
FM M
K (E14‐10)
n 1 r 12 n 2 r 22 r
2 2
K n 1 r1 n 2 r 2
500
Moment‐resisting joints
and for the rectangular arrangement shown in Figure E14‐9 (b):
2 2
M x max y max
FM n n
(E14‐11)
2 2
xi yi
i=1 i=1
The shear and normal force is uniformly distributed between the fasteners, Figure E14‐10,
meaning the following applies:
V N
FV and FN (E14‐12)
n n
where n = number of fasteners, n = n1 + n2 for two circles of fasteners, see Figure E14‐9 (a).
The overall loading of the respective fastener can be calculated by the vector addition of
FM, FV and FN, Figure E14‐10. The overall loading differs for each of the fasteners due to
the changing directions of FM. The greater the distance to the centre of rotation, the
higher the load exerted by FM on the respective fastener. Within a circle of fasteners
meanwhile, the load on a fastener peaks if the direction of the resulting force of FV and FN
corresponds to that of FM, as in Figure E14‐10 (b).
(a)
FV
FR FN FM FR
FVM FM
Fibre direction
member 1
Member 2
J
FV
FR D1
(b)
FM D2
FM FN FV FN D1
FVM FVM
F VM,max FM FR FM F V2 FN2
Member 1
Figure E14‐10 Loads on fasteners in a circular fastener arrangement with respect to member 1 = column.
(a) vector addition of forces, (b) maximal force on fastener if FR and FM have the same direction.
501
Moment‐resisting joints
Although the load‐bearing capacity of dowel‐type fasteners depends on the embedment
strength of the members to be connected, the influence of the angle between the force
and grain directions on embedment strength must be taken into consideration for fitted
bolts, dowels and, where applicable, screws. This means that the load‐bearing capacity of
each fastener varies depending on its position or load direction and, in theory, each
individual fastener of the joint would have to be verified. There is, however, scope to
simplify this verification with the theoretical maximum fastener load. With this in mind, it
is assumed that the resulting normal and shear forces and the force FM generated from
the moment act in the same direction (Figure E14‐10 (b)). This maximum load is as fol‐
lows for a circular fastener arrangement (index VM = fastener):
The angles between the force and grain directions are as follows in this case (Figure E14‐
10 (b)):
FV
1 arctan and 2 180 1 (E14‐14)
FN
where
1, 2 Angle between the force direction of the theoretical maximum load
and the grain direction of member 1 or 2
Angle between the axes of the beam and column
For a rectangular or trapezoidal fastener arrangement (Figure E14‐9 (b)), the theoretical
maximum value of FVM should be verified for the outermost fastener:
2 2
x max y max
F VM FN FM F V FM (E14‐15)
2
x max 2
y max 2
x max 2
y max
The corresponding angle between the force and grain directions results as:
y 2 2
max FM x max y max F V
1 arctan (E14‐16)
x 2 2
max FM x max y max FN
502
Moment‐resisting joints
E14.3 Timber stresses in the joint area
In addition to the loads on the fasteners, the stresses on the timber in the joint area must
also be taken into consideration. Figure E14‐11 shows two different fastener arrange‐
ments used to manufacture frame corners. The fastener arrangement marked with (b)
reveals a far less favourable combination of shear stresses and stresses perpendicular to
the grain compared to the circular arrangement (a), since higher tensile stresses perpen‐
dicular to the grain are generated over a larger member length and unfavourable interac‐
tion between tensile stresses perpendicular to the grain and shear stresses occurs
throughout the member length in question. For fastener arrangements along the mem‐
ber edges, meanwhile, higher tensile stresses perpendicular to the grain are observed
particularly in the vicinity of the end grain (Boult, 1988). The risk of splitting resulting
from these observations can be mitigated by reinforcement measures (e.g. glued‐on
wood‐based panels) or arranging fasteners with a high withdrawal capacity such as fully
threaded screws to secure corners (see Section E14.4). For non‐reinforced frame corners,
the load‐bearing capacity of the joint should be reduced by 15% if two circles of dowels
are chosen as a fastener arrangement (Heimeshoff, 1977; Kolb, 1970). It is also advisable
to select slender fasteners to further reduce the splitting risk.
(a) (b)
0 35 70 cm
0,2
σt,90 (N/mm 2)
τ (N/mm 2)
(b) -0,1
0,1
-0,3
35
0 -0,5
70 cm (b)
(a) (a)
-0,1
Figure E14‐11 Stresses in moment‐resisting joints. Above: fastener arrangements (a) and (b). Lower left: tensile
stresses perpendicular to the grain at 100 mm from the member end. Lower left: shear stresses in
the centre of gravity of the joint. (Racher and Gallimard, 1991. STEP 1995 Article C16)
503
Moment‐resisting joints
To verify the shear capacity of the timber members in the joint area, shear stresses
have to be derived, an example of which is shown using a joint with two circles of dow‐
els. To do so, the components in x‐direction of the forces transferred by the fasteners
and above the centre of gravity are added, governing section see Figure E14‐12 (a). For
this purpose, the forces in the fasteners are transformed in a line load f and here, the
following applies for the external circle of dowels (where 2 ∙ ∙ r1 = circumference and
n1 ∙ FM,1 = sum of the shear forces generated by the moment):
n 1 FM,1
f1 (E14‐17)
2 r1
and for the internal circle of dowels:
n 2 FM,2
f2 (E14‐18)
2 r2
As shown in Figure E14‐12 (b), based on the line load at a circle segment of size d, the
tangential force component f ∙ r ∙ d and the force component perpendicular to the grain
f ∙ r ∙ d ∙ sin (x‐direction in Figure E14‐12) can be determined. For both dowel circles,
the sum of all force components perpendicular to the grain lying in the semicircle above
the centre of gravity is calculated. This reveals the entire load component perpendicular
to the grain generated by the corner moment:
y y
Column axis r·dM
f1
M f·r·dM
n2 n1
r
f·r·dM ·sinM
r2 r1 dM r
Section M
x x
C C
Fibre direction
(a) (b)
Figure E14‐12 Derivation of stresses perpendicular to the grain in a circular arrangement of fasteners:
(a) governing section, (b) circle segment with division of forces.
504
Moment‐resisting joints
Equations (E14‐17) and (E14‐18) reveal that:
n 1 FM,1 n 2 FM,2
n 1 FM,1 n 2 FM,2
1
VM sin d sin d (E14‐20)
0 2 0 2
In addition to the load component from the bending moment in accordance with equa‐
tion (E14‐20), the component from the shear force must also be taken into consideration
to obtain the total shear force of the member:
V
V tot V M (E14‐21)
2
For the circular fastener arrangement, the design shear load can be calculated from
equation (E14‐21) where the shear force component VM generated by the moment,
equation (E14‐20), and the load FM generated by the moment, equation (E14‐10), are
inserted:
M n 1 r1 n 2 r 2 V
V tot (E14‐22)
n 1 r 12 n 2 r 22 2
For joint configurations with rectangular or trapezoidal fastener arrangements, the de‐
sign shear load can be derived accordingly. Various fastener arrangements were system‐
atically assessed by Kessel and Willemsen (1991).
For the rectangular arrangement shown in Figure E14‐11 (b), the design shear load in the
joint area can be calculated as follows:
V M 1 n V
V tot V M 2 2
yi (E14‐23)
2 2 x 1 y 1 i=1 2
505
Moment‐resisting joints
E14.4 Reinforcing measures for corners
The load direction of the fasteners varies according to their position, meaning the load‐
bearing capacity is not dependent on the number of fasteners arranged in a row. For
circular fastener arrangements, minimum spacings and minimum end and edge distances
of fasteners are not expressly regulated by design standards. Recommendations are
given in Table E14‐1 and the corresponding definitions are shown in Figure E14‐13.
If no reinforcements are arranged in a corner joint with multiple dowel circles, the load‐
bearing capacity of the joint should be reduced by 15%:
n ef 0.85 n (E14‐24)
In corner joints featuring only a single circle of dowels or where reinforcement measures
have been taken to prevent members splitting, nef = n may be assumed.
Table E14‐1 Recommendations for minimum spacings and edge and end distances for fasteners
arranged in circles.
506
Moment‐resisting joints
n1
r1
C r2 n2 a
a 2,t(c)
1,t
a2
a1
Figure E14‐13 Definition of minimum distances and arrangement of the corner reinforcement.
(Heimeshoff, 1977)
Reinforcing elements for beams and columns arranged to absorb tensile stresses perpen‐
dicular to the grain must be designed for a force Fax,d, which corresponds to the lateral load
FM,d of the fasteners of the external circle within a 30° sector (Heimeshoff, 1977):
30 1
F ax,d n 1 FM,d n 1 FM,d (E14‐25)
360 12
The reinforcements against tension perpendicular to the grain are to be arranged in the
external quarter of the corner in the column and beam as shown in Figure E14‐13. The
length of the reinforcing elements must be selected such as to prevent the risk of crack‐
ing tangential to the circle and parallel to the grain of the column or beam. Usual corner
designs thus include reinforcing elements with lengths 10 to 12 times the diameter of the
fasteners (dowels, fitted bolts) of the moment‐resisting joint.
507
Moment‐resisting joints
E14.5 Literature
P. Racher, original Article C16, STEP 1995.
Boult B.F. (1988). Multi‐nailed moment resisting joints. International Timber Engineering Conference,
Vol. 2:329‐338, Seattle.
Heimeshoff B. (1977). Berechnung von Rahmenecken mit Dübelanschluss (Dübelkreis). Arbeitsgemeinschaft
Holz, Holzbau‐Statik‐Aktuell, Folge 2, Düsseldorf.
Kessel H. and Willemsen T. (1991). Zur Berechnung biegesteifer Anschlüsse.
Bauen mit Holz 93:342‐352.
Kolb H. (1970). Festigkeitsverhalten von Rahmenecken. Bauen mit Holz 72:387.
Komatsu K. and Kawamoto N. (1992). Analysis of glulam semi‐rigid portal frames under long‐term load.
Paper 25‐8‐1, CIB‐W18 Meeting 25, Åhus.
Leijten A.J.M. (1988). Steel reinforced joints with dowels and bolts. International Timber Conference,
Vol. 2:474‐488, Seattle.
Ohashi Y. and Sakamoto I. (1989). Study on laminated timber moment resisting joint. 2nd Pacififc Timber
Engineering Conference, Vol. 2:37‐42, Auckland.
Racher P. and Gallimard P. (1992). Les assemblages de structures bois: a) comportement mécanique des
principeaux types assemblages; b) analyse du fonctionnnement d’une couronne boulonnée.
Annals ITBTP, No. 504, pp. 29‐40.
508
E15 Joints with multiple
shear planes
Original articles: H. Hartl, A. J. M. Leijten, B. O. Hilson
Trusses of various forms are primarily used in roof structures and bridges. Their ad‐
vantages include a low dead weight and resulting material saving, while downsides in‐
clude the significant construction height required in the centre of the truss compared to
simple beams. Stand‐out structural features include the in‐plane shape of a truss and the
joints between the chords, posts and diagonals, as well as any joints within the chords.
Since high forces and sometimes moments occur in the joints, there is a need for particu‐
larly high‐performance joints. In addition to joints with hinge bolts and steel nailing
plates, those with slotted‐in steel plates and connectors, double‐shear joints with lateral‐
ly loaded dowel‐type fasteners and step or bonded joints, this is the main area of applica‐
tion for joints with multiple shear planes. Given the distribution of fastener loads over
several shear planes, their compact joint areas make them the first choice to meet the
construction requirements for joints in trusses. Multiple‐shear steel‐to‐timber joints with
covered dowels and individually adapted gusset plates even meet the dual requirements
of visual appeal and effective fire resistance.
E15.1 Joints with multiple shear planes – construction types
In addition to the basic cases for fasteners in single‐ or double‐shear, trusses also feature
joints between members made of wood and wood‐based materials which comprise more
than three individual components (basic case of fastener in double‐shear). Such joints
arise, for instance, if the upper and lower chords comprise two or even three individual
components, depending on the area involved. The posts and diagonals are laid between
and/or outside the chord members. Furthermore, the limited length of sawn timber and
the subsequent need for splice joints lead to tension joints within the two‐component
tension chords with internal and external wood members. This results in joints involving
five or more different components arranged adjacent, while continuous fasteners have
four or more shear planes accordingly. Such trusses would typically use what was previ‐
ously a popular option but is now rarely seen, the nailed truss. The scope of trusses also
includes steel‐to‐timber joints, within which two or more gusset plates are inserted into
correspondingly tailored slots (or additionally externally attached), examples of which are
included in Figure E15‐1.
509
Joints with multiple shear planes
Figure E15‐1 Left: GREIM system: truss joint between one post, two diagonals and lower chord, respectively
made of glulam, nailed joint with two galvanised gusset plates. Right: connection BSB: slotted‐in
and predrilled glulam, galvanised predrilled 5 mm thick gusset plates and 6.3 mm thick dowels in
lengths as required.
E15.2 Basis of design
To determine the total load‐bearing capacity of joints with multiple shear planes, addi‐
tional points must be considered when applying the basic Johansen cases discussed in
Article E2. To calculate the total load‐bearing capacity, the initial step involves consider‐
ing plausible combination of possible individual failure modes (embedment failure in the
wood, plastic hinges of the fastener) over the entire fastener length. Figure E15‐2 pro‐
vides an overview of the possible failure modes of a fastener with four shear planes for a
timber‐to‐timber and a steel‐to‐timber joint. While the failure modes shown are practi‐
cally feasible, they may not all be preferable as far as structural aspects are concerned.
When using thin steel plates, an additional failure mode not shown in Figure E15‐2 may
also occur, since the thin steel plate cannot provide any clamping effect and the fastener
may deform as shown in Figure E15‐3 if the timber side members are thin.
510
Joints with multiple shear planes
1 3 5 1 3 5
2 4
2 4
Figure E15‐2 Failure modes of a fastener with four shear planes in a timber‐to‐timber (left) and
a steel‐to‐timber connection (right).
1 3 5
2 4
Figure E15‐3 Possible deformation pattern of a fastener with four shear planes in a steel‐to‐timber connection
with thin steel plates and thin timber side members (see case k in Figure E15‐4).
511
Joints with multiple shear planes
The member thicknesses are constant in the qualitative representation in Figure E15‐2. In
reality, however, the mechanisms shown are strongly influenced by the individual mem‐
ber thickness and the corresponding embedment strength. The shaded areas indicate
cavities which emerge after the respective displacement or rotation of a fastener relative
to the member, i.e. where the wood matrix was “squeezed” by the fastener. It is assumed
that the individual cross‐sections 1, 3 and 5 are loaded in one direction and 2 and 4 in the
other. The lower part of the illustration, meanwhile, shows the failure mechanisms with
two plastic hinges per shear plane. This mechanism facilitates the exceptionally ductile
nature of the joint, mitigates the overall splitting risk for solid and glulam timber and is
ultimately a favoured option.
The total load‐bearing capacity of a joint with multiple shear planes is ultimately de‐
termined as the sum of the minimum load‐bearing capacities of all shear planes. As
when designing the basic cases, likewise, ideal rigid‐plastic behaviour can be assumed for
fasteners under bending stress and the wood and wood‐based materials under embed‐
ment stress when determining the load‐bearing capacity per shear plane and fastener.
The latter is calculated assuming that each shear plane is part of a double‐shear joint. The
relevant basic cases in accordance with Johansen, as already explained in Article E2, are
shown in Figure E15‐4. In shear planes which connect edge members (cross‐sections 1 or
5), basically all failure modes are possible. The deformation pattern of a fastener, whose
end is rotated (Figure E15‐4: failure mode j for timber‐to‐timber joints and g and k for
steel‐to‐timber joints, k only for thin steel plates, see also Figure E15‐3), however, is only
mechanically possible for an actual edge member. This deformation pattern, with rotated
fastener ends, is not possible in centrally placed members (cross‐sections 2, 3 or 4),
which are “at the edges” only in the theoretical observation, since the deformed fastener
must continually penetrate all individual members. When designing joints, another as‐
pect to consider is the fact that total load‐bearing capacity may be limited by shear fail‐
ure along the external rows of fasteners or tensile failure of the wood (so‐called block
shear failure, Article E13). The step‐by‐step procedure used to determine the total load‐
bearing capacity is qualitatively shown in Figure E15‐5.
g h j k f g h j/l k m
Figure E15‐4 Basic cases in accordance with Johansen: failure modes of double‐shear joints.
512
Joints with multiple shear planes
Capacity
SP I (and IV):
t1 = b1
t2 = b2
t1 t2 t1
Capacity
SP II (and III):
t1 = b2 /2
t2 = b3
t1 t2 t1
Two different geometries
have to be considered when
calculating the capacity of
Capacity SP II resp. III.
SP II (and III):
t1 = b3 /2
t2 = b2
t1 t1
t2
E15.3 Literature
H. Hartl, A.J.M. Leijten, B.O. Hilson, original Articles D1, 7 (volume 3), STEP 1995.
513
F
Serviceability
F1 Deformations
Original article: S. Thelandersson
The overall performance of structures should always meet two basic requirements. The
first is structural safety, which is generally quantified by load‐bearing capacity and the
second is serviceability, which generally refers to the ability of the construction and its
individual parts to perform satisfactorily under normal usage conditions. The serviceabil‐
ity limit state also encompasses the wellbeing of users, e.g. restricting unpleasant floor
vibrations, and the visual appearance of the construction, e.g. impairments caused by
excessive deformations. As a general rule, insufficient structural safety is a threat to
human life and may also involve the risk of considerable damage costs. Conversely, ex‐
ceeding the serviceability limit state rarely has consequences which impact on personal
safety and any relevant economic risks are also likely to be low. Almost all cases of struc‐
tural damage which occur in practice are attributable to a lack of serviceability, reflecting
its importance when designing structures. For horizontal timber structures such as floors
or roofs, serviceability requirements such as limiting deflections or vibrations are very
often a decisive parameter when it comes to cross‐sectional dimensions.
F1.1 Deformations over the lifetime of a structure
The portion of variable actions (e.g. imposed loads on floors or snow loads) in timber
constructions tends to exceed the portion of permanent actions. This means that as the
actions change considerably over time, the degree of deformations does the same and
this fact must be taken into account appropriately when verifying serviceability. Fig‐
ure F1‐1 schematically shows the deflection history of a beam subject to permanent load
and snow loads (Mårtensson, 1992; Mårtensson and Thelandersson, 1992). The total
deflection is the sum of the deflection 1 from the permanent load directly after the load
is applied and the deflection 2, which changes over time. 2 can be considered irreversi‐
ble for the purpose of practical applications (Mårtensson, 1992). Short‐term peak loads,
such as those shown in Figure F1‐1, emerge as either snow loads or imposed loads in the
most common types of buildings.
517
Deformations
With respect to the type of time‐dependent behaviour shown in Figure F1‐1, distinctions
can be made between the following deflection portions (see Figure F1‐2):
0 is the precamber in the unloaded state (state 0),
1 is the deflection due to permanent load directly after the load is applied
(state 1),
2 is the deflection due to variable load 2,inst plus any time‐dependent portions
from permanent load creep (state 2) and
net is the deflection (sag) of the beam relative to a straight line connecting the
supports (= unet,fin).
As a general rule, portions 0 and 1 are fixed as soon as the construction is complete.
However, these portions may still be subject to change when the permanent load changes.
2 and net, conversely, change over time.
F
Qk
Gk
1 2 3 4 5 6 46 47 48 49 50
t
(a)
δ
δ2,inst
δ2
δcreep
δ1
1 2 3 4 5 6 46 47 48 49 50
t
Figure F1‐1 Illustration of the time‐dependent deflection of a beam under permanent (G) and variable (Q)
action. The curve (a) shows the resulting deflection when the beam is permanently loaded by
the characteristic values Gk + Qk. F is the load and t the time in years. (STEP 1995 Article A17)
518
Deformations
(0)
δ1 δ0
(1) δ2 δnet
(2)
l
Figure F1‐2 Deflection portions of a simply supported beam. (STEP 1995 Article A17)
F1.2 Load combinations for the serviceability limit state
The design principles valid for all Eurocodes define the various load combinations used to
verify the serviceability limit state. For timber constructions, the characteristic and quasi‐
permanent combinations are particularly significant and the former is most relevant
when exceeding the limit state triggers permanent damage or permanent and excessive
deformations. This combination is used to determine the initial deformation and the
symbolic notation of this combination is as follows (see also Equation (C2‐5)):
The quasi‐permanent combination applies for long‐term effects and is used to determine
the portion of creep deformation creep. The symbolic notation of the quasi‐permanent
combination is as follows (see also Equation (C2‐6):
519
Deformations
F1.3 Limitation of deformations
The most common reasons for limitations applied to deformations in structures are:
General usability and visual appearance (e.g. restricting disruptive visible
deflections or avoiding uneven ground),
Structural requirements (e.g. avoiding damage to non‐load‐bearing members such
as partition walls or façade claddings and guaranteeing problem‐free assembly,
water‐ and airtightness or allowing water to discharge properly from roofs),
Requirements concerning building equipment (e.g. guaranteeing the flawless
functioning of machines, cables, exhaust air ducts and their supports).
Modern design standards such as EC 5 only cover general functional requirements, e.g.
stipulating that structures must be designed and built to ensure the abovementioned
serviceability perspectives can be taken into consideration. Numeric values for defor‐
mation limits must be defined from case to case by the structural engineer in consul‐
tation with the building owner; EC 5 only provides recommendations.
F1.4 Deflection limits to avoid damage
Deflections should be limited, when there is a risk of permanent damage to partition
walls, installations, fittings and claddings if this is not done. In this case, it should be high‐
ly likely that the deflection limit will not be exceeded. For this purpose, the deflection is
calculated with the characteristic load combination (F1‐1) and any damage in this case is
generally due to deformations which occur once the construction is complete. The formal
condition applies to circumstances in which deflections can result in permanent damage:
whereby uinst represents the initial deformation for the characteristic combination, name‐
ly the portion 1 from Figure F1‐2 and the portions of the variable loads on the initial
deformation, the time‐dependent nature of which is reflected in the combination coeffi‐
cient 0,i. uinst,crit is the limit value, which, if exceeded, may cause damage. The value of
uinst,crit depends on the type and connection details of the members that can be damaged.
