Você está na página 1de 92

COST Action E34

Bonding of Timber
Working Group 1: Bonding on site

Core document

Edited by
Klaus Richter and Helena Cruz
First Edition March 2008

2008

COST E34 - WG1: Bonding on Site

THE AUTHORS
(in alphabetical order)
Broughton, James
Joining Technology Research Centre
School of Technology
Oxford Brookes University
Wheatley Campus
Gipsy Lane
Oxford OX3 0BP

jgbroughton@brookes.ac.uk

Brunner, Maurice
University of Applied Sciences
Solothurnstrasse 102
CH-2500 Biel 6

maurice.brunner@bfh.ch

Cruz, Helena
Laboratrio Nacional de Engenharia Civil
Av. Brasil, 101
P-1700-066 Lisboa
Custdio, Joo
Laboratrio Nacional de Engenharia Civil
Av. Brasil, 101
P-1700-066 Lisboa

helenacruz@lnec.pt

jcustodio@lnec.pt

Lavisci, Paolo
Legnopi srl
Via Borgo Valsugana, 11
I-59100 Prato
Lehmann, Martin
University of Applied Sciences
Solothurnstrasse 102
CH-2500 Biel 6

paolo.lavisci@legnopiu.biz

martin.lehmann@bfh.ch

Negro Joo
Dept. Civil Engineering U.C.
Polo II - Pinhal de Marrocos
P-3030-290 Coimbra

jhnegrao@dec.uc.pt

STAP
Rua Marqus de Fronteira, 8, 3 Dto.
P-1070-296 Lisboa

raquelpaula@stap.pt

Paula, Raquel

Pizzo, Benedetto
CNR-IVALSA
via Madonna del Piano
I-50019 Sesto Fiorentino (FI)

b.pizzo@ivalsa.cnr.it

Rautenstrauch, Karl
Bauhaus-University of Weimar
Faculty of Civil Engineering
Marienstr. 13 A
D-99423 Weimar

karl.rautenstrauch@bauing.uni.weimar.de

Richter, Klaus
EMPA, Wood Laboratory
berlandstasse 129
CH-8600 Dbendorf
Schober, Kay-Uwe
Bauhaus-University of Weimar
Faculty of Civil Engineering
Marienstr. 13 A
D-99423 Weimar

klaus.richter@empa.ch

kay-uwe.schober@bauing.uni-weimar.de

COST E34 - WG1: Bonding on Site

Serrano, Erik
SP Technical Research Institute of Swesen
Building Technology and Mechanics
Wood Technology
SE-351 96 Vxj

erik.serrano@sp.se

Rotafix Ltd
Rotafix House
UK-SA9 1UR Abercraf Swansea

daverotafix@aol.com

EMPA, Wood Laboratory


berlandstrasse 129
CH-8600 Dbendorf

rene.steiger@empa.ch

Smedley, Dave

Steiger, Ren

Van Leemput, Marc


CTIB-TCHN
Alle Hof ter Vleest 3
B-1070 Brussels

marc.vanleemput@ctib-tchn.be

Acknowledgement
We express our gratitude to all authors who voluntarily and with no specific funding have contributed to
this core document within their research area.
Klaus Richter and Helena Cruz
Chairpersons of COST E34 WG1 - Bonding on site
August 2007

10

COST E34 - WG1: Bonding on Site

TABLE OF CONTENTS
1

INTRODUCTION

13

2
2.1
2.2
2.3
2.4
2.5

TIMBER CONCRETE COMPOSITES


15
Timber-concrete composites with mechanical fasteners.......................................................... 15
Design methods........................................................................................................................ 17
Timber-concrete-composites with an adhesive bond ............................................................... 18
Research needs........................................................................................................................ 19
References ............................................................................................................................... 19

3
3.1
3.2
3.3
3.4
3.5
3.6
3.7

TIMBER WITH PASSIVE REINFORCEMENT


22
Overview................................................................................................................................... 22
Literature survey ....................................................................................................................... 22
Design methods........................................................................................................................ 24
Economic considerations.......................................................................................................... 26
Steel reinforcement................................................................................................................... 27
Research needs........................................................................................................................ 27
References ............................................................................................................................... 28

4
PRE-STRESSING OF TIMBER
29
4.1
Overview................................................................................................................................... 29
4.2
Calculation Methods ................................................................................................................. 33
4.2.1 Calculation example of pre-stressed timber beam ................................................................... 34
4.3
Research needs........................................................................................................................ 35
4.4
References ............................................................................................................................... 36
5
GLUED-IN RODS
37
5.1
Overview................................................................................................................................... 37
5.2
Design methods........................................................................................................................ 37
5.2.1 Basic assumptions.................................................................................................................... 37
5.2.2 Mechanics Failure modes and design philosophy................................................................. 39
5.2.3 Design codes and code proposals ........................................................................................... 40
5.3
Typical application methods ..................................................................................................... 43
5.3.1 Manufacturing principles........................................................................................................... 43
5.3.2 Examples .................................................................................................................................. 43
5.4
Applicable standards ................................................................................................................ 44
5.5
Research needs........................................................................................................................ 45
5.6
References ............................................................................................................................... 45
6
6.1
6.2
6.3
6.3.1
6.3.2
6.4
6.5
6.6

46
Overview................................................................................................................................... 46
Design methods........................................................................................................................ 47
Methods of application.............................................................................................................. 50
Repair of decayed ends using adhesives and steel reinforcement .......................................... 50
Repair of decayed ends using epoxy polymer concrete and GFRP reinforcement
(bars and plates)....................................................................................................................... 51
Applicable standards ................................................................................................................ 53
Research needs........................................................................................................................ 54
References ............................................................................................................................... 55

ON SITE INTERVENTIONS ON DECAYED BEAM ENDS

7
REPAIR OF GLUED LAMINATED STRUCTURES
56
7.1
Description of glued laminated components............................................................................. 56
7.2
Failure types and repair options ............................................................................................... 56
7.2.1 Wood decay due to inadequate construction details................................................................ 56
7.2.2 Glue line delamination and fissures.......................................................................................... 59
7.3
Research needs........................................................................................................................ 61
7.4
References ............................................................................................................................... 61
8
FACTORS INFLUENCING BOND PERFORMANCE
62
8.1
Introduction ............................................................................................................................... 62
8.2
Environment.............................................................................................................................. 63
8.2.1 Moisture content ....................................................................................................................... 63
8.2.2 Temperature ............................................................................................................................. 64
11

COST E34 - WG1: Bonding on Site

8.3
8.3.1
8.3.2
8.3.3
8.3.4
8.4
8.4.1
8.4.2
8.4.3
8.5
8.6

Materials ................................................................................................................................... 66
Surface preparation .................................................................................................................. 66
Age of surface........................................................................................................................... 67
Influence of wood species ........................................................................................................ 68
Treated wood ............................................................................................................................ 68
Stress........................................................................................................................................ 69
Influence of material stiffness on stress ................................................................................... 69
Influence of joint geometry on stress ........................................................................................ 70
Joint selection for the assessment of bond performance ......................................................... 74
Research Needs ....................................................................................................................... 75
References................................................................................................................................ 76

9
DURABILITY OF HISTORIC STRUCTURES REPAIRED WITH ADHESIVES
80
9.1
Overview ................................................................................................................................... 80
9.2
Examination of past interventions............................................................................................. 80
9.2.1 Structural type evaluations...................................................................................................... 81
9.2.2 Evaluations regarding the durability of the interface................................................................. 81
9.3
Considerations based on the inspections................................................................................. 82
9.4
Summary and recommendations.............................................................................................. 83
9.5
Research needs........................................................................................................................ 84
9.6
References................................................................................................................................ 85
10
QUALITY CONTROL ON SITE
86
10.1
Introduction ............................................................................................................................... 86
10.2
Quality control of materials ....................................................................................................... 86
10.2.1 Solid timber splice..................................................................................................................... 87
10.2.2 Epoxy adhesives and grouts..................................................................................................... 87
10.2.3 Metallic rods and plates ............................................................................................................ 87
10.2.4 FRP rods and plates ................................................................................................................. 87
10.3
Quality control of tools and equipment ..................................................................................... 87
10.3.1 Timber cutting and drilling slots ................................................................................................ 88
10.3.2 Surface cleaning ....................................................................................................................... 88
10.3.3 Mixing and application .............................................................................................................. 88
10.3.4 Tool maintenance ..................................................................................................................... 88
10.4
Quality control on site ............................................................................................................... 88
10.4.1 Contract Preparation................................................................................................................. 88
10.4.2 Removal of decayed timber ...................................................................................................... 88
10.4.3 Drilling and slotting.................................................................................................................... 89
10.4.4 Cleaning of bonded surfaces .................................................................................................... 89
10.4.5 Mixing........................................................................................................................................ 89
10.4.6 Installation of secondary adherends ......................................................................................... 90
10.4.7 Manufacture of the solid timber splice (TRS) ........................................................................... 90
10.4.8 Quality control for generic repair systems ................................................................................ 90
10.4.9 Health and safety ...................................................................................................................... 92
10.5
Quality plan ............................................................................................................................... 92
10.5.1 Responsibilities and records..................................................................................................... 92
10.5.2 Reception of materials .............................................................................................................. 92
10.5.3 Inspections and tests ................................................................................................................ 92
10.6
Certification of operatives ......................................................................................................... 94
10.6.1 Introduction ............................................................................................................................... 94
10.6.2 Training ..................................................................................................................................... 95
10.6.3 Certification procedure.............................................................................................................. 95
10.6.4 Theoretical examination............................................................................................................ 95
10.6.5 Practical examination................................................................................................................ 95
10.6.6 Evaluation ................................................................................................................................. 95
10.6.7 Inspection and Testing.............................................................................................................. 95
10.7
References................................................................................................................................ 96
11

12

SUMMARY OF ACHIEVEMENTS

97

COST E34 - WG1: Bonding on Site

INTRODUCTION
Richter K., Cruz H., Negro J.

Despite the development of key innovations for timber engineering and timber industries worldwide due to
the growing use of adhesives and adhesive bonding techniques, the requirements for durable high quality
bonds are demanding and remain a significant challenge; apart from the adhesive and timber quality, the
application of the adhesives and subsequent control of the curing processes are of critical importance to
bond durability.
In the glulam and general timber construction industries, the overwhelming majority of all structural
bonding processes are performed off site under factory controlled conditions. Here the control of
significant climatic factors, which are necessary for adequate adhesive spreading and curing, can be
relatively easily maintained within acceptable limits. Similarly, surface processes like sanding, planning
and pressing are almost entirely automated.
O the contrary, bonding is necessary when large structural timber parts need to be assembled at the
building site where process variables and the environment are difficult to control. In many cases such
assemblies are formed with steel connectors, but for special situations adhesive bonds are required to
meet the aesthetic or technical demands of the designer.
Other applications of structural on site bonding are related to the rehabilitation of buildings, which is an
area of increasing economical and social importance to most European countries. A great number of
historic buildings are either of common timber frame construction or incorporate complex timber
structures. Both types require specific interventions, including additional reinforcement or repair due to
overloading, insect attack or decay due to fungal activity, or bond line delamination in bonded structural
elements. Similar on site bonded reinforcement and repair techniques have been applied for some
decades following procedures developed for the repair or up-grading of other structures - the adhesives
being used either on their own or in conjunction with steel plates, rods, fibre reinforced materials and
even concrete. Such techniques are versatile, require less time and are more cost effective than
traditional carpentry methods, and, most importantly, help to minimise any disturbance to the building and
to its occupants during the intervention.
Yet, despite the fact that many historical timber structures require urgent and proper maintenance and
repair works, it is often the case that neither decision makers nor building contractors have the necessary
knowledge to apply these techniques appropriately and, as a result, damaging interventions are often
made.
Other concerns include a lack of suitable methods, whereby sufficient reliability of the bonded connection
can be guaranteed. One general reason for this is that to date a long service life has not been fully proven
for synthetic adhesives; the oldest bonded joints are only approximately sixty years of age. Furthermore,
reliable and representative accelerated ageing tests do not yet exist. Suitable on site quality control test
methods required for acceptance of the bonding process and conditions both of which are especially
difficult to control on the building site, are also missing.
The variability of materials and the insufficient quality control level, typical of on site conditions, also
contribute to reduce the confidence in glued systems. However, it should be remarked that, from all the
possible materials involved (timber, steel, GFRPs, concrete and the adhesive itself), timber is the one
from which most of the variability should be expected.
Of course, not all adhesives are suitable for on site bonding. It is therefore necessary to be able to select
suitable adhesives as well as to improve existing adhesives, or, develop alternative adhesive
formulations, suitable for indoor (and even outdoor) jobs on site where, for example, the use of pressure
is generally not available and clean regular bond lines are difficult to achieve.
In order to develop and assess new or improved adhesives, suitable test methods are needed. Yet
existing test methods, used to evaluate bond line performance of timber bonded joints, were originally
developed for phenolic or aminoplastic based products. These have been repeatedly proven to be
inadequate in the assessment of the more favoured epoxy type adhesives or epoxy bonded products.
Moreover, the application of these existing EN or national test and performance standards for epoxy
bonded products are much too penalising, since they merely impose severe conditions that are not
verified on site. The need for development of European standards for the evaluation of bond durability
(namely under high service temperature or humidity) as well as the long-term performance of epoxy
Introduction

13

COST E34 - WG1: Bonding on Site

adhesives has already been identified by CEN, but to date no work has been undertaken. The lack of
standards and calculation methods in this field impedes an objective evaluation of the safety level and
reliability of a glued connection, causing engineers to avoid this type of solution. Thus extensive prenormative research and thorough consideration of these problems is still required.
The brittle nature of most adhesives available for on site application needs also to be addressed. This
issue makes difficult the accommodation of stresses caused by slight relative movements of the glued
interfaces or withstanding intense shear stress gradients over long lengths.
In conclusion, the rehabilitation of timber structures has increasing economical, environmental and social
importance, and timber Bonding on Site (BoS) has an important role to play in this area. However, there
are a number of issues currently hindering the wider exploitation of BoS across Europe. These have
been identified as the following:

There is a lack of well-structured and concise knowledge on bonded reinforcement or repair


techniques for timber

BoS process and conditions are difficult to control (e.g. bondline thickness, surface properties,
bondline stresses and environmental conditions)

Adhesives for BoS are not specifically developed for timber

Appropriate test methods and standards for BoS adhesives are lacking

Rapid on site assessment methods (control of mixture and penetration, viscosity) are missing.

This core document therefore aims to:

14

Present qualitative and quantitative knowledge on structural BoS techniques

Enable the effective and safe application of reinforcement, repair and assembly of on site
bonding techniques

Disseminate knowledge to industry, research society and practitioners.

Introduction

COST E34 - WG1: Bonding on Site

TIMBER CONCRETE COMPOSITES


Brunner M., Negro J., Rautenstrauch K., Schober K.U.

2.1

Timber-concrete composites with mechanical fasteners

Given the low cost of concrete and its mechanical complementarity with timber, its use in timber-based
composites was one of the first to be considered. Mechanical fasteners were systematically used as
connecting devices. This chapter points out some relevant studies on this subject and attempts to provide
an overview. Though the scope of this document is focused on adhesive bonding, it is nevertheless
important to discuss those studies concerning mechanical connection (i.e., with fasteners), because of
the insight they provide on aspects such as ductile behaviour and partially composite behaviour.
Many aspects of timber-concrete composites behaviour have been investigated by a number of authors in
the last decade. The following list is not exhaustive and is an attempt to provide an overview of what has
been done and what is still missing.
Ceccotti [1] has written a very illustrative paper describing the main issues concerning the analysis and
design of timber-concrete composite elements. Gutkowski et al [2, 3] have investigated the performance
of composite elements with the interlayer connection consisting of a notch and a shear key (Fig. 2.1).
While the shear stresses were transferred by compression on the notch sloped surface, the role of the
shear key was to prevent the uplift force which tended to separate timber from concrete. This system led
to a very effective composite behaviour, with reduced interlayer slippage, but no long-term or moisture
content change effects were considered.

Figure 2.1: Notch and shear key connection

Bathon and Graf [4] have proposed the use of a two-dimensional steel mesh as a continuous shear
connector (Figure 2.2). The mesh is inserted and glued into a slot sawn in the timber. Although used as a
system component only, the adhesive plays a fundamental role in that it is the secondary stress transfer
media between timber and concrete. Therefore, from a conceptual point of view, this solution should be
placed half-way between conventional shear connector and an adhesive bond. More recently Bathon [5]
enhanced the concept by introducing pre-stressing, prior to casting the concrete layer, achieved by an
upward cambering of the timber beam. The tests were successful. The system failure was caused by the
plastic yielding of the steel mesh, whereas the adhesive bonding, timber or concrete all withstood the
load.

Timber Concrete Composites

15

COST E34 - WG1: Bonding on Site

Figure 2.2: Continuous mesh shear-connector


Ballerini, Crocetti and Piazza [6, 7] tested the performance of the notch plus shear key system for
different notch depths and angles, sizes of timber shear area and position of the screw (Figure 2.3). In
their conclusions, they underline the need to develop a standard test set-up. Though the sentence
referred to mechanical timber-to-concrete connection, such document should desirably be suited for
testing both fastener and adhesive bonding of timber to concrete.
Empty notch
Hydraulic
jack

Steel Plate

Concrete

107 mm

h
Timber

Lt
120 mm

LVDT

Teflon
plate

Figure 2.3: Test set-up proposed in [6, 7]


Benitez [8] performed comparative studies between three types (Figure 2.4) of mechanical shear
connectors: smooth steel dowels at an angle of 60 to the force direction, a ring-type connector (CHS,
circular hollow section) inserted into a fitted slot in the timber and fixed to it by a central screw) and an Ishaped connector made up from a UC (universal column) steel hot rolled section, fixed by screws to the
timber.

Figure 2.4: Shear connectors studied by Benitez [8]


A specifically designed shear connector was tested and FE modelled by Bou Said et al. [9]. The
connector is flower-shaped, with holes in the petals to enhance the anchoring with concrete (Figure 2.5).
Inoue et al. [10] used either deformed bars or special rods (Figure 2.6) as shear connectors.
Mungwa, Jullien et al. [11, 12] developed a new type of shear connector specifically designed for fast
installation while still maintaining good stiffness, strength and ductility characteristics.
Gelfi and Giuriani [13] tested the use of stud connectors obtained from ordinary steel bars in timberconcrete connections with and without an interlayer of planks.

16

Timber Concrete Composites

COST E34 - WG1: Bonding on Site

Figure 2.5: Shear connector

2.2

Figure 2.6: Shear connector

Design methods

A fundamental condition for the actual application of timber-concrete composite beams/floors is the
availability of design methods and recommendations.
The Pre-standard ENV 1995-2:1997 (Eurocode 5 Part 2: Bridges) proposed a simplified method for the
analysis of composite timber-concrete beams made up with metallic shear connectors. However, most of
these provisions were withdrawn in the newly approved version of this standard, in which still no
reference is made to the interface bonded with an adhesive.
The work of van der Linden [14] is a comprehensive attempt to set design rules for the analysis of
composite floors made up with shear connectors. He based his proposal in the assumption of linear
elastic constitutive laws for both the concrete and the timber, while on the steel connectors were
considered to provide a continuously distributed stiffness along the span and to follow an elastic-perfectly
plastic rule.
In Germany, the so-called "Gamma-Method" is widely used for the design of timber-concrete-composite
structures. The so-called "Gamma-factor" accounts for the fact that the shear joint between the concrete
and the timber is elastically deformable and the corresponding slippage leads to higher bending moments
in the timber and concrete components than would be the case for a stiff joint without slippage. In a
recent paper, Kaliske and Schmidt [15] point out that in many practical cases ductile shear connectors are
used. The use of the elastic design rules of the "Gamma-method" is quite conservative. The authors
argue for a corresponding change in the design of the ultimate limit state by taking into account the
favourable redistribution of the internal forces thanks to the plastic deformation of the connectors.
Grosse and Rautenstrauch [16] and Grosse et al. [17] proposed constitutive laws for the numerical
modelling of timber and timber-concrete composites. Such laws are based on the theory of plasticity, with
flow and hardening rules conveniently adapted to the well-known parallel- and perpendicular-to-the grain
behaviour of timber. Numerical models for composite connection elements, based in these constitutive
laws, were tested, showing good agreement with the experimental data.
Frangi and Fontana [18] have formulated design equations for the shear connectors. They studied both
an elastic model for the service limit state and elastic-plastic models for the ultimate limit state. In another
paper [19], the same authors discussed the fire behaviour of timber-concrete slabs.
Demarzo and Tacitano [20] proposed an approximate method for the analysis of composite timberconcrete elements, but one of the basic assumptions of the method is that of the linear elastic model for
the load-slip constitutive relation of the connectors, which is severely restrictive in conditions of ultimate
loading.
Dias et al [21] described an ongoing research on the development of non-linear finite element models to
simulate the behaviour of these structures.
Jorge et al [22-24] investigated the advantages of using lightweight instead of normal concrete for the
topping layer. This same type of concrete was used in the experiments of Raji and Rak [25] and Raji
and Zagar [26] who tested three different possibilities of interlayer bonding, including direct bonding
between timber and fresh lightweight concrete, with encouraging results supporting this latter solution.
Toratti et al. [27] have measured the response of timber-concrete floors in both laboratory and site
conditions and confirmed the good behaviour with regard to vibrations induced by walking. The opposite

Timber Concrete Composites

17

COST E34 - WG1: Bonding on Site

result was achieved by Lee et al. [28] however, who found that, in spite of the enhanced stiffness and
static response, the lower natural frequencies due to the increased mass could result in a stronger
vibration response.
One important aspect with timber-concrete composites is the long-term performance, due to the different
creep and mechano-sorptive behaviours of the component materials. Amadio et al. [29, 30] introduced
and discussed a finite-element based procedure for the evaluation of these long-term effects. Long-term
testing was also conducted by Ceccotti. A brief description of the tests and main conclusions may be
found in [1]. The same author and Fragiacomo [31] proposed a simplified method for accounting the longterm effects resulting from loading, shrinkage, creep and daily and yearly changes in environmental
conditions. An ongoing experimental program on the creep behaviour of composite timber-lightweight
concrete is referred to in [1]. The creep of timber-concrete bonding with metallic shear connectors has
also been monitored for a few years by Dias [32]. Kuhlmann and Schnzlin [33] proposed the use of
effective creep coefficients which account for the different creep rates for concrete and for timber. They
also propose that shrinkage could be accounted for by introducing a fictitious load. Flach and Frenette
[34] discussed the application of the timber-concrete composite concept to actual bridge designs and
point out that a deeper insight into the long-term behaviour is required in order to ensure reliability of such
solutions.