In the absence of any more accurate details, uinst,crit is assumed to be either a fixed abso‐
lute value, e.g. 20 mm, or a percentage of the span ℓ, e.g. ℓ/300 to ℓ/500 for simply
supported beams and ℓ/150 to ℓ/250 for cantilever beams (EC 5 Table 7.2). These values
are often recommended for beams in floors and roofs, which are directly connected with
partition walls or other non‐load‐bearing members.
520
Deformations
F1.5 General usability and visual appearance
With regard to general usability and the visual appearance, limiting excessive deflections
which may be permanent or occur over an extended period is often desirable. However, it
may be permissible to exceed the limit values at times, if the deformations are reversible
and/or limited to a short period. In this case, a higher probability of occurrence can be
accepted for the limit values, whereupon additional deformation due to creep is calculated
with the quasi‐permanent load combination (F1‐2) and added to the initial deformation uinst
(characteristic combination). The corresponding condition in this case is as follows:
Here, unet,fin denotes the sag of the beam relative to a straight line connecting the sup‐
ports (see Figure F1‐2, value net), while ufin,crit is the deflection limit value with respect to
the general visual appearance. The value of ufin,crit depends, among other things, on the
type or style of construction and the attitude of building users as well as whether or not
the beam is visible. The requirements for residential buildings tend to be far more strin‐
gent than those for industrial buildings. As a general recommendation for simply sup‐
ported beams, the value ufin,crit = ℓ/150 to ℓ/300 is specified; while for cantilever beams
ufin,crit = ℓ/75 to ℓ/150 applies (EC 5 Table 7.2).
521
Deformations
F1.6 Calculation of deflection in accordance with EC 5
The elastic initial deformations caused by bending um can generally be calculated based
on the simple beam theory and using formulas from corresponding textbooks. Given the
low shear modulus of the wood compared to the modulus of elasticity, it is not always
advisable to disregard shear deformations. Details used to calculate the shear defor‐
mation uv can also be taken from textbooks and the overall initial deflection uins is ulti‐
mately revealed as the sum of um and uv. To get an idea of the order of magnitude of the
shear deformations, a simply supported beam with rectangular cross‐section b ∙ h, span ℓ
and subject to a uniformly distributed load q is observed. In this case, the ratio between
deformation due to shear uv and that due to bending um at mid‐span of the beam can be
approximated as:
5 q 4
At mid ‐ span: um
384 E I
Q 1 q 1 q q
uv q x uv x x x 2
G As GAs 2 G As 2 2
2 2 (F1‐5)
q q
At mid ‐ span: uv x
2 8 G A s 8 G 5 A
6
2
uv E h
Ratio : 0.96
um G
522
Deformations
in moisture content tends to be far slower in the larger cross‐sections. Surface treat‐
ments which prevent the exchange of moisture between wood and the surrounding air
have the same effect (Mårtensson, 1992; Taylor et al., 1991).
As well as load‐related deformations, the serviceability of timber structures is also im‐
pacted by the swelling and shrinking of the wood. The deformations in the wood trig‐
gered by changing moisture content may well be of the same order of magnitude as
those caused by external loads. However, swelling and shrinking effects can be limited by
employing a corresponding construction; particularly in terms of fine details and ensuring
the use of dry wood. Deformations solely attributable to moisture changes will not be
covered in any further detail in the current article.
Rules to calculate deflections are specified as non‐binding in EC 5. In accordance with
these rules, the elastic initial deformations uinst are to be calculated based on the average
values of the corresponding stiffness parameters. For structures made of members with
the same creep properties, the final deformation ufin can be simplified as follows, assum‐
ing a linear relation between load and deformation:
u fin u inst,G 1 k def u inst,Q,1 1 2,1 k def u inst,Q,i 0,i 2,i k def (F1‐6)
i>1
where kdef is the creep coefficient, which takes into consideration the increase in defor‐
mations over time attributable to the surrounding climate and the load duration class.
Values for kdef are specified in EC 5 Table 3.2 for various construction materials and differ‐
ing service classes.
F1.7 Literature
S. Thelandersson, original Article A17, STEP 1995.
Mårtensson A. (1992). Mechanical behaviour of wood exposed to humidity variations. Report TVBK‐1006,
Department of Structural Engineering, Lund University.
Mårtensson A. and Thelandersson S. (1992). Control of deflections in timber structures with reference to EC 5.
Paper 25‐102‐2, CIB‐W18 Meeting 25, Åhus.
Taylor G.D., West D.J. and Hilson B.O. (1991). Creep of glued laminated timber under conditions of varying
humidity. International Timber Engineering Conference, London.
523
F2 Vibrations
Original article: S. Ohlsson
The vibration verifications introduced in this article are those at the serviceability limit
state, since most people perceive excessive vibrations as disruptive. The main focus in
this case is on floor vibrations. Vibrations caused by machines and floor vibrations in
rooms subject to rhythmical loading (e.g. dance or sports halls), must be examined in
more detail and are not covered in this article. Vibrations are dynamic phenomena and
the dynamic behaviour of floors is governed by the floors’ mass, stiffness, damping and
hence their natural frequency and eigenforms. To assess such complex dynamic behav‐
iour cost‐effectively, EC 5 replaces the actual dynamic load with equivalent vertical, static
loads (unit impulse and equivalent static load). DIN 1052 stipulated that when verifying
floor vibrations, the deflection due to quasi‐permanent load should not exceed 6 mm,
which limited the scope to a stiffness criterion and excluded any additional criteria, such
as natural frequency limitations. EC 5, conversely, includes a more complex verification
method, which means so‐called unit impulse velocity responses are reviewed as well as
stiffness and frequency criteria.
Although methods for a (simplified) vibration verification process already exist, in prac‐
tice, numerous problems remain with floor vibrations. Users dislike the latter, even if the
extent of such vibrations may be within the verification limit values specified in EC 5.
Newer research projects (e.g. Hamm and Richter, 2009) have re‐examined the vibration
behaviour of timber and timber‐composite floors and assessed the validity or relevance
of the EC 5 verifications. The influence of the floor build‐ups (e.g. with or without a
screed) was also examined. New design and construction rules are discussed to round off
this article, since they also pave the way to verify floor vibrations more reliably.
525
Vibrations
F2.1 Serviceability requirements
The fitness for purpose of a building depends on the extent to which it fulfils its pre‐
scribed functions in due form. All aspects of serviceability defined by the static system or
the load‐bearing members of a building can be summarised under the collective term
“structural serviceability”. Structural serviceability requirements are normally building‐
wide or on key building sections.
Most serviceability requirements are a factor of the following conditions:
Acceptable human comfort (comfortable structural behaviour),
Guaranteed functionality of the building, including installations and
Acceptable visual building appearance.
These conditions are set out in detail in EC 5. Damage to claddings or partition walls may
e.g. lead to a loss of functionality due to leaking floor decking in a bathroom or adversely
affect the visual appearance due to visible cracks.
Serviceability requirements differ significantly from load‐bearing capacity requirements
at times. Unlike ultimate limit states, where there is no possibility of reversing the pro‐
cess when exceeded, when serviceability limit states are reached, such as the unpleasant
feeling of excessive timber floor vibrations, reversal is possible. In both examples, the
limit state is reached due to deviations from the expected straight and level flooring. For
the serviceability limit condition, the key parameter is therefore the sum of all deviations,
collectively comprising a portion caused by external loads and one caused by climatic
influences. In this case, a precamber of the beams makes it possible to limit the total
deflection to tolerable values. A consideration of precamber is included in EC 5. However,
precamber has no influence on the beam vibration behaviour.
The limit values recommended in EC 5 for deflections and vibrations generally guarantee
compliance with serviceability requirements. The following section covers these require‐
ments with respect to vibrations in more detail. For all verifications of the serviceability
limit state, the rule is that the specified limit values are only recommended; the struc‐
tural engineer must define limit values or required verifications on a case‐by‐case
basis with the building owner.
526
Vibrations
F2.2 Serviceability concerning vibrations
Vibrations which limit serviceability can be caused by wide‐ranging loads and avoiding
vibrations perceived as unpleasant by building users is particularly important. The sensi‐
tivity of the people concerned is thus the main factor dictating the limit of what are
considered permissible vibrations. In timber frame constructions, human activities and
installed machines both constitute key sources of vibrations. with the former usually
comprising footfall from human traffic, running, or children playing around.
The key consequences of these causes are distinguished as follows:
Discomfort due to vibrations caused by steps and
Discomfort due to vibrations caused by machines.
The way different people react to vibrations is subjective and influenced by myriad fac‐
tors. One comprehensive presentation of the relevant issues is included in Griffin (1990).
The following relationships apply to most situations:
The human perception of vibration
Depends on the vibration acceleration for frequencies of less than around 8 Hz,
Depends on the vibration velocity for frequencies exceeding 8 Hz,
Exhibits a logarithmic characteristic, resembling the subjective perception
of volume,
Increases with the vibration duration,
Declines with proximity to and awareness of the vibration source, and
Declines with increasing physical activity or familiarity.
Two design goals can be derived from the above relationships: first of all, the extent of
vibrations in proximity to the vibration source should be limited, which can be ensured by
designing the timber floor in accordance with the approach described in the following
sections. In addition, transmission of vibrations to neighbouring construction sections
should be prevented.
527
Vibrations
With this in mind, a suitable static system should be chosen, including consideration of
the following points:
Timber frame structures allow vibrations to be transferred to neighbouring storeys,
which makes them less suitable for vibration‐prone constructions. For the same
reason, continuous floors between neighbouring residences are best avoided.
Partition walls should be located above each other and constructed right down to
the foundation. If a partition wall e.g. is only present in one storey and erected at
the mid‐span of a timber floor, this wall may structurally couple the two adjoining
floors and generate reciprocal vibrations. While these vibrations may be accepta‐
ble within the storey in which the vibration source itself is located, they often
cause discomfort in other storeys, where the vibration source is unknown.
F2.3 Human‐induced vibrations
The initial section already explained that in DIN 1052, the scope for limiting vibrations
only included deflection. This requirement included defining a stiffness criterion, since a
low bending stiffness of a floor leads to the deflection limit value being exceeded. The
greater the floor span, the higher the level of deflection and thus the lower the natural
frequency. A floor with a smaller span will thus oscillate with a smaller period (= higher
frequency) than one with a larger span. The natural frequency f can be represented de‐
pending on the stiffness and hence the static deflection ustat of a floor, see equation (F2‐1).
Equation (F2‐1) is based on a model of the floor as an undamped single degree of free‐
dom system with stiffness K, mass M and load F:
F
1 K 1 u stat 1 g 5
f (F2‐1)
2 2 M 2 F
g
2 u stat u stat cm
By limiting deflection, the natural frequency of a timber floor can be minimised. The first
natural frequency of a floor should be at least double the frequency of human‐induced
vibrations, since frequently repeated stimulation at a frequency of up to half that of the
natural frequency can result in resonance (Hamm and Richter, 2008). One step per sec‐
ond corresponds to 1 Hz. Activities such as running have a frequency of up to 4 Hz. Ac‐
cordingly, the desirable criterion is normally a natural frequency of at least 8 Hz.
528
Vibrations
The relationships observed shall now be explained in more detail. The vibration design of
timber floors in accordance with EC 5 takes into consideration everyday human activities
like the dynamic effects of footfall, with scope to determine the latter via tests (Ohlsson,
1982). Figure F2‐1 shows the course of the load induced by a person treading in place
over time. In residential buildings, the swiftly declining short‐term reaction prevails and
the load comprises two differing components:
A low‐frequency component (0 ‐ 8 Hz), caused by the step frequency and its
corresponding harmonic vibrations plus
A high‐frequency component (8 ‐ 40 Hz), primarily due to the contact between
the heel and floor, the so‐called heel drop.
If the natural frequency of the floor exceeds 8 Hz, the low‐frequency components gener‐
ate quasi‐static vibrations with amplitudes mainly determined by floor stiffness rather
than mass.
1,0
F (kN)
0,0
1,0
Fagg (kN)
0,0
0,0 2,0
t (s)
Figure F2‐1 Course of the contact forces generated by the steps of a walking person (top) and
corresponding load on the floor (below). (STEP 1995 Article A18)
529
Vibrations
The first natural frequency f1 of a rectangular timber beam floor with pinned supports
along all four edges can be determined with the approximate equation (F2‐2). Equation
(F2‐2) shows how the “natural frequency” of a simply supported beam is determined in
very simplified form; “natural frequency”, since a beam, unlike a single degree of free‐
dom system (simple harmonic oscillator), has not one but innumerable natural frequen‐
cies, the lowest thereof is designated as the first natural frequency:
E I
f1 2
f 1,beam (F2‐2)
2 M
where
M Mass per unit of area in kg/m2,
ℓ Floor span in m
(E ∙ I)ℓ Equivalent bending stiffness of the floor about an axis perpendicular to the
beam axis in Nm2/m
For biaxial floors (e.g. massive timber floors using cross‐laminated timber) however,
equation (F2‐2) is overly simplified, since the effective bending stiffness in the perpen‐
dicular direction increases the first natural frequency.
The natural frequency of such floors can be determined using equation (F2‐3):
1 b E I
f plate f beam 1 where 4 (F2‐3)
4
E I b
where
b Floor width in m
(E ∙ I)b Equivalent bending stiffness of the floor about an axis parallel to the
beam axis in Nm2/m
From this, the initial two criteria of EC 5 can be derived:
Frequency criterion:
f 1 8 Hz f lim (F2‐4)
Stiffness criterion:
u mm
a a lim in (F2‐5)
F kN
530
Vibrations
To avoid resonance, a limit value is defined for the first natural frequency of the floor and
for ratio a = u/F from the vertical deflection u due to a vertical static load F. According to
Kreuzinger and Mohr (1999), both criteria correspond to two of the three categories of
vibration behaviour depending on the type of stimulation involved. The stiffness criterion
limits the vibration due to a one‐time deflection caused by a footstep, while the fre‐
quency criterion helps mitigate the risk of resonance caused by repeated steps and pre‐
vents any build‐up of vibration amplitude. The final category, namely the impact of the
high‐frequency components caused by heel drop, is taken into consideration by a unit
impulse of 1 Ns. The resulting vibration velocity v is a characteristic of the structure. If we
take the floor as a freely movable and rigid body of mass M, it would be accelerated by
the unit impulse to a velocity of 1 (m/s) /M.
According to Ohlsson (1988), the initial maximum value of velocity under a unit impulse,
the so‐called unit impulse velocity response v, can be determined in simplified form for
the most common case in practice, a rectangular timber beam floor with pinned supports
along all four edges. For a span ℓ in m, a floor width b in m and a uniformly distributed
mass M in kg/m2, the following approximation applies:
4 0.4 0.6 n 40
v (F2‐6)
M b 200
In equation (F2‐6), n40 denotes the number of eigenmodes with natural frequencies of
less than 40 Hz and M ∙ b ∙ ℓ is the floor mass. An additional generalised mass of 50 kg is
included in the quotient 4/200, which represents a vibrational portion of the body of the
person supposedly disturbed by the vibration. n40 can be determined with the following
approximate equation:
0.25
2 b 4 E I
40
n 40 1 (F2‐7)
f 1 E I b
This simplifies the task of determining the maximum vibration velocity caused by a unit
impulse, whereby the number of eigenmodes having to be taken into consideration is
limited to natural frequencies below 40 Hz, since tests (Ohlsson, 1982) showed that
natural frequencies over 40 Hz no longer contribute significantly to the initial maximum
velocity vmax.
531
Vibrations
The favourable effect of a short vibration period is taken into consideration via a limit
value depending on damping. The key damping parameter in this interaction is the damp‐
ing coefficient 0, which indicates the decline in vibration amplitudes depending on time
rather than the number of periods. The damping factor 0 is defined as:
0 f (F2‐8)
f 1 1 m
vb v lim in (F2‐9)
Ns 2
150
130 1
110
b
90
2
70
50
0 1 2 3 4
a [mm/kN]
Figure F2‐2 Relationship between a and b according to EC 5, 1 = better behaviour, 2 = worse behaviour.
532
Vibrations
F2.4 Recent research
All floors examined in Hamm and Richter (2009) met the criterion for unit impulse veloci‐
ty response in accordance with equation (F2‐9), although a portion of the floors in ques‐
tion were classified by users as unpleasant, hence the questionable significance of this
criterion. Hamm and Richter were also able to show that floor build‐ups such as screed
use as well as the actual detailing, e.g. using a floating screed, have a key impact on the
vibration behaviour. Moreover, to date, no distinction has been made between differing
requirements for floors depending on their position, namely as a floor between or within
units of use respectively. Hamm and Richter thus propose disregarding the verification of
unit impulse velocity response and instead limiting the scope to the frequency and stiff‐
ness criteria, which, depending on usage, can be subdivided into two further categories.
Moreover, requirements for structural detailing are also imposed and a value for the limit
acceleration in the event of non‐compliance with the frequency criterion. All details are
summarised in Table F2‐1. In Hamm (2012), a design example is included.
533
Vibrations
Table F2‐1 Relationship between usage and requirements for design and structural detailing,
from: Hamm and Richter, 2009, Table 3.
F2.5 Literature
S. Ohlsson, original Article A18, STEP 1995.
Chui Y.H. (1988). Evaluation of vibrational performance of light‐weight wooden floors. International Conference
on Timber Engineering 1:707‐715, Forest Products Research Society, Madison.
Griffin M.J. (1990). Handbook for human vibration. Academic Press, London.
Kreuzinger H. and Mohr B. (1999). Gebrauchstauglichkeit von Wohnungsdecken aus Holz; Abschlussbericht.
TU München, Fachgebiet Holzbau.
Hamm P. and Richter A. (2009). Bemessungs‐ und Konstruktionsregeln zum Schwingungsnachweis von
Holzdecken. Fachtagungen Holzbau 2009. Leinfelden‐Echterdingen, 15‐29.
Hamm P. (2012). Schwingungen bei Holzdecken – Konstruktionsregeln für die Praxis. 2. Internationales Forum
Holzbau, Beaune.
Ohlsson S. (1982). Floor vibration and human discomfort. Dissertation, Chalmers University of Technology,
Göteborg.
Ohlsson S. (1988). Springiness and human‐induced floor vibration. Document D12, Swedish Council for Building
Research, Stockholm.
534
G
Accidental loads
and additions
G1 Reaction to fire and fire design
Original article: H. Hartl
Describing the way materials behave in response to a fire simply is an almost impossible
task. First of all, fire can be divided into two phases, the developing phase and the fully
developed fire. Material behaviour, meanwhile, must be classified taking both these
phases into consideration, where the reaction to fire involving a material describes its
behaviour in the developing phase. The developing fire is influenced by a series of differ‐
ing parameters, such as the combustibility of the material, the degree of inflammability,
the propagation velocity of the fire over the surface and the rate at which heat is re‐
leased. The fully developed fire represents the phase after the flash over, in which all
combustible materials become involved in the fire. Desirable material properties during
this phase include the ability to retain load‐bearing capacity, restrict the spread of fire to
the area in which it originated to avoid any propagation of flames, hot gases or excessive
heat on the side of a wall or floor opposite a fire, which would indirectly help the fire
spread to neighbouring areas. The ability to withstand the fully developed fire is generally
known as fire resistance, but can only be attributed to a member and not a material. The
behaviour of simple structural members such as columns or beams also depends on
factors like the support conditions, magnitude of load and load distribution.
Observing how solid timber and wood‐based materials behave in a developing fire, it
becomes clear that wood‐based materials burn and are thus classified as combustible.
While this combustibility can be modified by applying a coating or impregnating with fire
retardants, these measures cannot render wood‐based materials non‐combustible, de‐
spite the fact of more energy is required to make the material burn. Solid timber is diffi‐
cult to ignite and there are few cases in which wood ignites earlier than other materials.
Solid timber requires a surface temperature exceeding 400°C within a short to medium
time frame for spontaneous combustion, namely without the presence of an ignition
source. Even with an ignition source, the surface temperature must be maintained at
more than 300°C for a specific period, whereupon the material will ignite. Wood is nor‐
mally used as a benchmark material for comparison when classifying other materials,
since wood is seen as a material, which, in most areas of use, presents an acceptable
ignition risk. The actual values depend on the density, wood species, moisture content
and the ratio of the circumference to the cross‐section. The fire propagates over the
surface of the combustible wood, and each ignition in the surrounding area triggers
further ignitions. Since wood is difficult to ignite, the speed with which the flames spread
537
Reaction to fire and fire design
is acceptably low for a combustible material. Virtually all countries allow the use of un‐
treated wood in instances involving a low fire risk. The degree to which wood gives off
heat depends on the type of temperature curve and the availability of oxygen as well as
the density, circumference and size of the observed timber piece.
If wood‐based materials are exposed to a fully developed fire, they respond with a series
of favourable characteristics. The surfaces will ignite once the heat flow has reached a
sufficient level and initially burn fairly vigorously; an insulating charcoal layer soon forms
(see also Figure G1‐9). However, since wood and charcoal in particular are poor thermal
conductors, very little heat is transferred to the remaining unburnt material. For this
reason, the construction is not damaged by any excess thermal expansions. If wood‐
based panels are used as load‐bearing elements or as sheathing in partitioning construc‐
tions, meanwhile, the low thermal conductivity prevents the swift transfer of heat from
the hot side of the construction to the colder side.
In tests, the fully developed fire is described in terms of the standard temperature‐time
curve in accordance with ISO 834 (see Figure G1‐1) or in accordance with the corre‐
sponding national standard. The key (so‐called REI) criteria include compliance with the
Load‐bearing capacity – criterion R,
Integrity – criterion E,
Insulation – criterion I.
These REI‐criteria also govern design. Structures must be dimensioned such as to ensure
they can retain their load‐bearing capacity during fire of a certain duration (R) as well as,
when required, their integrity (E) and insulation capacity (I).
The standard temperature‐time curve in Figure G1‐1 corresponds to a fictional fire, which
differs significantly from a natural fire, since the latter also includes cooling phases. The
favourable behaviour of wood in a cooling phase, however, is not taken into account
when loading with the standard temperature‐time curve. Moreover, a natural fire also
tends to reach considerably higher temperatures much more rapidly and over a shorter
time period, before quickly recooling. However, for most applications, a statement con‐
cerning reaction to fire under a standard temperature‐time curve will suffice, which is far
more feasible from an experimental perspective and facilitates comparisons between
differing construction materials or components.
538
Reaction to fire and fire design
1200
T(˚C)
1000
800
600
400
200
0
0 30 60 90 120 150 180 210 240 270 300 330 360
t(min)
Figure G1‐1 Standard temperature‐time curve in accordance with ISO 834. (STEP 1995 Article A13)
Load‐bearing capacity criteria apply in addition to the actual failure characteristics of
critical deflection and deflection velocity. Integrity is generally determined using the
formation of excessive cracks and splits (national definitions) or by the ignition of a ball of
cotton wool placed on the side opposite the fire. The insulation capacity is deemed to be
compromised when a mean temperature rise of 140°K is determined on the side oppo‐
site the fire or a maximum increase of 180°K is exceeded.
Wood only loses its load‐bearing capacity when the remaining cross‐section not yet de‐
stroyed by fire becomes so small that the stresses imposed by the loads attain the
strength of the remaining cross‐section. Wood‐based materials will not rupture in the
event of a fire or shrink to such an extent that cracks appear, but the surface will remain
intact, until the wood is so thin that the fire ends up burning through. The increase in
temperature will only exceed the limit value, if the thin and fire‐stressed area reaches the
opposite side, whereupon the burn‐through soon follows. Under the conditions of a fully
developed fire, the behaviour of wood can be predicted very reliably. In this article, the
reaction to fire of wood and wood‐based materials is explained, followed by the various
design concepts for members and joints.
G1.1 Fire protection regulations
Before the actual “fire design” when exposed to fire (as opposed to “cold design” under
normal temperature of the EC 5) is covered, the requirements and fire protection regula‐
tions will be summarised. On a European level, the impact in the event of fire is regulated
in EC 1 Part 1‐2, which provides the structural engineer with calculation methods to
determine design fires and temperature distributions in structural members. However, it
does not include requirements stating which building types must withstand a fire follow‐
ing the standard temperature‐time curve and for how long. These requirements for
539
Reaction to fire and fire design
buildings are regulated in Germany by the State Building Codes (LBO) (see also Article
A1). The LBO classifies buildings into building classes (GK); of which there are five in Ba‐
den‐Württemberg:
Building class 1: Detached buildings up to 7 m high, comprising a maximum of two
units of use covering a collective area not exceeding 400 m² as well as detached
buildings used for agriculture and forestry purposes,
Building class 2: Buildings up to 7 m high, comprising a maximum of two units of
use covering a collective area not exceeding 400 m²,
Building class 3: Other buildings up to 7 m high,
Building class 4: Buildings up to 13 m high and comprising units of use each with
an area not exceeding 400 m²,
Building class 5: Other buildings, including underground buildings.
The LBO further stipulates the need to configure and erect structures, “which allow the
formation of fires and the spreading of fire and smoke (fire propagation) to be prevent‐
ed” (§ 15 LBO BW). Additional statements concern, among other things, escape routes,
accessibility, spacing and lightning protection systems; more accurate provisions on
building design (fire resistance, selection of construction material) are included in the
“General design specification of the Ministry of Transport and Infrastructure for the State
Building Code” (LBOAVO BW). The LBOAVO BW sets out the fire protection regulations
for various building classes. For individual building components such as load‐bearing walls
and columns, floors, partition and fire walls, fire resistance classes are defined, see Ta‐
ble G1‐1. Building components of fire resistance class F90 must be able to withstand a
fire for at least 90 min without losing their load‐bearing capacity and, if applicable, retain
their integrity. This means, for example, according to LBOAVO BW, floors in buildings of
building class (GK) 5 must be fire resistant, in GK 4 highly fire retardant and in GK 2 and 3
fire retardant. Additional definitions such as those of the fire resistance classes are in‐
cluded in the Construction Products List A (national document, published by DIBt Berlin).
A selection of designations and classes on a European and federal level is included in
Table G1‐1.
540
Reaction to fire and fire design
Table G1‐1 Fire resistance classes of building components, Construction Products List A Annex 0.1.
541
Reaction to fire and fire design
On a European level, EN 13501 stipulates the classification of construction products in
classes A to E, which also include further sub‐classifications specifying smoke development
and burning droplets. The additional classification, discussed above, in line with REI‐criteria
closely resembles that of the State Building Codes (see Table G1‐1). Accordingly, the Euro‐
pean fire resistance class REI30 is comparable to the German class F30‐B. For a non‐load
bearing but partitioning wooden wall, conversely, EI30 or F30‐B may be required.
In general, the applicable regulations governing the requirements in the event of fire are
very complex and relatively obscure, since European regulations are applied alongside
those on a state or national level. The numerous relevant documents and frequent revi‐
sions made also hamper efforts to get a handle on which individual regulations apply. A
good overview is provided by Scheer and Peter (2009) as well as the Construction Prod‐
ucts List A. One stand‐out feature of timber construction is the “Guideline on fire protec‐
tion requirements for highly fire retardant components in timber construction”, HFH‐
HolzR. This guideline paves the way for timber buildings of building class 4.
G1.2 Principles
Wood and wood‐based materials mainly comprise cellulose and lignin, which, in turn,
consist of carbon, hydrogen and oxygen (see Article B1). This means they are combustible
and removing this property is practically impossible. However, absolute non‐combus‐
tibility is only seldom required.
Influences on reaction to fire
The shape, surface, circumference and size of the building components made of wood
and wood‐based materials significantly influence their reaction to fire and the combusti‐
bility depends on the ratio of surface to volume. The greater this ratio, the more rapidly
the flames will spread. Many sharp edges and rough surfaces increase this ratio and lead
to unfavourable reaction to fire. Meanwhile, cracks and splits also increase the risk of fire
breaking out, which explains why the charring rate of virtually crack‐free glulam timber is
lower than for solid timber.
The duration up to the point of the wood igniting and the fire spreading depends on the
density (oven‐dry density), meaning different wood species react differently to fire. The
relation between density and the rate of combustion RC is schematically shown in
Figure G1‐2. The relationship between density and ignition resembles the ‐RC relation‐
ship: the higher the density, the later the point at which the wood will ignite.
542
Reaction to fire and fire design
The wood moisture content is another key influence parameter for the reaction to fire of
wood. In timber structures, the moisture content generally ranges between 8 and 15%,
which means 80 to 150 kg of water have to be extracted from each tonne of wood before it
will burn. The influence of the moisture content on the charring rate can thus be disregard‐
ed within the low bandwidth for the equilibrium moisture content of between 8 and 15%.
200
RC(%/min) 180
160
140
120
100
80
60
40
20
0
100 300 500 700 900 1100 1300
ρ (kg/m3)
Figure G1‐2 Relationship between density ρ and rate of combustion RC. (STEP 1995 Article A13)
Chemical and physical processes during the combustion of wood
When wood and wood‐based materials burn, it sets off a chemical decomposition reac‐
tion, during which charcoal and combustible gases form. The spontaneous combustion of
a thin strip of wood takes place at a temperature of between 340 and 430°C. The ignition
temperature, however, may be far lower (e.g. 150°C), if the wooden piece has already
been heated for an extended period. Temperatures under 100°C, but above room tem‐
perature, add heat to the wood and promote its drying out, whereupon the strength and
modulus of elasticity decline.
At 100°C, the water begins to vaporise and the steam escapes via the route of least re‐
sistance, namely corners, edges, joints, open pores and cracks. These are places where
the wood dries faster than normal. The temperature remains constant, until the water
has completely vaporised. Figure G1‐3 shows the temperature underneath the so‐called
pyrolysis layer depending on time, when the wood is exposed to fire corresponding to
the ISO temperature‐time curve. The curve clarifies the rise in temperature after the
water vaporises at 100°C. The pyrolysis layer is the area between the charcoal and the
unchanged wood. In this layer, although the wooden substance has already been chemi‐
cally changed by the fire, it is not yet completely decomposed.
543
Reaction to fire and fire design
At between 150 and 200°C, gases are emitted, 70% of which comprise non‐combustible
carbon dioxide (CO2) and 30% combustible carbon monoxide (CO). From 200°C mean‐
while, more and more combustible gases form and the CO2 portion declines. As soon as
the gases ignite, the surface temperature rises significantly and the wood begins to de‐
compose in the pyrolysis layer, which is around 5 mm thick. At temperatures exceeding
500°C, the gas formation declines very sharply and the charcoal formation rises, which
explains the typical visual appearance of wood after a fire.
T (˚C)
300
200
100
0
0 40 80 t (min)
Figure G1‐3 Temperature in the heated wood below the pyrolysis layer, wood is exposed to the
ISO‐temperature‐time curve. (STEP 1995 Article A13)
The thermal conductivity of charcoal is only around a sixth of that of solid timber. The
charcoal layer thus acts as a form of insulation and delays the decomposition of the
deeper remaining cross‐sectional areas. For this reason and due to the low thermal con‐
ductivity of wood, the temperature in the middle of the cross‐section remains far lower
than at the surface, making wood far more fire‐resistant than is generally assumed. Fig‐
ure G1‐4 shows beams or columns, which have been exposed to fire on three or four sides.
(c)
(a) (a)
(c)
(b) (b)
(a) (a)
(b) (b)
Figure G1‐4 Beams and columns before and after exposure to fire. (a) remaining cross‐section,
(b) charcoal layer, (c) fire protection layer. (STEP 1995 Article A13)
544
Reaction to fire and fire design
Charring rates (rates of combustion)
Numerous tests with wood and wood‐based materials reveal a linear dependency be‐
tween the charring depth and fire duration. Accordingly, when calculating the fire re‐
sistance of a cross‐section, a constant charring rate is assumed. Simple approaches to fire
design without taking fillets into consideration include the use of one‐dimensional char‐
ring rates 0 in accordance with Table G1‐2 (determined under standard exposure to fire
in accordance with ISO 834). The remaining cross‐section is assumed to be sharp‐edged
for the fire design. The effective charring rate n includes edge fillets and cracks and is
therefore higher. For wood‐based panels, the one‐dimensional charring rates depend on
density and panel thickness and must be calculated for characteristic densities k other
than 450 kg/m3 and thicknesses hp smaller than 20 mm:
0,ρ,t 0 k ρ k h (G1‐1)
where
450
kρ
k
20
kh
hp
and 0 in accordance with Table G1‐2.
Table G1‐2 Charring rates 0 and n. (n includes edge fillets and cracks), EN 1995‐1‐2, 2010.
545
Reaction to fire and fire design
Figure G1‐5 One‐dimensional charring depth dchar,0 and effective charring depth dchar,n.
Using the charring rate and the duration t of the fire (namely 30, 60 or 90 minutes), the
charring depth dchar of an unprotected surface is determined, Figure G1‐5.
One‐dimensional charring depth:
d char,0 0 t (G1‐2)
Effective charring depth:
d char,n n t (G1‐3)
If the one‐dimensional charring depth is used, the corner fillets must be taken into consid‐
eration as shown in Figure G1‐5, whereby dchar,0 is used as the fillet radius of the corners.
Normally, a simpler method using the effective charring depth is used (see Section G1.4)
and the use of the one‐dimensional or effective charring depth depends on the initial
width of the building component. If the initial width bmin in equation (G1‐4) applies, the
one‐dimensional charring depth may be used; but if the initial width is smaller than bmin,
the effective charring depth must be used.
546
Reaction to fire and fire design
For protected surfaces, the start of combustion of the wooden components is initially
delayed. This only begins at time tch; for example at time tf, after the protective layer has
failed (tch need not be identical to tf). The effect of the fire on the protective layer means
the wood arranged behind is preheated, which, in turn, accelerates the rate at which it
burns. Only after a charring depth of 25 mm or the charring depth of unprotected timber –
whichever is lower – does the charring speed revert to the simple level, time ta. Fig‐
ure G1‐6 shows this relationship and how components with initial protective cladding
burn up faster once the protective layer fails, see curve 3. Accordingly, the protective
layer is only of benefit when the failure time tf of the protective layer exceeds the time
required to reach a charring depth of 12.5 mm in an unprotected component.
40
Charring depth dchar,0 or dchar,n [mm]
30
20
1
10
3
0
tf ta
Time t
Figure G1‐6 Illustration of charring depth depending on time for tch = tf,1. 1: Course for unprotected surface;
3: course for initially protected surface; tch = delayed combustion of the wooden component,
tf = failure of protective layer, from time ta = normal charring speed of wooden component.
The relationship from Figure G1‐6 does not apply to all protected components; Fig‐
ure G1‐6 only applies to components with a protective layer that delays the combustion
of the wooden component until the layer fails at tf. For some types of protective layers,
however, combustion of the wooden component may occur even before time tf, if the
layer e.g. cracks or lacks sufficient thermal insulation properties.
Temperature‐dependent strength and stiffness values
The mechanical properties of wood are temperature‐dependent, with stiffness and
strength both declining with increasing temperature. The heated part of an unburnt
remaining cross‐section thus has lower strength or stiffness than at room temperature,
while the magnitude of the decline varies for differing mechanical properties. The exam‐
ple shown in Figure G1‐7 features tension strength parallel to the grain of solid timber
depending on temperature, while Figure G1‐8 shows compression strength parallel to the
grain. The decrease in compressive strength under high temperatures is more prominent
547
Reaction to fire and fire design
than the decrease in tensile strength. In addition, both figures clearly show the scattered
nature of the test results. The dependency of mechanical properties on temperature
declines at varying rates depending on the boundary conditions during testing, since
among other things, ensuring constant temperature and humidity conditions in each case
is far from easy, particularly with large test bodies.
1
0.8
0.6
ft,0(T)/ft,0
0.4
0.2 EC 5 1‐2:2010
literature
0
0 50 100 150 200 250 300
Temperature T in °C
Figure G1‐7 Relation between temperature and tensile strength ft,0. (according to Scheer and Peter, 2009)
1 EC 5
0.8 1‐2:2010
literature
0.6
fc,0(T)/fc,0
0.4
0.2
0
0 50 100 150 200 250 300
Temperature T in °C
Figure G1‐8 Relation between temperature and compressive strength fc,0.
(according to Scheer and Peter, 2009)
Temperature distribution in the cross‐section
The temperature distribution in the cross‐section depends on the charring speed, cross‐
sectional geometry, fire duration, fire exposure (on one or multiple sides) and the ther‐
mal conductivity of wood as well as on any existing metal parts (= high thermal conduc‐
tivity). In the event of fire, surfaces ignite and initially burn quickly, but a heat‐insulating
charcoal layer soon forms, see Figure G1‐9. Since wood, particularly in the form of char‐
coal, is a poor thermal conductor, very little heat is transmitted to the remaining unburnt
material. In the solid timber, the temperature in the remaining cross‐section in the prox‐
imity of the combustion area (base of the pyrolysis layer = area between charcoal and
unchanged wood) is already unchanged. The temperature of 300°C at the combustion
limit declines over a thickness a0 (of around 25 to 40 mm) to the unchanged surrounding
548
Reaction to fire and fire design
temperature level (e.g. 20°C), Figure G1‐10. Even within the combustion area, the tem‐
perature declines, from that of the fire (800 to 1000°C) to around 300°C at the combus‐
tion limit. This effect of low thermal conductivity can be viewed positively, since among
other things, the mechanical properties of the remaining cross‐section, reduced by the
heat‐affected zone a0, remain largely unchanged.
(a)
(b)
(c)
Figure G1‐9 The changes in wood in the event of a fire: (a) charring layer, (b) pyrolysis layer,
(c) unchanged wood. (STEP 1995 Article A13)
θ (˚C) 20 + Δ θm
20
a0
dchar
br /2
Figure G1‐10 Temperature distribution for br > 2 ∙ a0. (STEP 1995 Article B17)
Reaction to fire of joints with metal fasteners
The load‐bearing capacity of metal fasteners, which are not protected against fire, is
swiftly lost when heat is applied. In addition, the high thermal conductivity of the metal
parts heats up the surrounding wood; causing both embedment strength and withdrawal
capacity to decline. Covering the metal fasteners with wood or wood‐based materials
delays any heating of the metal parts, while the area of unprotected surfaces of metal
parts is the decisive criterion governing how metal fasteners behave in the event of fire.
Figure G1‐11 shows the relevant yield point of steel depending on temperature. Alt‐
hough steel is non‐combustible, it features high thermal conductivity and a yield strength
that declines swiftly with increasing temperature.
549
Reaction to fire and fire design
1,00
ψ
0,80
0,60
0,40
0,20
0,00
0 200 400 600 800 1000
˚C
Figure G1‐11 Relevant yield point depending on temperature. (STEP 1995 Article C19)
G1.3 Safety concept for fire design
As a general rule, the applicable rules for fire design are the same principles as those
applied for verifications in a “cold” state, whereby actions and material properties are
determined using characteristic values. However, fire testing has usually been performed
using deterministic methods, limiting the scope to average strength values. To reconcile
test results with the calculation principles to be applied, EC 5 Part 1‐2 contains specially
tailored approximation methods for fire design. However, design for exposure to fire is
not only possible via calculation but also via tests, while in Germany, classification via
DIN 4102‐4 is also possible. In Part 4 of DIN 4102, a set of classified construction materi‐
als and components is regulated, allowing a specific fire resistance class to be obtained
without calculations or tests. For example, DIN 4102‐4:1994 sets out that load‐bearing,
non‐partitioning walls must be made of studs of a specific grading class and dimension
and of sheathing with a minimum thickness and density to achieve the fire resistance
duration F30‐B.
If a fire design by calculation is performed, other partial safety factors and load combina‐
tions (actions and design material properties) apply compared to the case for normal
design without exposure to fire. The effect of actions E(t) and the resistance of timber
members R(t) during fire exposure are shown in Figure G1‐12. The fire resistance is
reached after time tf, when R(t) is smaller than E(t).
Verification on the design level for the governing load duration t is thus as follows:
where
Ed,fi Design effect of actions in the event of fire
Rd,t,fi Design resistance in the event of fire
550
ZĞĂĐƚŝŽŶƚŽĨŝƌĞĂŶĚĨŝƌĞĚĞƐŝŐŶ
35,00
ĞƐŝŐŶǀĂůƵĞƐĨŽƌĂĐƚŝŽŶƐ
ĐĐŽƌĚŝŶŐ ƚŽ Ϭ ĂŶĚ ϭ WĂƌƚ ϭͲϮ Žƌ /EϰϭϬϮͲϮϮ͕ ƚŚĞ ĨŝƌĞ ĚĞƐŝŐŶ ƐŚŽƵůĚ ďĞ ŝŵƉůĞͲ
ŵĞŶƚĞĚǁŝƚŚĂĐŽŵďŝŶĂƚŝŽŶŽĨĂĐƚŝŽŶƐĨŽƌĂĐĐŝĚĞŶƚĂůĚĞƐŝŐŶƐŝƚƵĂƚŝŽŶƐ͗
ǁŚĞƌĞ
Ě ĐƚŝŽŶƐŝŶƚŚĞĞǀĞŶƚŽĨĨŝƌĞ
J'͕ũ WĂƌƚŝĂůƐĂĨĞƚLJĨĂĐƚŽƌĨŽƌƉĞƌŵĂŶĞŶƚĂĐƚŝŽŶƐŝŶƚŚĞĞǀĞŶƚŽĨĨŝƌĞ͕J'͕ũсϭ͘Ϭ
hŶůŝŬĞĐŽŵďŝŶĂƚŝŽŶƐŽĨĂĐƚŝŽŶƐĨŽƌŶŽƌŵĂůĚĞƐŝŐŶ͕ŝŶƚŚŝƐĐĂƐĞ͕ƚŚĞƉĂƌƚŝĂůƐĂĨĞƚLJĨĂĐƚŽƌƐ
ĨŽƌƉĞƌŵĂŶĞŶƚĂŶĚǀĂƌŝĂďůĞůŽĂĚƐ J'ĂŶĚ JYĂƌĞƐĞƚĂƚϭ͘Ϭ͘,ŽǁĞǀĞƌ͕ĨŽƌĨŝƌĞĚĞƐŝŐŶƉƵƌͲ
ƉŽƐĞƐ͕ ƚŚĞ ĂĐƚŝŽŶƐ ĐĂŶ ďĞ ĐĂůĐƵůĂƚĞĚ ĨƌŽŵ ƚŚĞ ŶŽƌŵĂů ĚĞƐŝŐŶ ǀĂůƵĞƐ ďLJ ƚŚĞ ĨŽůůŽǁŝŶŐ
ĞƋƵĂƚŝŽŶ͗
Ě K Ĩŝ Ě ;'ϭͲϳͿ
KŝƐƚŚĞƋƵŽƚŝĞŶƚŽĨƚŚĞĂĐƚŝŽŶƐĨƌŽŵƚŚĞĂĐĐŝĚĞŶƚĂůĐŽŵďŝŶĂƚŝŽŶ;ĨŝƌĞͿĂŶĚƚŚĞĨƵŶĚĂŵĞŶͲ
ƚĂůĐŽŵďŝŶĂƚŝŽŶ͗
dŚĞƌĞŝƐƐĐŽƉĞƚŽƐŝŵƉůŝĨLJŝŶĂĐĐŽƌĚĂŶĐĞǁŝƚŚEƚŽϱWĂƌƚϭ͘ϮƚŽƉƵƚ KсϬ͘ϲϬĐŽŶƐĞƌͲ
ǀĂƚŝǀĞůLJ͘
ϱϱϭ
Reaction to fire and fire design
Design values of material properties
Values other than for normal design apply to the partial safety coefficients on the materi‐
al side, since the material properties depend on temperature, resp. during an accidental
design situation, such as in the event of fire, where the partial safety factor M assumed is
1.0. To verify load‐bearing capacity, the design strength fd,fi and stiffness Sd,fi are as follows:
f 20 S 20
f d,fi k mod,fi or S d,fi k mod,fi (G1‐9)
M,fi M,fi
where
kmod,fi Modification factor in the event of fire, see below
M,fi Partial safety factor for material properties, M,fi = 1.0
f20 or S20 20%‐quantiles of a strength or stiffness property at normal temperature
The 20%‐quantiles are determined via the coefficient kfi from the characteristic values:
f 20 k fi f k or S 20 k fi S 05 (G1‐10)
kfi‐values are specified in Table G1‐3 and vary for the different construction materials,
since the scatter varies for the individual materials (namely the 20%‐quantile may, de‐
pending on the scatter of a material property, differ to a greater or lesser extent from the
5%‐quantile).