2.3

Timber-concrete-composites with an adhesive bond

There are relatively few publications dealing with wood-based composites made up with an adhesive
bond between timber and other materials. Besides the aforementioned work of Raji and Rak [25], in
which adhesives were considered as one of the alternative bonding systems under study, regular
research in this domain has only been pursued by Rautenstrauch et al. [35-39], which will be referred to
later on in this section, and Brunner et al. [40-43]. In [40] the first tests in structural size composite beams
connected with this system are described. The theoretical failure loads were estimated with an elastic
calculation model assuming a stiff bond with no slippage between the timber and the concrete layers.
There was remarkable agreement between the ultimate test loads and the expected failure loads
(theoretically). A simplified evaluation of the long-term response was undertaken, in which the
contribution of timber was neglected, giving some simplifying yet restrictive assumptions. Furthermore,
they state that most of the long-term deformation is apparently caused by concrete shrinkage rather than
by creep effects. In order to clarify the effect of the interaction of materials with different shrinkage and
creep behaviours, there is an ongoing research on the topic. In [41], Brunner and Gerber conclude that
bonded joints can withstand severe climatic changes - particularly with regard to moisture content without significant strength loss, which is a point in favour of the reliability of this bonding system.
Brunner, Romer and Schnueriger [44] have published some recent work on timber-concrete-composite
slabs. They studied the wet-in-wet process, whereby the concrete is poured onto the still wet adhesive on
the timber component. They made a parameter research to find out the most favourable conditions for the
adhesive bond and found out, among other things, that the best shear bond was attained when there was
a short time interval to permit the wet adhesive to stiffen before the concrete was poured. They cast
several test specimens, without and then with openings in the concrete, to simulate the many conduits
which are often placed in concrete slabs in Switzerland. All the test specimens failed at loads which
corresponded very well to the calculation results.
The ability of timber to be adhered to concrete is also being investigated by Negro [45], with the
expectation of the application of such technique to composite timber-concrete beams and slabs.
Another development using timber-concrete composites for structural applications is the replacement of
cement bounded concrete by epoxy-resin bounded polymer concrete (PC). Polymer concrete is a
composite material formed by combining mineral aggregates such as sand or gravel with a monomer.
Rapid-setting organic polymers are used in PC as binders. Studies on epoxy polymers have shown that
curing method, temperature and strain rate influence, the strength and stress-strain relationships. PC is
increasingly being used as an alternative to cement concrete in many applications. Today, polymer
concrete is used for finishing work in cast-in-place applications, precast products, highway pavements,
bridge decks and waste water pipes, thereby developing better PC systems and characterizing the
compressive strength in terms of constituents are essential for the efficient utilization of PC. However, the
data on epoxy PC are rather limited, and there is an increasing interest in the deformation characteristics
under working conditions in combination with other materials such as wood for composite structures.
Epoxy resin-based polymer concrete can be combined with timber floors for upgrading without

18

Timber Concrete Composites

COST E34 - WG1: Bonding on Site

necessitating the removal of the suspended ceiling. This technique is likely to be suitable for the
reinforcement of timber floors and historical structural wood components [46].

2.4

Research needs

Timber-concrete-composites with mechanical connections have been widely researched. The technique
is frequently used in engineering practice. New research will always be necessary when new connections
are proposed. The current calculation methods are based on elastic methods like the Gamma-Method.
Some recent publications indicate that plastic design methods may lead to a more economic use of the
connectors, which are often rather expensive. This new research field needs to be intensified. TimberConcrete-Composites with an adhesive connection are a relatively new research field which has not yet
attracted the attention of many researchers. There is a need for further research activities in this field in
order to confirm and to validate the available findings. There are several open questions which need to be
addressed in an interdisciplinary approach by timber engineers and chemical engineers, e.g. with regard
to an optimization of the adhesive types, the best adhesive formulations and quantities needed, as well as
surface preparation. Furthermore, cyclic and durability requirements under diverse environmental
conditions need to be further analysed, this is a current research activity at Bauhaus-University of Weimar
till 2008.
Apart from technical questions, research is also needed on economic issues such as:

2.5

Timber-concrete composite structures with natural bond (no mechanical fasteners)

Prefabrication: factory type of manufacturing to assure quality and to cut down costs.

Connections on site.

On site production: quality management, cost effective techniques.

References

[1]

Ceccotti, A. 'Composite concrete-timber structures', Progress in Structural Engineering Materials 4


(2002) 264-275

[2]

Gutkowski, R., Balogh , J., Natterer , J., Brown, K., Koike, E. and Etournaud, P. 'Laboratory tests of
composite wood-concrete beam and floor specimens'. WCTE, Whistler, Canada, 2000, paper 8-2-1

[3]

Gutkowski, R., Thompson, W., Brown, K., Etournaud, P., Shigidi, A. and Natterer, J. 'Laboratory
tests of composite wood-concrete beam and deck specimens'. 1st RILEM Symposium on Timber
Engineering, Stockholm, Sweden, 1999, 263-271

[4]

Bathon, L. and Graf, M. 'A continuous wood-concrete-composite system'. WCTE, Whistler,


Canada, 2000, paper 8-2-2

[5]

Bathon, L. and Clouston, P. 'Experimental and numerical results on semi pre-stressed woodconcrete composite floor systems for long span applications'. World Conference Timber
Engineering, Lahti, Finland, 2004, 339-344

[6]

Ballerini, M., Crocetti, R. and Piazza, M. 'An experimental investigation on notched connections for
timber-concrete composite structures '. WCTE, Malasya, 2002, paper 4.4.4

[7]

Piazza, M. and Ballerini, M. 'Experimental/numerical results on timber-concrete composite floors


with different connection systems'. WCTE, Whistler, Canada, 2000, paper P50

[8]

Benitez, M.F. 'Development and testing of timber/concrete shear connectors'. WCTE, Whistler,
Canada, 2000, paper 8-3-2

[9]

Bou Said, E., Jullien, J. and Siemers, M. 'Non-linear analysis of local composite timber-concrete
behaviour'. WCTE, Malasya, 2002, paper 2.2.5

[10]

Inoue, M.e.a. 'Development of connecting system between reinforced concrete and timber'. WCTE,
Malasya, 2002, paper 10.4.1

[11]

Jullien, J.F., Michel, G., Mungwa, M.S. and Siemers, M. 'A new shear connector for wood-concrete
composite structures'. 1st RILEM Symposium on Timber Engineering, Stockholm, Sweden, 1999,
563-570

Timber Concrete Composites

19

COST E34 - WG1: Bonding on Site

[12]

Mungwa, M.S., Jullien, J.F., Foudjet, A. and Hentges, G. 'A new shear connector for woodconcrete composite structures'. WCTE, Montreux, Switzerland 1998, 518-525

[13]

Gelfi, P. and Giuriani, E. 'Stud shear connectors in wood-concrete composite beams'. 1st RILEM
Symposium on Timber Engineering, Stockholm, Sweden, 1999, 245-254

[14]

Linden, M.v.d. 'Timber-Concrete Composite Floor Systems', PhD Thesis, Technical University of
Delft, The Netherlands, 1998

[15]

Kaliske, M. and Schmid, J. 'A new design approach for timber-concrete-composite beams'. COST
C12 Final Conference, Innsbruck, Austria, 2005,

[16]

Grosse, M. and Rautenstrauch, K. 'Numerical model of timber and connection elements used in
timber-concrete composite construction'. CIB W18 Meeting 37, Edinburgh, UK, 2004,

[17]

Grosse, M., Lehmann, S. and Rautenstrauch, K. 'Testing connector types of laminated timberconcrete composite elements'. CIB W18 Meeting 34, Venice, Italy, 2001,

[18]

Frangi, A. and Fontana, M. 'Elasto-Plastic Model for Timber-Concrete Composite Beams with
Ductile Connection', Structural Engineering International-IABSE 1 (2003)

[19]

Frangi, A. and Fontana, M. 'Fire behaviour of timber-concrete composite slabs'. WCTE, Montreux,
Switzerland, 1998, 76-83

[20]

Demarzo, M.A. and Tacitano, M. 'Alternate method to elastically coupled timber-concrete beams'.
WCTE, Malasya, 2002, paper 3.2.2

[21]

Dias, A., Kuilen, J.d., Cruz, H. and Lopes, S. 'Non-Linear FEM models for timber-concrete joints
made with dowel type fasteners'. WCTE, Lahti, Finland, 2004, 371-376

[22]

Jorge, L., Cruz, H. and Lopes, S. 'The Use of Lightweight Concrete in Composite Timber-Concrete
Floors (in Portuguese)'. ENCORE, Lisboa, Portugal 2003, 901-906

[23]

Jorge, L., Cruz, H. and Lopes, S. 'Tests in Timber-LWAC Composite Beams with Screw-type
Fasteners'. WCTE Lahti, Finland 2004, 559-564

[24]

Jorge, L., Cruz, H. and Lopes, S. 'Experimental research in timber-LWAC composite structures'.
Int. Symp. Adv. Timber and Timber-Composite Elements Build, Florence, Italy, 2004,

[25]

Raji, V. and Rak, M. 'Continuous shear connecting - The best way to compose timber and
lightweight (eps) concrete'. WCTE, Malasya, 2002, paper 4.4.5

[26]

Raji, V., Zagar, Z. 'FEM models of composite timber-lightweight concrete floor systems'. WCTE,
Whistler, Canada, 2000, paper P45

[27]

Toratti, T., Talja, A. and Jrvinen, E. 'Classification of human-induced floor vibrations in buildings: a
wood-concrete composite floor example'. WCTE Malasya 2002, paper 4.1.3

[28]

Lee, P., Chui, Y.H. and Smith, I. 'Dynamic and static performance of wood floor with concrete
topping'. WCTE, Malasya, 2002, paper 11.2.2

[29]

Amadio, C., Ceccotti, A., Di Marco, R. and Fragiacomo, M. 'Numerical evaluation of long-term
behaviour of timber-concrete composite beams'. WCTE, Whistler, Canada, 2000, paper 8-2-4

[30]

Amadio, C., Di Marco, R. and Fragiacomo, M. 'A linear finite-element model to study creep and
shrinkage effects in a timber-concrete composite beam with deformable connections'. 1st RILEM
Symposium on Timber Engineering, Stockholm, Sweden, 1999, pp. 747-756

[31]

Fragiacomo, M. and Ceccotti, A. 'A simplified approach for long-term evaluation of timber-concrete
composite beams'. WCTE, Lahti, Finland, 2004, 537-542

[32]

Dias, A. 'Mechanical behaviour of timber-concrete joints', PhD Thesis, Delft, The Netherlands,
2005

[33]

Kuhlmann, U. and Schnzlin, J. 'Time dependent behaviour of timber-concrete composite


structures'. WCTE, Lahti, Finland, 2004, 313-318

[34]

Flach, M. and Frenette, C. 'Wood-Concrete Composite Technology in Bridge Construction'. WCTE,


Lahti, Finland, 2004, 289-294

[35]

Rautenstrauch, K., Grosse, M. and Lehmann, S. 'Forschungsvorhaben Brettstapel-Beton-Verbund,


Teil 1: Untersuchung des Tragverhaltens von Brettstapel-Beton-Verbunddeckenplatten mit

20

Timber Concrete Composites

COST E34 - WG1: Bonding on Site

neuartigen Verbindungsmitteln
Eigenverlag, 2001

aus

Flachstahlschlssern',

Bauhaus-Universitt

Weimar.

[36]

Rautenstrauch, K., Grosse, M. and Lehmann, S. 'Forschungsvorhaben Brettstapel-Beton-Verbund,


Teil 2: Flachstahlschlsser in Rundhlzern und Derbstangen, Betonnocken in aufgestellten
Lamellen und verzinkte Lochbleche.' Bauhaus-Universitt Weimar, Eigenverlag, 2002

[37]

Rautenstrauch, K., Grosse, M., Lehmann, S. and Hartnack, R. 'Baupraktische Dimensionierung


von Holz-Beton-Verbunddecken'. 6. Informationstag des Instituts fr Konstruktiven Ingenieurbau
(IKI), Bauhaus-Universitt Weimar, 2003,

[38]

Rautenstrauch, K., Grosse, M., Lehmann, S. and Hartnack, R. 'Baupraktische Dimensionierung


von Holz-Beton-Verbunddecken'. 6. Informationstag des Instituts fr Konstruktiven Ingenieurbau
(IKI), 2003,

[39]

Rautenstrauch, K., Grosse, M., Lehmann, S. and Hartnack, R. (2004) 'Modellierung und
baupraktische Bemessung von Holz-Beton-Verbunddecken mit mineralischen Deckschichten unter
Bercksichtigung neuartiger Verbindungsmittel. Beitrag der Fachtagung Holz-Beton-Verbund
Leipzig 2004'. In: Knig, G. and Holschemacher, K. (eds) Holz-Beton-Verbund. Bauwerk Verlag
GmbH, Berlin.

[40]

Brunner, M. and Gerber, C. 'Composite decks of concrete glued to timber'. WCTE, Malasya, 2002,

[41]

Brunner, M. and Gerber, C. 'Long-term tests on a glued timber-concrete composite'. WCTE


Malasya, 2002, paper 10.4.2

[42]

Brunner, M. and Schnriger, M. 'Towards a future with ductile timber beams'. WCTE, Malasya,
2002, paper 11.2.3

[43]

Brunner, M. and Schnriger, M. 'Timber beams strengthened with pre-stressed fibres:


Delamination'. WCTE, Lahti, Finland, 2004,

[44]

Brunner, M., Romer, M. and Schnriger, M. 'Timber-concrete-composite with an adhesive


connector (wet on wet process)', Materials and Structures 40 (2007) 119-126

[45]

Negro, J. 'Shear testing of glued timber-concrete small-size specimens. Unpubl. Report', Dept.
Civil Eng., Univ. Coimbra, Portugal, 2005

[46]

Schober, K.U. and Rautenstrauch, K. 'Upgrading and repair of timber structures with polymer
concrete facing and strengthening'. WCTE, Portland OR, USA, 2006,

Timber Concrete Composites

21

COST E34 - WG1: Bonding on Site

TIMBER WITH PASSIVE REINFORCEMENT


Schober K. U., Brunner M., Negro J., Rautenstrauch K., Lavisci P.

3.1

Overview

It is a well-known feature of timber beams that they usually fail suddenly due to the breaking of fibres on
the tensile face. This behaviour may be changed by an adequate strengthening of the tensile face, which
may lead to a more ductile failure induced by the gentle local buckling of the fibres in the compressive
face. Such strengthening may be achieved through a number of ways:

by using a larger tensile flange

by using a much better timber grade on the critical tensile face

by bonding a passive (slack) or active (pre-stressed) reinforcement with adhesives

A number of materials may be used as reinforcement:

CFRP (carbon fibre reinforced plastics)

GFRP (glass fibre reinforced plastics)

AFRP (aramid fibre reinforced plastics)

Kevlar, steel bars and plates

When timber beams are strengthened with the above materials, the result is a composite structural
element. In most practical cases adhesives are used for the bond line. Epoxy based adhesives have been
used in most cases for on site repair jobs, but most formulations were developed for other materials.
These adhesives are generally too rigid for bonding timber and there is no chemical bonding or suitable
mechanical anchorage in wood. The bond line is prone to fail because of dimensional changes in the
wood induced by moisture content variations, even under Service Class 2 applications (moisture content
of timber up to 18 %). However, this group of adhesives has the potential to be the most suitable one for
on site bonding. Therefore, it is necessary to improve the existing adhesives or to develop alternative
adhesive formulations, suitable for indoor (and even outdoor) jobs on site where the use of pressure is
generally not available and clean regular bond lines are difficult to achieve. Moreover, the existing test
methods to evaluate bond line performance were developed for other types of industrially used
adhesives, especially phenol or aminoplastic based products, and have repeatedly proved to be
inadequate to assess the behaviour of epoxy type adhesives or epoxy bonded products. On site
application of adhesives is somewhat difficult and the consequent quality of adhesive bond is not easy to
evaluate. Since properties of reinforced elements very much depend on the care put into the work, such
difficulties have to be overcome. Procedures for applying and controlling are needed and must be
established [1]. These procedures generally involve the use of materials compatible with and able to
perform in accordance with the physical characteristics of the timber.

3.2

Literature survey

The last two decades have seen a lot of research work done on the strengthening of timber beams with
fibre reinforced plastic laminates. Tingley [2] is an important pioneer. Since the 1980s, he has
systematically carried out a large number of scientific tests on glulam beams strengthened with fibre
reinforced plastic laminates and also formed a company which successfully sells the technology to glulam
companies, mostly in North America. The technology is widely used to strengthen and upgrade new
beams for construction. In the FIRP system a flat CFRP laminate is inserted as reinforcement between
the two bottom lamellas of the glulam element, in order to protect it and to enhance the contact area for
shear transmission.

22

Timber With Passive Reinforcement

COST E34 - WG1: Bonding on Site

Figure 3.1: FIRP System


European researchers started intensive research on the topic in the 1990s. One such pioneer was van de
Kuilen [3], who studied the strengthening of timber beams on the tensile face alone, and then on both the
tensile and compressive faces. The tests were performed with larch wood reinforced with glass fibres
glued with phenol-resorcinol type adhesive. Smedley has written several papers on the work he has
done, which are presented on the website of the LICONS project [4]. German researchers have also been
quite active. Bla and Romani have written papers on strengthening of glulam beams using CFRP [5].
Unlike the case in North America, there is some scepticism that the technology will be economical for new
timber structures in Europe. Many authors such as Rautenstrauch and Schober see a good market
opportunity for the technology in the strengthening of existing timber structures [6].They have also
investigated two basic approaches for the use of reinforcement materials embedded in the wood, and the
use of external reinforcement resulting in a system of composite type [7]. The first results and insights
have been used successfully for renovation and reconstruction of historic roof and floor constructions as
shown in the next figures.

Figure 3.2: Damaged spire of the Merseburg


cathedral (D) due to high wind loads
and weathering. Cantilever joint after
strengthening with CRPFs

Figure 3.3: CFRP and polymer concrete


strengthening of an historic ceiling
joist in waffle slab, Mansfeld castle
(D)

Several experimental beam tests by Borri et al. [8], Triantafillou [9], Schober and Rautenstrauch [10]
showed that the most frequent fracture mechanism is caused by the failure of the traction zone without
the complete plasticization of the compression region, depending on the quality of the wood. However,
under particular conditions it is possible to note the other failure mechanism, which is theoretically
preferable for several reasons. First, the section shows a more ductile behaviour, while the stresses in the
FRP material with reinforcement are highly increased and therefore the composite material is more
involved. Initially the load defection is shown to be linear elastic up to local failures induced by the
presence of defects e.g. knots and cracks. Wood yield produced a non-linear response terminated by a
sudden drop of the load as a result of CFRP rupture. CFRP rupture was immediately followed by wood
fracture in the tension zone, resulting in the collapse of the beams. Significantly good results, though,
have been found for the reinforcement of old floor beams where strength and stiffness needed a
moderate improvement (15-20%) in order to fulfil the requirements of new codes and/or higher imposed
loads. Figure 3.4 illustrates such an example, which is quite typical in Italy.
Timber With Passive Reinforcement

23

COST E34 - WG1: Bonding on Site

Figure 3.4: Reinforcement with CFRP of floor beams on the tensile side. Photos: P. Lavisci.

3.3

Design methods

The service limit state of strengthened beams can generally be calculated with sufficient accuracy using
the usual linear elastic methods. The stiffness of the composite beam can be determined with Steiner's
rule. With an adhesive bond there is no significant slippage to be taken into account. The calculation of
the ultimate limit state is rather more difficult. The load-bearing capacity of strengthened timber beams
can be estimated by assuming linear elastic behaviour when the strengthening is relatively little. When
the tensile face is adequately strengthened however, the compressive face may suffer higher stresses
and thus "yield" before the tensile face breaks. The failure mode of the beam will be ductile,
corresponding to timber failure under compressive loading, and the load-bearing capacity can be more
accurately estimated with plastic design methods.
There is as yet no universally recognized and accepted calculation method for strengthened timber
beams in Europe. The new Eurocodes for example do not treat the topic. The determination of the
flexural strength of the composite structure can been done in the elastic range and the increasing of the
bending stiffness by applying CFRP reinforcement can be defined by a fictitious modulus of elasticity,
calculated from the cross-section data of the un-reinforced specimen according to EN 408:

E fict =

a 21 (F2 F1 )
16 I (w2 w1 )

(1)

where F2-F1 is the load increase in elastic range and w2-w1 the equivalent deflection values. The
reinforcing scheme increased the capacity (Efict) in comparison to the values measured for the unreinforced wood beams. Different plastic calculation models have been proposed for the ultimate limit
state. Most of the models make a clear distinction between the elastic-plastic stresses on the
compressive face of the timber on the one hand, and the purely linear-elastic stresses on the tensile face.
One of the older models, proposed by Tingley [2], assumes a stress distribution as shown in Fig. 3.5 with

linear strain distribution over the height of the beam

constant compressive stress over the entire compression zone

linear stress distribution on the tensile face

Some European authors such as Brunner and Bla use a slightly modified stress distribution in the
compression zone to permit a transition from zero to the compressive stress as shown in Fig.3.6. Van de
Kuilen uses the more complex stress distribution shown in Fig.3.7. He assumes the hyperbolic curve (ex)
proposed by Glos [11] to describe the relationship between the stresses and strains in an unreinforced
timber beam.