In the design case of fire, consequently, not only are the partial safety factors on the action
and material side reduced, the modification factor kmod is also replaced with kmod,fi and a
20%‐ rather than 5%‐quantile value is used for the material properties, since it is assumed
that the weakest cross‐section is not simultaneously the one most stressed by fire.
Table G1‐3 Values for kfi, EC 5:2010 Part 1‐2.
Material kfi
Solid timber 1.25
Glulam, wood‐based materials 1.15
Laminated veneer lumber 1.1
Laterally loaded timber‐to‐timber joints 1.15
Laterally loaded joints with external steel plates 1.05
Joints with axially loaded fasteners 1.05
552
Reaction to fire and fire design
Coefficient kmod,fi
As explained above, the mechanical properties depend on temperature to a greater or
lesser extent, whereby different strength properties are influenced to varying extents by
temperature. The most significant decline is in compressive strength, whereas tensile
strength, for example, is less affected by temperature (see Figure G1‐7 and Figure G1‐8).
This influence is considered with the coefficient kmod,fi. The value of kmod,fi depends on the
design method. With the reduced cross‐section method, kmod,fi = 1.0 is assumed. With the
reduced properties method, which can be applied for softwood cross‐sections exposed to
fire on three or four sides as well as for general fire exposure of softwood logs, other
values for kmod,fi ≤ 1.0 must be determined, which take into consideration the reduction in
mechanical properties of the construction material at elevated temperatures (for more
details, see section “reduced properties method”).
G1.4 Design methods for structural members
EC 5 Part 1‐2 includes two methods for the fire design of members resp. to determine
cross‐sectional values in the event of a fire:
Simplified method: Reduced cross‐section method (RCSM),
More accurate method: Reduced properties method (RPM).
In addition to both these methods, more complex methods can also be applied to calcu‐
late the charring depths (parametric fire exposure, EC 5 Part 1‐2 Annex A) or determine
all required design steps considering a parametric design fire (EC 5 Part 1‐2 Annex B and
EC 1 Part 1‐2). However, in the vast majority of cases, these methods are infeasible for
the structural engineer and simply represent alternatives for special constructions.
553
Reaction to fire and fire design
Reduced cross‐section method
With this method, the fire resistance duration of a structural member depends on the
load‐bearing capacity of the remaining unburnt cross‐section. Figure G1‐13 shows this
effective cross‐section, where the following applies:
d ef d char,n k 0 d 0 (G1‐11)
where
dchar,n Effective charring depth corresponding to equation (G1‐3)
d0 = 7 mm
k0 See Table G1‐4
(a) dchar
(b)
def
Figure G1‐13 Effective cross‐section. (STEP 1995 Article B17)
The additional layer thickness k0 ∙ d0 takes into consideration the fact that the wood near
the combustion limit is very hot, and hence has only low strength and stiffness. The actual
combustion layer is thus further increased by this “heat‐affected zone” (a0 in Figure G1‐10).
d0 was determined as follows:
The integration of the temperature distribution in accordance with Figure G1‐10 reveals an
average temperature of 80°C. If we assume the average strength at 80°C is around 70% of
the initial strength and conclude that 70% of a0 has wood strength unaffected by tempera‐
ture and 30% can no longer bear any load, d0 = 7 mm results in 30% of a0 = 25 mm.
The coefficient k0 corresponding to the required fire resistance duration t can be taken
from Table G1‐4. Using the remaining cross‐sections, the cross‐sectional values required
for design can now be determined.
554
Reaction to fire and fire design
Table G1‐4 Definition of coefficient k0.
Reduced properties method
This method is also derived from the temperature distributions shown above. The load‐
bearing capacity is, in turn, calculated for a remaining cross‐section, and here, the me‐
chanical properties of the remaining cross‐section are also reduced. This method is in‐
cluded in EC 5 Part 1‐2 for softwood cross‐sections exposed to fire on three or four sides
as well as for general fire exposure of softwood logs. The reduction in mechanical proper‐
ties is performed by the coefficient kmod,fi, which is specified as depending on the load, on
the circumference p and the area Ar of the remaining cross‐section, see also Figure G1‐14:
1 p
Tension strength and modulus of elasticity : k mod,fi 1.0
330 A r
1 p
Bending strength: k mod,fi 1.0 (G1‐12)
200 A r
1 p
Compression strength: k mod,fi 1.0
125 A r
where
p Circumference of remaining cross‐section subject to fire in m
Ar Area of remaining cross‐section in m2
555
Reaction to fire and fire design
1,0
0,9
mod,f
kkmod,fi
t,E
0,8
0,7
m
0,6
0,5
0,4 c
0,3
0,2
0,1
0,0
0 20 40 60 80 100
p/A (m-1)
Figure G1‐14 kmod,fi for tension (t), bending (m), compression (c) and modulus of elasticity (E).
(STEP 1995 Article B17)
Figure G1‐14 and equations (G1‐12) show how the influence of temperature on the me‐
chanical properties varies depending on the property. The compressive strength declines
more than bending or tensile strength with increasing temperature.
With the above rules, cross‐sectional values are determined to establish all relevant load‐
bearing capacity and stability verifications for the members exposed to fire. One key
point which has to be noted is the applicable scope of the calculation methods intro‐
duced. Both simplified methods apply to members with constant cross‐section and ho‐
mogenous structure (namely no combined glulam). Combined glulam or beams with
variable cross‐section are not explicitly regulated in EC 5 Part 1‐2, additional considera‐
tions are therefore required. Mechanically jointed beams, conversely, are regulated; the
slip modulus Kfi can be determined from Ku using a conversion factor.
G1.5 Design methods for joints
Since metal has far higher thermal conductivity than wood, it is very effective at transmit‐
ting heat to the inside of members, which means yield strength declines very rapidly with
increasing temperatures, Figure G1‐11. Accordingly, metal fasteners should be protected.
Such protection can be provided very easily for dowels, e.g. by applying wooden plugs.
Unprotected joints with side members made of wood
Presuming the proper design of joints and fasteners in accordance with EC 5 Part 1‐1
(spacings and distances of the fasteners, etc.) and when complying with specific mini‐
mum thicknesses of joints and side members, unprotected joints with side members
made of timber (also including steel‐to‐timber joints with slotted‐in steel plates) have a
fire resistance duration of td,fi = 15 min, class R15 (20), without further verification. Higher
556
Reaction to fire and fire design
fire resistance duration for the specified joints (not with bolts and fasteners in accord‐
ance with EN 912), but not exceeding R30, can be achieved by increasing the member
thickness and the end and edge distances by the dimension afi:
a fi n k flux t req t d,fi (G1‐13)
where
n Effective charring rate in accordance with Table G1‐2
kflux Coefficient for considering the elevated heat flow through the metal fasteners,
generally kflux = 1.5
treq Required fire resistance duration
td,fi Fire resistance duration in accordance with Table G1‐5
Table G1‐5 Fire resistance duration of unprotected joints with side members made of wood.
Reduced stresses method
This method allows higher fire resistance durations to be achieved for unprotected joints,
without having to increase cross‐sections by the afi. A maximum fire resistance duration
of 40 min can be achieved for timber‐to‐timber joints with dowels d ≥ 12 mm using the
cross‐sections for normal design. By combining the different methods (cross‐sectional
increase, protecting joints and reduced stresses method), a maximum fire resistance
duration of 60 min can be achieved (Scheer and Peter, 2009).
Additional rules for joints with slotted‐in steel plates
A fire resistance duration of 30 or 60 minutes (R30 or R60) for a joint with unprotected,
non‐protruding steel plates (plate thickness ≥ 2 mm) is achieved, when the widths bst of
the steel plates are complied with in accordance with Table G1‐6.
557
Reaction to fire and fire design
Table G1‐6 bst of steel plates with unprotected edges.
Unprotected edges One edge or two opposite edges unprotected
R30 ≥ 200 mm ≥ 120 mm
R60 ≥ 280 mm ≥ 280 mm
Protected joints with side members made of wood
Cladding or wooden plugs can protect joints with side members made of wood. If wood‐
en plugs are used, these should be of at least thickness afi in accordance with equation
(G1‐13). When using a cladding, it is important to ensure that the start of combustion of
the protected member tch does not go below a certain time duration, which must be
guaranteed by the cladding. Accordingly, the following applies for a cladding with wood‐
based panels or gypsum plasterboards of type A or H:
and for cladding with gypsum plasterboards of type F:
where
tch Time to the start of combustion, need not be identical to tf, the time of failure
of the protective cladding
treq Required fire resistance duration
td,fi Fire resistance duration of the unprotected joint in accordance with Table G1‐5
Fire resistance classes of protected joints exceeding R30 can be reached in a simplified
manner by increasing the end and edge distances by 2 ∙ afi in accordance with equation
(G1‐13).
Joints with external steel plates
Unprotected joints
Such joints require a fire design in accordance with EC 3 Part 1‐2.
Protected joints
Joints are considered protected, if the external steel plates are completely equipped with
cladding made of wood or wood‐based materials of minimum thickness afi with
td,fi = 5 min in accordance with equation (G1‐13).
558
Reaction to fire and fire design
Axially loaded screws
In the case of axially loaded screws exposed to fire, the design withdrawal resistance in
accordance with EC 5 Part 1‐1 is reduced by a factor depending on the edge distances
of the screw as well as the required fire resistance duration.
G1.6 Literature
H. Hartl, original Articles A13, B17, C19, STEP 1995.
Kollmann F.F.P and Coté W.A. (1968). Principles of wood science and technology. Volume I, Solid Wood.
Springer Verlag, Berlin, 592 S.
Scheer C. and Peter M. (2009). Holz Brandschutz Handbuch. ed. Deutsche Gesellschaft für Holzforschung,
Ernst und Sohn Verlag, Berlin.
559
G2 Joints subject to seismic loads
Original article: A. Ceccotti
Earthquake design of structures according to EC 8 takes two different earthquake loads
into consideration. In the event of a light to moderate earthquake with a peak ground
acceleration value corresponding to a return period of 95 years (corresponding to a
probability of exceedance of 10% in 10 years), structures should not sustain any damage
or excessive deformations. Structures should also be able to withstand a severe earth‐
quake with a return period of 475 years (probability of exceedance of 10% in 50 years). In
the event of a severe earthquake, although any structural collapse must be prevented,
any larger scale of damage is acceptable. During a severe earthquake, a structure “sof‐
tens”, the natural frequency i declines and the kinetic energy conveyed by the earth‐
quake is dissipated. Reducing the stiffness “slows down” the structural response to cyclic
loading; the structure then has sufficient time to reverse its direction of motion prior to
developing deformations triggering collapse.
The ability of a structure or parts thereof to withstand plastic deformations and hence
allow energy to dissipate before reaching the load‐bearing capacity is key to explaining
why the impact of an earthquake does not lead to the structure collapsing (Ceccotti,
1989). Timber structures with plastic joints can withstand a stronger earthquake than
corresponding constructions with rigid joints, which are prone to brittle failure. Wooden
members generally behave in a linear‐elastic manner when subject to cyclic load, where‐
by brittle failure ensues, due largely to growth irregularities like knots. Except for com‐
pressive stress perpendicular to the grain, very little energy is dissipated during this pro‐
cess. The same applies to glued joints, the behaviour of which can also be characterised
as linear‐elastic and brittle and which have similarly small impact on the energy dissipa‐
tion within a timber construction. Timber constructions made of glued members or joints
are therefore not considered dissipative, i.e. as linear‐elastic up to failure and are as‐
signed to a lower ductility class.
The ability to dissipate energy via plastic deformations, however, can be achieved by
joints with mechanical fasteners, which behave “semi‐rigidly”, unlike rigid glued joints.
Well‐designed joints with mechanical fasteners generally show prominent plastic behav‐
iour before the load‐bearing capacity is reached. This plastic behaviour of joints is taken
into consideration when verifying the structure, by classifying it in a ductility class which
takes its energy dissipation ability into consideration. The more ductile the joints behave
before the load‐bearing capacity is reached, the greater the structure’s resistance against
earthquake effects.
561
Joints subject to seismic loads
If a structure exposed to an earthquake with a peak ground acceleration ay reaches its
elastic limit (“yield point”), it can also withstand a q‐times stronger earthquake with a
peak ground acceleration au without collapsing. The peak ground acceleration of the
stronger earthquake is au = q ∙ ay. The values for au correspond to the seismic action,
which is defined in standards depending on the geographical location of the structure to
be designed. In EC 8, the coefficient q is known as “a factor reducing seismic actions” or
“behaviour factor”. The above‐specified ductility classes are identified by the value of the
behaviour factor q.
For design purposes, therefore, it is sufficient to verify that a structure will remain in the
elastic area during an earthquake with a peak ground acceleration ay. The ability of the
structure to withstand q‐times stronger earthquakes, even when damaged, is then im‐
plicitly taken into consideration. For structures, which remain linear‐elastic up to failure,
q = 1.5. This value takes overstrength into consideration, meaning structures show higher
strength under a dynamic load than under a static load. Larger values for q mean lower
values for the calculated seismic actions, although this requires sufficient ductility and
energy dissipation. If ductile behaviour cannot be guaranteed, the structure should be
designed with q = 1.5.
Although q = 1.5 must only be assumed for structures such as three‐hinged frames with
glued frame corners, which lack any appreciable energy dissipation capability, there are
times when designing even plastic structures with q = 1.5 makes sense. This may apply
e.g. if a variable action other than the earthquake must be taken into consideration. If,
for a comparatively lightweight structure, dead weight plus snow is the governing load
combination, it is possible that all verifications can be complied with, even if the seismic
action is considered with q = 1.5, since snow and earthquake actions need not be simul‐
taneously considered with their characteristic values. Assuming q = 1.5 eliminates the
need for the structural engineer to consider specific detailed rules or perform tests to
determine ductility.
Apart from the described cases including specific load combinations, consideration of the
behaviour factor q can also be a decisive factor dictating the economic efficiency of struc‐
tures in seismic areas. When using a value q > 1.5, sufficient ductility of the joints with
mechanical fasteners should be ascertained. This can be done either by corresponding
tests or, when joint configurations with known plastic behaviour are used, by ensuring
compliance with the detailing rules specified in the following section.
562
Joints subject to seismic loads
G2.1 Ductility
Provided sufficient distances are maintained between fasteners and the timber member
edges, timber joints with mechanical fasteners, unlike glued joints, show ductile and
plastic behaviour prior to failure. The plastic behaviour is caused both by plastic embed‐
ment deformations of the wood underneath the fasteners as well as the plastic behav‐
iour of the metal fasteners subject to bending (see Article E2). The load‐deformation
curve of a joint under static load initially shows a steep rise (see Figure G2‐1 (a), area I).
As soon as either the steel or the wood under embedment stress can no longer show
elastic behaviour, “yield point” Fy, the load‐deformation curve flattens out, until a hori‐
zontal area finally emerges, which determines the load‐bearing capacity Fmax of the joint
(see Figure G2‐1 (a), area II). Subsequently, the load decreases with increasing defor‐
mations (see Figure G2‐1 (a), area III). In this area, the joint has already failed due to
splitting, withdrawal or failure of the fastener. This part of the load‐deformation curve
can only be recorded, if the test is performed in displacement‐control.
The definition of ductility Ds is specified in Figure G2‐1, where two cases are distin‐
guished. In the initial case (a), the load‐deformation curve can be approximated by two
straight lines, while in the second case (b), the curve is completely non‐linear. In case (b),
the ratio of gradients of both areas was defined as 1/6. This definition allows the yield
deformation vy to be determined objectively and comparably and thus allows the so‐
called static ductility Ds to be specified at vu/vy. Even when the load‐bearing capacity of
the joint has been reached, it can still carry a portion of the maximum load. With de‐
creasing load, the deformation vu needed to define the static ductility may correspond to
the deformation reached at a minimum of 80% of the maximum load.
F F
Fmax Fmax β
III
Fy II
0,4Fmax
I v α tgβ= 1 tgα
Ds = vu 6
vy vu v y vy vu v
(a) (b)
Figure G2‐1 Determining the ductility under static load having different load‐deformation behaviour.
(STEP 1995 Article C17)
563
Joints subject to seismic loads
G2.2 Cyclic behaviour and energy dissipation
The cases shown in Figure G2‐1 describe the behaviour under monotonic loading. In the
following section, the behaviour under earthquake load is examined where the load
direction changes within a short time. With this in mind, a nailed joint is observed, which
is subject to a quasi‐static load with changing direction (see Figure G2‐2 (b)). When the
elasticity limit is exceeded for the first time, the wood under the nail is compressed. This
causes a cavity, within which the nail is unsupported during subsequent loading cycles
within the same deformation range. The remaining load‐bearing capacity is thus based on
the nail acting as a cantilever over the length of the cavity or on a – mechanically identi‐
cal – connection with a gap between the members. As soon as the deformation attained
to date is exceeded, the nail, in turn, is supported by the surrounding wood and the load‐
deformation curve again follows the envelope curve, which would have formed had a
monotonic load been applied. However, any withdrawal of the nail during alternate load‐
ing (see Figure G2‐2 (b)) may reduce the load‐bearing capacity.
u u
(a) (b)
Figure G2‐2 Cavities in the wood of a nailed joint, caused by cyclic load and plastic deformations.
(STEP 1995 Article C17)
The typical hysteresis loops resulting from the behaviour described are shown for small,
moderate and large deformations in Figure G2‐3. The shape of the hysteresis loops varies
significantly from those which emerge for an elastic‐plastic material such as steel. For
elastic‐plastic material, the force required to restore plastic deformations is identical to the
force originally required to generate the deformations in question (see Figure G2‐4 (c)). In
timber joints with mechanical fasteners, the cavities formed due to the wood being com‐
pressed mean hardly any hysteresis loops emerge in the second and fourth quadrant,
leading to narrower and more “pinched” loops.
564
Joints subject to seismic loads
F F F
v v v
v v v
t t t
F F M
c c
v v v
t t
t c t c
Figure G2‐4 (a) shows load‐deformation loops of a dowelled joint with slender dowels, in
which both the embedment deformations of the wood and the plastic bending defor‐
mations of the dowels help dissipate energy. Sturdy dowels are not subject to plastic
deformations and the energy dissipation is solely based on the plastic embedment de‐
formation of the wood. Figure G2‐4 (b) shows the corresponding load‐deformation
curves. As a general rule, the envelope curve for cyclic load matches the load‐defor‐
mation curve for monotonic load, hence the load‐deformation behaviour is assumed to
be independent of the load history. Differences exceeding 10% can only be observed if
the fastener arrangement in the joint changes, e.g. due to withdrawal or brittle fatigue
failure of the fasteners. Unlike in masonry or concrete constructions however, timber
joints are not usually prone to such low‐cycle fatigue failures.
565
Joints subject to seismic loads
One exception are joints with punched metal plate fasteners, in which failure under cyclic
load is often caused by the punched teeth being suddenly withdrawn from the timber or
brittle failure of the metal plate itself. Other examples of the detrimental effect of cyclic
load on the load‐bearing capacity are joints with cold‐formed steel connectors where
nails can be withdrawn or joints in timber frame walls with brittle sheathing, where the
cyclic embedment stresses can lead to sheathing material breaking out, which dramati‐
cally reduces the load‐bearing capacity.
To obtain comparable information about how timber joints subject to cyclic loading be‐
have, a CEN test standard was devised, which specifies a simple method to define the
load‐deformation loops. Tests should be implemented quasi‐statically and under defor‐
mation control. The displacements are increased in steps with triple cycles of amplitudes,
Figure G2‐5. Any difference between the load achieved in the initial and third cycle sub‐
sequently represents the strength impairment under cyclic loading (see Figure G2‐6).
4,00 vy
0,25 vy
v
0,50 vy
2,00 vy
1,00 vy
0,75 vy
Figure G2‐5 Test method for timber joint subject to cyclic loads, EN 12512. (STEP 1995 Article C17)
F a
ΔF
v a
vc
vt
vc v t
vt
Figure G2‐6 Strength impairment F between the initial (a) and the third cycle at the same displacement
level. (STEP 1995 Article C17)
The variable F in Figure G2‐6 clarifies the strength impairment between the envelope
curve, which corresponds to the initial and third cycles at the same displacement level.
566
Joints subject to seismic loads
The energy dissipated by plastic deformations in a cycle is shown as a shaded area Ed in
Figure G2‐7, while the ratio between the dissipated energy Ed and the potential energy Ep
is designated as the equivalent viscous damping ratio eq. With increasing deformation,
the energy dissipated in a hysteresis loop Ed increases, whereby eq remains around the
same level. Well‐designed dowelled joints and nailed plywood‐to‐timber joints in timber
frame constructions can achieve values of around 8 to 10% for eq.
F Ed
Ep
Ed
veq =
2 π Ep
v
Figure G2‐7 Energy dissipation through hysteresis. (STEP 1995 Article C17)
Provided the joints remain in the elastic area, the equivalent viscous damping ratio eq is
equal to zero (see Figure G2‐3 (a)). However, still in the elastic area, when cyclic loading
is applied, a certain amount of energy is dissipated. The “viscous” damping ratio of the
construction material wood itself is in the order of magnitude of only 1%. Through fric‐
tion between members or compression perpendicular to the grain, particularly in con‐
structions with numerous non load‐bearing members, e.g. timber frame constructions,
the damping ratio in the elastic area may reach values of around 5%. Accordingly, for
calculations in the elastic area, a damping ratio of 5% is often assumed.
G2.3 Behaviour of different joint typologies
The favourable behaviour of timber joints with mechanical fasteners under cyclic loads is
characterised by a high level of ductility, the ability to withstand repeated loading and the
ability to dissipate energy by plastic deformations. Premature failure due to timber split‐
ting can often be avoided by ensuring sufficient spacings and end and edge distances.
EC 5 specifies minimum values for fastener distances, which were selected to avoid the
risk of the wood splitting under a static load. Despite the lack of indications that cyclic
loads promote any splitting of the wood, greater fastener distances than the specified
minimum distances in EC 5 may further reduce the tendency toward splitting and en‐
hance the joint ductility as a result.
567
Joints subject to seismic loads
Splitting of the wood can also be prevented by gluing on wood‐based panels or introduc‐
ing fully threaded screws in the area of the joint. As well as preventing splitting, such
measures may also enhance load‐bearing capacity due to the generally higher embed‐
ment strength of the reinforcement materials (see Article E12).
The energy dissipation characteristics of a timber joint with mechanical fasteners can also
be achieved using slender fasteners. Here, the slenderness ratio is defined as the ratio of
member thickness to fastener diameter. Unlike sturdy fasteners which do not deform
plastically, slender fasteners form plastic hinges, which dissipate energy during this pro‐
cess. In addition, the use of slender fasteners also helps guard against the splitting risk
(see Figure G2‐4).
The risk of losing load‐bearing capacity due to cyclic loading can be largely prevented by
using slender fasteners with a high withdrawal resistance, which are made of steel with a
large deformation capacity. In addition, joints between wood and brittle materials should
not be used to dissipate energy. As a general rule, the behaviour of normal fasteners in
terms of the response under cyclic loading can be described as follows (see also Article
G3 for indications of structural detailing).
Dowel‐type fasteners
Nails, staples and screws generally show pronounced plastic behaviour in timber joints.
The risk of withdrawal under repeated load means the length of the fasteners should be
greater than for a joint under static load. For this reason, smooth shank nails should not
be used for load‐bearing joints in seismic areas. A slenderness ratio exceeding 8 generally
already guarantees ductile behaviour under seismic loads (see Figure G2‐8). As shown
with tests on nailed wall panels (Yasumura, 1988), a high degree of ductility and favoura‐
ble energy dissipation characteristics in nailed joints between plywood and wood in tim‐
ber frame construction can already be observed with nails with a slenderness ratio of
four in plywood (see Figure G2‐9).
568
Joints subject to seismic loads
a
F (kN)
F 0,4
0,2
-6 -4 -2 4 6 v (mm)
-0,2
v
-0,4 t
Figure G2‐8 Typical behaviour of a nailed joint under repeated load (Slenderness ratio of the nails = 8.5).
(STEP 1995 Article C17)
200
F (kN)
F
100
v
-100
-200
-150 0 150 300
v (mm)
Figure G2‐9 Ductile behaviour of a plywood sheathed timber frame house. (STEP 1995 Article C17)
569
Joints subject to seismic loads
The load distribution in bolted joints is never uniform due to the oversized drilled holes.
This may result in individual bolts in the joint being overloaded and potentially, splitting
of the wood, preventing any redistribution which would allow more uniform loading of
the individual bolts. Bolted joints in seismic areas are accordingly only recommended if
slender bolts are used and the joints are carefully manufactured.
Connectors and punched metal plate fasteners
Split ring and shear plate connectors are not generally recommended for dissipating
energy in structures in earthquake‐prone areas due to their generally brittle failure (see
Article E6). In contrast, toothed‐plate connectors exhibit plastic behaviour. However, the
prerequisite is ensuring sufficient end and edge distances and spacings, which guarantee
embedment failure under the bolts and dowel teeth and prevent any timber splitting.
Although the load‐deformation curves of joints with punched metal plate fasteners often
show a certain degree of plastic deformations, given the risk of possible failure of the nail
plate and withdrawal of punched nails, cyclic tests should be performed on the joints
before considering any ductility.
G2.4 Seismic behaviour of joints with mechanical fasteners
Considerations to date relate to the behaviour of joints with mechanical fasteners under
quasi‐static and regularly changing loads. In the event of an actual earthquake, however,
both the frequency and load cycles would vary from those of the above‐described tests.
Since both these variables differ for each earthquake, which makes them impossible to
forecast, important statements on basic joint behaviour subject to seismic loads can be
made although the tests cannot simulate the exact seismic behaviour. Since the stiffness
and load‐bearing capacity of timber and timber joints rises with decreasing load duration,
we can assume that joints subject to seismic loads have a higher load‐bearing capacity
and stiffness than joints subject to quasi‐static testing. Likewise for ductility, the same
value can be assumed during seismic loads as was determined during quasi‐static tests.
Quasi‐static tests thus suffice for the purpose of determining the key parameters outlin‐
ing the response to seismic loads with sufficient accuracy. If the form of the hysteresis
loops of the joint under cyclic load is known, the load‐bearing capacity of the construc‐
tion, taking into consideration the non‐linear joint behaviour, can be calculated using a
calculation program for any earthquake load (RILEM, 1994). Since load cycles from actual
earthquakes occur very irregularly, unlike those of quasi‐static tests, the number of cycles
at which the deformation peaks is generally small, while numerous load changes occur at
comparatively low deformation levels. Figure G2‐10 shows the result of a non‐linear
calculation showing the moment‐rotation curve of a frame corner joint using dowels
under the impact of the El Centro earthquake, whose acceleration values were increased
by 50% for this calculation.
570
Joints subject to seismic loads
If, when designing structures, the seismic action is reduced by dividing by the behaviour
factor q and hence the ability to dissipate energy is taken into consideration, it is crucial
to ensure that the joints in which energy dissipation should take place can actually un‐
dergo plastic deformation, before connected timber members or additional joints, such
as anchors in concrete, fail. Timber members or additional joints must accordingly with‐
stand greater loads before failing than those joints intended to function as dissipative
areas. The higher load‐bearing capacity of the members and remaining joints thus pre‐
vents brittle failure before plastic deformation in the dissipative joints.
M (Nm) x104
3,2 m
2,0 4
t = 0,2 s t = 2,0 s
1,0 3
2,4 m
1 5
0
2
-1,0 6
-2,0
A -0,02 -0,01 0 0,01 0
t
ϑ (rad)
28 s
G2.5 Literature
A. Ceccotti, original Article C17, STEP 1995.
Ceccotti A. (Hrsg.) (1989). Structural behaviour of timber constructions in seismic zones.
CEC DG III ‒ University of Florence Workshop.
RILEM TC 109 TSA (1994). Timber structures in seismic regions: RILEM State‐of‐the‐Art Report, Material and
Structures 27:157‐184.
Yasumura M. (1988). Experiment on a three‐storied wooden frame building subjected to horizontal load.
International Timber Engineering Conference, Seattle, pp. 262‐275.
571
G3 Earthquake‐compliant
structural details
Original article: A. Ceccotti, P. Touliatos
When designing structures in areas at risk of earthquakes, one of the key tasks is to di‐
mension load‐bearing members: the corresponding rules set out in the provisions are
particularly important. However, the structural engineer must also acknowledge that
merely following the specified design rules to the letter alone will not ensure an earth‐
quake‐proof construction. The arrangement of members and structural details are also
important to ensure a secure construction. As experience shows, for smaller structures
with more regular plan and elevation, compliance with certain minimum member sizes
and detailing rules of joints may suffice to achieve earthquake‐proof construction with‐
out calculation. In addition, non‐load‐bearing components such as lightweight partition
walls or floor build‐ups impact on damping due to friction and hence affect the entire
dynamic behaviour of the construction. This effect is only quantifiable in exceptional
cases, but is normally favourable since it involves additional energy being dissipated. In
many countries, code provisions for design in seismic areas hence include minimum
dimensions and other requirements with corresponding examples, which eliminate the
need to verify smaller structures.
In EC 8, this scope was not considered due to the diversity of construction techniques
used among European countries. For the following reasons, in addition to calculation
rules, EC 8 nevertheless includes a series of recommendations for structural details.
There is a need to comply with specific details
To guarantee the validity of the calculation rules, which are based on specific
assumptions of structural details,
To ensure that the assumed level of ductility is guaranteed, whereupon the
correct behaviour factor q for calculating the inertial forces can be selected.
One relevant example is detailing rules for joints with mechanical fasteners,
which should show plastic behaviour before the maximum load is reached.
However, EC 8 does not impose any constraints on structural engineers in terms of possi‐
ble timber construction designs. In principle, any type of timber joint can be used, pro‐
vided it meets specific ductility requirements, which can be verified by tests (see Article
G2). In most cases, few rules concerning structural detailing need to be considered which
may eliminate the need to perform such tests. However, solely complying with a range of
detailing rules in itself does not spawn a satisfactory design in most cases. A list including
573
Earthquake‐compliant structural details
all possible details could never cover all conceivable cases. Accordingly, it is crucial to
understand the significance of the recommendations on structural detailing and apply
them accordingly. With this in mind, the current article addresses the key structural
details of a timber construction subject to earthquakes in generic form. The structural
engineer must transfer the solutions shown here to individual cases.
G3.1 Structural continuity
Earthquake actions can be considered horizontal, which, unlike vertical actions that only
impact on a comparatively small part of the structure, generate loads throughout the
entire structure (see Figure G3‐1). Although the same also applies for wind loads, in
Europe, earthquake loads are often decisive, particularly with heavier constructions
(Ceccotti and Larsen, 1989).
Figure G3‐1 Different structural effects of vertical and horizontal loading. (STEP 1995 Article D10)
The continuous connection of the different members and their effectiveness in different
load directions (e.g. tension and compression) are thus key prerequisites to ensure a fa‐
vourable response in the event of an earthquake. For shear walls in timber frame construc‐
tions, the individual frames must be interconnected, to activate as many dissipative areas
as possible and render the structure more effective overall. Figure G3‐2 shows the key
details and solution examples required to achieve interaction of the entire structure. Floors
should always be designed as diaphragms. For this purpose a ring beam must be installed
that absorbs the tension forces arising when the floor is loaded in plane (Figure G3‐2b). The
necessary continuity and force transfer at the floor corners can be guaranteed using a
574
Earthquake‐compliant structural details
smaller spacing between the sheathing and ring beam (Figure G3‐2a), where attention
must be paid to avoid choosing overly small spacing as this may lead to premature failure.
The vertically arranged members must also be interconnected through the floors, so tensile
forces can be transferred, which may also be generated from the vertical accelerations in
combination with the other internal forces (Figure G3‐2c). The connection between upper
and lower load‐bearing timber frame walls by nailing the respective sheathing to the ring
beam does not suffice when it comes to transmitting tensile forces, since the timber ring
beam is subject to tensile stress perpendicular to the grain during an earthquake. The forc‐
es perpendicular to the grain can be transferred e.g. by using a nailed‐on reinforcement of
wood‐based panels or crossbanded LVL for the ring beam.
Figure G3‐2 Details assuring structural continuity under horizontal actions. a corner reinforcement,
b tension joint of the ring beam, c tension joint of vertical members, d prevention of sliding
and uplifting from foundations. (STEP 1995 Article D10)
The joints connecting the timber structure with the foundation are particularly crucial, to
avoid uplift and sliding (Figure G3‐2d). Openings in horizontal diaphragms and shear walls
are weak points and should be reinforced. Since diaphragms have to distribute loads,
openings such as stairwells are particularly critical and their impact on the overall behav‐
iour must also be taken into consideration. Figure G3‐3 clarifies those diaphragm and
shear wall areas which require specific focus. Tensile stresses perpendicular to the grain
575
Earthquake‐compliant structural details
should be avoided at all costs. If tensile stresses perpendicular to the grain cannot be
avoided, EC 8 specifies that “additional measures to avoid splitting must be incorporated”
(e.g. punched metal plate fasteners or plywood panels, Figure G3‐4, or a reinforcement
with fully threaded screws inserted perpendicular to the grain).
Figure G3‐3 Detailing examples for a timber frame building. a prevention of uplift, b tensile joint in vertical
members, c reinforcement of wall opening by adding studs and lintels, d reinforcement of
diaphragm openings by adding girders and trimmer joists, e stiffening of floor beams by adding
blocks, f prevention of sliding. (STEP 1995 Article D10)
Figure G3‐4 Detailing against tensile stresses perpendicular to the grain in joints. a, c: poor, b, d, e: good.
(STEP 1995 Article D10)
576
Earthquake‐compliant structural details
In addition, joints must also be capable of withstanding alternating loads, since the load
direction varies on multiple occasions during an earthquake. For these reasons, simple
contact joints, which are only capable of transferring compressive forces, are unsuitable.
In Figure G3‐5, options for additional measures are shown to secure these joints. To
prevent differing displacements of foundations, individual foundations should be avoided
or interconnected. For homes constructed on softer ground in particular, the foundation
must be configured such that it is sufficiently rigid and acts like a “raft” when exposed to
earthquake loads.
Figure G3‐5 Possible provisions against the loosening of support in old and modern constructions.
(STEP 1995 Article D10)
G3.2 Regularity of the building
Regularity in the plan and elevation of a building are crucial to ensure optimal behaviour
in the event of an earthquake. Regular and symmetrical structures minimise torsional
effects, which are difficult to estimate using normal calculation methods. Calculations of
the effects of eccentrically exerted inertial loads on irregular structures are often unreal‐
istic, due to the global action of earthquakes and impreciseness in stiffness distribution.
Instead of such calculations, the building should be regularly constructed and the brac‐
ings designed to withstand earthquakes should be uniformly distributed (Figure G3‐6).
577
Earthquake‐compliant structural details
This means torsional stresses can be virtually prevented and the design results obtained
are also more reliable. If larger openings are unavoidable, the resulting eccentricity
should be limited by additional bracing measures such as frames or truss‐type compo‐
nents (see Figure G3‐6f). Another option would be to reinforce existing shear walls
choosing larger panel thicknesses, smaller fastener spacings or double‐ rather than sin‐
gle‐sided sheathing. Non‐load‐bearing partition walls also respond positively in an earth‐
quake, since they help dissipate energy via hysteresis and friction.
Figure G3‐6 Schematic examples of distribution of bracings. a, e: poor, b, f: fair, c, d: good.
(STEP 1995 Article D10)
G3.3 Structural compatibility
Problems may occur when members or structural parts of different stiffness are inter‐
connected. Examples include joints between a timber construction and a chimney or a
masonry wall. If larger differences in stiffness exist between connected members, both
structures could either be constructed such that they could transfer their portion of
seismic load independently of the other structure, or both structures could be connected
to form a single structural unit. In the example shown in Figure G3‐7, the glass façade is
independent of the timber structure.
The joints between external masonry walls and an internal timber construction are often
not designed to ensure the construction as a whole is capable of withstanding earth‐
quake loads. Since less rigid timber constructions are generally subject to larger defor‐
mations than more rigid masonry constructions, the timber construction tends to rest
against the masonry walls, which means they have to withstand the horizontal inertial
forces of the entire construction and must be designed to take this into consideration
(Figure G3‐8). If the masonry is incapable of withstanding the entire load, the following
578
Earthquake‐compliant structural details
approach is also possible: the masonry can be considered a mass without rigidity (dam‐
aged due to cracks) and the entire load is transferred by the timber construction. The
construction in Figure G3‐9 is designed such that in the event of an earthquake, the
masonry would collapse without causing excessive damage to the timber construction.
The failure of the masonry leads to a great deal of energy being dissipated, which means
the timber construction and its far lower mass can withstand the earthquake without
excessive damage.
Figure G3‐7 Timber structure with specially designed joints to withstand earthquakes allowing for relative
displacements between load‐bearing and non‐load‐bearing members. Detail p: a beam
reinforcing the frame, b spring elements, c joint of bracing system, d steel tension rod with
springed support, e steel column, f independent movements of the main three‐hinged frame
and the external façade frame, g glulam beam, h glass façade. (STEP 1995 Article D10)
579
Earthquake‐compliant structural details
Figure G3‐8 Mixed masonry‐timber building (Greek islands, 1500 B. C.). The timber construction only carries
vertical loads, while the masonry also accommodates horizontal loads. The floors cannot act as
diaphragms, due to their workmanship. (according to Touliatos, 1993, STEP 1995 Article D10)
G3.4 Ductility and energy dissipation
In historic buildings, such that shown in Figure G3‐9, the required energy dissipation was
achieved by friction between wood and masonry as well as plastic compression defor‐
mations perpendicular to the grain in the timber. In modern, lighter structures, the ener‐
gy dissipation is usually guaranteed by using plastic deformations in joints with mechanical
fasteners (see Article G2).
When verifying structures with reduced inertia forces in accordance with EC 8 (q > 1.5;
see Article G2) the required ductility and energy dissipation characteristics must be con‐
firmed with tests. For specific configurations however, where experience suggests that
ductile and plastic behaviour can be expected, no tests are required.
580
Earthquake‐compliant structural details
Timber frame panels with sheathing nailed to the studs show far more ductile behaviour
than when using diagonals to ensure the lateral load‐bearing capacity. According to EC 8,
plastic behaviour of timber frame panels can be assumed, when, among other things, the
sheathing comprises wood‐based panels at least 4 ∙ d thick and the nail diameter d does
not exceed 3.1 mm. This provision applies to particleboards with minimum density k of
650 kg/m3, plywood at least 9 mm thick and particle‐ or fibreboards at least 13 mm thick.
Timber‐to‐timber and steel‐to‐timber joints with nails or dowels can be considered suffi‐
ciently ductile, if the timber thickness of the connected members is at least 8 ∙ d and the
dowel diameter does not exceed 12 mm. This specification is imposed because joints
respond favourably to cyclic loads when designed such that failure mechanism 3 in ac‐
cordance with Johansen governs (two plastic hinges per shear plane).
Of course, the use of other mechanical fasteners and timber member sizes in construc‐
tions at risk of earthquakes is also possible. However, if a behaviour factor q > 1.5 is used
in the design, the plastic behaviour of joints must be verified with tests (see Article G2).
For joints with dowel‐type fasteners, failure mechanism 3 should always be the goal, so
that the plastic embedment deformations of the wood and the plastic bending defor‐
mations of the fasteners can boost the joint ductility.
Even when the structure is designed for q = 1.5, meaning the plastic behaviour is no
longer required computationally, the use of slender fasteners with plastic deformation
potential is recommended. This means a safety reserve, without additional costs for the
structure. The recommendations for detailing timber constructions in accordance with
EC 8 should thus be considered independently of the underlying behaviour factor q as far
as possible.
581
Earthquake‐compliant structural details
Figure G3‐9 Example of a mixed masonry‐timber building (Greek islands, 1800 A. D.). a timber diagonal,
causing compression perpendicular to the grain in the lower corner, b anchoring of the timber
construction in the masonry, c masonry as support for the upper timber construction,
d timber columns to accommodate vertical loads after the masonry fails (see detail lower right),
e bent timber corner piece to reinforce the timber frame wall, f joint that is simply to replace
after damage (see detail lower left), g corner piece made of wood to reinforce the roof construc‐
tion, h tie beam to prevent horizontal forces on walls, i brick infill. The floors cannot act as
diaphragms, due to their workmanship. (according to Touliatos, 1993, STEP 1995 Article D10)
582
Earthquake‐compliant structural details
G3.5 Literature
A. Ceccotti, P. Touliatos, original Article D10, STEP 1995.
Ceccotti A. and Larsen H.J. (1988). Background Document for specific rules for timber structures in EC 8.
Report EUR 12226 for the Commission of the European Communities, Brussels.
Touliatos P. (1993). Seismic disaster prevention in the history of structures in Greece. National Technical
University of Athens.
583
G4 Damages in hall structures
Authors: Matthias Frese, Ann‐Kathrin Grün and Hans Joachim Blass
In January 2006, Germany and neighbouring countries saw a series of collapses of what
were, at times, decades‐old timber roof structures. Although most failed due to the
weight of snow, it was not possible to conclude that an exceptionally high snow load was
the only possible cause of failure. Several recent damage cases have generally shown a
combination of multiple causes to be considered or a need to differentiate between
causes and catalysts, to explain structural failures; this is what applies to failures of tech‐
nical systems in general, barring a few exceptions (Schmitt‐Thomas, 2005).
At that time, there was no uniform statistical and systematic consideration of damage
cases involving timber‐hall structures, which could be used to draw conclusions for those
planning and designing such structures. Damage cases, some of which are described in
specialist literature, tend to be individual considerations, which makes it difficult to form
concrete conclusions about their significance. This is largely what spurred on the re‐
search project in 2006 named “Damage analysis of timber‐hall structures”. It aimed to set
out the basis for an overall consideration (constructing a database with damage cases,
developing a system for classifying, describing and analysing damage), citing the causes
of damage and formulating conclusions. The project results are shown in a research
report (Blass and Frese, 2010). In the meantime, there was scope to expand the database
and include certain damage cases and it seemed expedient to re‐analyse data material
compiling over 700 damage cases. The conclusions saw general requirements for reliable
structures emerge, some of which had already resurfaced multiple times.