24

Timber With Passive Reinforcement

COST E34 - WG1: Bonding on Site

Strains

Stress distribution

Internal forces:

fc,T

Legend:
Fc,T

e1

Neutral axis

fc,T: axial compressive strength


of timber
Fc,T= internal compression force
ft,T: bending strength of timber
Ft,T= internal tension force in
timber
ft,F: tensile stress in FIRP
Ft,F= force in fibre

e2

Ft,T
ft,F
Ft,F

ft,T

Figure 3.5: Plastic model proposed by Tingley

Strains

Stress distribution

Internal forces:

fc,T

Legend:
Fc,T

Neutral axis

e1

fc,T: axial compressive strength


of timber
Fc,T= internal compression force
ft,T: bending strength of timber
Ft,T= internal tension force in
timber
ft,F: tensile stress in FIRP
Ft,F= force in fibre

e2

Ft,T
ft,F
Ft,F

ft,T

Figure 3.6: Plastic model used by Brunner and Bla

Strains

Stress distribution

Internal forces:

fc,T

Legend:
Fc,T

ex
Neutral axis

e1

e2

Ft,T

fc,T: axial compressive strength


of timber
Fc,T= internal compression force
ft,T: bending strength of timber
Ft,T= internal tension force in
timber
ft,F: tensile stress in FIRP
Ft,F= force in fibre

ft,F
ft,T

Ft,F

Figure 3.7: Plastic model used by Kuilen

Timber With Passive Reinforcement

25

COST E34 - WG1: Bonding on Site

Many calculation methods make the following common assumptions:

Elastic-plastic distribution of the compressive stresses in the timber. The failure stress is
assumed to be equal to the axial compressive strength.

Linear-elastic distribution of the tensile stresses in the timber. Many authors assume the
maximum failure stress to be equal to the bending strength of timber. Brunner argues that the
so-called bending strength of the timber beam is no true material property because it is
calculated from experiments by assuming a linear distribution of the stresses. In high grade
timbers, the compressive face will experience some slight plasticization before failure occurs.
Brunner proposes that for higher grade timbers the true failure strength of the tensile face must
be somewhat higher than the bending strength.

In cases of very great strengthening, such as with pre-stressing, the contribution of the timber
tensile force is relatively small and may be neglected.

All the calculation methods are iterative and in many cases they lead to similar results. An
assumption is made for the position of the neutral axis. The failure strain of the tensile face can
be estimated from the bending strength and the Modulus of Elasticity: the characteristic strains in
the compression zone of the timber and the strain in the artificial fibre can be calculated
accordingly. From the stress-strain diagrams, the stress distributions in the composite materials
can be calculated. The internal forces are calculated from the stress distribution in the crosssection and they must fulfil the following equilibrium equation:

No external axial force. Sum of internal forces must be zero:


Fc,T + Ft,T + Ft,F = 0

(2)

The calculation is repeated for different assumptions of the position of the neutral axis until the above
condition is fulfilled. The bending resistance is then calculated by multiplying the internal forces on the
tensile face in both timber and laminate with the corresponding distances from the compressive force in
the timber:
MR = Ft,T . e1 + Ft,F . e2

(3)

The Italian National Research Council has issued a guideline for the design of timber reinforcement with
FRPs, in order to disseminate best practices among the many professionals that already deal with this
subject [12].

3.4

Economic considerations

The use of these new reinforcement materials in a passive, not pre-stressed state raises a number of
important questions:
1.

2.

3.

4.

26

Does the effectiveness of the strengthened system compensate for the increase in costs resulting
from the additional workmanship, use of complex adhesives and the still quite high (though
decreasing) cost of the aforementioned reinforcement materials?
Whilst the use of reinforcement may dramatically improve the performance of a concrete
structure, much more moderate results are to be expected with timber, because this material can
withstand both tensile and compressive stresses quite well. If, for instance, a perfectly composite
behaviour of a CFRP-reinforced timber beam is assumed and the cross-section is homogenized,
one may easily conclude that a 1 mm-thick CFRP-laminate is roughly equivalent to an extra
20mm timber lamella. Does it make economic sense to use reinforcement?
In most practical cases, the control of the serviceability limit state, in particular with regard to
deflections and vibrations, is the most important factor governing the required size of the timber
beam. An analysis suggests that the use of reinforcement may not always be very helpful in this
case. While the ratio between the design strengths of many engineered reinforcement materials,
particularly CFRP and timber may be as high as 100, the ratio of their modulus of elasticity is only
of about 20.
Ultimate limit state calculations of reinforced elements show that the stress in the reinforcement
material is well below the strength. Timber tensile failure occurs at strains of about 0.2 0.3%,
which is well below the yield strain of many steel types or the service strain of most plastic fibres
(> 0.5%), which means that we cannot make full use of the potentially high strength of the
reinforcement materials.
Timber With Passive Reinforcement

COST E34 - WG1: Bonding on Site

5.

Given the high cost of the new plastic reinforcement materials, the economy of such
strengthening of timber does not seem to justify the use in new structures. Many European
authors such as Schober and Rautenstrauch suggest that the only really feasible market for
strengthening and upgrading existing structures.

3.5

Steel reinforcement

A possible alternative to the high-performance fibres is the use of the common and less expensive
construction rebars. In most systems steel rebars are inserted into longitudinal holes. After the rebars are
in place, adhesive is injected under pressure through small holes in the top surface of the timber element.
The ribbed surface of the rebars improves the bonding with the adhesive. Figure 3.8 depicts two
commercial systems i.e. Tasbeam and Aralam.
One cannot help wondering whether the increase in strength compensates for the cost, because this
process seems to be even more complex than that of bonding fibre reinforcement strips to the tensile face
of the timber beam. Some preliminary studies indicate that the answer is no, for reasons similar to those
pointed out for the new engineered materials.

Figure 3.8: Steel reinforcement systems (left systems Tasbeam, right: Aralam)

3.6

Research needs

The technical aspects have been widely researched, but some difficult problems remain:

Practical solutions to premature debonding: with special adhesives, or with mechanical devices
such as clamps, bonding in stages, etc.

Appropriate calculation methods, particularly for the ultimate state, when the timber compressive
face may plastify.

Practical application methods on site

Long-term behaviour of reinforced timber structures

The economic aspects need to be more carefully analysed:

Possibilities of mass production

Appropriate tools and technologies to rationalise and speed up work

Timber With Passive Reinforcement

27

COST E34 - WG1: Bonding on Site

3.7

References

[1]

CEN 'TC, 193/SC1/WG11. Adhesives for on site assembling or restoration of timber structures. On
site acceptance testing:
Part 1: Sampling and measurement of the adhesives cure schedule. Doc. N20.
Part 2: Verification of the shear strength of an adhesive joint. Doc. N21.
Part 3: Verification of the adhesive bond strength using tensile proof-loading. Doc. N22.' (2003)

[2]

Tingley, D. 'FIRP Reinforcement Technology Information Packet', Science and Technology


Institute, Corvallis OR, USA, 1995

[3]

Kuilen, J.v.d. 'Theoretical and experimental research on glass fibre reinforced laminated timber
beams'. International Timber Engineering Conference, London, England, 1991, 3.226-3.233

[4]

Anonymus 'Low Intrusion Conservation Systems for Timber Structures.' 2006, Website:
http://www.licons.org/

[5]

Bla, H.J. and Romani, M. 'Design model for FRP reinforced glulam beams'. International Council
for Research and Innovation in Building and Construction. Working Commission W18 Timber
Structures, Venice, Italy, 2001,

[6]

Rautenstrauch, K. and Schober, K.U. 'Verstrkung von historischen Holzbauteilen mittels CFKLamellen'. Europische Messe fr Restaurierung, Denkmalpflege und Stadterneuerung, Leipzip,
Germany, 2004,

[7]

Schober, K.U. and Rautenstrauch, K. (2005) 'Strengthening of timber structures in-situ with an
application of fiber-reinforced polymers'. In: Seracino. (eds) FRP Composites in Civil Engineering CICE 2004. Taylor & Francis Group, London, 697-704. ISBN 90 5809 638 6.

[8]

Borri, M. and et al. 'FRP reinforcement of wood elements under bending loads'. Structural Faults
and Repair, London, England, 2003,

[9]

Triantafillou, T.C. 'Shear reinforcement of wood using FRP materials', Journal for Materials in Civil
Engineering 9 (2) (1997) 65-69

[10]

Schober, K.U. and Rautenstrauch, K. 'Experimental investigations on flexural strengthening of


timber structures with CFRP'. International Symposium on Bond Behaviour of FRP in Structures,
Hong Kong, China, 2005,

[11]

Glos, P. 'Zur Modellierung des Festigkeitsverhaltens von Bauholz bei Druck-, Zug- und
Biegebeanspruchung. Berichte zur Zuverlssigkeitstheorie der Bauwerke 61', Universitt Mnchen,
1981

[12]

CNR 'Studi Preliminari finalizzati alla redazione di Istruzioni per Interventi di Consolidamento
Statico di Strutture Lignee mediante lutilizzo di Compositi Fibrorinforzati', CNR-DT 201/2005
(2005) 60 p

28

Timber With Passive Reinforcement

COST E34 - WG1: Bonding on Site

PRE-STRESSING OF TIMBER
Negro J., Brunner M., Lehmann M.

4.1

Overview

The idea of strengthening timber beams by reinforcing the tensile side with non-pre-stressed materials is
basically good, but in practice the economy may place a limit on its use because the high strength of the
expensive reinforcement is not fully utilised. An increasing number of European researchers believe that
the only truly feasible market may be in the repair and strengthening of existing structures.
A small but increasing number of researchers believe that the economic efficiency of the reinforcement
could be improved by pre-stressing instead of using it as a passive reinforcement only, because its final
(service) stress may be substantially increased. This research field, which is relatively new for timber
structures, will be discussed in this section.
The mechanics and calculation of pre-stressing has been clearly understood for almost a century.
However, the concept has been empirically used since ancient times. The Egyptians, for example, bound
wooden pieces together with heated metal rings. As the metal cooled, it firmly bound the wooden pieces
to form a wine barrel. The basic concept is that of imposing onto the structure or element a stress state
which is opposite to that expected to result from the service loading. This leads to an increase in the
loading that the structure can withstand or, alternatively, to a reduction in structural material needed for a
specified loading.
The only successful practical application of pre-stressing to timber structures up to the present has been
the case of stress-laminated decks. In this system, however, initial stresses due to pre-stressing are not
an objective in its own, as in the general case. Instead, the main goal is to use these perpendicular-tograin compressive stresses to enhance the friction between planks, thus allowing shear stresses to be
transmitted through the interface, resulting in a better two-dimensional behaviour. The stress-laminated
deck has been subjected to intensive international research and there is wide agreement on the loadbearing behaviour.
Besides the stress-laminated panels, only a few references are made, in the literature, to the use of prestressing with timber (Fig. 4.1, 4.2). Luggin and Bergmeister [1] consider that this is a promising
technology. Galloway et al. [2] investigated the behaviour of Kevlar-reinforced timber beams, with and
without the use of pre-stress.

Figure 4.1: Post-tensioned timber beams

Figure 4.2: Post-tensioned timber beams

Krahemann and Fontana [3] refer to a research project in timber-concrete decks with the timber beams
pre-stressed with steel bars.

Figure 4.3: Pre-stressing-induced


camber

The strengthening effect may be significantly improved if active (pre-stressed) instead of passive
reinforcement is used. The ultimate limit state design can take into account the possibility of a favourable
ductile failure of the structural element of the compressive face instead of the brittle tensile face of the
Pre-Stressing Of Timber

29

COST E34 - WG1: Bonding on Site

beam. Pre-stressing also offers a further advantage with regard to service state design. Since the
deflections are often the critical design criteria for timber beams, the upward deflection induced by the
eccentrically placed pre-stressed laminate (Fig. 4.3) may prove to be of major help [4].

1 - Jacking

Reinforcement

2 - Gluing/curing

Figure 4.4: Pre-stressing by


cambering

3 - Jack release

In principle, any of the previously mentioned reinforcement materials may be used for pre-stressing. In
fact, fabric or laminate composite (plastic) materials are the obvious choice for strengthening or
rehabilitation of existing structures, because they can be applied on the element surface, thus avoiding
drillings or any other weakening procedures. The preferred method for the actual application of prestressing on site is an issue still under investigation, but a possible approach could consist of bonding the
reinforcement after imposing a camber to the beam and to unload after the curing of the adhesive, as
schematized in Figure 4.4. This method was investigated by Lehmann [5]. In this investigation a
calculation model was developed and verified using small specimens and structural sized GL24h beams
which were reinforced using the method described in Figure 4.4. The calculations and the measurements
with strain gauges showed that the pre-stress force in the Carbon Fiber Reinforced Plastic (CFRP)
lamella is not constant over its length. It peaks in the middle of the beam where it is mostly needed and is
zero towards the ends where high stress would causes delamination (Fig. 4.5). The shear stress in the
glueline is constant and quite low. The pre-stress force is related to the force present in the prop used for
jacking, which is itself limited to the bending strength and ability of holding the beam ends in place. The
pre-stress force introduced with this method is much lower than could be attained by stretching the
CFRP-lamella before attaching it to the timber. The major benefits of this method are a significant
contribution to the service limit state (40%) and the simplicity of putting it in place as well as the reduced
danger of delamination.
Distribution of the pre-stress stress in the CFRP lamella
300
Stress [MPa]
250
200
150
100
50
0
0

500

1000

1500

2000

2500

Position of the strain gauges [mm]

3000

3500

4000

Figure 4.5: Pre-stress distribution in the


CFRP lamella

A major problem still to be satisfactorily solved is delamination. When the usual stiff adhesives on the
market are used, the shear stresses in the interface between the timber and the plastic laminate attain
high peaks concentrated at the beam ends. Luggin [1] reported that most of his test specimens failed
because of delamination, when the plastic laminate was suddenly peeled off the timber. The delamination
problem has been studied by Brunner and Schnueriger [6]. They proposed the development of a ductile
adhesive to distribute the shear stresses over a certain beam length. Unfortunately, the adhesives they
worked with proved to be incapable of withstanding the required pre-stressing force. Nevertheless,
30

Pre-Stressing Of Timber

COST E34 - WG1: Bonding on Site

provided that this problem can be managed with further research work, the latter approach with ductile
adhesives may prove to be the most economical and suitable for practical use.
Brunner and Schnueriger have also studied alternative methods to attach pre-stressed FRP to strengthen
glulam beams [7]. They used special gradiented anchoring equipment developed by CarboLink, a spin-off
company of the EMPA, Switzerland, to attach the fibres in stages starting at the beam centre. After each
step the pre-stressing force was slightly reduced: thus the force was effectively anchored over a
considerable length at both ends of the beam. They have performed two test series to date. In the first
test series, only one pre-stressed CFRP-laminate was used per beam. There was no delamination: in all
the test specimens, the load-bearing behaviour corresponded quite well to the calculated predictions.
They observed however, that the pre-stressing force was not strong enough to induce a significant
plasticization of the compressive face of the timber beam. In a follow-up project, they sought to augment
the pre-stressing force by attaching up to three pre-stressed CFRP-laminates on top of each other. The
laminate closest to the beam had the same length as the beam and therefore extended to the supports.
The outermost laminated were shorter in length and did not extend to the supports. There were no signs
of delamination during the long storage period of about three months before the beams were tested.
During the testing however, the outermost laminates debonded suddenly. Curiously, even with this
sudden failure mode, the load-bearing capacity calculated corresponded well to the measured test values
(Figures 4.6 and 4.7).
320-1 Kraft 1+2
100
90
80
Last F1+F2 in kN

70
60
50
40
30
20
10
0
-20

20

40

60

80

100

120

140

160

Durchbiegung wm in mm unter Pressen

Figure 4.6: Typical load-deflection behaviour of the glulam


beams pre-stressed with multi-CFRP [7]

Figure 4.7: Debonding of the outermost pre-stressed


CFRP in phases [7]

However, the use of these new reinforcement materials raises a number of questions still to be answered:

Does the effectiveness of the strengthening system compensate the increase in costs resulting
from additional workmanship, adhesives and mostly the high (though decreasing) cost of the
aforementioned reinforcement materials?

While the use of reinforcement may improve dramatically the performance of a concrete
structure, a much more moderate result is to be expected with timber, because this material can
withstand either tensile or compressive stresses. If, for instance, a perfectly composite behaviour
of a CFRP-reinforced timber beam is assumed and the cross-section is homogenized, one may
easily conclude that a 1mm-thick laminate is roughly equivalent to an extra 20mm lamella. Is this
worthwhile?

While the ratio between design strengths of most engineered reinforcement materials
(particularly CFRP) and timber may be as high as 100, that of their modulus of elasticity is only
of about 20.

When the design of the reinforced element under the assumptions of the previous topic is made, one
concludes that the stress in the reinforcement material is well below that leading to its effective use.
Given the high cost of such materials, one may wander whether its use in timber reinforcement is worthy
or not. The effectiveness of fibres may be improved by pre-stressing instead of using them as a passive
reinforcement only, because their final (service) stress may be substantially increased. The problem with
this procedure is, again, the risk of delamination, because the shear force in the interface rises in the

Pre-Stressing Of Timber

31

COST E34 - WG1: Bonding on Site

same proportion of the tensile stress in fibres, while the interface effective shear length does not, because
of the limited ductility of current adhesives.
Krahemann & Fontana [3] remark that the timber collapse by tension occurs for strains of about 0.2-0.3
%, well below the yield strain of most steels or the service strain of most fibres (>0.5 %), which means
that the reinforcement materials shall remain under-explored up to the collapse, unless pre-stressing is
used.
The use of pre-stressing may as well be considered for new structural elements. An interesting possibility,
in this case, given the high cost of C/GFRP materials, is their replacement by ordinary pre-stressing steel
wire. In such case, a longitudinal slot should be drilled in the timber element and the wire inserted in it
and involved with an appropriate adhesive. Although the process seems at first glance to be somewhat
complex, it could be performed at a relatively small additional cost within the framework of the traditional
production scheme of glued-laminated timber. The procedure could be used for medium-to-large spans.
However, problems concerning element stability, the equipment required for application of large prestressing forces and the small number of elements produced in a typical design (making expensive any
adaptation of the production line) are likely to make such an application uncompetitive. It is in the range of
4-6m and with mass production that the system looks more promising. These are typical housing
dimensions and such beams could be used in flooring, similarly to those made up with pre-stressed
concrete, which have been (and still are) successfully used for decades. Figure 4.8 [8] compares the
critical condition as a function of the span, for a GL24h beam with a cross-section of 90x180 and either
subjected or not to a pre-stressing eccentric force of 47.5kN. Value 1 in vertical axis corresponds to the
onset of ultimate (ELU) or serviceability (ELS) limit state. As can be seen, pre-stressing allows for a span
close to 5m, comparing to less than 4m with the non-pre-stressed beam.

Figure 4.8: Influence of pre-stressing in critical condition of small beams


Regarding the pre-stressing system, either pre-tension or post-tension could be used. The latter poses
delicate problems of perpendicular-to-the-grain brittle fracture, requiring some type of transverse
reinforcement close to the ends (anchorage sections), which increases the complexity and cost of the
system. Pre-tension avoids this inconvenience by spreading the shear stresses along the bonding length.
Besides, it allows pre-stressing to be applied to part of the span only, by adequate positioning of the glue
lines, or controlling the ratio between compressive and tensile service stresses, by the position of the
applied force across the beam depth. Unfortunately, these apparent benefits have their counterpart in
what is probably the major weakness of the system: the lack of ductility of the currently available
adhesives. This problem, though requiring some attention in any application involving bonding of timber to
other material, is particularly concerning when pre-stressing is used, given the intense permanent shear
stress it produces at the bonded interface.

32

Pre-Stressing Of Timber

COST E34 - WG1: Bonding on Site

Figure 4.9: Timber test specimen for withdrawal tests


of glued-in pre-stressing steel wire

Figure 4.10: Epoxy specimen testing

In a feasibility study, Barroso et al [8] undertook withdrawal tests of glued-in pre-stressing steel wire
(Figure 4.9). The bonding between the steel wire and timber was made up of either epoxy resin or
phenol-resorcinol-formaldehyde. In all the tests, failure occurred at the interface between the steel wire
and the adhesive, with the clean wire being pulled out. Significant values for the average withdrawal
strength (1.21 and 1.35 MPa, respectively, with COV of 19% and 16%) were found, though far below
those expected/needed. In order to clarify this issue, pull-out tests of pre-stressed steel wire glued within
epoxy specimens were made, as shown in Figure 4.10. Values in the range 5-9 MPa were obtained, with
large COV.

Figure 4.11: Force-displacement diagram


The force-displacement diagrams were as depicted in Figure 4.11, in which the sinusoidal pattern is most
likely due to the helical slight indentation of the wire and to the brittleness of the adhesive, which prevent
the effective participation of a long glueline in the process of stress transfer between materials. This is an
ongoing research.
The problem of brittleness of adhesives is also pointed out by Kemmsies and Streicher [9], though in the
context of research involving bonding of timber with steel in end connections.