G4.1 Data collection
Origin of data
The data on the damage cases came from expert reports, an old database of the
Studiengemeinschaft Holzleimbau e. V., the collections of the Technical Universities of
Munich and Graz as well as the archive of renovation projects of a glulam company based
in Northern Germany, Gebr. Schütt KG. Some cases were taken from literature or passed
on to the authors directly. Accordingly, the data was inconsistent and exposed qualitative
585
Damages in hall structures
differences. The data compiled was the result of a non‐controllable survey: the material
used was consequently that which was available and accessible. This may lead to con‐
straints concerning the representative nature of timber‐hall structures in a general sense
and the significance of the results. Most of the damage descriptions assessed concerned
timber hall structures based in Germany.
Data classification
Structures generally comprise recurring individual members. These, in turn, are assem‐
bled based on recurring construction principles. Structures of a specific group, timber hall
structures in this case, can thus be easily compared. This simplifies the data collection
and common consideration of their characteristics, damages and their causes. A system
was developed, allowing damage cases to be reliably recorded in a database. According
to this, the features of a damage incident are assigned based into five thematic groups:
construction, member, material and damage features as well as error sources. Figure G4‐1
shows the systematic approach, with a selection of features and related or exemplary
characteristics. Based on quantitative or metric parameters such as the height above sea
level, qualitative features such as district or heating are collated, with a specific vocabu‐
lary consisting of keywords. The ongoing classification system, see Blass and Frese (2010),
includes over 60 qualitative and quantitative features as well as multiple keywords, to
describe a damage case consistently.
The building features convey an overview of the damaged hall structures. Member and
material features are detail‐oriented and directly linked to damage. They provide infor‐
mation on structural systems, damaged members and the materials used. Damage pa‐
rameters describe the damage pattern, also in terms of consequences for the structural
safety of members or the overall construction. Initial damage is first to occur. This is in
contrast to secondary damage, when damage is e.g. inflicted through damage on another
member. Not every incident of initial damage necessarily represents a type of failure;
initial damage in the form of critical deformation, moisture penetration, rot, blue stain
and mould fungi, corrosion or harmless cracks in the grain direction (cf. Figure G4‐2) does
not generally lead to a member or structure failing. Tensile failure includes bending fail‐
ure, since this is normally triggered by a local tensile failure. Significant cracks in the grain
direction (Figure G4‐3 and Figure G4‐8), shear (Figure G4‐2) and tensile failures are par‐
ticularly significant and allow the strength properties of solid timber and glulam to be
demonstrated based on real member size. This “testing” independent of laboratory
methods shows the strengths and weaknesses of the material against a background of
changing climate, load history and member size.
586
Damages in hall structures
Source Expert 1, 2, 3…
District e.g. Karlsruhe city
Height above sea level e.g. 115 m above sea level
Year of construction e.g. 1979
heated
Building features Heating unheated
Construction phase
Ice sports
Storage
Production
Usage Equitation
Sw imming
Sport
Warehouse
Assembly
Arch
Continuous beam
Single span beam
Truss
Affected structural system Beam w ith internal hinges
Cantilever
Frame
Girder grid
Member features Span of structural system e.g. 23 m
Bending member
Bending w ith compression
Column
Affected member Column w ith bending moment
Truss
Tensile member
Glulam
Material Solid timber
Strength class e.g. GL28
Damage data Material features Adhesive e.g. Resorcinol formaldehyde
Manufacturer Manufacturer 1,2,3…
Critical deformation (def)
Buckling (buck)
Moisture accumulation (moistacc)
Rot
Blue stain or mould fungi (fungi)
Corrosion (Corr)
Initial damage Compression w rinkles (w rink)
Compr. fail. perp. to grain (com perp)
Cracks parallel to grain (crack par)
Shear failure (shear)
Tensile failure (tens)
Damage features Block shear failure (block shear)
Unknow n (unknw n)
Guaranteed
Still guaranteed
At risk
Structural safety Failure member
Failure structure
Collapse member
Collapse structure
Planning (plan)
Execution (exec)
Assembly (ass)
Building physics (b phys)
Load (load)
Error sources Construction (constr)
Material quality (mat qual)
Moisture (moist)
Insects (ins)
Changing climate (clim)
Sw elling shrinking (sw eshr)
Maintenance (maint)
Figure G4‐1 Selection of keywords and features. Abbreviations in brackets for Table G4‐12.
587
Damages in hall structures
Figure G4‐2 Initial damages – rot (top left), blue stain or mould fungi (top right), cracks in grain direction
(minor shrinkage cracks, lower left), shear failure (part of beam in compression zone protruding
at end grain, lower right).
The characteristics of structural safety are assessments of experts in their reports. They
relate to damaged members or entire structures. Guaranteed indicates sufficient struc‐
tural safety; still guaranteed means that rehabilitation works are recommended soon; at
risk are members or constructions including serious initial damage that leads to less
structural safety than required. Significant repair works are required in this case. The
keywords failure member or failure structure denote an obvious loss of structural safety,
but not a collapse. The member or construction can no longer fulfil its function (Fig‐
ure G4‐3, left). To meet subjective safety requirements, e.g. propping or other safety
measures must be implemented immediately. Collapse structure denotes the complete
loss of a structure, collapse member a partial loss. A repair involves considerable ex‐
pense, which is why the characteristics of structural safety are subject to an order of
priority, identified by a spectrum between safe and collapse.
588
Damages in hall structures
Figure G4‐3 Cracks in pitched cambered beams, immediate propping/safeguarding required (left and centre),
professionally retrofitted beams (right)
The characteristics of the error sources correspond to 12 categories, which are then
allocated to the causes of initial damage. A selection of such causes is included in Ta‐
ble G4‐1. The error source planning involves the technical planning of a construction or
member. Typical for the error source building physics is generally an unfavourable ther‐
mal transport. Construction involves constructive characteristics, which lead to directly
related circumstances in a damage case. Although originating in technical planning, build‐
ing physics and construction are classed as sufficiently important to be individual error
sources. Execution involves qualitative performance of works on the construction site
and identifies unfavourable deviations from planning. Assembly involves defects when
assembling members and constructions. Load refers to any type of overload. Moisture
refers to the ingress of humidity, which is always due to external causes and thus ex‐
cludes damaging humidity caused by e.g. thermal radiation (in indoor ice rinks) or a lack
of detailing for durability. Changing climate is the cause of resulting residual stresses in
glulam cross‐sections (see Möhler and Steck, 1980; Häglund, 2010). Swelling or shrinking
refers to the long‐term directional volume changes affecting an entire member. Swelling
or shrinking is only cited in combination with the error source construction, when con‐
structive characteristics prevent shrinking (a lack of shrinking does not lead to damage,
even if shrinking is prevented). For certain error sources, responsible persons can be
nominated: Planning, building physics and construction concern the entire planning
phase of a construction and hence the designer, execution and assembly the building
contractor. Material quality is almost always associated with manufacturers of glulam
and maintenance to owners or authorised representatives.
589
Damages in hall structures
Table G4‐1 Categories of error sources.
Keyword Cause
Planning Infringements of provisions/generally accepted technical rules;
Failure to heed engineering expertise, e.g. in specialist literature;
static calculation errors
Execution Use of green wood; lack of drainage due to incorrectly installed outlets;
incorrect number or diameters of fasteners; non‐compliance with building
execution provisions; unfavourable changes of the structural system,
the cross‐sectional dimensions or strength class
Assembly Transport damage; insufficient member protection against weathering impacts;
a lack of support measures during the construction phase
Building physics Thermal radiation (indoor ice rinks); solar radiation; members exposed to
indoor and outdoor climates (e.g. gable beams)
Load Overload with regard to permanent loads, imposed loads, snow and wind loads
as well as accumulation of water on flat roofs
Construction Notched beams; dowel circles in frame corners; beams with openings; defective
detailing for durability; tensile stresses perpendicular to the grain in joints;
prevention of shrinking (e.g. slotted‐in steel plates); curved or kinked
members subject to bending; unwanted restraints; structures that are difficult
to manage in terms of design and execution
Material quality Defective quality of wood, finger joints or bond lines concerning required
characteristics; preliminary damage
Moisture High moisture content due to roof leaks or sprinkler systems in indoor
riding arenas
Changing Cyclic changes in moisture content in glulam due to changing climate;
climate leading to changing humidity gradients in the cross‐section
Swelling or Long‐term increase or decrease in moisture content leading to
shrinking swelling or shrinking
Maintenance Failures during inspection, maintenance and repair
G4.2 Presentation of the structures and their damage
General points
The database includes 709 datasets. Each concerns a single independent incidence of
initial damage within a construction and concerning the construction, member and mate‐
rial features as well as the error sources to which the initial damage is attributed. Multi‐
ple attributions usually apply to error sources. Since many structures are subject to two
or more incidents of damage, the 709 incidents of damage affect a total of 529 struc‐
tures. In 44 cases, it was not possible to attribute a clear expression of characteristics to
initial damage. In such cases, the initial damage was described as unknown. Due to the
590
Damages in hall structures
lack of details in the damage descriptions, not all the corresponding characteristics were
defined for all features or error sources. The overall figures in the following tables and
diagrams thus differ.
Building features
Most of the structures are located in the former West German states (Figure G4‐4). Data
on damage cases in the new German states was only available in individual cases. This is
due to the divide up until 1989, which meant dialogue between experts was very limited
up to the time of reunification and even for some considerable time thereafter. It is inac‐
curate, therefore, to conclude that timber‐hall structures in the new German states did
not suffer comparable damage. Most of the damage originated from the four most popu‐
lous West German states: Baden‐Württemberg, Bavaria, Lower Saxony and North Rhine‐
Westphalia as well as Schleswig‐Holstein, the active sphere of the directly cooperating
glulam construction company Gebr. Schütt KG (Table G4‐2). More concerning is the fact
that extending the database of 550 cases (values in brackets) in Blass and Frese (2010)
with 159 cases has led to the damage tripling in Schleswig‐Holstein as well as a clear
increase in Lower Saxony. If a similarly rigorous survey examined damage incidents in the
other German federal states, similar findings could be expected. Accordingly, the damage
actually accessible or located in the course of the damage survey and recorded here is
very likely to be less than a third of the actual extent of damage which could be described
and analysed as part of such assessment. With the snow load in mind, the height above
mean sea level of the structures was recorded (Figure G4‐5). It ranges from ‐1 m in the
North German lowlands up to 1145 m above sea level in the low mountain range, foot‐
hills of the Alps or the Alps themselves. The year in which the structures were construct‐
ed comprises the period from 1912 to 2006, the average year of which was 1980
(Figure G4‐6). 81% of the timber‐hall structures had a sealed building envelope. The
distribution of the service classes shows that service class 1 prevails (Table G4‐3). Most of
the structures are heated. In terms of usage, the damaged structures frequently include
sports halls, warehouses and production facilities and those used for conferences.
591
Damages in hall structures
Table G4‐2 Initial damage per federal state. In brackets: original analysis of 550 damage cases.
1 2 3 4
5 6 7-8 >8
Figure G4‐4 Distribution of initial damage over a German map with district borders.
592
Damages in hall structures
35
N 490
30 Mean 206
25 Min -1
Percent
20 Max 1145
15
10
5
0
0 200 400 600 800 1000
Height above sea level [m]
Figure G4‐5 Frequency distribution of height above sea level.
25
N 386
20 Mean 1980
Min 1912
Percent
15 Max 2006
10
0
1912 1924 1936 1948 1960 1972 1984 1996 2008
Year of construction
Figure G4‐6 Frequency distribution of the year in which the structure was built.
Table G4‐3 Service classes of the structures.
593
Damages in hall structures
Member and material features
Damage was mainly observed in single and multi‐span beams (including those with hinges)
as well as in frames (Table G4‐4). The 19 listed trusses denote non‐specific structural
systems, in which only individual members like chords (beams), compression or tension
members were affected. Most affected structural systems were statically determinate;
there were rarely statically indeterminate systems (Table G4‐5). The portions in both
tables thus correspond to each other.
The distribution of the spans of the structures is shown in Figure G4‐7. The modal value is
20 m and corresponds to the typical span for sports halls and indoor riding halls. Damage
to very wide‐spanned halls was relatively infrequently observed, but cf. Hansson and
Larsen (2005).
Table G4‐4 Affected structures.
Table G4‐5 Statical determinacy of structures.
594
Damages in hall structures
25
N 411
20 Mean 23
Min 4
Percent
15 Max 58
10
0
0 5 10 15 20 25 30 35 40 45 50 55 60
Span [m]
Figure G4‐7 Frequency distribution of span.
Table G4‐6 Affected members.
595
Damages in hall structures
Table G4‐7 Member shapes.
Table G4‐8 Construction materials.
Damage features and further considerations
Table G4‐9 shows the allocation of initial damage to the defined damage patterns. Over
70% is due to cracks parallel to the grain, which were mainly discovered in glulam mem‐
bers (cf. Table G4‐8). 5‐6% is allotted to tensile or shear failure and rot. Overall, 5% af‐
fected the serviceability or appearance of the members. Buckling and corrosion were
both in the region of 1%. Block shear and compression failure perpendicular to the grain
are individual cases. Compression wrinkles are not included in the database. In 6% of cases,
the damage descriptions lacked sufficient details to define the precise nature of the dam‐
age. The overwhelming frequency of cracks parallel to the grain in curved glulam beams
with opening moments (Figure G4‐3) triggered closer scrutiny into their origin (Blass and
Frese, 2010, Frese, 2011). Considering the anisotropy of the degrees of swelling and
596
Damages in hall structures
shrinkage in the curved areas (longitudinal/tangential‐radial ≈ 1:24), changes in moisture
content lead to curvature changes and, under unfavourable support conditions, to
stresses perpendicular to the grain. Different longitudinal shrinkage of the lamellae in the
edge and core areas due to moisture content changes also generates stresses perpendic‐
ular to the grain in combined curved beams, which exacerbate the formation of cracks in
curved glulam beams. To calculate the deformation of glulam members based on changes
in moisture content, reference is made to Blass and Frese (2010), the American Institute
of Timber Construction (1994) and Larsen and Riberholt (1983).
Expert reports in particular contain precise details of the relative crack depth (Figure G4‐8)
and moisture content of affected members. The diagram in Figure G4‐9 shows the rela‐
tions between both values. The illustration does not contradict the observation that
cracks parallel to the grain occurred more frequently in members that were attributable
to service class 1 or occurred at times when particularly low moisture content was rec‐
orded. Figure G4‐10 shows the distribution of member depth (in the area of the sup‐
ports) of glulam beams with shear failures. For 21 beams of over 29 registered shear
failures, the member depths are known and ranged between 650 and 2400 mm. The
distribution reflects that deep members with greater volume subject to shear have less
effective shear strength than less deep members. In the damage descriptions shown, no
shear failures of glulam members under 650 mm in depth were documented.
Figure G4‐8 Relative crack depth in glulam: partial (left) and continuous crack (right).
597
Damages in hall structures
Table G4‐9 Distribution of initial damage.
1.1
1.0
Relative crac k depth
0.9
0.8 Service class
0.7 1
0.6 2
0.5 3
0.4
0.3
0.2
0.1
0.0
5.0 7.5 10.0 12.5 15.0 17.5 20.0 22.5
Mean moisture c ontent [%]
Figure G4‐9 Relation between crack depth and moisture content.
598
Damages in hall structures
50
N 22
40 Mean 1352
Min 650
Percent
30 Max 2400
20
10
0
800 1200 1600 2000 2400
Beam depth at end support [mm]
Figure G4‐10 Frequency distribution of the depth of glulam members with shear failure at supports.
Table G4‐10 shows the distribution of consequences inflicted as a result of the initial
damage on the structural safety of the members or constructions. In just under a quarter,
structural safety is guaranteed (including still guaranteed), and at risk in another quarter.
Another quarter succumbed to failure or collapse. No assessments were available for the
remainder. The characteristic values guaranteed and at risk, the cumulative portions of
which amounted to almost 50%, show that after professional inspections, apparently
assessments can be made regarding the structural safety of members and constructions.
This can be used as the basis for suitable measures, which ensure structural safety in
future and extend the service life. The cumulative portion of the assessments still guar‐
anteed, threatened and failure member or structure (54%) shows that a large part of the
timber‐hall structures and their structural systems meet the basic requirements of struc‐
tures in the case of failure (EC 0). A distribution without such assessments, which gradu‐
ally subdivide the predominantly discontinuous transition between safe and collapsing
structure, would indicate that structures would become unsafe and collapse without
advance notice. As part of efforts to prevent a collapse, the above‐specified assessments
set out the scope for action within which prompt action can still make a difference (cf.
Figure G4‐3 on the left). Structures are subject to natural and man‐made wear. The in‐
spection, cf. Studiengemeinschaft Holzleimbau e. V. (2015), maintenance and repair of
structures, including those made of timber, thus becomes highly significant. Accordingly,
the planning must take into consideration the accessibility to load‐bearing members for
inspections as far as possible. However, the decisive factor dictating the success of a
maintenance measure is not only the expertise, but also the independent nature of the
individuals and companies commissioned to provide such service.
599
Damages in hall structures
Table G4‐10 Assessments of structural safety.
For 127 damage cases, the actual year – and for 100 the actual month – is known in
which the initial damage relevant to the structural safety took place. These damage cases
particularly include cracks parallel to the grain, tensile and shear failures. They occur in
relation to internal forces and moments, the values of which generally peak under the
impact of snow and accordingly show seasonal dependencies. While in Figure G4‐11
above, the years in which damage occurred are uniformly distributed (except 1956), in
Figure G4‐11 below, there is a far higher likelihood of occurrence during the January to
March period than for the remaining months. This therefore accurately reflects the as‐
sumption that a dry (internal) climate in the winter and early spring together with the
impact of snow is at least part of the trigger for the initial damage affecting structural
safety, particularly for cracks parallel to the grain. This fact notwithstanding, in the re‐
vised snow load standard of 2005 (DIN 1055‐5) or 2010 (NA to EC 1 Part 1‐3) − in contrast
to the 1975 version (DIN 1055‐5) − the characteristic value of the snow load to be applied
for higher elevations had significantly increased case‐by‐case, cf. Schroeter (2007). If
those timber‐hall structures from the database are isolated for which an effective snow
load according to present standards of more than 25% higher would apply than com‐
pared to the 1975 version, the portion of members and constructions suffering failure or
collapse would be 2 to 3 times as many compared with the other hall structures. In this
respect, excessive snow load from today’s perspective was previously responsible for
failure and collapses in certain cases. Increasing the characteristic values for the snow
load at higher elevations is the right step to help eliminate damage concerning structural
safety in specific geographic locations. In over 35% of the damage affecting structural
safety, at least five years had elapsed between the year of construction and damage
event, and even ten in most cases (Figure G4‐12). Premature emergence of damage,
dominated by cracks parallel to the grain, is what happens after timber is dried e.g. by
swift heating, which results in unfavourable residual stresses or tensile stresses perpen‐
dicular to the grain.
600
Damages in hall structures
30
N 127
25
20
Percent
15
10
0
1956 1962 1968 1974 1980 1986 1992 1998 2004 2010
Year of damage occurrence
30
N 100
25
20
Percent
15
10
0
1 2 3 4 5 6 7 8 9 10 11 12
Month of damage occ urrence
Figure G4‐11 Frequency distribution of the year (above) and month (below), in which initial damage affecting
structural safety took place.
40
N 114
35
Mean 12
30
25
Percent
20
15
10
5
0
2.5 7.5 12.5 17.5 22.5 27.5 32.5 37.5
601
Damages in hall structures
Error sources
A total of 1282 error sources could be allocated to 709 damage incidents. This means
that each damage incident tends to be associated with two sources on average. If a caus‐
al link is established (cf. Figure G4‐13), an error source was allocated without weighting
of damage, namely it was either linked completely or not at all. Here too, there is no
order of priority within the sources, which would reveal a specific quantitative signifi‐
cance or allow the individual sources to be compared. The individual portions of error
sources are shown in Table G4‐11. In this table, the key causes for the damage incidents
analysed here are shown: The main damage occurs in relation with the construction
(29%). These cases primarily involve incidents where stresses perpendicular to the grain
are exerted or deficient detailing for durability becomes apparent. Changing climate
(12.6%), swelling or shrinking (9.8%), only observed in isolation, load (9.5%), material
quality (7.1%), planning (9.0%), building physics (6.3%) and execution (5.2%) are of mod‐
erate importance. Unfavourable influences of assembly, moisture and maintenance play
a minor role.
Figure G4‐13 Circumstances, which may lead to damage. Error source planning/execution: torsional restraints
at supports not built (top left), error source building physics: possible crack formation in glulam
by local heating/drying behind planned glazing (top right), error source assembly: weathering of
glulam (lower left) and construction of dowel circles, in which residual stresses are caused by
shrinking (lower right).
602
Damages in hall structures
Table G4‐11 Distribution of the general error sources.
603
Damages in hall structures
Consequences
Given the very limited scope for engineers to have a direct influence on exceptionally
high loads, e.g. those exerted by snow loads far beyond characteristic values, and on
material quality, the key to avoiding damage according to this analysis is linked to improv‐
ing the construction process and design via technical calculations and considerations. The
VDI guideline 2221 (1993) includes cross‐industry indications for successful construction
The key conclusions, which are individually drawn for those responsible for the construc‐
tion process, are as follows:
Designer: Avoiding tension stresses perpendicular to the grain or arranging rein‐
forcements; consideration of detailing for durability and avoiding of exposed
spruce glulam; awareness of the unfavourable influence of changing climates and
high temperatures on glulam members, more in‐depth details included in Möhler
and Steck (1980), Häglund (2010) and Gamper et al. (2013); consideration of
swelling and shrinking also concerning the anisotropy of the degree of swelling
and shrinking artificially generated in glulam; design of structures with only limited
scope for damage from impacts beyond standards, closer details, concerning gen‐
eral building construction, in Kersken‐Bradley (1992), Pötzl (1996) and Harte et al.