4.2

Calculation Methods
Strains

Stress distribution

Internal forces:

Legend:
f c,T : Axial compressive strength of timber
F c,T : Internal compressive force timber
f m,T : Bending strength of timber
F t,T : Internal tensile force in timber
ft,L : Tensile stress in FRP-laminate
F t,L : Tensile force in FRP-laminate

fc,T

F c,T

Neutral axis

e1

e2

F t,T
ft,L
ft,T

Ft,F

Figure 4.12: Strain and stress distributions in a strengthened timber beam

Pre-Stressing Of Timber

33

COST E34 - WG1: Bonding on Site

The calculation models described in Chapter 3 need only a slight modification in order to permit the
calculation of timber beams strengthened with pre-stressed CFRP laminates. Brunner and Schnringer
[10], for example, refer to the calculation of pre-stressed concrete and modifies the calculation model
depicted in figure 4.12, by adding the initial pre-stressing force in the CFRP to the additional force
corresponding to the strain level. The distribution of the stresses in the timber remains essentially the
same as depicted in fig. 4.12, though of course the neutral axis will be shifted downwards to
accommodate the larger compressive zone needed to counter-balance the greater force in the CFRP
laminate.
4.2.1

Calculation example of pre-stressed timber beam

The calculation model described above was used to predict the load-bearing capacity of a strengthened
timber beam which was later tested in bending (figure 4.13).
The timber has the following material properties:

Dimension 140 mm width, 200mm height

GL 32h: Em=14 kN/mm2

5% fractile values according to Eurocode EN 1194: fc,k = 29 N/mm2, fm,k = 32 N/mm2

Medium values expected in loading tests are about 1/3 higher than the 5% fractile values:
fc = 39 N/mm2, fm= 43 N/mm2

Characteristic strains:

Tensile failure at t = fm / E = 43 / 14000 = 3,07 .

Yielding of compressive face at c = fc / E = 39/14000 = 2,79


The FRP laminate used has the following properties:

S&P-carbon laminate type 150/2000

Cross-section 1.4x50mm

E=165 kN/mm2

Pre-stressing force: 60 kN (simplification: neglecting of losses due to creep and elastic


deformations)

It is hereby assumed that failure of the tensile face of the timber beam at a strain of 3.07 resp. a stress
of 43 N/mm2 will induce collapse.
The calculation is iterative. Assuming a height z1=91mm of the tensile face of the timber leads for
example to the following results:

34

z2 = (39/43)  91 = 83 mm

Maximum strain on the compressive face: O = 2.79  (26 + 83) / 83 = 3.66

Strain in the FRP (additional to pre-stressing) = (91,7/91)  3,07 = 3,09


Additional stress in FRP = x E = 3,09  165.000 = 510 N/mm2

z3 = 200 (91 + 83) = 26 mm

Pre-Stressing Of Timber

COST E34 - WG1: Bonding on Site

S T R A IN S
%o

IN T E R N A L
FO R C ES

STRESSES
N /m m 2

39

D1
Z3

2 .7 9

200 mm

Z2

D2

Z1

Z2

3 .0 7

43

50 m m

Z1 + Z0
510

1 40 m m

Figure 4.13:Calculation example of pre-stressed glulam beam


Internal forces:

Timber compressive face:


D1 = (26 x 140) x 39 / 1000 = 142.0 kN
D2 = 0,5 x (83 x 140) x 39 / 1000 = 226,5 kN
D(total) = D1 + D2 = 368,6 kN

Timber tensile face:


Z2 = 0,5 x (91 x 140) x 43 / 1000 = 273,9 kN

CFRP:
Pre-stressing force Z0 = 60 kN
Additional force Z1 = 510 x (1,4 x 50) / 1000 = 35,7 kN

Total tensile force: Z0 + Z1 + Z2 = 369,6 kN

Since the compressive forces and the tensile forces are (nearly) equal, the iterative process can be
ended. The distances between the forces can be found from a consideration of the geometry. The
resultant compressive force D (total) for example has the following distance from the top of the beam:
e1 = (142,0 x 13 + 226,5 x 54) / 368,6 = 38,1mm
Similarly, it can be shown that the resulting total tensile force acts at a distance of about 178 mm from the
top of the beam. The distance between the resulting compressive and tensile forces is therefore 140 mm
and the expected failure moment of the pre-stressed glulam beam can be calculated as:
MU = 368,6 x 0,140 = 52 kNm
It is worth remarking here that the calculated maximum strain on the compressive face is only about 3.66
. Although the failure strain of structural timber under compressive loading is not listed in any norms
and standards known to the authors, literature studies indicate that it may be close to the better
researched values for small clear specimens, which many authors suggest lies at about 12 . Hence the
compressive face of the timber specimen could readily accommodate larger forces. In other words, the
pre-stressing force in the FRP laminate could be greatly increased before there would be any real danger
of timber compressive failure.

4.3

Research needs

Though the pre-stressing technique seems to have some potential for application to timber structures,
further developments are strongly constrained by the debonding problem. The preliminary tests suggest
that, in spite of the reasonably high shear strength, the current adhesive formulations lack ductile
behaviour. As a consequence, peak shear stresses concentrate in a short length and it is not possible to
mobilize the interface length required to withstand the large forces introduced by pre-stressing.

Pre-Stressing Of Timber

35

COST E34 - WG1: Bonding on Site

Another relevant aspect concerns the evaluation of instantaneous and long-term pre-stress force losses.
It is well documented that a second or even a third pre-stressing operation is required with stresslaminated decks. This should not be possible with pre-stressing wires or laminates glued to the timber.
Besides, the creep behaviours in either the parallel or perpendicular to the grain directions are not
necessarily the same, given the orthotropy of timber.
Finally, the development of a calculation method is an essential condition for the use of the technique. A
linear elastic approach will be adequate to the range of stresses in which both the timber and the prestressing material are expected to work, but the real issue is the development of universally accepted
design rules which take the ductile behaviour of the compressive face of the timber into account.

4.4

References

[1]

Luggin, W. and Bergmeister, K. 'Carbon Fiber Reinforced and Prestressed Timber Beams.' 2nd Int.
PhD Symposium in Civil Engineering, Budapest, Hungary, 1998,

[2]

Galloway, T.L., Fogstad, C., Dolan, C.W. and Pucket, J.A. 'Initial Tests of Kevlar Prestressed
Timber Beams', Nat. Conf. Wood Transport. Structures Gen. Tec. Rep. FPL-GTR-94, Madison, WI,
USA, 1996

[3]

Krahemann, P. and Fontana, M. 'Vorgespannte Holz-Beton-Verbundtrager. Schlussbericht zum


KTI-Projekt Nr. 4617.1', IBK, ETH Zurich, Switzerland, 2002

[4]

Lehmann, M., Properzi, M. and Pichelin, F. 'Prestessed FRP for the in -situ strengthening of timber
structures'. WCTE, Portland, USA, 2006,

[5]

Lehmann, M. 'Renforcement sur site de poutres en bois avec du carbone prcontraint', HSB
Bienne, 2006

[6]

Brunner, M. and Schnriger, M. 'Timber beams strengthened with prestressed fibres:


Delamination'. WCTE, Lahti, Finland, 2004,

[7]

Brunner, M. and Schnriger, M. 'FRP-Prestressed Timber'. FRPRCS-8 Conference, Patras,


Greece, 2007,

[8]

Barroso, D., Negro, J. and Cruz, P. 'Evaluation of the behaviour of prestressed glulam beams'.
CIMAD 04 Timber in Construction, Guimares, Portugal, 2004,

[9]

Kemmsies, M. and Streicher, R. 'Glued timber-steel plate joints'. 1st RILEM Symposium on Timber
Engineering, Stockholm, Sweden, 1999, 389-398

[10]

Brunner, M. and Schnriger, M. 'Towards a future with ductile timber beams'. WCTE, Malaysia,
2002, paper 11.2.3

36

Pre-Stressing Of Timber

COST E34 - WG1: Bonding on Site

GLUED-IN RODS
Serrano E., Steiger R., Lavisci P.

5.1

Overview

Glued-in rods (GiR) are an effective way of producing stiff, high-capacity connections in timber structures.
Glued-in rods have been successfully used for almost 30 years for in-situ repair and/or strengthening of
structures, as well as for new construction works. Relevant research reports are compiled in Section 2.1
of [1]. Extensive information can also be found in the proceedings of Working Commission W18 "Timber
Structures" of the International Council for Research and Innovation in Building and Construction CIB.
GiR are used for column foundations, moment-resisting joints in beams and frame corners and as shear
connectors. Early examples of their use include also the connection of windmill blades made from gluedlaminated timber. Most applications have used the glued-in rod connections with metal bars glued into
softwood. In practice, glulam made from softwood in combination with rods with metric threads are the
most commonly used combinations for new construction, and great experience has been gathered in the
repair and strengthening of beams made of solid timber, both softwood and hardwood, and in connecting
concrete slabs to floor beams with bent steel rebars. For applications where corrosion or weight could be
of concern, the use of pultruded FRP rods is also quite common. Some investigations have also aimed at
the use of reinforcement bars.
Basically all types of adhesive useful for wood bonding have also been tried for glued-in rods, but oneand two-component epoxies, PUR and resorcinol types are those most frequently used in practice.
Specific adhesive products have been formulated to fulfil the needs of GiR joints with timber, which offer
far better performances with respect to strength and durability.

5.2

Design methods

5.2.1

Basic assumptions

Depending on the degree of detailing in the mechanical modelling different parameters influence the
behaviour of the joint. However, these parameters can be influenced by other parameters, normally not
included in the modelling, such as the curing conditions of the adhesive. Other parameters (duration of
load effects, moisture effects and density) can be included in the modelling, but are normally only
accounted for by reduction factors or empirical relations. The parameters of influence are depicted in
Figure 5.1.
Mechanical behaviour

Geometry

Material

Loading and
boundary
conditions

Size of adherends, bond line


thickness, shape (slenderness),
edge distances, number of rods in
group, rod-to-grain angle, etc

Strength, stiffness, plasticity,


creep (DOL), moisture and
temperature dependency, failure
criteria, etc

Tension, compression, axial,


transverse, pull-pull, bending,
mechanical loading, moisture
induced loads, static, dynamic,
load duration, monotonic, cyclic
t

Figure 5.1: Factors affecting the mechanical behaviour of a glued-in-rod joint with timber.
Thus, the following parameters influence the load-bearing capacity of glued-in rods:

Glued-in rods

37

COST E34 - WG1: Bonding on Site

Joint geometry
Ratios of area of wood/adhesive thickness/rod diameter

Absolute size of joint (represented by drill-hole diameter dh and anchorage length ).


Note that the commonly used parameter , which is the slenderness ratio =/dh, does not
represent the absolute size of the joint.

Edge distances

Number of rods

Rod-to-rod distances

Rod-to-grain angle (including unintentional deviations from planned angle due to production
process, definition of a tolerance-range)

Material stiffness
Ratios of rod/adhesive/wood MOE and shear modules

Ratios of MOE/shear modulus for each material (especially important for the wood material, this
being strongly orthotropic)

Material strength
Strength of the wood

Strength of the adhesive (cohesive strength)

Yield and ultimate strength of the steel

Fracture mechanical properties


Stochastic properties, irregularities and coefficient of variation
Loading and boundary conditions
Loading direction in relation to the rod

Boundary conditions (pull-pull, pull compression, pull beam, bending, and pull-pile foundation")
see figure 5.2.

Load duration (static)

Number of load cycles, frequency and amplitude (dynamic)

Other parameters
Wood species

38

Wood density/strength class

Manufacturing practice (curing time and pressure, surface characteristics etc)

Glued-in rods

COST E34 - WG1: Bonding on Site

pull-pull
pull-compression

pull-beam
pull-pile foundation
Figure 5.2: Different types of loading configurations
Data analysis is suggested to make use of the Method of Maximum Likelihood (MLM) to account not only
for the failure loads but for the additional information provided by the survivors and by different failure
modes. Making use of the censored data information leads to higher characteristic values (and thus
economizes the design of timber structures), especially for small samples or samples with a bigger
scatter of data [2].
The influence of the density of the wood has been discussed to some extent previously. One can surely
find several reasons why the wood density should influence the load bearing capacity of glued-in rods.
Using the wood density alone as a factor of influence, does however complicate things since several
other factors are related to it.
In fact, the influence of density on the joint strength has been somewhat debated over the last years. In
the GIROD project [3], no clear influence on joint strength was found. In contrast, according to Swiss test
results on glued-in rods [4, 5] the pull-out strength of glued-in rods glued depends on timber density by c
with c= 0.5 to 0.6 in the case of rods set parallel to the grain and c = 0.25 for rods set perpendicular to the
grain. These contradicting results could be explained by test configurations, wood selection and it is thus
possible that design models would have to take into account this influence, at least for rods set parallel to
the grain.
5.2.2

Mechanics Failure modes and design philosophy

The following failure modes are relevant for a single rod. Although such joints are of little interest in
practice, they form the basis for the research and the design of group of rods.
1. Failure of the rod due to
a. material failure
b. buckling of the rod outside the wood in case of compression loading
2. Pull-out of rod due to
a. adhesive failure at steel-adhesive interface
b. cohesive failure in the adhesive
c. adhesive failure at wood-adhesive interface
d. cohesive failure in wood close to the bond line
3. Pull-out of wood-plug
4. Splitting failure due to
a. short edge distance
5. Tensile failure in the net or gross wood cross section
Glued-in rods

39

COST E34 - WG1: Bonding on Site

6. Splitting of the wood due to stresses perpendicular to the grain at the bottom of the hole (in cases of
rods being pulled axially when inserted perpendicular to the grain)
7. Splitting of the wood perpendicular to the grain at transverse loading
In addition to these failure modes for single-rod connectors, the following can be of interest for multi-rod
joints.
8. Splitting failure due to short rod-to-rod distance
9. Group pull-out / effective load distribution among single rods
In practice, the failure load for each of the above failure modes must be assessed and the design
philosophy set, in order that a chosen failure mode can be ensured. Here, several approaches have been
suggested. One approach could be to ensure that a joint fails by a ductile failure mechanism, such as by
failure in the steel, which of course must allow large plastic strains to develop. By a ductile failure mode of
the joint is meant a joint showing a force-deformation response with large deformations and with constant
or monotonically increasing load capacity until final collapse [6] . Some design codes (e.g. Swisscode SIA
265:2003) prescribes this type of ductile failure, which is favourable for any design case, regardless of
materials in use and regardless of the possibility of seismic actions.
It is worthwhile mentioning that no matter what failure mode is intended the engineer has to be able to
assess all of the above failure modes, in order to perform the design. The adhesive used, in any case,
shall not be the weakest link. There is thus no contradiction in performing large test series for the pull-out
of glued-in rods, even if the practising engineer would choose a failure mode based on plastic failure
taking place in the rod. The failure of the steel rod itself is very easy to predict, given the lower variability
of the material.
5.2.3

Design codes and code proposals

ENV1995-2:1997
This design code included formulae for glued-in rods in Annex A (Normative). The basic expression for
the pull-out of a single rod was according to:

R = d equ i f v

(1)

with the equivalent diameter of the rod, dequ, equalling min(dhole, 1,25 drod), i being the glued-in length
and fv denoting the formal shear strength of the wood-adhesive interface. Here a dependency of the
shear strength on the density of the wood was assumed:
0 .2 1 .5
f v = 1.2 10 3 d equ

(2)

with being the density of the wood.

40

Glued-in rods

COST E34 - WG1: Bonding on Site

Figure 5.3: Minimum distances according to ENV1995-2:1997


The GIROD-approach (EC5-code proposal)
The equation proposed was based on quasi-non-linear fracture mechanics and reads

= f

tanh

(3)

where is the mean shear stress along the rod at joint failure and f the local bond line shear strength.
The parameter is determined by the geometry of the joint and by the stiffness of the adherend materials
and the strength and fracture energy of the bond layer. Experimental calibration and verification was
made by means of a large number of ramp loading pullout tests. As a result of the GIROD project, the
following design formula for the characteristic pull-out load, Rax,k, was originally suggested:

Rax ,k = f ax ,k d equ tanh( ) /

where

= 0.017 / d equ

(4)

with being the glued-in length, dequ the equivalent rod diameter (both in mm) and with the formal shear
strength set to fax,k=5.8 N/mm2. This was further modified in the final suggestion put forward to, CEN, such
that was set to 0.016 and fax,k=5.5 N/mm2.
No influence of density was found in the GIROD-project; so therefore, no such influence was included in
the code proposal.
The final suggestions regarding the minimum distances for rods were according to Figure 5.4: .

Glued-in rods

41

COST E34 - WG1: Bonding on Site

Figure 5.4: Minimum distances according to the GIROD-proposal.


Riberholt CIB-design code proposal
In [7] a CIB-code proposal for glued-in bolts was presented. The axial load resistance was expressed in
terms of two different cases, one for short glued-in lengths, which was a linear relation, and one for longer
glued-in lengths, which included a square root relation. Thus:

F = f ws d
F = f wl d i

for i 200mm

(5)

for i < 200mm

(6)

The density, , is the relative density of the wood (derived form oven dry mass and volume in humid
conditions), and the strength parameters fws and fwl are given as 520 or 650 N/mm1.5 and 37 or 46 N/mm2
for brittle and non-brittle adhesives respectively. According to Riberholt, resorcinol and some epoxies
were classified as being brittle while 2-component PUR was classified as being non-brittle. It is known
nowadays that the adhesive rheology may be adjusted within a wide range, independently of the chemical
nature of the resins used. The minimum edge distances for adhesives adhering to the rod were set
according to the definition in Figure 5.5.

a1
a2
a4
a5

Mutual
Edge
End grain
Side

2d
1.5d
2d
2.5d

Figure 5.5:Minimum edge distances according to CIB-code proposal.


Swisscode SIA 265:2003
Being aware of the fact that the design of a glued-in rod joint is governed strongly by the system used, the
Swiss design code abstains from giving an overall valid design formula but rather sharpens the eye of the
designer to some important facts:

42

Joints with glued-in profiled rods shall be confined to members assigned to Service Classes 1
and 2.

Glued-in rods

COST E34 - WG1: Bonding on Site

It shall be verified that the properties of the adhesive (resin adhesive/mortar) and its bond with
the profiled rod and timber remain constant within the assumed temperature and moisture
ranges over the entire lifespan of the structure.

Attention shall be paid to the impact of changes in timber moisture content and associated
restraint stresses or cracking.

concerning axially loaded rods:

Mechanical force transmission shall be assumed between adhesive and profiled rod; any
adhesive bonding shall not be taken into account.

Uniform load distribution in tension joints with several rods acting together is only achievable
where the individual rod joints exhibit adequate ductility. Otherwise, a non-uniform load
distribution shall be assumed.

The behaviour of glued-in profiled rods may generally be taken to be ductile where the tensile
failure of the rod, involving significant deformation, is certain to occur prior to any other failure
mode.

Type approvals
In some countries, like Sweden and Germany, instead of giving general rules for design, the GiR
connection is dealt with using product type approvals, being exclusively valid for the respective products.
Such type approvals in turn refer to other supporting documents containing documentation from tests on
joints and adhesives, descriptions of manufacturing principles, design formulae and production control
plans. The manufacturing principles describe what materials to use wood, steel and adhesives, and
regulations concerning the production environment, diameter of the holes and adhesive application
principles etc.

5.3

Typical application methods

5.3.1

Manufacturing principles

Several installation principles have been used in the production of glued-in rod assemblies. The principal
difference relates to the diameter of the hole in relation to the diameter of the rod. In some countries, the
most common way of gluing in the rods has been to pour the adhesive into the drilled hole - alternatively,
by applying the adhesive on to the rod and placing the rod into the drilled hole. In an installation where
the diameter of the hole is less than the diameter of the rod, then the rod is screwed into the hole.
Other principles have been used. One principal method includes the drilling of small holes perpendicular
to the diameter of the hole to be used for the rod. One small hole is drilled at the blind end of the main
hole and is used to inject liquid adhesive under pressure, usually using a hand pump. The adhesive fills
the hole and the excess exudes through the other small hole adjacent to the end of the rod, see Figure
5.6.
Adhesive out

Adhesive in

End sealing

Figure 5.6: Injection of adhesive for glued-in rods.


A list of repair configuration, with drawings and extensive description of the manufacturing steps, is
available on the website of the LICONS projects [8].
5.3.2

Examples

Figure 5.7 and Figure 5.8 illustrate some typical applications. Other examples, with calculation reports
and cost estimations, can be found in the LICONS website [8]. In all cases the joint was designed for
wood failure.

Glued-in rods

43

COST E34 - WG1: Bonding on Site

The GSA system [9] is designed in a way that steel failure occurs (instead of wood- or adhesive failure)
in order to achieve a ductile rather than a brittle rupture. This aim is reached by reducing the crosssection of the steel rods within a certain length v. By removing the rods thread within the length v the
anchorage zone is shifted away from the surface more to the interior. Stress concentrations are reduced
and local splitting due to shear forces and stresses perpendicular to the grain is prevented. Although the
drill-hole on its whole length + v is filled with glue, any contribution of the length v to the pull-out
resistance is disregarded due to the lack of mechanical indenting of rod and adhesive in that zone. Tests
on single rods glued in Norway spruce glulam parallel and perpendicular to the grain showed the good
performance of the system [4, 5]. Nominal shear strengths of up to 9 N/mm2 in case of axial pull-out were
reached. The pull-out strength of rods set perpendicular to the grain was found to be 20 to 50% higher
compared to the rods bonded in parallel to the grain. The influence of the length of the glued zone and
the diameter of the drill-hole dh could be summarized well by the so-called slenderness ratio which is
defined by = / dh, and which for 7.5 15 was found to be related to the mean shear strength in
both cases parallel and perpendicular to the grain by an exponent of approximately -.

5.4

Applicable standards

At present there are no EN standards covering glued-in rods. Work is ongoing within CEN/TC193/SC1
Wood adhesives, organized into WG6 Adhesives for glued in rods in timber structures and WG11
Adhesives for on site assembling or restoration of timber structures. WG6 develops a proposal for
testing procedures and requirements for adhesives for glued-in rods, aimed for application in factory
conditions, while WG 11 is working on test methods and acceptance criteria for adhesives to be used on
the building site.

Figure 5.7: Application of glued-in rods for shear connectors in floors (left) and glulam grids (right).

Figure 5.8: Application of glued-in rods for restoration of old beams and trusses.

44

Glued-in rods

COST E34 - WG1: Bonding on Site

5.5

Research needs

Although much research work has been developed on the subject of GiR, future research should be
addressed on the following topics:

5.6

Improving the ductility of the joint.

Optimizing application conditions for an improved quality control both on site and in factory
conditions.

References

[1]

Aicher, S., Reinhardt, H.W. (eds) 'Joints in Timber Structures. Proceedings PRO 22', RILEM
Publications S.A.R.L. (2001)

[2]

Steiger, R. and Khler, J. 'Analysis of censored data - Examples in timber engineering research'.
CIB W18-Meeting 38, Karlsruhe, Germany, 2005

[3]

Bengtsson, C. and Johansson, C.-J. 'GIROD - Glued in rods for timber structures', SP Swedish
National testing and Research Institute, 2002

[4]

Steiger, R., Gehri, E. and Widmann, R. 'Pull-out strength of axially loaded steel rods bonded in
glulam parallel to the grain', Materials and Structures 40 (2007) 69-78

[5]

Widmann, R., Steiger, R. and Gehri, E. 'Pull-out strength of axially loaded steel rods bonded in
glulam
perpendicular
to
the
grain',
Materials
and
Structures
40
(2007)
http://dx.doi.org/10.1617/s11527-006-9214-9

[6]

Gehri, E. 'Ductile behaviour and group effect of glued-in steel rods'. In: Aicher, S.Reinhardt, H.W.
(eds) Joints in Timber Structures. RILEM Publications, Paris, 333-342.

[7]

Riberholt, H. 'Glued bolts in glulam - Proposal for CIB code'. CIB-W18, Parksville, Canada. , 1988,
21-7-2

[8]

Anonymus 'Low Intrusion Conservation Systems for Timber Structures.' 2006, Website:
http://www.licons.org/

[9]

Anonymus 'n'H. Neue Holzbau AG, Switzerland', 2006, Website: http://www.nh.lungern.ch

Glued-in rods

45

COST E34 - WG1: Bonding on Site

ON SITE INTERVENTIONS ON DECAYED BEAM ENDS


Pizzo B., Schober K.U.

6.1

Overview

In the past, wood has been the material used mainly for the construction of horizontal load bearing
structures, for practically all types of buildings. It is a relatively high durable material per se, and it is
known that the simple passing of time is not responsible for any decrease of its mechanical
characteristics [1]. But wood is a biological material and hence it is naturally subjected to decay if some
precautions are not taken for its preservation. Unfortunately, this situation is scarcely encountered in old
buildings and decay of wooden elements has been experienced by engineers and people familiar with
structural restoration. Nevertheless, unless related to original design problems or new structural
requirements, the majority of diseases are concentrated at the end of beams (Figure 6.1), often in direct
contact with the masonry, whereas the rest of the element is in generally sound condition.

Figure 6.1: Two examples of decayed beam-end: on the left the end is already cut for the subsequent
intervention.
In several cases, mainly in the recent past, wooden structures have been integrally or partially
substituted, even without an effective need, due to a superficial technical and /or economical evaluation.
Some exceptions have been made only for a minor amount of historically important buildings. Nowadays,
notably starting from the last three decades, an ever increasing number of designers try to protect the
original wooden structures, even for buildings less interesting from a historical point of view, on the basis
of aesthetical or economical considerations. In fact it is by now accepted that restoring ancient wooden
structures allows applying all the selected designing solutions without substantially increasing the cost of
the intervention due to the use of products and techniques adaptable to the several service conditions.
However, the specific characteristics of wood make the approach to the diagnostic and the design
different from that one requested for the other building materials. The experiences of past interventions
on ancient wooden structures demonstrate that some precautions are necessary in order to assure the
greatest durability of the joints [2]. Some general rules can be described as follows:
1.

A visual and instrumented diagnostic survey must always precede the interventions, in order to
obtain objective data and limit the intervention to the barest essentials;

2.

Adhesives specific for structural intervention on wood have to be used ([3, 4]) observing the
specific suggestions of the adhesives producer;
3.
The wood has to be allowed to exchange moisture with the surrounding ambient, and a certain
thermal stability should be favoured as far as can be possible;
4.
Products should be rationally used and limited to the essential (e.g. the filling of shrinkage cracks
should be preferably avoided except in specific cases).
The interventions on single elements are generally carried out when a beam-end is biologically decayed
due to insect and / or fungal attacks, or when it is physically missing. It is indispensable that the rest of
the beam is in good conditions, which needs to be approved by a diagnostic evaluation, and hence only
46

On Site Interventions On Decayed Beam Ends

COST E34 - WG1: Bonding on Site

the replacement of the attacked part is required. This process can be carried out by using a prosthesis,
made for example of solid wood or glulam, that is connected to the sound part of the beam by means of
connecting elements: they could be either bars (made of steel or GFRP) or plates (often made of steel,
most recently of carbon-fibre). This type of technique avoids the dismantling of the entire wooden
structure, thus considerably limiting the scale of the intervention. Additionally, in this way the original load
distribution in the whole structure is not changed.
Several companies all over the world have developed similar techniques, since the early seventies. The
most diffused have surely been:

the W.E.R. (Wood Epoxy Reinforcement) System, developed in Canada [5];

the Beta-System developed by Renofors originally in The Netherlands and extensively used in
the United Kingdom and in Germany ([6], Figure 6.2 taken from [7]).

Recently, some other techniques, have been used mainly in Italy but also in the U.K., based on a
minimum intervention approach and making use of the design rules described in the design code
ENV 1995-2:1997 [8]. More details are available on the Licons Project web site [9].

Figure 6.2: Schematic steps of Beta-System


technique

6.2

Design methods

In general terms, the repair is different, often more complicated, than a new construction. The design of
the repair interventions on decayed ends of wooden elements has frequently followed an approach based
on the basic construction rules, probably simplistic but evidently successful.
For example, the one used for the Beta-System [6] considers a strain distribution in repaired (or
reinforced) beams as similar to that one of sound wood and moreover requires that the Moment of
Resistance of the sound timber is equal to that one of the rods used for the intervention (Figure 6.3). In
such a way the total area of rods can be calculated. The length of anchorage is calculated simply based
on the resin-to-rods allowable bond stress and on shear strength of sound timber (so, rigid bond between
resin and wood is assumed relative to the design safety value of shear strength for wood).

On Site Interventions On Decayed Beam Ends

47

COST E34 - WG1: Bonding on Site

Figure 6.3: Design method used for the Beta-System


A slightly more sophisticated approach is that of the W.E.R.-System based on the philosophy that the
reinforcement must have a load bearing capacity which is equal to the allowable capacity of the timber
itself, when it is in sound condition. No hypotheses are made on the stress or strain distribution, and the
design of the reinforcement is undertaken with an expected load obtained by considering the allowable
bending stress of timber of the same species of the beam and the dimensions of the real element. The
size of the reinforcement is determined by the existing maximum moment of the beam. The effective
dimensions for the reinforcement (says for example the thickness and the height for a plate, Figure 6.4)
are then stated by verifying it in relation to the horizontal shear and for the crimpling and buckling forces.
Also in this case, the design of the length of the connection between prosthesis and sound wood is based
on the assumption of rigid bond between resin and wood at the allowable value of shear strength for
wood, even if other factors (e.g. wood conditions, load conditions, service environment, etc.) are taken in
account. An interesting feature of this approach is that a fire safety condition enters into the design of the
reinforcement, even if in an empirical way: a minimal distance from the edges of the beam is suggested in
order to ensure a sufficient wood cover.

Figure 6.4: Decayed end reinforcement (W.E.R.-System)


A most recent approach, broadly used for example in Italy, is that one in which connection bars are
calculated with reference to the ENV 1995-2:1997, where the Eurocode 5 is applied in this case not to
new timber structures but to the old. The design of rods is therefore based on the evaluation of their pullout strength (Limit States approach).
According to Lauriola [8] it is convenient to subdivide the procedure in two parts:
1.

48

quantifying of parameters in the beam after the cutting of the element (Fig. 6.5)

On Site Interventions On Decayed Beam Ends

COST E34 - WG1: Bonding on Site

2.
designing of the connection bars.
As for phase 1 the procedure considers that, after the estimation of the loads (distributed load on the
interested beam, constraining force at the support, and shear force at the section of calculation, bending
moment at the section of calculation) the tensile forces on each of the two lower connection bars can be
calculated as:

Nd =

N1 N 2
+
4
2

(1)

where
N1 is the load due to the shear force at the section of calculation

Tcalc

tan

(2)

M calc
hu

N2 is the load due to the bending moment at the section of calculation

(3)

Figure 6.5: Shear forces and bending moment at the beam section
Regarding part 2, the pull-out strength, Fax,Rd, is given by considering the minimum value among the joint
strength related to the:

rupture of the bond between wood and adhesive, F1

tensile cohesive rupture of wood, F2

rupture of the rod, F3

with:

k mod d equ a f v ,k

F1 =

m,w

(4)

F2 =

kmod f t ,0,k Aeff

m,w

(5)

F3 =

f y Ares

(6)

where:
kmod

coefficient of modification (depending on the duration of the accidental load)

dequ

equivalent diameter (minimum of hole diameter and 1.25d)

anchorage length of the rod

fv,k

shear strength at the bond interface


is given by:

1.2 k1.5
, k being the density of wood.
0.2
1000 d equ

(7)

The new version of ENV 1995-2 (2004) presents a different expression for fv,k
ft,0,k

characteristic tensile stress of wood parallel to grain

Aeff

effective area for tensile rupture of wood

fy

characteristic tensile stress of the rods

Ares

resistant area of rods

partial safety coefficient of materials,

m ,w

being the partial safety coefficient for wood

By this way the diameter of the rods can be derived from the expression:
On Site Interventions On Decayed Beam Ends

49

COST E34 - WG1: Bonding on Site

Fax , Rd = min( F1 , F2 , F3 ) >

(8)

Nd

The minimal length of anchorage of rods is given by:

0.4d 2
l a ,min = max
8d

(9)

The effect of adhesive rheology on the performance of the joint is not considered in all these approaches,
and probably a too precautionary value of the shear strength of the adhesive on wood is used. Moreover,
all the calculations are based on the hypothesis of conservation of the plane sections.

6.3

Methods of application

6.3.1

Repair of decayed ends using adhesives and steel reinforcement

Considering the variability of real cases and the variability of their requirements the following systems
illustrated should only be considered as examples, and intended as general models. According to the
more recent orientation and in order to make the interventions as less intrusive as possible, the
connecting elements are confined to the four corners of the beam, and their length is reduced to the
minimum necessary. As already provided in the new version of Eurocode 5, some experimental tests
should preferably be carried out in order to evaluate such optimised length (see chapter 5, Glued-in rods).
The operating procedures depend on the specific requirements of the site, but they are generally of a
common approach, that can be described as follows:

propping of the beams;

removal of the decayed portion of the timber, usually terminated in inclined cut (Figure 6.6);

substitution of the decayed wood with a new section of wood, shaped in such a way that external
dimensions are the same (Figure 6.7);

preparation of internal holes or external grooves between the original wood and the new
introduced wood for the positioning of the connecting elements. (Figure 6.8);

partial filling of the grooves or holes by an adhesive (generally but not necessarily thixotropic)
and introduction of the connecting elements inside the grooves or holes (Figure 6.9);

insertion of a final wood fillet, hiding the grooves (Figure 6.10), not necessary in the case of
holes;

removal of the beam supports after the complete curing of the adhesive.

The final appearance of an intervention is shown in Figure 6.11.

Figure 6.6: Inclined cut of decayed end

50

Figure 6.7: Shaping of integration element

On Site Interventions On Decayed Beam Ends

COST E34 - WG1: Bonding on Site

Figure 6.8: Linking grooves between integration


and original timber

Figure 6.9: Partially filling of the grooves with


adhesive

Figure 6.10: Insertion of a final wood wedge

Figure 6.11: Repaired beam

Several real interventions realised in the recent years in historic monuments in Italy are described and
documented in [10].
6.3.2

Repair of decayed ends using epoxy polymer concrete and GFRP reinforcement (bars and
plates)

Epoxy resins have been used in specific cases to repair timber that has deteriorated from decay or insect
attack and certain structural deficiencies in existing construction. The compressive strength and filling
capabilities of epoxies can aid in repairing timber structures. The tensile, shear, and bond strength of
epoxies as structural wood adhesives are, however, limited and are further subject to variability due to
conditions of use. Mechanical reinforcement should be used in conjunction with epoxies for repairs
intended to develop shear capacity. For structural rehabilitation and restoration the decayed ends were
cut and replaced by timber or polymer concrete where the shear and tension forces resulting from loads
are transmitted by GFRP (glass fibre reinforced plastic) reinforcement with a synthetic resin filling
compound between the new and existing part of the structure (Figure 6.12).

Figure 6.12: Shear reinforcement with GFRP

On Site Interventions On Decayed Beam Ends

51

COST E34 - WG1: Bonding on Site

In some cases the decayed part of the structure will be replaced by a polymer concrete (PC) supplement
using rapid-setting organic polymers as binders and small grade aggregates such as sand or gravel. PC
is increasingly being used as an alternative to cement concrete in many applications. Today, PC is used
for finishing work in cast-in-place applications, precast products, highway pavements, bridge decks and
waste water pipes.
The application of PC in structural rehabilitation and restoration has several advantages

Historically significant structures can be protected by a minimum disturbance of the construction


and minimum replacement of decayed parts.

The appearance of the timber structure will not be changed.

The section design can be easily executed by a timber formwork on the level of the necessary
construction height.

All work can be undertaken from the top side; the suspended ceiling will remain unaffected.

The floor below the reconstruction work can be used with some restrictions

Suitable for inhabited floors and restoration of complicated timber joints

The full load-carrying capacity is achieved after one day.

Research work has been carried out at the Bauhaus-University of Weimar to improve the Beta-technique
for glued-in rods and replacement of the original Beta resin with bonding adhesive PC Compono to
obtain a higher quality of adhesion and the load-carrying capacity. Compared to the Beta-system an
increase of the structural performance of 140% for timber-on-timber supplements and 30% for timber-PCsupplements could be achieved with the new epoxy concrete.
Compression repair
Deterioration caused by decay or insect attack can be repaired with epoxy PC by removing the damaged
wood and filling the void with an epoxy or epoxy / wood fibre mixture, thus restoring strength sufficient for
the particular design. Because the curing of epoxies is typically an exothermic reaction, it is
recommended that larger voids be filled with wood prior to epoxy injection to minimize the amount of
epoxy used. Shrinkage and swelling properties and modulus of elasticity of wood and epoxy differ, and
these differences must be considered in large volume epoxy fillings. Consideration should be given to the
suitability of compression area epoxy repairs if there will be exposure to alternating wetting and drying
cycles. Some epoxy formulations can provide adequate bearing in areas of compression at bolts and
shear connectors and can be used to repair such areas.
Tension repair
The use of epoxy PC is not recommended for the repair of wood stressed in tension parallel to grain,
whether occurring on the tension side of bending members or for wood stressed in axial tension. Radial
reinforcement repairs for curved bending members can be accomplished by placing GFRP rods or rebars
in oversize holes filled with epoxy PC. In this type of repair the radial stresses are transmitted through the
epoxy PC in shear.
Shear repair
If longitudinal checks or splits in the wood occur that require re-establishment of the horizontal shear
strength of bending members, it is possible to use GFRP reinforcement in conjunction with epoxy PC for
this purpose in a similar fashion as described for radial tension repair. For horizontal shear reinforcement,
the FRP reinforcing elements in epoxy-filled holes act to transmit longitudinal shear forces between
adjacent beam sections. In all cases involving the use of mechanical reinforcement in combination with
epoxy adhesive, the net section remaining after the required holes for the steel are drilled should be used
when checking for resulting bending stresses.
Splits are openings which travel from one side of the timber to the opposite side or from one side and
through to the adjacent side. Seasoning checks can develop into splits under certain conditions. Splits
can also be caused by shrinkage of the wood in connection details with widely spaced bolts in tight-fitting
holes. Radial tension in curved members or notches or tapered cuts on the tension side of bending
members may also cause splits. The conditions causing the split should be determined and remedied.
This may require re-design of connection details, providing protection to keep the wood dry, or designing
appropriate radial reinforcement. Tight-fitting fasteners, such as fully-threaded lag screws or epoxy-

52

On Site Interventions On Decayed Beam Ends

COST E34 - WG1: Bonding on Site

bonded GFRP bars, should be used in conjunction with epoxies in repairing splits, especially for long
splits. Epoxies without mechanical reinforcement are not recommended for repairing splits.

6.4

Applicable standards

As already evidenced in the above text, no specific standards exist at European level for the repair of oldtimber beam ends. For such reason, usually designers refer to the general approach described in ENV
1995-2:1997, valid only for the specific case of interventions carried out by using glued-in rods. Designing
with glued-in, not-nailed plates for example is still a sort of challenge assigned to the skills or expertise of
the individual designer.
Nevertheless, some general rules exist for projects of structural interventions in the specific context of old
buildings that belong to the Cultural Heritage. The Italian organization for standardization has developed
a significant activity in this field and produced a standard [11]. A brief summary of the contents of this
standard, with a generic reference to the repair of beam ends, is therefore given in the following.
The standard gives some general indications on the criteria that should be followed for the preliminary
evaluation, the planning and the eventual interventions on wooden structures in old buildings. Some
preliminary verifications and evaluations have to be ever considered, including for example the state of
conservation and the service conditions for each wooden element of the structure by means of a
diagnostic survey. The accuracy of this diagnostic phase depends on the complexity of the structure and
on the requested level of detail. The preliminary procedures include:

historical analysis: if the original construction techniques or working abilities would be preserved,
this information could even suggest the choice of the eventual intervention

characterisation of materials: this may even include strength grading of wood members,
following specific standards

geometric characterization: for relieving the dimensions of structural members but also for
having a geometric survey of joints (carpentry or mechanical devices) and of deformations

characterisation of decay: the possible interactions between decay and service environment
have to be also analysed (for example, the heads of beams inserted in walls)

structural analysis: the evaluation of the static conditions and of the state of stress is considered
as mandatory. In the definition of the structural scheme an addition to the static scheme of single
timber members, including those of the structural units and joints etc have to be considered. In
order to identify the contribution of the timber structure in relation to the static performance of the
building. Only in simple cases, for example the substitution of the decayed beam end is it
possible to refer to models with single timber elements whilst neglecting the overall static
analysis. The preliminary structural analysis includes the verification of the repaired elements of
structure in order to demonstrate that timber replacement can guarantee an acceptable level of
safety.

In respect of the planning of interventions, several aspects of the conservation project need to be
considered, including some recommendations for a caution on the compatibility between the used
materials and for the correct periodic maintenance works (type and space time after intervention have to
be clearly fixed). The fact that repair is not even necessary is also evidenced, and in any case the
concept of reversibility is considered as a tendency which should be preferably pursued.
The project has to respect, whenever possible and if the original configuration is compatible with actual
levels of safety, the pre-existing overall static model. Therefore, for example, joints and constraints have
to be restored by maintaining their original stiffness (unless improper) and longitudinal shrinkage splits not
treated in any way that could prevent the transversal movements of wood. Emphasis is placed on the fact
that important modifications to the timber frame have to undergo an overall static analysis in order to
include the interactions between wooden and non-wooden elements. This induces an adequate
justification of interventions bringing to a significant alteration of the original overall static performance.
Caution in the choice of the relevant safety factors to apply on loads is also recommended, by
considering the importance of the structure and the consequent acceptable level of risk. In this context,
interventions on ends of beams are considered as reintegration of missing timber, and hence the use of
materials other than wood (for example epoxy polymer concrete or grouts) has to be evaluated very
carefully and is not generally recommended. Moreover, adhesives specifically formulated for structural
use with wood are only permitted. Conversely, indirect improving expedients such as facilitating microclimatic exchanges with the surroundings, mainly at the beam ends, are strongly favoured. In case of
On Site Interventions On Decayed Beam Ends

53

COST E34 - WG1: Bonding on Site

using permanently pre-stressed elements, it is necessary for an evaluation of the performance of the
timber structure in terms of long-term duration and, additionally, a plan of the procedures for future
inspections have to be described. A check on the efficacy of a proposed intervention is required before its
implementation, either through experimentation, numerical simulation or by some factors that are of
unique knowledge to the individual case. Such an approach can be also used in order to qualify new
products and/or methods of intervention. Finally, requirements are provided regarding the periodic
inspection of the interventions which have to be clearly specified (by indicating typology, time and space)
and reported in a maintenance plan. The inspections serve the purpose of confirming the compliance with
the specifications of the plan. This includes factors concerning ambient conditions, permanent and/or
moving loads, and a visual examination with the aim of relieving new or accentuated damage on single
elements, with a special attention to the possibility of new biotic attacks. Additionally, the results of these
inspections have to be registered in a specific paragraph of the maintenance plan.

6.5

Research needs

Research in the field of repair of ancient timber structures involves different aspects, most of them
overlapping with the relevant topics of other issues presented in this document (e.g. glued-in rods, timber
reinforcement etc.). Particularly important are:

54

Elaboration of a specific design concept for on site applications of old timber structures, including
consideration of long-term behaviour of connections

Further improvement of structural wood adhesives designed for on site repair with the following
characteristics:

easy-to-use and apply on site

optimised compatibility for applications in wood

curing at room temperature but characterised by values of Tg higher than 60-70C

high values of interfacial fracture toughness, such that a high long-term durability could be
guaranteed

values of shear strength and modulus of elasticity comparable to those of the actual
commercially available products, normally characterised by good creep behaviour.