(2007); Ensuring accessibility to members of structural systems for inspections as
well as consideration of the scope to replace members at risk and those subject to
significant wear; planning of monitoring methods (Riedner, 2007; Fellmoser, 2011;
Pawlowski et al., 2013); finally measures to offset time pressure during the plan‐
ning and design phase.
Glulam manufacturer: Timber drying up to the equilibrium moisture content to be
expected in the building, see Gamper et al. (2013); where applicable, surface and
especially end grain protection, to delay any exchange of humidity between wood
and the surroundings, see Möhler and Steck (1980).
Building contractor: Careful protection of the wood during the construction phase.
604
Damages in hall structures
Table G4‐12 Initial damage and attributed error sources. Abbreviations in Figure G4‐1.
mat not
plan exec ass B phys load const moist clim sweshr maint Σ
qual spe.+
1 9 2 2 9 2 1 0 0 1 0 1 28
def *
3.57 32.1 7.14 7.14 32.1 7.14 3.57 0 0 3.57 0 3.57
2 2 0 0 3 0 0 0 0 0 0 1 8
buck
25 25 0 0 37.5 0 0 0 0 0 0 12.5
1 0 1 7 0 2 0 3 0 0 1 0 15
moistacc
6.67 0 6.67 46.7 0 13.3 0 20 0 0 6.67 0
2 1 0 6 1 22 1 11 0 0 7 4 55
rot
3.64 1.82 0 10.9 1.82 40 1.82 20 0 0 12.7 7.27
0 1 0 5 0 0 0 2 0 0 0 0 8
fungi
0 12.5 0 62.5 0 0 0 25 0 0 0 0
0 2 0 3 0 1 0 3 0 0 0 0 9
corr
0 22.2 0 33.3 0 11.1 0 33.3 0 0 0 0
comp 4 0 0 0 1 0 0 0 0 0 0 0 5
perp 80 0 0 0 20 0 0 0 0 0 0 0
crack 87 29 22 51 51 325 61 4 155 119 1 60 965
par 9.02 3.01 2.28 5.28 5.28 33.7 6.32 0.41 16.1 12.3 0.1 6.22
7 5 0 4 11 12 9 1 6 5 0 4 64
shear
10.9 7.81 0 6.25 17.2 18.8 14.1 1.6 9.38 7.81 0 6.25
8 10 0 3 14 2 13 2 0 0 1 2 55
tens
14.6 18.2 0 5.45 25.5 3.64 23.6 3.6 0 0 1.82 3.64
shear or 1 0 0 0 8 3 2 0 1 0 0 0 15
tens 6.67 0 0 0 53.3 20 13.3 0 6.67 0 0 0
block 0 0 0 0 0 0 0 0 0 0 0 2 2
shear 0 0 0 0 0 0 0 0 0 0 0 100
2 8 2 0 24 3 4 1 0 0 2 7 53
unknwn
3.77 15.1 3.77 0 45.3 5.66 7.55 1.89 0 0 3.77 13.2
115 67 27 81 122 372 91 27 162 125 12 81 1282
Σ#
8.97 5.23 2.11 6.32 9.52 29 7.1 2.11 12.6 9.75 0.94 6.32 100
Error sources (1st row) and initial damage (1st column); number (top) and percentages (bottom)
* Example: 1/28 ∙ 100 = 3.57%
+ Not specified
# Totals and percentages in Table G4‐11
605
Damages in hall structures
G4.3 Summary
Around 70% of the independent damage incidents analysed here are cracks parallel to
the grain. Rot, shear and tension failures respectively each comprise around 6%, while
the remaining 12% concern serviceability. Damage is particularly observed in association
with constructions subject to stresses perpendicular to the grain. Changing climates
meanwhile, which trigger cyclical changes in moisture content, are significant for the
formation of cracks parallel to the grain in glulam. The shrinking of entire cross‐sections
as well as error sources concerning load, material quality, planning, building physics and
execution are of minor significance for damage. An even less important role is played by
detrimental influences from assembly, maintenance and moisture. To avoid damage
during construction, it is imperative to focus more closely on the design, calculation and
detailing. Tension stresses perpendicular to the grain, as required for local or global equi‐
librium within members, should be avoided or accommodated by using reinforcements.
The damage analysis supports the following findings in particular: Only glulam members
with a large volume subject to shear were prone to shear failures; increasing the charac‐
teristic values for snow load in the National Annex of the EC 1 for higher elevations in
comparison with the snow load values of 1975 is one correct step to ensure structural
safety, independently of any geographical feature.
606
Damages in hall structures
G4.4 Literature
This article has already been published (in German): Frese M, Grün A.‐K. and Blass H.J. (2015). Schäden an
Hallentragwerken aus Holz: Beschreibung – Ursachen – Vermeidung. KIT Scientific Working Paper No. 31,
ISSN 2194‐1629.
Schmitt‐Thomas K.G. (2005). Integrierte Schadensanalyse – Technikgestaltung und das System des Versagens.
2. Aufl., Springer‐Verlag, Berlin.
Blass H.J. and Frese M. (2010). Schadensanalyse von Hallentragwerken aus Holz. Karlsruher Berichte zum
Ingenieurholzbau Band 16, KIT Scientific Publishing, Karlsruhe.
Möhler K. and Steck G. (1980). Untersuchungen über die Rißbildung in Brettschichtholz infolge
Klimabeanspruchung. Bauen mit Holz 82:194‐200.
Häglund M. (2010). Parameter influence on moisture induced eigenstresses in timber. European Journal
of Wood and Wood Products 68:397‐406.
Hansson M. and Larsen H.J. (2005). Recent failures in glulam structures and their causes. Engineering Failure
Analysis 12:808‐818.
Frese M. (2011). Wechselwirkung zwischen der Anisotropie der Schwind‐ und Quellmaße sowie
Holzfeuchteänderungen in der Ebene von gekrümmtem Brettschichtholz. European Journal of Wood and Wood
Products 69:359‐367.
American Institute of Timber Construction (1994). Timber Construction Manual. 4th Ed., John Wiley & Sons,
New York.
Larsen H.J. and Riberholt H. (1983). Trækonstruktioner, Beregning. SBI‐Anvisning 135. Statens
Byggeforskningsinstitut, Denmark.
Studiengemeinschaft Holzleimbau e.V. (2015). Leitfaden zu einer ersten Begutachtung von Hallentragwerken
aus Holz. Online‐Publikation, 9.4.2015,
http://www.brettschichtholz.de/publish/binarydata/pdfs/aktuelles/stghb_leitfaden‐hallentragwerke‐
2014_print_140218.pdf.
Schroeter H. (2007). Erläuterungen und Beispiele zur Lastnorm DIN 1055 neu. Bautechnik 84:559‐571.
VDI‐Richtlinie 2221 (1993). Methodik zum Entwickeln und Konstruieren technischer Systeme und Produkte.
Gamper A., Dietsch P., Merk M. and Winter S. (2013). Gebäudeklima – Langzeitmessung zur Bestimmung der
Auswirkungen auf Feuchtegradienten in Holzbauteilen. Bautechnik 90:508‐519.
Kersken‐Bradley M. (1992). Unempfindliche Tragwerke – Entwurf und Konstruktion. Bauingenieur 67:1‐5.
Pötzl M. (1996). Robuste Tragwerke − Vorschläge zu Entwurf und Konstruk on. Bauingenieur 71:481‐488.
Harte R., Krätzig W.B. and Petryna Y.S. (2007). Robustheit von Tragwerken – ein vergessenes Entwurfsziel?
Bautechnik 84:225‐234.
Riedner W. (2007). Sicherheit und Überwachung von weitgespannten Hallensystemen. Bautechnik 84:78‐80.
Fellmoser P. (2011). Monitoring von Holzkonstruktionen. Bauingenieur 86:541‐543.
Pawlowski R., Henke K., Schregle P. and Winter S. (2013). Überwachung von Bauwerksverformungen mittels
digitaler Figureverarbeitung. Bauingenieur 88:214‐221.
607
H
Annexes
Annex 1: Dynamic modulus
of elasticity
In the following section, exemplarily for the case of longitudinal vibrations, the equation
is derived with which the dynamic modulus of elasticity can be calculated from the first
Eigenfrequency measured. The equation (1) derived here corresponds to equation (B5‐1).
Required fundamental equations, for definitions see Figure A1:
F u
E F AE
A x
F u
F F ( x ) F ( x dx ) dx A E dx (a)
x x x
2
u
F ma m 2
(b)
t
Area A,
Modulus of elasticity E
u
x x + dx
Figure A1 Definitions.
Assumption: Area A and Young’s modulus E are constant → no transverse contraction.
Equating (a) and (b):
2u u
m A E dx
t 2
x x
2 2
u A dx E u
2
2
t m x
611
Annex 1: Dynamic modulus of elasticity
With A ∙ dx = volume V, density = m/V:
2 2 2
u E u 2 u
2 c long 2 (c)
t
2
x x
For long, thin bars (negligible transverse contraction), the following thus applies:
E
c long (d)
where clong = longitudinal sound velocity and the wavelength :
c long T where T period (e)
Now, the case of resonance is considered:
resonance bar length n n 1,2,3,... (f)
2
From equations (e) and (f), the following first mode of vibration emerges:
2 2
T c long
c long n c long n T
Inserting in equation (d) with T = 1/f:
2 E 4 2 E
E 4 f
2 2
(1)
T T 2
612
Annex 2: Stress interactions
Multiaxial stresses
The anisotropy of the wood means stresses exerted at an angle to the grain lead to mul‐
tiaxial stresses. This can be seen in Figure A2, where compression compr exerted at an
angle to the grain generates normal stresses parallel and perpendicular to the grain
parallel and perpendicular and shear loads . In EC 5, this fact is taken into consideration via
interaction equations that depend on the angle (see e.g. equations (D1‐2) or (D1‐6)).
The following section shows the origin of these equations, to verify the equations given in
EC 5 and improve understanding of their mechanical background. The basis is the as‐
sumption of a linear interaction criterion (see also Figure D1‐7).
Fc
Fperp
Fparallel
Fv Fv
D Fv Fv
Fperp
Fparallel
Fc
Figure A2 Multiaxial stresses in a specimen compressed at an angle to the grain.
Derivation Hankinson equation
In an initial step, only normal stresses are taken into consideration: Fv = = 0.
Ft,a Ft,a
h
D
b
D
p q h
Ft,a
Ft,0 Ft,90
Figure A3 A piece of wood tensioned at an angle to the grain.
613
Annex 2: Stress interactions
Trigonometric relationships (Figure A3):
h h
sin p (a)
p sin
h h
cos q (b)
q cos
F t,90 F t,90
sin F t,α (c)
F t,α sin
F t,0 F t,0
cos F t,α (d)
F t,α cos
From forces to stresses:
F t,0 (c)
(iii)
F t,α b h cos
(d)
t,α
bh F t,90 (iv)
b h sin
Failure criterion ‒ assumption of a linear interaction without shear stresses:
t,0 t,90
1 (I)
f t,0 f t,90
Inserting:
t,0
(i) in (iii): t,α t,0 t,α cos 2 (II)
cos
2
t,90
(ii) in (iv): t,α t,90 t,α sin 2 (III)
sin
2
(II) + (III) in (I):
614
Annex 2: Stress interactions
The equation (2) thus obtained is now solved for ft,:
Equation (3) is the so‐called Hankinson equation. The derivation applies to compressive
or tensile stress; namely the index “t” for tension can be replaced with the index “c” for
compression. Accordingly, Equation (D1‐1) corresponds to the central expression in
equation (3), while the last expression is in EC 5 in the interaction Equation (D1‐6), where
the coefficient kc,90 is considered additionally.
Extension with shear
The following expression is added (for definitions see Figure A3):
t,0 t,90
1
f t,0 f t,90 fv
Inserting (I), (II) and (IV)
and solving (where t, = ft,):
f t,0
f t,α (4)
f t,0 f t,0
cos 2
sin
2
cos sin
f t,90 fv
Equation (4) is used in Equation (D1‐2) to verify tension at an angle to the grain.
615
Annex 2: Stress interactions
Quadratic interaction
If a quadratic rather than linear interaction is now assumed, equation (4) changes as
follows:
2 2 2
t,0 t,90
1 (quadratic interaction)
f t,0 f t,90 fv
2
f t,α cos 4 2
f t,α sin 4 2
f t,α cos 2 sin 2
2
2
1 (inserting (I), (II), (IV))
f t,0 f t,90 f v2
Solving for ft,:
f t,0
f t,α (5)
2 2
f f t,0
cos t,0 sin cos sin
4 4 2 2
f f
t,90 v
Equation (5) is reflected in the verification of a step joint, equation (E9‐3) (with compres‐
sive rather than tensile strengths). The difference between both equations exists in the
form of factor 2, with which compression strength perpendicular to the grain and shear
strength in equation (E9‐3) is increased. This factor 2 is empirical in nature. Tests con‐
ducted with step joints show that the quadratic interaction criterion in accordance with
equation (5) with doubled compression strength perpendicular to the grain and doubled
shear strength correlates better with the test results.
Case of a tapered edge
In Figure A3, the stress used for derivations was parallel to the edge. In the case of ta‐
pered edges, however, the stress is parallel or perpendicular to the grain, as in Figure A4.
D 1 ℓ = 1/tanD ℓ
W Vm
D 1
D
W W Vm
V W
Figure A4 Tapered edge with stresses parallel and perpendicular to the grain. Left: Internal forces and
moments at the considered infinitesimal element. Right: Shear stresses are at right angles to
each other and of equivalent size, which means that due to equilibrium conditions, tension
perpendicular to the grain t,90 is developing.
616
Annex 2: Stress interactions
The equilibrium of forces in both horizontal and vertical directions results in:
H 0 m 1 m m tan (x)
tan
t,90
V 0 1 t,90 t,90 tan (y)
tan
Inserting (x) in (y):
Inserting in a quadratic interaction with shear:
2 2 2 2 2
m t,90 m m tan 2 m tan 2
1
fm f t,90 fv fm f t,90 fv
And solving for t,0:
1 tan 4 tan 2
m2
2 1
f2 f t,90 f v2
m
1 1
m
1 tan tan 1
4 2
2 f m2 f m2
2
1 tan 4
tan 2
2
fm f t,90 fv
2
fm 2
f t,90 fv 2
Additional transformations result in:
fm
m (6)
2 2
f f
1 m tan 4 m tan 2
f f
t,90 v
Equation (6) is used to verify beams with a tapered edge, see equation (4‐5) with equa‐
tions (D4‐6) and (D4‐7). There too, additional factors are used to minimise or increase
shear strength. For tensile stresses along the tapered edge, the factor is 0.75. Therefore,
only 75% of the shear strength is considered, since the shear strength declines with sim‐
ultaneously applied tension perpendicular to the grain (Equation (D4‐6)). For simultane‐
ously applied compression perpendicular to the grain, the opposite applies and shear
strength increases. This is reflected in a factor of 1.5 for the shear strength, which is
specified in the case of compressive stresses along the tapered edge (Equation (D4‐7)).
617
Annex 3: Elastic buckling load
and buckling lengths
Derivation of elastic buckling load for Euler case II
The elastic buckling loads for the four Euler cases shown in Figure D2‐7 can be analytical‐
ly derived. This is shown as an example for the Euler case II, which is shown in Figure A5;
a slender column hinged at both ends.
F N FN u
crit
w(x)
x
Figure A5 Column hinged at both ends and of length L, E∙I = const. (STEP 1995 Article B6)
For the column shown in Figure A5 both the following equations apply:
M( x ) F w( x ) (a)
M
w ( x ) (differential equation of deflection curve) (b)
E I
In this case, it is assumed that the material concerned is homogeneous and elastic and
the load is applied at the centre of gravity. Now, (a) is inserted in (b):
F w( x ) F
w ( x ) w ( x ) w( x ) 0
E I E I
Substitution with 2 = F/(E∙I) and the base function for the differential equation:
w( x ) A cos( x ) B sin( x )
619
Annex 3: Elastic buckling load and buckling lengths
With the boundary conditions w(x = 0) = w(x = L) = 0 it follows that:
w( x 0) A 1 0 0 A 0
w( x L) 0 B sin( L) 0 sin( L) 0 L n mit n 1,2,3,...
F 2
E I 2
2 2 F F crit (7)
E I L L
2
Equation (7) corresponds to equation (D2‐19).
Buckling lengths of some structural systems
Connected columns
If columns hinged at both ends are braced by a column with a clamped support, the
critical buckling load of the clamped column is reduced by the normal forces Ni in the
hinged columns, since they generate horizontal forces in the deformed system. Consider‐
ing the effect of rotation in the semi‐rigid joint at the column base, the effective length
factor for buckling in the system plane (see Figure A6) is approximately:
2 E I
4 1
r K r
In this case, is specified in Figure A6. For the columns hinged at both ends, meanwhile,
a buckling verification must be performed with a buckling length corresponding to their
real length.
Nr N1 N2 Ni
li
l2
l1
lr
EI
Kr lr Ni
α= ∑
Nr li
Figure A6 Connected columns. (STEP 1995 Article B7)
620
Annex 3: Elastic buckling load and buckling lengths
Arches
For three‐ and two‐hinged arches (see Figure A7) with ratios h/ℓ between 0.15 and 0.5
and an essentially constant cross‐section, the effective length for buckling in the arch
plane may be assumed at:
ef s 1.25 s
whereby s is equivalent to half the arch length. For the buckling verification, the normal
force in the quarter point should be used.
h
l
Figure A7 Two‐hinged arch. (STEP 1995 Article B7)
Two‐ and three‐hinged frames
For two‐ and three‐hinged frames with normal forces NS and NR in columns (NS) or rafters
(NR) and angles of inclination of S ≤ 15° (see Figure A8), the following approximate equa‐
tion can be applied for the buckling length of columns:
2 E I S 1 s E I S NR s 2
ef S h where S 4
h K φ 3E IR E I N h 2
R S
The corresponding buckling length of the rafter (if R ≤ 20°) is:
E IR N S h
ef R s where R S
E I S NR s
If the second moments of area change (tapered rafters or columns), the cross‐sections at
0.65 ∙ s or 0.65 ∙ h can be inserted in the above equations (for geometric definitions, see
Figure A8).
Figure A8 Three‐hinged frame.
621
Annex 3: Elastic buckling load and buckling lengths
Columns or rafters with knee bracing
The effective lengths of the columns shown in Figure A9 on the left and of the rafters
shown in Figure A9 on the right for buckling in the frame plane can be estimated as:
ef 2 s l 0.7 s o
sl
so
so
so
sl
sl
Figure A9 Frame with truss as rafter (left) and three‐hinged frame with V‐shaped columns (right).
(STEP 1995 Article B7)
Torsional buckling of spatial frames
For rotationally symmetrical structures (see Figure A10), two types of buckling are to be
examined in principle. As well as buckling within the plane of the half‐frame, rotational
buckling of the spatial structure represents one possible type of stability failure. The
latter is characterised by a rotation of the compression ring about the vertical axis of
symmetry. For 1 < < 2 and a/s < 0.2, the following approximate solution applies for the
effective length factor for the rafter:
2a 3 2 a E I
1
s
4 s 2 1 a K r
s
Here, E∙I is the bending stiffness of the rafter about the vertical axis and Kr the rotational
stiffness of the connection between the compression ring and rafter, equally for bending
about the vertical axis. Similar to the procedure for two‐ and three‐hinged frames with
variable second moments of area, the cross‐sections at 0.65 ∙ s can be inserted if the
rafters are tapered.
a
Figure A10 Rotationally symmetrical spatial frame. (STEP 1995 Article B7)
622
Annex 4: Derivations lateral
torsional buckling
Critical moment Mcrit
The equation used in EC 5 for the critical moment can be analytically derived. This is
shown in the following section, whereby the following assumptions are important for
understanding:
The beam is subject to a constant moment and elastic (not plastic).
The beam has a constant rectangular cross‐section and constant stiffness.
The beam cannot rotate at the supports (torsional restraints).
The load application is in the centre of gravity of the beam cross‐section.
The beam cross‐section does not warp and any twist is minimal.
Equations Euler‐Bernoulli beam theory:
dQ dM M M
q , Q, , w , w
dx dx E I E I
M0 M0
x Mx y My
z z
ℓ
Mz
x
dy
y
[
dz
dx
z dy
I M]
dy
K
M0 dx=w'y(x)
x
I M[
MK
Figure A11 Definitions of laterally deviated beam with rectangular cross‐section and constant moment M0.
623
Annex 4: Derivations lateral torsional buckling
From Figure A11, the following relations can be derived, assuming a small angle of twist :
M
M y : wz ( x ) (a)
E I y
M
M z : wy ( x )
E I z
(b)
M
M x (no warping): ( x ) (c)
G I T
M
sin M M 0
M0
(d)
M
cos 1 M M 0
M0
(e)
M w y ( x ) M 0
(f)
insert (d), (e), (f) in (a), (b), (c):
M0
wz ( x )
E I y
(i)
M0
wy ( x )
E I z
(ii)
M0
( x ) w y ( x )
G I T
(iii)
Differentiate equation (iii) with respect to x:
M0
( x ) w y ( x )
G I T
and insert in equation (ii):
G I T M M 02
( x ) 0 ( x ) ( x ) (x) 0
M0 E I z E I z G I T
624
Annex 4: Derivations lateral torsional buckling
Now, solve the differential equations:
Initial value problem: (x = 0, x = ℓ) = 0 (→ torsional restraints and no twist)
2
M0
Be: = A ∙ sin( ∙ x) + B ∙ cos( ∙ x) where 2 (iv)
E I z G I T
(0) = A ∙ 1 + 0 =0 → A = 0
(ℓ) = B ∙ cos( ∙ ℓ) =0 ↔ cos( ∙ ℓ) =0 ↔ ∙ ℓ = n ∙ ↔ = (n ∙ )/ℓ
Key criterion (smallest critical moment): n = 1 → = /ℓ
Insert in (iv):
M0
2
2
2 2
E I z G I T
Solve for M0 = Mcrit:
2 E I z G I T
M 02 M crit E I z G I T (8)
2
Equation (8) corresponds precisely to the equation specified in EC 5 for Mcrit, see also
Equation (D2‐21), which therefore only applies for systems under the requirements ini‐
tially specified.