Evaluation of the quality of bonding for existing and repaired structures by NDT techniques

On Site Interventions On Decayed Beam Ends

COST E34 - WG1: Bonding on Site

6.6

References

[1]

Kuipers, J. 'Effect of age and/or load on timber strength'. CIB W18, Meeting 18, Florence, Italy,
1986,

[2]

Berti, S., Pizzo, B., Lauriola, M., Lavisci, P. and Rizzo, G. 'Consolidamento con adesivi epossidici
delle strutture lignee. Parte 1: Indagine sulla durabilit degli interventi', Arkos 3 (1) (2002) 32-37

[3]

Pizzo, B., Lavisci, P., Macchioni, N. and Misani, C. (2002) 'A new approach to the design of epoxy
adhesives for the restoration of timber structures'. In: (eds) European Timber Buildings as an
Expression of the Technological and Technical Cultures. Elsevier, Heritage series, Paris, 245-254.

[4]

Schober, K.U. and Rautenstrauch, K. 'Upgrading and repair of timber structures with polymer
concrete facing and strengthening'. WCTE, Portland, USA, 2006,

[5]

Stumes, P. 'The W.E.R.-System Manual: structural rehabilitation of deteriorated timber',


Association for Preservation Technology Pubs, 1979

[6]

Waterhouse, M. 'The restoration of structural timbers'. Annual Convention of British Wood


Preserving Association, 1981, 26-31

[7]

Mettem, C.J. and Davis, G. 'Resin bonded repair systems for structural timber', Construction
Repair March/April (1996)

[8]

Lauriola, M. (2002) 'Esempi di calcolo secondo i criteri dellEurocodice 5'. In: Mapei (eds) Il
restauro delle strutture lignee. 19-25.

[9]

Anonymus 'Low Intrusion Conservation Systems for Timber Structures.' 2006, Website:
http://www.licons.org/

[10]

Pizzo, B., Lavisci, P., Macchioni, N. and De Ciechi, M. (2004) 'On site consolidation systems of old
timber structures'. In: (eds) Interaction between Science, Technology and Architecture in Timber
Construction. Elsevier Heritage Series, Paris, 323-352.

[11]

UNI 11138 'Cultural Heritage. Wooden Artefacts. Load bearing structures of buildings: Criteria for
the preliminary evaluation, the design and the execution of interventions' (2004)

On Site Interventions On Decayed Beam Ends

55

COST E34 - WG1: Bonding on Site

REPAIR OF GLUED LAMINATED STRUCTURES


Smedley D.

7.1

Description of glued laminated components

Glued laminated structures by their very nature and method of design obviate the problems that occur
with cut to size timber components which almost certainly will contain knots and variations in grain that
may affect the individual performance of a construction component.
Lengths of glued laminated component(s) are usually made in a continuous factory process involving
individual pre-dried lamella (2% moisture content). Various sizes are used but typically 32 or 35 mm thick
by 200 mm wide. Bonding of sets of lamella can produce beam depths are of up to 1 metre plus.
Thermosetting Adhesives used are based on phenol, resorcinol, melamine, or urea formaldehyde. These
are essentially thin film adhesives and require that the timber to be bonded needs to be smooth and well
regulated. Additionally due to the no gap filling characteristics of these materials pressure is required,
typically 0.8-1.2 MPa and, most important, a temperature in excess of 20 o C.
Glued laminated components have the added advantage that the designer is able to construct large
unsupported spans which are ideal for warehouse areas, swimming pools and other sports facilities
where large open spaces are required.
Designing structures with glulam therefore brings about a number of consistencies and predictabilities
which are very useful for the engineer.
However, as in all construction disciplines the quality of the final construction is dependent not only upon
design but also on the quality and consistency of the initial factory-produced glued laminated sections, in
this case glulam components manufactured and supplied by a specialist factory and the quality of the
work by the erection team. If and when there is a problem with glulam structures suffering from decay,
deterioration or lamella de-bonding, then the causes fall into two specific categories, as explained below.

7.2

Failure types and repair options

7.2.1

Wood decay due to inadequate construction details

Indicated below are just two of many examples of design detail scenarios whereby damage or decay can
occur to a particular type of glued laminated structural component.
There may be some weathering detail on exterior-based components whereby the component is either
not fully protected from the elements and/or the construction detail enhances the possibility of premature
decay to the timber. A particular example is the difficulty of designing for the foot of exterior structural
columns. The column may sit on solid ground which may be concrete or paviours. This basic design may
be theoretically protectively enhanced by encapsulating the foot of the column in a metallic shoe and
base plate. This is a common design for not only semi weather protected patios but also fully exposed
exterior supports and structural columns within leisure areas such as swimming pools. The post may be
supported on a metallic tube in such a configuration that the end grain of the timber is not subject to direct
ground contact. Figure 7.1 illustrates some principles of foot designs and figure 7.2 indicates an example
of both good and bad design.

56

Repair Of Glued Laminated Structures

COST E34 - WG1: Bonding on Site

Rod with/without plate


Concrete floor
Metal shoe
Figure 7.1: Examples of designing the foot of exterior structural columns

Figure 7.2: Typical damaged foot due to water retention in the shoe (left) and example of good water
shedding design (right)
Glued laminated hogged bridge construction has specific weathering detail problems. Firstly, it is
necessary to consider the deck design and run-off. The decking needs to over-sail the hogged span
members in such a way that water is shed away from the top sections of the supports, otherwise in the
medium to long term the top lamellae of the bridge supports suffer and finally decay can reach a stage
where the decking is not fully supported. This problem is further highlighted by the fact that the timber rail
systems are often designed with vertical posts where the horizontal fixing bolt detail gathers moisture and
decay occurs in the area of the fixing and weakens the rail support system, see Figure 7.3 left.

Andre rubber
bearing

Figure 7.3: Typical critical situation in glued laminated bridge constructions

Repair Of Glued Laminated Structures

57

COST E34 - WG1: Bonding on Site

Glued laminated bridges are usually situated in a critical environment in that they are fully exposed to the
weather but also in the case of river crossings they are subject to a micro-climate and constant humidity.
Frequently the owners of these structures do not carry out annual inspections and often individual
components of the structure reach a state of deterioration where either wholesale or partial replacement
becomes necessary. As well as the areas indicated above, there is often a major problem with the
underside of the abutment area which, in the case of river bridges, is subject to constant high humidity
and condensation (Fig. 7.3 right)
Methods of repair include the in-situ wholesale replacement of individual lamellae, see Figure 7.4.

Figure 7.4: Lamella damage and decay (above), removed damage areas (below left) and
moisture matched adhesively bonded replacement lamella (below right)
In really critical situations the bridge structure requires to be supported, the abutment ends completely
removed and a new timber prosthesis installed with enhanced protection to the supporting and bearing
surfaces, see schematic repair method in Figure 7.5 and Figure 7.6.

Figure 7.5

58

Scheme of replacement of abutment ends with new timber prosthesis

Repair Of Glued Laminated Structures

COST E34 - WG1: Bonding on Site

Figure: 7.6 Scheme of expansion joint detail


7.2.2

Glue line delamination and fissures

Unlike standard cut timber sections, glued laminated components are subject to a factory-made regime
and associated quality control. However, due to the nature of manufacture there are occasions when the
application of the adhesive to the individual laminates and the subsequent bonding process may be
inadequate. Despite certain in factory testing, when these deficiencies do occur they generally only
become apparent when the component, in particular a curved stress element of the construction, is put
under load, or the elements take their toll on a wholly exterior based component such as a major curved
beam roof support or bridge beam. If there is an in factory deficient gluing operation then a typical interlaminate adhesion failure occurs relatively quickly, either early in the erection stage or not long after
(weeks or months) following final commissioning of the structure. If these fissures are discovered at an
early stage then there are standard procedures for reinstating the integrity of the bond line by injecting
suitable pressure-less adhesives. Firstly it is necessary to identify the nature and cause of what are
generally called splits, fissures or checks. Occasionally fissures appear within the lamellae due to the
drying out process. On most occasions the appearance of the fissure is at the interface between
individual lamellae and this may be due to a number of causes, adhesive starvation, adhesive failure of
adhesive to the timber surface or, very rarely, cohesive failure for example, within the adhesive bond line.
In practise typical lengths of fissure exhibit a combination of these faults. Figure 7.7 indicates a typical
fissure extending from 0.5 to 3mm in width and 100mm in depth in a glulam structure some 3-4 weeks
after erection. One of the principal causes for inter-laminar failure in glulam structural components
eventually erected in doors is, and has been, due to inadequate storage and protection of the component
between leaving the glulam manufacturer and erection within the building. Unprotected glulam
components intended for interior use are, because of contractual restraints, left unprotected and open to
the weather. In cases where these elements have been erected and the subsequent inter-laminar
deterioration has been discovered it has been possible to re-instate the integrity of a structural beam or
column by first of all injecting the fissures to seal the inter-lamellae. This surface sealing needs to be fully
cured before the secondary and important strengthening operation takes place. This is carried out by
drilling appropriate diameter holes through the full set of lamellar and thereafter injecting the drilled holes,
usually with an epoxy adhesive, into which the metallic or more usually GFRP strengthening rods are
inserted (Figure 7.7).
In Germany, an epoxy resin system was evaluated in an extensive test program [1, 2]. The study included
both strength tests on a large dimension beam as well as laboratory scale tests using mechanical and
climatic parameters. Based on these results the epoxy system (Wevo EP 20) has been recently
certificated for the repair of open glue lines in Germany and has been used widely [3]. Currently a 2-part
polyurethane adhesive [4] has been produced and marketed for the use in the repair of timber structures.
In contrast the adhesive lamella bonding techniques shown in figure 7.7 and the rod tensile strengthening
operation (Figure 7.8) utilise a slow curing adhesive with a pot life in excess of 4-5 hours. This 2 part
aliphatic epoxy adhesive, namely Rotafix CB10TSS, has a modulus of elasticity of 2.4GPa (ductile failure

Repair Of Glued Laminated Structures

59

COST E34 - WG1: Bonding on Site

mode) compared with conventional 2 part epoxy adhesives 4.5-6GPa. This material is thixotropic for ease
of horizontal and upward vertical placement and is ideally suited to bond lines of 1.5 to 6mm thickness

Feeler gauge

Fissure

Figure 7.7: Glulam structure after 2 months with delamination and fissures (left) and glulam by repair
micro injection of separated lamella (right)
There may be other reasons for deterioration of the integrity of the bonded laminates even within the
confines of the interior of a building. On large structures that are subject to environmental controls such
as air conditioning, it is possible for laminated beams to have the moisture content reduced by the drying
out of the lamellae in such a way that either cohesive or adhesive failure occurs in the structural beam
and in cases of extreme temperatures fissures within the lamellae may also be observed. Again, the
general procedure for restoration is to use suitable adhesives for in-situ repair of the splits, shakes and
adhesive bond lines and reinstate the strength of the beam with adhesively-bonded reinstatement of
beam strength using GFRP rods. Figure 7.8 illustrates part of the methodology used in reinstating the
integrity of a set of 18 metre span beams which have a maximum depth of 1,2 metres in the centre and
0,75 metres at the supporting ends. The fissures between the lamellae are first injected as described
previously and illustrated in figure 7.7. The secondary operation is to drill a series of 16 mm diameter
holes at 300 mm centres from the underside of the beam through each and every lamellae, each hole at
o
30 to the horizontal. Finally all of the drilled holes are cleaned, injected with the thixotropic 2-part epoxy
adhesive and the reinforcing rods rotated into the adhesive filled holes.
Prior to the development of this retro-fit strengthening system the solution used was a combination of
removing the existing roof and installing large steel plates. This required screwing the plates on to either
side of the large span laminated timber beams. Therefore, the bonding on site approach is superior in
cost saving. A typical contract value for the conventional repair of 5 X 18 metre span gymnasia beams
was 30.000; internal rod solution approximately in the region of 8.000 to 12.000. An additional
advantage was the aesthetic value of a hidden repair.

60

Repair Of Glued Laminated Structures

COST E34 - WG1: Bonding on Site

Figure 7.8: Preparation of glulam repair by drilling 12mm holes for the GFRP rods and position of
the rods in the reinforced beams.

7.3

Research needs

An indication from researching the current BoS techniques for this piece of work is that further research is
required into the use of environmentally and operative friendly adhesives. Unlike the controlled factory
bonding operations, on site operatives currently may be, or are, exposed to noxious chemicals such as
cleaning solvents and in particular epoxy adhesive formulations which, out of necessity to cope with sub
ambient temperatures, utilise the more carcinogenic multifunctional diluents and aromatic curing agents.

7.4

References

[1]

Radovic, B. and Goth, H. 'Entwicklung und Stand eines Verfahrens zur Sanierung von Fugen im
Brettschichtholz. Teil 1', Bauen mit Holz 9 (1992) 732-742

[2]

Radovic, B. and Goth, H. 'Entwicklung und Stand eines Verfahrens zur Sanierung von Fugen im
Brettschichtholz. Teil 2', Bauen mit Holz 10 (1992) 816-818

[3]

Geidner, T. 'Risse in BS-Binder', Bauen mit Holz 9 (2003) 33-36

[4]

Purbond 'VN 3064. Zweikomponenten-Polyurethan Harz zur Sanierung von tragenden


Holzbauteilen nach DIN 1052/EN301', 2006

Repair Of Glued Laminated Structures

61

COST E34 - WG1: Bonding on Site

FACTORS INFLUENCING BOND PERFORMANCE


Custdio J., Broughton J.G.

8.1

Introduction

Adhesive bonding technology has been studied for several years and it has played an essential role in the
development and growth of the rehabilitation and repairing techniques of timber structures involving
adhesives joints instead of mechanical joints. The widespread use of the former technique can be
attributed to its inherent advantages, such as, the fact that an adhesive joint can distribute the applied
load over the entire bonded area and avoids points of stress concentration, it requires no holes, adds very
little weight to the structure, has a superior fatigue resistance, is suitable for joining dissimilar materials,
and can reduce manufacturing costs. Although adhesives offer a number of advantages, they also suffer
from a number of limitations, such as, the need for careful surface preparation of the adherends is
required, joints must be designed to eliminate peel and cleavage stresses, the adhesive joints have a
limited high-temperature resistance, they are affected by service environments and are difficult to
disassemble for inspection and repair [1].
Rehabilitation and repairing techniques of timber structures involving structural adhesives has proved to
be both structurally efficient and economically competitive when compared with alternative repair
procedures. One of the most promising repair techniques involves the use of epoxy adhesives. Epoxy
adhesives, whilst not ideal, are currently the best generic adhesive type, particularly as a family of
adhesives, for in-situ repair operations. The epoxy adhesive families include a wide variety of products
with quite different properties, and suitable formulations need to be identified for every application. They
do not require high pressure during their application and curing and they can be reasonably tolerant with
regard to glue line thickness variations. They also exhibit strong adhesion to several materials, little or no
shrinkage during cure, dimensional stability after hardening, excellent mechanical resistance and high
resistance to chemical products and water. Epoxy adhesives are available in a great variety of formulated
products, and correspondingly exhibit different properties, in terms of adhesion to wood, viscosity,
reaction and cure time, strength and elasticity. Unlike traditional generic adhesive types, epoxy adhesive
families can be produced to cure under a wide variety of ambient conditions an essential requirement
for in-situ use. Due to their ability to adhere strongly to other materials other than wood, they can also be
used in conjunction with steel plates, rods or fibre-reinforced materials for the consolidation or
reinforcement of wood structures [2, 3].
Although obtaining high initial bond strength is relatively easy, obtaining good bond durability is
comparatively more difficult. The ability of a joint to maintain satisfactory long-term performance, often in
severe environments, is therefore an important requirement of a structural adhesive joint since the joint
has to be able to support design loads, under service conditions, for the planned life time of the structure
[4].
The effects of moisture and heat, especially in combination with an applied stress have a considerable
influence on the durability of structural adhesive joints. In the short term, the mechanical properties of
timber, adhesives and bonded products vary accordingly to the specific environments where they are
applied. In general, all properties decrease as the temperature and moisture levels increase although, if
the yield points of the materials have not been exceeded during service, their load strength and stiffness
may return to their original levels. In the long term, wood adhesives and bonded products will degrade at
a rate determined by the temperature, moisture and level of stress [1, 5].
Well-designed and well-made joints with any of the normally used woodworking adhesives should retain
their mechanical properties indefinitely if the wood moisture content stays below approximately 15% and
if the temperature remains within the range of human comfort. However, when adhesives are exposed
either intermittently or continuously to high temperatures for long periods they will eventually deteriorate
[3, 6].
A number of factors determining the durability of structural adhesive joints have been identified and can
be grouped in three categories: environment, material and stress. The environment is dominated by
temperature and moisture. The material category includes the adherend, the adhesive, and the interphase between them. The last group refers to the stress to which the bond is subjected during or after
exposure, since it affects its lifetime or residual strength. This chapter will focus briefly on each one of the
aforementioned factors, and so it will have the following structure:
62

Factors Influencing Bond Performance

COST E34 - WG1: Bonding on Site

Environment
Moisture content
Temperature

Materials
Surface preparation
Age of surface
Influence of wood species
Treated wood

Stress
Influence of material stiffness on stress
Influence of joint geometry on stress
Joint selection for the assessment of bond performance

8.2

Environment

The environment to which joints are exposed plays an important role in their durability. In particular, when
exposed to the service environment, moisture and heat are the most important factors in determining the
loss of joint strength. The results of natural and accelerated ageing tests have shown that adhesive joints
lose strength on exposure to high temperatures and humidities [7].
8.2.1

Moisture content

Water is often regarded as one of the most worrying agents that may affect the properties of an epoxy
adhesive and the interface between it and timber. Most bonded structures, when exposed to water or
humidity will lose strength over a period of time and in extreme cases will fail completely, but this effect is
limited to extreme conditions [1].
It is almost impossible to keep water out of an adhesive joint. There are a number of possible
mechanisms by which water may enter a bond: (a) diffusion through the adhesive from the exposed
edges; (b) transport along the adhesive/adherend interface; (c) migration via capillary action along cracks
and crazes in the adhesive, and (d) diffusion through permeable adherends, such as composites or wood
[8].
Water ingress into a structural adhesive joint can decrease the bond performance by several reversible
and irreversible mechanisms. The effect of water, at least initially, can be reversible in nature. This is
especially true when corrosion-resistant adherends, good surface preparation and treatment, and
hydrolytically stable adhesives are involved. The loss of joint strength caused by reversible processes,
such as plasticization and swelling of the adhesive and its resulting induced stresses need not be a
problem in itself, provided that any bond degradation has not proceeded too far. If the joint is dried out,
the bond can regain some of its lost strength. However, with time, irreversible processes become more
important. There appears to be a critical water concentration above which the weakening becomes
irreversible. This critical water concentration is dependent on the materials used in the joint
(adhesive/adherends), the temperature and applied stress. At higher moisture levels, the various
irreversible processes that can occur become a serious threat to the long-term durability of the joint. For
instance, water may physically damage the adhesive by hydrolysis, cracking, or crazing. The presence of
water may also produce an unstable adhesive/adherend interface in which will be gradually displaced
from the surface by water. Ultimately, water may penetrate the adherend surface and produce a loss in
the strength of the adherend itself. Stress can enhance the effect of these irreversible processes,
because, internal stresses produced by swelling will add to the existing externally applied stresses on the
joint. The damage produced by these mechanisms will accumulate in the bond, until a threshold state is
reached that produces a rapid propagation of the damage leading to complete failure of the structural joint
[8, 9].
In general, the amount of moisture in the wood combined with the water in the adhesive will greatly affect
wetting, flow, penetration and cure of aqueous wood adhesives. However, research studies carried out in
the last years have proved that this is not always the case for epoxy adhesives [10-12]. The tests carried
out by the previous authors demonstrated that higher moisture contents at the time of bonding, at least up
to 22%, do not compromise the bond strength developed initially at the interface.

Factors Influencing Bond Performance

63

COST E34 - WG1: Bonding on Site

Surveys, recordings and long-term monitoring of repairs undertaken in the last 25 years, regarding resin
methods of repair to timber structures, have given information on the ways in which resin-bonded joints
behave. Even with the greater understanding of this technology, doubts over structural performance, the
effects of moisture and long-term durability, still persists. The effect of high timber moisture contents, both
prior to bonding and following bonding, has minimal effect on the integrity of epoxy-bonded joints. Bond
enhancement is not required to increase initial strength, but where connections may be subjected to
repeated wetting and drying it may be necessary to improve long-term durability of the structural joint with
primers, coupling agents and other surface treatments. All the research undertaken so far and the
apparent absence of reports on problems seem to indicate that efficient, high strength joints can be made
with epoxy resin adhesives. It may be concluded that that the durability of adhesive bonds and resinbonded timber repairs could be comparable to other timber repair solutions, provided that: the joints are
correctly designed using an appropriate structural approach; suitable methods, materials and
specifications are adopted; the work is undertaken by experienced operatives trained in resin methods,
and strict quality control is exercised both off and on site [4, 10, 13].
8.2.2

Temperature

Epoxy adhesives have been used for several years, but, because epoxies soften and lose strength at
relatively low temperatures, epoxy repaired timber exposed to fire could be quite hazardous. The lack of
information relative to its behaviour under the action of fire and under high service temperatures has
restrained the use of these reinforcement systems [14].
At abnormally high or low temperatures the adhesive in a bonded joint may experience internal stresses
that develop when different parts of a joint have different coefficients of thermal expansion. Polymers also
tend to get soft at elevated temperatures and brittle at low temperatures. Long-term exposure to elevated
temperature could cause their oxidation or pyrolysis [8, 15, 16].
The effect of temperature variation on the strength of epoxy-repaired structures can be divided in two
categories. One category considers the effect of temperature changes due to natural environmental
causes. In this category temperature changes from -18 to 65C are reasonable expected variations. The
second major effect to be considered is fire resistance, where extreme temperatures (higher than 200C)
are reached [1, 17].
Todays prevailing belief of practitioners is that since the bond lines in a structural joint are hidden in the
interior of the timber element they experience, due to the low thermal conductivity and specific heat of
wood, considerably lower temperatures compared to ambient conditions. However, some studies have
proved otherwise [2, 18-23].
Aicher et al. [2, 18] and Cruz et al. [19, 20] performed several experiments to assess the influence of the
temperature in the behaviour of rods bonded into glulam or timber with an epoxy adhesive. The
investigations revealed a strong strength reducing influence at elevated temperatures, especially in the
duration-of-load behaviour of the connections, and that a relatively low damping of the maximum
temperatures and pronounced phase shifts occurred. The results showed also an influence of the crosssectional thickness and the speed of temperature variation to the amount of damping and phase shift; the
phase shift and temperature damping were found to be more or less independent of wood species and of
rod type. It was concluded that the service temperature to which the timber structures were exposed
strongly influenced the temperature reached inside the elements and at the glue line. Because of this,
although the service temperature, even in extreme conditions, will not lead to complete timber structural
failure, or even a structural efficiency loss, it can limit the performance and durability of bonded structural
joints where adhesives with low glass transition temperature are used. The maximum temperature level
acting in service on the glue line thus sets the performance requirements on the glass transition
temperature of appropriate adhesives.
More recently, a research study was performed at the EMPA Wood Laboratory where several epoxy
adhesives were subjected to an increased temperature regime [24]. Post-curing the epoxy adhesives at
high temperatures led to an increase in the glass transition temperature, but lowered the shear modulus.
For the adhesives cured at normal conditions, the mean shear modulus of high viscous and filled
products/modifications was significantly higher than that of non modified low viscosity epoxies. The fillers
did not influence the glass transition temperature either for normal or post-cured adhesives, but did
significantly increase the shear modulus of the resins. Over all the tested temperature range, high viscous
or filled adhesives showed greater strengths than that of low viscosity. Up to 50C, high viscous
adhesives/modifications performed better than the low viscous products. The shear strength of all
adhesives at 50C was lower than at ambient temperature. Above a temperature level of ~70C all
products/modifications exhibited losses in shear strength. The effect of post-curing at high temperatures
64

Factors Influencing Bond Performance

COST E34 - WG1: Bonding on Site

didnt improve the strength and stiffness of the joints for both short and long-term loading. It was
concluded that the use of epoxy adhesives to bond carbon-FRP and wood strongly depended on the
temperature experienced. Up to 50C there was no significant loss in strength and stiffness and only
minor creep deformations occurred. However, for higher temperatures, due to the results obtained it is
advisable to test the epoxy adhesives before they can be considered suitable for that specific application.
The effects of temperature variation on joint strength fall into two categories. The shear strength of a joint
is a function of temperature (as already mentioned) and it is also a function of the time during which a
given temperature is sustained or has been sustained. This last relation is important when considering the
case of an epoxy repaired timber structure exposed to fire, or in situations where the bond line would be
subjected to prolonged and repeated exposure to hot environments, e.g., in roof trusses in countries with
hot summers.
Little information is available on epoxy-repaired timber structures exposed to fires, although there has
been some investigation about the effects of fire on wood structures in general. Avent has done one of
the first and few investigations in the field of the fire effect on epoxy-repaired timber [14]. In his study the
behaviour of epoxy repaired timber trusses was evaluated in both experimental and theoretical
investigations. The approach consisted of experimentally determining the epoxy-to-wood bond strength
on small shear block specimens over a wide range of temperatures (21C to 260C); then experimentally
determining the time required for full sized epoxy-repaired joints (both single and double shear specimens
of southern pine) to fail when subjected to a sudden elevated temperature (93C to 538C) and with these
results develop a mathematical model to predict the strength and duration of load for joints exposed to a
sudden high heat condition. The shear tests conducted on both epoxy adhesives have shown that the
shear bond strength decreased significantly as the temperature increases above 66C. With the
experimental data, the mathematical model was developed and a fire rating curve, using a step function
increase to 538C, was obtained for epoxy repaired joints with glue lines at various distances from the
exposed surface. The model used had some assumptions, for instance only the heat effects were
investigated. Heat was assumed to penetrate through the sides and penetration through the edges of the
members was neglected. For example, the two epoxy adhesives used in a 538C temperature increase
on a joint (loaded at its ultimate capacity) resulted in a failure after approximately 10 minutes. Trough
the fire rating curve a 30 minutes fire rating was obtained with the glue line at 76 mm from the surface
and 60 minutes rating was obtained with a glue line at 100 mm from the surface. This investigation was
the first and also the only one found in the literature, concerning the effect of fire on epoxy repaired timber
joints. More investigations are needed to complement and corroborate the conclusions, that suggest that
because of the high insulating qualities of wood, members with interior glue lines will not fail immediately
and that the approximate fire rating prediction can be extend to other woods, joints and repair techniques.
Frangi et al. [21] conducted a comprehensive research project on the shear behaviour of bond lines in
glued laminated timber beams at high temperatures (20-170C), where the performance of seven
different adhesives (five types of one-component polyurethane adhesives, one resorcinol-formaldehyde
adhesive and one epoxy adhesive) was compared. The tests demonstrated that the epoxy adhesive used
for the slab test was sensitive to heat and that the shear strength reduction was greater than the other
adhesives. Between 50C and 60C, the epoxy adhesive bonds, exhibit a change from failure of the
adhesion between adhesive and timber to cohesion failure. The mean residual shear resistance at 60C
was only 60% of the shear resistance at room temperature and at 90C it was only 30% of the shear
resistance at room temperature.
The temperature range between 80C and 100C that has been recommended as safe surface
temperatures for wood exposed for long periods may not be true for all adhesives types [1], because of
the sensitivity of epoxy to heat at temperatures in the range of 30-80C (depending on the epoxy
formulation), the load duration of an epoxy-repaired joint during a fire would depend primarily on the
insulating properties of the wood and the distance of the glue line from the surface [20].
Timber construction can be designed to have good fire resistance for controlling fire spread and resisting
structural collapse in fully developed fires. Light timber frame construction must be protected with fire
resistant sheet materials with careful design of junctions and penetrations. Heavy timber construction has
good fire resistance resulting from the predictable rate of charring and strength of the residual crosssection [1]. Due to the low fire resistance of epoxy connections, the glue line should be protected from
high temperatures by an adequate section of wood member.
Advanced calculation methods may be applied for the determination of the char depth, the development
and distribution of the temperature within structural members (thermal response model) and the
evaluation of the behaviour of the structure or any part of it (structural response model). This informations
can be obtained in the Fire Parts of the Eurocodes. Eurocode 5 [25, 26] applies to the design of buildings
Factors Influencing Bond Performance

65

COST E34 - WG1: Bonding on Site

and civil engineering works in timber or wood-based panels joined together with adhesives or mechanical
fasteners. It complies with the principles and requirements for the safety and serviceability of structures
and the basis of design and verification given in [27]. Eurocode 5 should be used in conjunction with
Eurocode 1 [28-30], Eurocode 2 [31], Eurocode 6 [32], Eurocode 8 [33] and other ENs for construction
products relevant to timber structures.
Although, Eurocode 5 deals with requirements for mechanical resistance, serviceability, durability and fire
resistance of timber structures, epoxy adhesives are not considered. In fact it only alludes briefly to
adhesives, referring only to phenolic and amino plastic adhesives. In the Eurocode 5, only three
requirements regarding structural adhesives are made: (1) the adhesives shall produce joints of such
strength and durability that the integrity of the bond is maintained in the assigned fire resistance period;
(2) PF and amino plastic Type 1 adhesives according to [34] may be used in all service classes to bond
wood to wood, wood to wood-based materials or wood-based materials to wood-based materials, and
also that adhesives according to [35] may be used for plywood and LVL; (3) adhesives which comply with
Type II specification as defined in [34] should only be used in service classes 1 or 2 and not under
prolonged exposure to temperatures in excess of 50C.
In summary, the rehabilitation and repairing techniques of timber structures involving structural adhesives
should always take into consideration the service temperature effect on the adhesive performance, being
necessary cautions in the structural joint design and in the epoxy adhesive selection. Extensive prenormative research and thorough consideration of these effects are still required for the development of
European standards for the evaluation of bond durability and long-term performance under high service
temperature of epoxy adhesives joints. This will enable the effective and safe application of reinforcement
techniques based on the use of structural adhesives, especially in highly demanding situations where the
present lack of knowledge and reliability of these products hinder their widespread.

8.3

Materials

Besides the environmental factors mentioned above, the materials involved in a structural joint also
influence bond strength and durability. The factors in the material category include: the adherend; the
choice of the adhesive; the design of the joint; freedom from contamination (including extractive
contamination) and stability of the adherend surface; the ability of the adhesive to wet the surface, and
entrapment of air/volatiles. The condition of the adhesive/adherend interface becomes a decisive factor
affecting the initial bond strength and the long-term durability of the bonded joint [8, 36, 37].
8.3.1

Surface preparation

Gluing wood is not normally difficult and it is generally possible to obtain a good bond, provided that
adequate surface preparation is undertaken before bonding. The main reasons for preparing the wood
surface before bonding are: (1) to produce a close fit between the adherends; (2) to produce a freshly cut
or planed surface, free of machine marks and other surface irregularities, from extractives and other
contaminants; (3) to produce a mechanically sound surface, without crushing and burnishing it, which
would inhibit adhesive wetting and penetration [4, 38, 39].
Usually, surface preparation involves some form of machining, either to create flat surfaces or specific
mating shapes. In both cases, tight control of the fit between the mating surfaces is necessary to produce
a bond line with uniform thickness. The quality of the surface varies with the type of machining process as
well as with how carefully the process is controlled [38].
During machining, the disruption of the chemical structure may leave residual charge on the timber
surface, also making it very receptive to wetting by the adhesive. This promotes the development of
strong adhesion forces, an important requirement for good bonds. Unfortunately, water, gases,
microscopic dust and dirt particles, extractives in the wood, are also likely candidates for open bond sites.
The longer freshly machined wood is exposed to the atmosphere, the more of these bonding sites will be
taken by gases and pollutants, and less will be available for the adhesive. This, as in the case of many
materials, is why wood loses its wettability over time and the wood surface becomes inactivated [4, 6, 40].
In this situation, although adhesive penetration may still take place, filling the voids on the wood surface,
the adhesive is not molecularly attracted to the wood and as a result a weak bond will occur at the
adhesive/wood interface [36, 38].
Wood cells can be damaged by the machining, usually, involved in the preparation of wood surfaces prior
to gluing. The nature and extent of damage varies with the type and severity of machining. The several
machining processes can be ordered by their increase in damage to the wood: microtoming and hand
66

Factors Influencing Bond Performance

COST E34 - WG1: Bonding on Site

planing (which are not industrially practical); machine joining; machine planning; jointer sawing; and
conventional rip sawing. Machine planing is one of the most common methods, and its effect on both
wood surface and on the strength of glued joints has been extensively investigated, e.g., it has been
shown, that the wood failure after planing with sharp knives occurred at a considerable distance from the
glue line but was found to occur very close to the glue line after planing with dull knives. The surface
quality obtained with this process is strongly affected by the number of cutter marks per inch and the
direction of cut. A quality planed surface can be obtained with carefully sharpened knives, a shallow
depth of cut, and a moderate feed rate. This should produce a clean cut, leaving both early wood and
latewood cells exposed without compression damage, and with a maximum exposure of fresh wood with
the least damage to the surface and subsurface cells [1, 41, 42].
Besides mechanical damage, thermal damage should also be avoided during any surface preparation, for
that matter. When the wood is overheated or over dried, the chance of an inactivated surface is
increased, because heat increases the movement of wood extractives increasing the chance that they will
move to the wood's surface and attach to open sites. In addition, severe heat can actually alter the
chemistry of wood components, destroying available bond sites [43].
If surface inactivation is detected, several actions may be taken to increase the chance of good bonding.
As has already been mentioned, the most effective and traditional method to revitalize wood surfaces for
bonding has been through planing. Another way to re-activate the wood surface is to sand the surface to
be glued, because sanding raises the critical surface tension of wood improving its wettability. However,
sanding is generally not recommended for preparing wood surfaces for bonding. Abrasive and heavy
sanding leads to poor performance of the bonded joint in service, because they can cause severe surface
and subsurface damage to the cells and can destroy the flatness of a surface by preferentially removing
the softer early wood tissue and leaving the harder latewood tissue, with the result that the bond will have
a poor and uneven quality [44]. On the other hand, light hand or machine sanding with a fine grit may be
acceptable as long as not enough wood is removed to affect the flatness and soundness of the surface. A
better result is obtained when surface preparation combines an abrasive stage followed by a chemical
cleaning stage. If timber is oily, e.g., teak, the surface can be wiped with a volatile solvent such as
acetone before applying the adhesive after the solvent has flashed off. If timber is resinous, like Douglas
fir or Oregon pine, a 10% sodium hydroxide solution could be used to wash the face to be glued, followed
with a fresh water wash and the bonding can be performed as soon as the surface dries. Wiping with
solvent also removes sanding dust. Another approach is to incorporate additives in the adhesive to
increase its wettability and penetration on the timber [1, 39, 45].
8.3.2

Age of surface

Because adhesives bond by surface attachment, the physical and chemical condition of the adherends
surface is extremely important to obtain good joint performance. Directly after preparation, all wood
surfaces undergo an inactivation process. The severity and rate of this inactivation depends on wood
species, wood moisture content (MC), temperature level and time of temperature exposure. An
inactivated wood substrate bonds weakly with adhesives, because the inactivation process reduces the
ability of an adhesive to properly wet, flow, penetrate and, in some cases, influence the adhesive cure,
e.g., phenolic and urea adhesives [6].
Wood surfaces can be chemically inactivated by external contamination and self-contamination. External
contamination includes airborne chemical contaminants, oxidation and pyrolysis of wood bonding sites
from over drying or exposure to high temperatures, impregnation with preservatives, fire retardants, and
other chemicals. Self-contamination results from a natural surface inactivation process where the
hydrophobic wood extractives might migrate with time to the wood surface and can there undergo
chemical reactions [13, 45, 46].
To achieve optimal adhesion conditions it is important to control the time-dependence of the inactivation
process. It is recommended that no more time than necessary should be allowed to elapse between final
surfacing and bonding. The prepared surfaces should be kept covered with a clean plastic sheet or other
material to maintain cleanliness prior to the bonding operation. Experimental studies have demonstrated
a substantial reduction in wettability during the first 24 hours after preparing the surfaces of several wood
species. So it is commonly accepted that wood should be surfaced or resurfaced within 24 h before
bonding to remove extractives and other physical and chemical contaminants that interfere with bonding
[46, 47].

Factors Influencing Bond Performance

67

COST E34 - WG1: Bonding on Site

8.3.3

Influence of wood species

Wood is a porous, permeable, hygroscopic, orthotropic, biological composite material of extreme


chemical diversity and physical intricacy. Because of that, its properties vary between species, between
trees within a species, and even within a tree. This variability can lead an adhesive to produce bonded
joints, which will perform inconsistently and with different levels of performance among the several types
of timbers used for structural bonding [6].
Wood extractives exist as complex mixtures of related compounds. Timber can contain a large number of
these mixtures that include groups of chemicals such as tannins, anthocyanins, flavones, catechins,
kinos, lignans and volatile hydrocarbons. Each of these groups can contain many individual chemical
substances. Due to this diversity, countless opportunities exist for chemical reactions between extraneous
materials and the atmosphere, and between these materials and adhesives or other chemicals that may
contact the materials at the wood surface. Extractives can move by several mechanisms, such as
diffusion, migrating with the existing water, or vaporisation. The pH, buffering capacity, and acid content
of the wood can be strongly affected by the type and amount of extractives. The setting or curing
reactions of some adhesives have been reported to be sensitive to these factors. Wood extractives are
then extremely important because of their often undesirable and unpredictable effect upon adhesive
bonding [36].
As a general rule, hardwoods contain more extractives than softwoods. In many species there is a
measurable difference between the gluing characteristics of sapwood and heartwood, possibly due to the
nature and the amounts of extractives found in the heartwood. Most timbers are easy to glue and it is
generally possible to obtain good bonding. Some species have characteristics which make them less
easy to glue, but that does not mean that those timbers should be avoided when adhesives are used, it
merely indicate that a special surface preparation or another adhesive should be used. In addition, most
adhesive manufactures give specialist advice concerning their products and they may also produce
variants of the standard adhesives which have been formulated to overcome specific problems. The
simplest way of minimizing the negative effect of the extractive presence is to remove them before
bonding.
Epoxy resins are capable of reacting with suitable hardeners to from cross-linked matrices of great
strength and with excellent adhesion to a wide range of substrates. This makes them ideally suited to
adhesive applications in which strength under adverse conditions is a requirement. Their unique
characteristics include negligible shrinkage during cure, an open time equal to the usable life, excellent
chemical resistance, ability to bond nonporous substrates and great versatility. This can attributed to the
great variety of epoxy resins that are available with a wide range of molecular weights and to the diversity
of shapes and sizes of hardeners. Epoxy systems can be modified incorporating various additives,
reactive diluents, plasticizers, fillers, solvents and elastomers to achieve specific effects. Epoxy adhesives
specially formulated used alone or in conjunction with primers, can give more reliable bonds on difficult
species, such as very dense, resinous or oily timbers [36].
8.3.4

Treated wood

Depending on the species and the end-use, timber can be treated with chemicals to enhance its
performance against biological agents, fire and weather. Wood can be protected from the attack of decay
fungi, harmful insects, or marine borers by applying chemical preservatives. Timber protection against fire
can be achieved through its impregnation with fire retardants. Dimensional stabilizers and water
repellents are used to improve timber resistance against the weather. All these treatments should be
considered as contaminants as far as adhesion is concerned. Types of surface chemical treatments and
adhesives, and conditions of joint assembly and adhesive cure have varied, interacting, and even strong
effects on the strength and durability of bonds. Certain combinations of these factors can lead to excellent
bonds, despite the interference from chemical treatments [1, 40, 48].
Several studies [49-60] concerning the influence of different preservative treatments, including oil- and
water-borne preservatives on the strength and durability of bonds were conducted on joints bonded with
traditional adhesives, such as phenol-resorcinol-formaldehyde (PRF), resorcinol-formaldehyde (RF),
melamine-formaldehyde (MF), melamine-urea-formaldehyde (MUF), emulsion-isocyanate (EPI).
However, studies about the preservative treatments effects on bonds using modern adhesives like
epoxies are practically inexistent.
Tascioglu et al. [55]conducted a study where the effects of various wood preservative systems (copper
naphthenate, pentachlorophenol, creosote, CCA and CDDC preservatives) and treatment processes (preand post-treatment) on wood/FRP composite adhesive bond durability and shear strength were
evaluated. It was concluded that pre-treatment of individual wood laminations with oil-borne and water68

Factors Influencing Bond Performance

COST E34 - WG1: Bonding on Site

borne preservatives increased the delamination between the wood and FRP composite reinforcement
with respect to untreated control samples. Post-treatments of wood/FRP composite laminated beams with
the same preservative systems had more limited effects than pre-treatments on delamination. In most
cases, for both pre- and post- treatment the wood/FRP delamination was beyond the limit allowed for
wood-wood gluelines by industrial standards for resistance to delamination in accelerated exterior service
tests.
Experiments have shown that, in spite of the negative effects that chemical preservatives constitute to the
adhesion, many types of preservative-treated lumber can still be successfully bonded with specially
formulated adhesives and by closely followed procedures [6, 61].
Vick et al. [56, 60] observed that preservative type, retention level, and length of assembly time
significantly affected the durability of bonds formed by thermosetting wood adhesives, and consequently
the adhesives didnt adhere to treated lumber well enough to consistently meet rigorous industrial
standards for resistance to delamination. He conducted studies at the Forest Products Laboratory that led
to a discovery that trihydroxymethyl resorcinol and its dimers, trimers, and higher oligomers could
enhance structural adhesion of various epoxy formulations to softwood and hardwood species, as well as
to wood in fibre-reinforced plastic composites. The developed hydroxymethylated resorcinol (HMR)
coupling agent greatly improved adhesion to CCA-treated wood when HMR was applied as a dilute
aqueous primer on lumber surfaces before bonding. The HMR physicochemically coupled PR, EP, EPI,
polymeric methylene diphenyl diisocyanate (pMDI), and melamine-urea adhesives to treated wood so that
the bonds could meet rigorous industrial standards for strength and durability like the delamination
requirements defined in ASTM D2559.
Fire retardants increase the hygroscopicity of the timber, and can affect the adhesion of the adhesive to
the timber surface. The most common fire-retarding chemicals used for wood are inorganic salts based
on phosphorous, nitrogen, and boron. The acidity of the fire-retardant-treated wood, particularly at the
elevated temperature and moisture conditions of hot-press curing, can hinder the cure of neutral or
alkaline curing adhesives like RF or PRF. Through the combination of washing the surfaces to be bonded
with certain alkaline aqueous solutions, e.g. 10% caustic solution, followed by a drying stage before
bonding with carefully selected resins of appropriate molecular-size distribution, strong and durable bonds
can be made to certain fire retardant-treated woods [1, 40].
Dimensional-stabilizer treatments are also a source of surface contamination, as they decrease the
hygroscopicity of the timber, which can affect the application of the adhesive to the surface. One of the
most common treatments is acetylation, which is a chemical modification of wood that drastically reduces
moisture-related dimensional changes and rate of bio-deterioration. Resins that are water-soluble and
depend on a hydrophilic substrate for penetration will be less efficient due to the decreased hydrophilic
nature of the cell wall resulting from modification. Adhesion is reduced to varying degrees among
thermoplastic and thermosetting adhesives in proportion to their compatibility with the amount of nonpolar, hydrophobic acetate groups formed in the acetylated wood. Many different types of adhesives have
been studied in the gluing of acetylated lumber, but only room-temperature-curing resorcinolic adhesives
and an acid-catalyzed phenolic hot-press adhesive have been found to develop durable bonds to
acetylated wood. All other wood adhesives develop poorer bonds to acetylated wood than to untreated
wood [1, 36, 40, 62].