Textbooks such as Timoshenko (S.P. Timoshenko (1961). Theory of elastic stability.
McGraw‐Hill, New York.) provide examples of critical moments for other systems.
Derivation of the equation for kcrit
With lateral torsional buckling, other than bending about the major principal axis, bend‐
ing about the minor principal axis and torsion also take place. The normal stresses can be
taken into consideration here with a linear interaction:
My Mz y
z 1
W y f m,y W z f m,z f m,y f m,z
Transformation and definition kcrit:
y z Mz
1 1 k crit (9)
f m,y f m,z W z f m,z
625
Annex 4: Derivations lateral torsional buckling
Design format:
y My
1 f m,y M y k crit W y f m,y (10)
k crit f m,y k crit W y
Using equations (9) and (10), coefficient kcrit can now be derived.
The base function in this case reads as follows:
2
My
E I z M crit
2
Mz e (11)
2 My
2
1
M crit
The initial factor corresponds to the Euler buckling load about the z‐axis (see equation
(7)), the second factor is the so‐called Dischinger factor. Mcrit corresponds to the critical
moment from equation (8) and e is the initial deformation about the minor axis of the
system. The background to equation (11) lies in second order theory, its derivation is
complex and will not be explained here in further detail5.
Equations (11) and (10) are now inserted in equation (9):
2
k crit W y f m,y
Mz E I z
2
M crit e
k crit 1 1
W z f m,z 2
k crit W y f m,y W z f m,z
2
1
M crit
Now, crit = Mcrit/Wy is inserted:
2 2
2 E I z k crit f m,y
e
2 crit
2
k crit 1
W z f m,z 2
k crit f m,y
2
1
crit
2
5
Any readers who are interested can study the derivation in textbooks. A detailed derivation is also included in
“O.P. Hörsting (2008). Zum Tragverhalten druck‐ und biegebeanspruchter Holzbauteile. Dissertation Universi‐
tät Braunschweig (in German)“; equation (2.158) given there corresponds to equation (11) if initial rotation of
the beam axis is disregarded.
626
Annex 4: Derivations lateral torsional buckling
2 E I z
e 2
2
k crit rel
4
k crit 1
2
W z f m,z 1 k crit rel
4
Now, multiply and substitute the expression in brackets for a better overview:
e 2 E b
k crit rel k crit rel 1 k crit 1 k crit rel a k crit rel k crit 1 0
3 4 2 4 3 4 2 4
2
2
f
m,z
This equation is solved for rel to be able to draw it. In EC 5, the resulting curve was sim‐
plified, see Figure A12 (for GL24h with fm,z = 24 MPa, E0,g,05 = 9400 MPa, e/ℓ = 1/500 and
b = ℓ /40):
k crit 1
rel 4 3 2
(12)
k crit a k crit
1
0.8 EC 5
0.6
kcrit
0.4 Equation (12)
0.2
0
0 0.5 1 1.5 2 2.5 3
rel
Figure A12 Coefficient kcrit depending on the relative slenderness rel. See also Figure D2‐10.
627
Annex 5: Mechanically
jointed beams
Derivation ‐method
The derivation of the differential equations shown here is performed exemplarily for a T‐
cross‐section comprising two components, see Figure A13. The following prerequisites
apply: For each cross‐section, Euler‐Bernoulli beam theory applies and shear defor‐
mations are disregarded. The connection is assumed to be spread out along the axis of
the beam, over which the cross‐sections and joint stiffnesses are constant.
b1
p
E1,A1,I1
h1
k h2
z
b2 E2,A2,I2
x
V1 V1+dV1
N1 A N1+dN1
M1
v M1+dM1
M2+dM2
N2 A
M2
N2+dN2
V2 V2+dV2
dx
Figure A13 Mechanically jointed cross‐section comprising two components: Dimensions, forces and
moments. A‘ = dA/dx. (STEP 1995 Article B11)
Deformation relationships (see Figure A14):
h h
u u 2 u 1 w 1 2 u 2 u 1 w a
2 2
u1, u2 Longitudinal displacements of cross‐sections 1 and 2
w Common deflection
u Relative displacement of the individual cross‐sections at
the position of the fasteners
629
Annex 5: Mechanically jointed beams
Here, we see that u is independent of the fastener position, while the distance of the
axes a of the cross‐sections is decisive instead. The derived equations thus apply not only
to cross‐sections arranged vertically, like the T‐cross‐section here, but also for those
arranged adjacent to each other. However, this only applies if we disregard shear defor‐
mations of the individual cross‐sections.
Elasticity relations according to beam theory (' = differentiation with regard to x):
N 1 E 1 A 1 u1 N 2 E 2 A 2 u2
M 1 E 1 I 1 w M 2 E 2 I 2 w
V 1 E 1 I 1 w V 2 E 2 I 2 w
Shear flow in joint v k u k u 2 u 1 w a
σ1 σm1
u = u2-u1+
=w'(h1/2+h2/2)
w' M1
u1 v1 0,25v v1-0,25v
N1 h1/2
v v (2v2-v)2
a 4v2-3v
w' M2 h2/2
u2 a2 v2
N2 h 0,75v v2-0,75v
σ σm2 σ2 vv v Σv
Figure A14 Cross‐section from Figure A13: Deformations, longitudinal stresses, shear flow.
(STEP 1995 Article B11)
Equilibrium of forces in x‐direction (no external load p in x‐direction), see Figure A13
(N' = dN/dx):
dN 1
N1 N1 v dx N 1 dx v dx N 1 v 0 (a)
dx
dN 2
N2 N2 v dx N 2 v 0 (b)
dx
in z‐direction (p = external load in z‐direction), Figure A13 (V' = dV/dx):
dV dV
V 1 V 1 1 V 2 V 2 2 p dx 0
dx dx
dV dV
1 2 p dx V 1 V 2 p 0 (i)
dx dx
V 1 V 2 V p
630
Annex 5: Mechanically jointed beams
Equilibrium of moments at point A, Figure A13 (M' = dM/dx):
A h1 dM 1
M 0 M 1 V 1 dx v dx M1
2 dx
h1
M1 dx V 1 dx v dx (ii)
2
h1
M1 V 1 v
2
h2
M2 V 2 v (iii)
2
Now, equations (ii) and (iii) are added, differentiated once with respect to x and
V‘1 + V‘2 = V‘ is expressed with –p (equation (i)), where h1/2 + h2/2 = a (Figure A14):
M1 M2 v a p 0 (c)
Accordingly, the three equilibrium conditions (a), (b) and (c) for the three deformations
u1, u2 and w are formulated. If the internal forces and moments are replaced by the elas‐
ticity relations, we obtain a system of differential equations for the x‐direction:
E 1 A 1 u1 k u 2 u 1 w a 0
(13)
E 2 A 2 u2 k u 2 u 1 w a 0
and for the z‐direction:
The variation in elastic energy
1
E 1 A 1 u12 E 2 A 2 u22 E 1 I 1 E 2 I 2 w 2 k u 2 u 1 w a 2 p w dx
2
2
also elicits these equations.
In equation (14), the equilibrium in z‐direction, elastic foundation kw or the influence of
second order theory can also be taken into consideration, whereby the expression
kw ∙ w ‐ N0 ∙ w" is added to the left side.
The system of differential equations is particularly easy to solve for a single‐span beam
with sinusoidal load distribution (L = beam length):
p p 0 sin x
L
631
Annex 5: Mechanically jointed beams
Base function for the deformation:
u 1 u 10 cos x
L
u 2 u 20 cos x
L
w w 0 sin x
L
Inserted in the system of differential equations (equations (13) and (14)) and cancelling
out sine or cosine, we obtain:
u 10 u 20 w0 p0
2
2
E 1 A1 k k k a 0
L L
2
k 2
E 2 A2 k k a 0
L L
4
2
k a k a 4
E 1 I 1 E 2 I 2 k 2
a 2 1
L L L L
The solving of the system of equations delivers:
L4 1 L4 1
w 0 p0 p0
4
E 1 A1 1 a 2
4
E I ef
E 1 I1 E 2 I 2
E A
1 1 1 1
E 2 A2
a 1 E 2 A2
u 10 w 0
L 1 E 1 A1 E 2 A 2
a 1 E 1 A1
u 20 w 0
L 1 E 1 A1 E 2 A 2
where:
1
1 (15)
E A 2
1 2 1 1
L k
Equation (15) is coefficient .
632
Annex 5: Mechanically jointed beams
Normal stresses
As was explained in Article D7, the effective bending stiffness (E ∙ I)ef and hence stress
distribution is largely dependent on joint stiffness, which is recorded by the coefficient .
For (E ∙ I)ef and where n = number of individual cross‐sections, the following applies
(equation (D7‐1)):
E I ef E i I i i E i A i a i2
n
(16)
i=1
In the following, the determination of stresses using the three cross‐sections A, B and C
shown in Figure A15 is explained.
A VA
B VB
C VC
hi hi
h hi hi
hi hi
q
uA
uB
uC
Figure A15 Deflection and bending stress distribution of a full cross‐section (A), a cross‐section comprising
three mechanically jointed individual cross‐sections (B) and a cross‐section comprising three
non‐connected individual cross‐sections (C). Corresponding to Figure D7‐1.
Full cross‐section A
Now we consider beam A as a rigidly connected beam with three components and with
E1 = E2 = E3 = const. and hi =h1 = h2 = h3 = heights of individual cross‐sections. The entire
external moment M is then distributed over the three individual cross‐sections as shown
in Figure A16.
A VN VM VA
N1 M1
a1
h M1
a1
N1 M1
Figure A16 Beam A, comprising three rigidly connected individual cross‐sections.
633
Annex 5: Mechanically jointed beams
The external moment M must be in equilibrium with the internal moments and forces Mi
and Ni of the individual cross‐sections i:
M M i N i a i
This means for the beam shown in Figure A16:
1 M 3M1
M 3M1 2N1 a1 or N 1 M 3 M 1 1 (17)
2a1 2a1 M
The following applies to the stress A at the edge of the beam:
M 1 h1 N 1
A (18)
I 1 2 A1
For the considered „full cross‐section“, the following also applies:
I1
3
M1 M and I I i A i a i2 3 I 1 2 A 1 a 12
I i 1
Insert in equation (17) and considering the Steiner part (I ‐ 3 ∙ I1):
M 3I1 M M M
N1 1 I 3 I 1 2 A 1 a 12 A 1 a 1
2a1 I 2a1 I 2a1 I I
The stress component from the normal force N1 is thus as follows:
N1 M
1 a1
A1 I
Accordingly, the overall stress A at the beam edge from equation (18) evolves into the
bending stress at the edge of a full cross‐section of height htot = 3 ∙ h1:
M 1 h1 M M I h M M h M h
A a 1 1 1 a 1 1 a 1 tot
I1 2 I I1 I 2 I I 2 I 2
634
Annex 5: Mechanically jointed beams
Cross‐section C comprising three non‐connected individual cross‐sections
C VC
hi
hi
hi
For beam C, the normal force is Ni = 0 and the bending stress C at the edge of the beam
is thus
M i hi E I i
C where Mi M and I ges I i
Ii 2 E I ges
Mechanically jointed cross‐section B comprising three individual cross‐sections
B VB
hi
hi
hi
To simplify matters, we define Ei = E1 = E2 = E3 = const. and 1 = 3. The derivation of the
normal force is complex for the case in question, since the differential equations for
deformations considering shear deformations in the joints need to be solved. This was
done by Möhler (1956) and a reduction coefficient was defined, as explained in Article
D7. In the case of semi‐rigid joints, equation (17) can be transformed similarly to beam A,
using I ef 3 I 1 2 1 A 1 a 1 (Ei = const.) and M 1 M I 1 I ef :
2
M 3I1 M M M
I ef 3 I 1 2 1 A1 a 1 1 A1 a 1
2
N1 1
2a1 I ef 2 a 1 I ef 2 a 1 I ef I ef
The stress component from the normal force N1 is thus as follows:
N1 M
1 1 a1
A 1 I ef
The overall stress B at the beam edge can then be calculated as:
M 1 h1 M M I h M M h
B 1 a1 1 1 1 a1 1 1 a1 (19)
I 1 2 I ef I 1 I ef 2 I ef I ef 2
635
Annex 5: Mechanically jointed beams
Equation (19) corresponds to equations (D7‐6) and (D7‐7) or the bending stress at the
edge of the overall beam, whereby in the initial summand of equation (19), the bending
stress at the edge of cross‐section 1 is calculated and in the second summand, the nor‐
mal stress in the centre of gravity of cross‐section 1. While the effect of the semi‐rigid
joints (and the potential differing moduli of elasticity) on distribution of the internal forc‐
es and moments is considered via the effective stiffness when calculating the moment
component i,m = M/(E ∙ I)ef ∙ Ei ∙ hi/2, the normal force component depends directly on
how the Steiner part contributes and hence the coefficient .
The normal forces Ni generated by an external moment in composed beams with rigid
joints exceed those for beams with semi‐rigid joints. For non‐connected individual cross‐
sections, the external moment does not generate any normal force, Ni = 0. Consequently,
the normal stress (tension t or compression c) is determined as in equations (19) and
(D7‐7) by considering i:
M
i,t(c) E a
E I ef i i i
One reminder here is that all internal forces and moments transferred via a semi‐rigid
joint must be reduced by the coefficient . This also applies to the shear stresses max,
since these are determined via the joint between cross‐sectional parts 2 and 3, which is
why 3 must be applied.
636
Annex 5: Mechanically jointed beams
Shear stresses and loads on fasteners
b1
A1, E1, I1
h1
a1 Neutral axis
S
s1, K1, F1 y a2
z h2
a3 h = a2 + 2
h2
x b2
A2, E2, I2
s3, K3, F3 h3
A3, E3, I3
b3
For the cross‐section shown comprising three components, the maximum shear stress
max shall be determined. The shear stress peaks at the neutral axis at z = 0. To determine
max, the first moment of area Sef of the lower beam part for z = [0, h + h3] is needed,
whereby the semi‐rigidly connected cross‐section 3 is to be reduced with 3 as explained
above (zsi = distances from the centres of gravity of the individual cross‐sections to the
neutral axis):
1 1
S ef z s2 A 2 z s3 A 3 h b 2 h 3 a 3 b 3 h 3 b 2 h 2 3 a 3 A 3
2 2
Sef can also be determined from the upper beam part for z = [0, ‐(a1 + h1/2)] and with 1.
From Sef, we obtain the following maximum shear stress, taking differing moduli of elas‐
ticity into consideration:
max
V max E S ef
V max 3 E 3 A 3 a 3 0,5 E 2 b 2 h
2
E I ef b 2 E I ef b 2
This corresponds to equation (D7‐8).
With max, it is now possible to calculate the force exerted on the individual fasteners. The
shear flow is:
V max 3 E 3 A 3 a 3 s 3
F3
E I ef
637
Annex 6: Derivation
Johansen equations
The following section derives certain Johansen cases for single‐shear timber‐to‐timber
joints.
Failure mechanism b – embedment failure in member 2:
Fv,Rk
t1 t2
f h,2,k
Fv,Rk
Figure A17 Failure mechanism b.
In the event of failure in accordance with mode b, as in Figure A17, the embedment
strength in member 2 is decisive. The equilibrium of forces V = 0 results in:
Equation (20) precisely corresponds to the equation for failure mechanism b in EC 5.
639
Annex 6: Derivation Johansen equations
Failure mechanism c – embedment failure in both members:
Fv,Rk
f h,1,k
a2 a2
f h,1,k
b2
t2
a1a1
b1
d
t1 f h,2,k
f h,2,k
Fv,Rk
Figure A18 Failure mechanism c.
In the event of failure in accordance with mode c, as in Figure A18, the embedment
strength has been reached in both members. The following applies:
f h,2,k
f h,1,k
The equilibrium of forces V = 0 results in:
F v,Rk f h,1,k d b 1 f h,2,k d b 2 f h,1,k d b 2
Consequently:
b 2 b1
The bending moment in the shear plane results in:
b1 a 1 2 a b2 2
Member 1 : M SP f h,1,k d f h,1,k d a 1 b 1 a 1 1 f h,1,k d 1 a 1
2 2 2
b2 a 2 2 a 2 b 22
Member 2 : M SP f h,2,k d f h,2,k d a 2 b 2 a 2 2 f
h,2,k d a2
2 2 2
Equating and replacing fh,2,k with ∙ fh,1,k:
b2 2 b2 2 b2
f h,1,k d a 2 2
2
f h,1,k d 1 a 1 f h,2,k d a 2 2
2 2 2
640
Annex 6: Derivation Johansen equations
Now replace b2 with b1/:
b2 2 b2
f h,1,k d 1 a 1 f h,1,k d a 2 2
2
2 2
b1
2 2 b
2
a1 a2 1 2
2
2 2
b1 1
2
a2 a1
2 2
2
Replacing a1 = (t1‐b1)/2 and a2 = (t2‐b2)/2 = ( ∙ t2‐b1)/(2 ∙ ):
b 12 1 t 2 b 1 t 1 b 1
2 2
2 4 4
1 b 12 2 t 1 t 2 b 1 t 12 t 22 0
The following is revealed when solving the quadratic equation:
t
t
2 2
t1 t t
b1 2 2 1 2 2 3 2 1 2 (21)
1 t1 t1 t1 t1
In addition, the following applies for the equilibrium of forces V = 0 at member 1:
F v,Rk f h,1,k d b 1
Inserting equation (21):
t
t
2 2
f h,1,k d t 1 3 t t
2 1 2 2 2 1 2
2
F v,Rk (22)
1 t1 t1 t
1 t1
Equation (22) is included in the equation for failure mechanism c in EC 5, where equation
(22) corresponds to the Johansen portion, while the rope effect is missing.
641
Annex 6: Derivation Johansen equations
Failure mechanism e – embedment failure in both members and
one plastic hinge in member 1:
Fv,Rk
t2
t2
b1 b2 b2
a2 a2 a2 a2
f h,1,k
My,k f h,2,k
A
My,k
t1 f h,1,k
d
f h,2,k b1
f h,2,k
f h,2,k
Fv,Rk
Figure A19 Failure mechanism e. Right: free‐body diagram of fastener.
In the event of failure in accordance with mode e, as in Figure A19, both members reach
the embedment strength and a plastic hinge forms. The following applies:
f h,2,k
f h,1,k
The shear force in the fastener is equal to zero at the point where the bending moment
peaks:
f h,2,k d b 2 f h,1,k d b 1 f h,1,k d b 2 f h,1,k d b 1 0
Consequently:
b 2 b1
The equilibrium of moments at the fastener MA = 0 is:
b1 b a2 a
M y,k f h,1,k d b 1 f h,2,k d b 2 a 2 b 1 2 f h,2,k d a 2 b 1 b 2 a 2 2
2 2 2
Now fh,2,k is replaced with ∙ fh,1,k, b1 with ∙ b2 and a2 = (t2‐b2)/2 is replaced and the
equation is transformed:
2
2 2t 2 t2 4 M y,k
b2 b2 0
2 1 2 1 2 1 f h,1,k d
642
Annex 6: Derivation Johansen equations
Solving the quadratic equation reveals:
t2
2
t2
2
t2 4 M y,k
b2 (23)
2 1 2 1 2
2 1 f h,1,k d 2 1
In addition, the following applies for the equilibrium of forces V = 0 at member 2:
F v,Rk f h,2,k d b 2 f h,1,k d b 2
Inserting equation (23):
f h,1,k d t 2 4 2 1 M y,k
2 1
2
F v,Rk (24)
2 1 2
f h,1,k d t 2
Equation (24) is included in the equation for failure mechanism e in EC 5, where equation
(24) corresponds to the Johansen portion, while the rope effect and the pre‐factor 1.05
are missing (the pre‐factor takes into consideration the differing partial safety factors and
kmod for steel and timber).
Failure mechanism f – embedment failure in both members and two plastic hinges:
Fv,Rk
t1
b1 b2
f h,1,k
My,k f h,2,k
My,k
My,k
My,k
f h,1,k A
b1
f h,2,k
b2
t2
Fv,Rk
Figure A20 Failure mechanism f. Right: free‐body diagram of fastener.
643
Annex 6: Derivation Johansen equations
In the event of failure in accordance with mode e, as in Figure A20, both members reach
the embedment strength and two plastic hinges form. The following applies:
f h,2,k
f h,1,k
and
b 2 b1
The equilibrium of moments at the fastener MA = 0 is:
b b
M y,k M y,k f h,1,k d b 1 b 2 1 f h,2,k d b 2 2
2 2
Replacing b2 with b1/ and fh,2,k = ∙ fh,1,k and solving for b1:
2M 2
b1
f h,1,k d 1
The equilibrium of forces V = 0 at member 1 results in:
F v,Rk f h,1,k d b 1
Inserting b1:
2
F v,Rk 2 M f h,1,k d (25)
1
Equation (25) is included in the equation for failure mechanism f in EC 5, again without
rope effect and pre‐factor 1.15.
644
This comprehensive book provides in-depth
knowledge and understanding of design
rules according to Eurocode 5. It is based on
the first edition of the STEP series (Structural
Timber Education Programme), which was pre-
pared in 1995 by approximately 50 authors
from 14 European countries. Since its release,
knowledge has advanced significantly in areas
such as construction materials, structural ele-
ments and joints. The present work updates
and extends the STEP compilation and is aimed
at students, structural engineers and other
timber structure professionals.
ISBN 978-3-7315-0673-7
9 783731 506737