8.4

Stress

The main purpose of a structural adhesive joint, or any structural joint for that matter, is to transfer load
from one component (substrate) to another. In the case of adhesive joints, this frequently translates into a
connection between dissimilar substrate materials that have very different material properties. In order to
characterise the mechanical behaviour of an adhesive joint these material properties together with the
means in which the load is transferred, i.e. joint geometry, need to be thoroughly understood. The
following text discusses these factors and how they influence the stress distribution in timber joints and
how this may influence the joints long-term durability. Much of the following discussion has been taken
from a detailed review of Mechanical Testing of Adhesive Joints compiled by Broughton and Hutchinson
[63].
8.4.1

Influence of material stiffness on stress

In transferring load the stiffness of the various materials involved is a key contributor in the development
and magnitude of resultant stress concentrations. In the particular case of timber connections for bonding
on site, this usually involves metallic or fibre-reinforced plastic (FRP) rods, or plates, bonded into either
Factors Influencing Bond Performance

69

COST E34 - WG1: Bonding on Site

new and/or aged timber. Thus, most bonded on site joint configurations consist of a combination of
dissimilar materials exhibiting a large property mismatch between the rod and other materials. The
resultant stress peaks arising from the externally applied load are at a minimum when stiff adherends
(e.g. steel or FRPs) and/or flexible adhesives (e.g. some acrylics and polyurethanes) are involved, but
can become very large when more flexible substrates (e.g. timber) or rigid adhesives (e.g. stiff epoxides)
are present.
8.4.2

Influence of joint geometry on stress

Another key parameter in the evaluation of bond performance is the configuration of the joint itself.
Generally joints, whether used for adhesive characterisation or durability testing, fall into two groups,
where: (1) the test configuration closely matches that of the final connection and (2) the joint geometry is
very different to that of the final connection design (in this case only certain aspects are focused upon). In
the latter case (2), adhesion is often the main focus and is assessed by employing test configurations that
stress the bonded interfaces highly in peel (e.g. 90 or 180 peel tests), cleavage (fracture tests) or direct
tension (e.g. butt test, pull-off test). In the former case (1), lap shear or co-axial bonded joints are often
preferred for bonded timber connections, whereby the loading is supposedly acting in shear, as in the
case of the on site connection.
It is common for designers and engineers to prefer joint configurations from Group 1, as they provide joint
strength data. On the other hand, adhesive technologists and scientists prefer those from Group 2, as
they tend to separate adhesion from joint strength. Indeed, adhesion is often discussed in relation to the
strength of joints, but the force required to fracture a joint is resisted by a complex interaction of internallygenerated stresses and strains. Attempts to use joint strengths as a measure of adhesion can therefore
be extremely misleading.
Single Lap joint
The tensile single lap shear joint is one of the most commonly occurring joints in practice. It is therefore
the joint configuration which is most extensively studied and often used for testing adhesives. The topic of
lap shear joint behaviour has been addressed in detail by Adams et al. [64], where the shear stress
distribution, adh, is described by the linear-elastic Volkersen relationship:

adh = A1 cosh x + A2 sinh x

(1)

A1 and A2 are arbitrary constants, defined by the boundary conditions. For two identical substrates the
parameter 2 is defined by the relationship;

2 Ga 2
=
E ts ta
2

(2)

The key joint parameters are thus:

Adhesive shear modulus, G a

Overlap length,

Substrate modulus,

Substrate thickness(es), ts

Adhesive layer thickness, t a

According to equation (2) decreasing the overlap length or shear modulus of the adhesive, or increasing
the substrate stiffness or adhesive layer thickness, will decrease the shear stress in the adhesive layer,
so providing a stronger joint.
As shown in Figure 8.1, an important aspect of this type of joint is that the applied loading between the
substrates is misaligned, which causes the joint to rotate. In practice this generates high peel stresses
due to the induced bending moment caused by the substrate misalignment, which subsequently lead to
tensile rupture of the adhesive and/or yielding of the substrate. This means that in joints constructed with
relatively stiff adhesives, failure of the shear joint is usually precipitated indirectly in tension. As a result
of this the potential for any increase in joint strength by designing an increased overlap length is limited.
It follows that the adhesive layer is subjected to both shear and tearing (peel) stresses at the ends of the
joint; the substrates, too, are subjected to shearing, stretching and bending. Likewise, in joints involving
70

Factors Influencing Bond Performance

COST E34 - WG1: Bonding on Site

highly anisotropic substrates, such as timber or fibre-reinforced polymer composites, interlaminar


delamination may be experienced as a consequence of high through-thickness stresses. Attempts to
reduce the peel stresses in the single lap joint have led to a proliferation of alternative configurations
including the incorporation of adhesive spew fillets or subtle geometric tapering of the substrates at their
extremities.

Figure 8.1: Shear and peel stress distribution in a single lap shear joint
In the case of timber substrates only, an alternative compressive lap shear test is preferred that
incorporates a large bond area. The main consideration for the development of this test configuration was
the problematic means by which timber can be gripped in a tensile-test without it being crushed. Also, the
use of relatively thick substrates (10mm) and confined boundary conditions imposed by the test fixture
(providing lateral constraint) ensures peel stresses are less pronounced than in the typical tensile case.
This joint configuration has been used by [12], for assessing the performance of joints bonded with epoxy
and PU adhesives and is currently being explored as a quality control (QC) test for on site bonding of
timber [65].
Co-axial joint
A co-axial joint, such as a bonded-in rod, has the desirable effect of eliminating gross joint rotation and
thus reducing peel stresses, though significant shear stresses remain at the ends of the joint (Figure 8.2:
). Additionally, the timber and the adhesive generally have similar modules, resulting in a much more
uniform stress distribution at the adhesive/timber interface. It is for these reasons that, unlike single lap
shear joints, the strength of bonded co-axial joints in timber depends much more significantly upon the
anchorage (overlap) length, as well as the rod diameter (width) and the shear properties of the bondline.
2
Equation (1) can be adapted to suit the coaxial joint configuration, where is substituted for:

2 =

Ga r 1
1
+

t a E w Aw E r Ar

(3)

In this case r is the rod diameter, Er the rod modulus and Ar the cross-sectional area of the rod. Ew is the
modulus of the timber parallel to the applied load and Aw the effective cross-sectional area of the timber.
There are a number of analyses based upon this and similar approaches, which are discussed in more
detail in Chapter 5.
This type of joint configuration has been preferred by many investigators for the assessment of pull-out
strengths under static, cyclic and aged conditions. It clearly falls into Group 1, as it is highly
representative of the actual joint configuration used for on site structural bonding, varying only in rod
diameter and length.
Broughton and Hutchinson [66] modified the joint geometry in Figure 8.2 by including a counter-bore
(internal profile) at the location where the rod enters the timber. Experimental results using epoxy-bonded
steel rods in Glulam and LVL timber showed between 20% and 40% improvements in pull-out loads
respectively. Further work by Broughton and Gardelle [67] developed a simple design rule for estimating
the effective depth of the counter-bore. It was noted that this technique could be particularly important for
Factors Influencing Bond Performance

71

COST E34 - WG1: Bonding on Site

cases where short embedment lengths are necessary and increased embedment length is not
practicable.

Figure 8.2: Shear stress distribution in a co-axial joint at rod/adhesive and timber/adhesive interface
Peel joint
Peel joints are generally employed for one of two reasons. The first is to evaluate the cohesive resistance
of adhesives to peel loading, for applications where this extreme mode of loading is likely - e.g. in bonding
flexible components. The second is to subject adhesive bonds deliberately to their weakest mode of
loading, in order to assess adhesion. Various forms of the peel test exist, with the peel angle as the main
variable. The load measured by the peel test is influenced by the adhesives strain to failure, the bondline
thickness, and the stiffness of the peeling adherends. This type of joint configuration is not generally used
for structural timber applications due to the relative inflexibility of the timber and reinforcing materials
typically employed.
Modified butt joint (pull-off test)
The modified butt joint geometry is depicted in Figure 8.3, i.e. pull-off test. This method is often used as a
quick and easy test for on site QC acceptance of epoxy-bonded FRP-reinforced concrete and steel
structures (using portable hydraulic equipment) to discriminate between poor and good surface
preparation [68]. The joint geometry could readily lend itself to timber substrates just as easily. If the
substrate and adhesive modules differ greatly, as they often do, a complex interaction of axial and radial
bondline stresses will result when the joint is under load. Joint strength increases with a decrease in
bondline thickness, and the presence or absence of a circumferential spew fillet of adhesive affects the
joint performance significantly.

72

Factors Influencing Bond Performance

COST E34 - WG1: Bonding on Site

Figure 8.3: Modified butt joint or pull-off test


Fracture tests
Adhesive joints usually fail by the initiation and propagation of flaws and, since the basic tenet of fracture
mechanics is that the strength of most real solids is governed by the presence of flaws, these theories
have proved to be extremely useful in their application to adhesive joints. The theory of brittle fracture is
as applicable to polymers as it is to metals, but with greater emphasis on the development of a plastic
zone around the tip of the growing crack. One major feature of the fracture mechanics argument is that
the fracture energy, GC , for a given joint, tested at a stated rate and temperature, is independent of the
test geometry employed. Another major feature is that it is applicable equally to interfacial failures as it is
to cohesive failures as long as the failure mode is noted. Three separate fracture modes-of-loading are
possible, though Mode I (tension) and Mode II (shear) are generally accepted as the more important for
adhesive joints (Figure 8.4).

Figure 8.4: Simplified mode I and II fracture energies


Mode I (tensile-opening) is the lowest energy fracture mode for isotropic materials, such that a preinduced crack always propagates along a path normal to the direction of maximum principal stress. In
joint fracture this is not necessarily the case since crack propagation is constrained to the adhesive layer,
and mixed-mode effects may be more important including Mode I and Mode II. Mode I is often used in
Group 2 type testing to assess surface preparation techniques, adhesive toughness, adhesion and
durability, where parallel (or the preferred tapered) cantilever beam configurations are common for
investigating fracture energy, GIC, and fracture toughness, KIC.

Factors Influencing Bond Performance

73

COST E34 - WG1: Bonding on Site

In the case of the co-axial type bonded-in rod configuration Mode II is perceived as the dominant mode of
loading. It can be shown from linear-elastic fracture analysis that the total strain energy, Ua, required to
produce a shear strain is equal to Ga 2 per unit volume, where is the shear strain and is also equal to
v / Ga , where v is the shear resistance of the bondline. Substituting the latter into the former, the brittle
fracture energy of the bondline, Gf, can be determined for a given bondline thickness ta:

Gf =

v2 t a
2Ga

(4)

Unfortunately, due to the empirically derived term v, the fracture energy in this case is a function of joint
geometry. Expressing equation (4) in terms of the adhesive modulus Ga, equation (3) can be re-written to
provide a combined continuum/fracture mechanics expression from equation (1), where:

2 =

v2 r 1

1
+

2 G f E w Aw E r Ar

(5)

This method was originally developed for timber pull-out behaviour by [69], and is currently the preferred
method for determining the pull-out load of rods bonded into timber, using small-scale pull-through tests
to determine , [70].
8.4.3

Joint selection for the assessment of bond performance

As shown in Figure 8.1 to Figure 8.4 the majority of test joint configurations develop bondline stresses
that are far from uniform. Failure loads are therefore related to stress concentrations at the ends (or
perimeter) of the joint and the toughness of the adhesive. Notably, if the bonded area is sufficient to
enable stress redistributions within the adhesive layer, the peak stress concentrations can be further
affected due to changes in the adhesive material properties as a result of environmental exposure, initially
giving apparent increases in joint strength and toughness.
Thus the size of test joints, as well as their geometrical configuration and the type and nature of the
substrates involved, clearly affects bond performance. It generally takes longer for environmental effects
to become apparent in large joints unless initial adhesion is poor. This is because moisture-induced
problems are related to diffusion through the adhesive layer from the end, or perimeter, of the joint.
If the substrates are permeable to water, such is the case with timber, then relatively large bond areas
can be utilised as the time taken to saturate the bondline is somewhat shorter and environmental effects
will become apparent sooner. In the case of impermeable substrates test joints with small bonded areas,
or with large bondline perimeters compared to their bonded areas, are generally preferred for durability
testing. In particular, unstressed or stressed perforated lap shear specimens find favour, not just because
the bonded area is small but also because the joint failure load is little affected by a lack of adhesive in
the central regions due to the majority of load being carried by the ends of the joint. This is an important
issue as, for example, water may plasticize the bondline perimeter, reducing stiffness and relaxing
internal stresses; this may give rise to short-term increases in joint strength [71, 72]. Similarly, waterinduced plasticization, particularly for polyurethane materials, may give rise to dramatic improvements in
joint fracture toughness, provided that there is no significant reduction in adhesion. Alternatively, placing a
joint constructed with, say, a cold-cured epoxy or polyurethane into a hot exposure environment may
further cure the adhesive and increase its strength but reduce its toughness, yet, a test at elevated
temperature may indicate that joint strength is reduced but toughness increased. Any changes in
adhesion may therefore be masked by the changes in overall joint stiffness for some considerable time,
depending upon the joint geometry. The post-curing of joints is thus also an important consideration in
testing where exposure environments will subject joints to temperatures above those at which they were
cured initially.
Transient loading is another factor that might be detrimental to bond performance, again depending upon
the materials, environments and loading conditions involved. Slow cycle fatigue loading generally allows
adhesive materials to creep and is more detrimental than rapid cyclic loading. As a test technique, Mode I
load cycling of double cantilever beams has been shown to be highly discriminating in durability trials
because the crack tip is continuously sharpened during each load cycle and not allowed to blunt as in a
quasi-static test [73].
Relatively simple test procedures may suffice for, say, assessing the relative effectiveness of pretreatments although more elaborate techniques, particularly including induced stress, are often required
for judging the difference between treatments resulting in very good and excellent adhesion. Lap shear,
74

Factors Influencing Bond Performance

COST E34 - WG1: Bonding on Site

tensile butt and peel joint configurations have traditionally dominated the methods adopted in many
industrial sectors. Changes in strength are recorded as a function of exposure time, and the mode and
locus of failure noted. In particular, relatively short experimental timescales are affected both by utilising
small (and even perforated in the case of non-permeable substrates) bond areas, and by stressing the
joints. Such trials may usefully be employed to rank the effectiveness of surface pre-treatments, but the
data produced are not quantitative and extrapolation to real-scale joints is questionable [74].
Stressed durability test methods which can provide data on tensile opening (Mode I) fracture energy, GIC ,
were originally developed in the aerospace industries. Quantitative information on the bonded system can
be obtained in terms of fracture energy versus period of exposure. Qualitatively, the cheap self-stressed
wedge cleavage joint simulates the forces and effects on bondlines in airframes, and has proved to be
correlated with service performance. In other industries, several studies [4, 75, 76] have found the wedge
test and other cantilever beam geometries to be most useful for examining durability phenomena and for
evaluating the effectiveness of pre-treatments, though the choice of experimental load level is very
important. Some workers have even found good correlation between the results of durability trials
involving (perforated) stressed lap joints and wedge cleavage joints [74].

8.5

Research Needs

Essentially, whatever the joint configuration, a bonded joint represents a layered system comprising
different materials and interfaces, all of which respond in different ways to an externally-applied load
and/or change in environmental conditions. As a result, a number of competing mechanisms are taking
place during service life. This review has outlined the current strategies for long-term durability testing
and accelerated ageing, these being:

changes due to environmental exposure tend to be more detectable in smallish joints (small
bond area).

no single joint configuration and test procedure can provide all of the information necessary to
provide a definitive measure of bond performance.

several procedures are required to separate environmental effects on adhesion, the adhesive
and the adherend, much being the result of the resultant stress distribution within the chosen
joint configuration.

temperature is a vital component of the durability assessment procedure of adhesive joints but
excessive temperature durations are not generally considered realistic and should be used with
special caution for accelerated durability evaluation.

stress is a crucial component for the durability determination of adhesive joints but excessive
levels of stress should be avoided and should be used with special caution for accelerated
durability evaluation.

moisture is a critical and especially harsh component for the durability estimation of adhesive
systems and as such is a necessity for any durability testing. High levels of moisture can affect
bonds and, in this case, it is generally accepted that excessive levels can be used to accelerate
degradation.

many joint configurations have been employed, see Figure 8.5 for joint types without particular
reference to substrate limitations or exposure environments, and it is clear that induced peel or
cleavage loading (Group 2 type) provides the more discriminating joint configurations to assess
bond performance, but relating this quantitatively to long-term service life prediction is difficult
without also assessing joints similar to that employed in practice, i.e. Group 1 type.

There is a need for the successful prediction of long-term performance and suitable accelerated ageing
tests throughout the many fields and industries that involve adhesively bonded structures. It is essential
that future research effort for bonded timber connections should encompass the lessons from previous
work and focus upon joint combinations and configurations that will provide realistic data for the specific
materials under consideration.

Factors Influencing Bond Performance

75

COST E34 - WG1: Bonding on Site

Figure 8.5: General suitability of current test specimen geometries to investigate interfacial bond stability
and environmental durability [63]

8.6

References

[1]

Anon. 'Wood Handbook. Wood as an engineering material. General Technical Report FPL-GTR113' Forest Products Laboratory, Madison, 1999

[2]

Aicher, S., Wolf, M. and Dill-Langer, G. 'Heat flow in a glulam joist with a glued-in steel rod
subjected to variable ambient temperature', Otto-Graf-Journal 9 (1998) 185-204

[3]

Plecnick, J.M. and Bresler, B. 'Temperature effects on epoxy adhesives', Journal of the Structural
Division 106 (1980) 99-113

[4]

Mays, C.G. and Hutchinson, A.R. 'Adhesives in Civil Engineering' Cambridge University Press,
1992

[5]

Leafblad, J.L. 'The weathering process: understanding the science when developing a
conservation approach'. Conservation of Historic Wooden Structures, Florence, Italy, 2005, 11-15

[6]

Dunky, M., Pizzi, A. and van Leemput, M. (2002) 'Cost Action E 13. Wood Adhesion and Glued
Products. State of the Art Report - Volume 1. Glued Wood Products'. Luxembourg, European
Communities.

[7]

Kinloch, A.J. 'Durability of structural adhesives' Elsevier Applied Science Publishers, Ltd, 1986

[8]

Pizzi, A. and Mittal, K.L. (2003) 'Handbook of Adhesion Technology. 2nd Rev&Ex edition'. New
York, Marcel Dekker.

[9]

Gledhill, R.A., Shaw, S.J. and Tod, D.A. 'Durability of adhesive-bonded joints employing
organosilane coupling agents', International Journal of Adhesion & Adhesives 10 (3) (1990) 192198

[10]

Broughton, J.G. and Hutchinson, A.R. 'Effect of timber moisture content on bonded-in rods',
Construction and Building Materials 15 (2001) 17-25

[11]

Surleau, J. 'Dure de vie des assemblages par goujons colls sous sollicitations mcaniques et
hygrothermiques', PhD, I'cole Doctorale de Sciences Physiques et de L'Ingnieur, l'Universit
Bordeaux, 2004

76

Factors Influencing Bond Performance

COST E34 - WG1: Bonding on Site

[12]

Wheeler, A.S. and Hutchinson, A.R. 'Resin repairs to timber structures', International Journal of
Adhesion & Adhesives 18 (1998) 1-13

[13]

Back, E.L. 'Oxidative activation of wood surfaces for glue bonding', Forest Products Journal 41
(1991) 30-36

[14]

Avent, R.R. and Issa, C.A. 'Effect of fire on epoxy-repaired timber.' Journal of Structural
Engineering 110 (12) (1984) 2858-2875

[15]

Kelsey, R.G. 'Thermomechanical properties of epoxy mortars', Journal of Materials in Civil


Engineering 5 (2) (1993) 187-196

[16]

Plecnick, J.M. 'Behaviour of epoxy repaired beams under fire', Journal of Structural Engineering
(1986)

[17]

Winandy, J.E., TenWolde, A., Falk, R.H. and Barnes, H.M. 'Temperature histories of plywood roof
sheathing and roof rafters as used in North American light-framed construction. Proceedings'.
WCTE, Shah Alam, Malaysia, 2002, pp. 114-121

[18]

Aicher, S., Kalka, D. and Scherer, R. 'Transient temperature evolution in glulam with hidden and
non-hidden glued-in steel rods. ' Otto-Graf-Journal 13 (2002) 199-213

[19]

Cruz, H., Custdio, J. and Machado, J. 'Efeito da temperatura no desempenho de colas epoxdicas
usadas em reforo estrutural'. 2 Congresso Nacional da Construo, Repensar a Construo Construo 2004, Porto, Portugal, 2004, pp. 907-912

[20]

Cruz, H., Custdio, J. and Machado, J. 'A temperatura de servio elevada e o recobrimento no
desempenho de colas epoxdicas usadas em reforo estrutural', Revista Internacional Construlink
3 (10) (2005) 4-11

[21]

Frangi, A., Fontana, M. and Mischler, A. 'Shear behaviour of bond lines in glued laminated timber
beams at high temperatures', Wood Science Technology Journal 38 (2004) 119-126

[22]

Kemmsies, M. 'Comparison of bond lines in glulam beams adhered with a PRF and one
component PUR after fire exposure'. Symposium on timber engineering, 1999,

[23]

McMurray, M.K. and Amagi, S. 'The effect of time and temperature on flexural creep and fatigue
strength of a silica particle filled epoxy resin', Journal of Materials Sciences 34 (1999) 5927-5936

[24]

Richter, K. and Steiger, R. 'Thermal stability of wood-wood and wood-FRP bonding with
polyurethane and epoxy adhesives', Advanced Engineering Materials 7 (5) (2005) 419-426

[25]

EN 1995-1-1 'Eurocode 5: Design of timber structures. Part 1-1: General - Common rules and rules
for buildings', (2004)

[26]

EN 1995-1-2 'Eurocode 5: Design of timber structures Part 1-2: General rules - Structural fire
design', (2003)

[27]

EN 1990 'Eurocode: Basis of structural design', (2002)

[28]

EN 1991-1-1 'Eurocode 1: Actions on structures. Part 1-1: General actions - Densities, self-weight,
imposed loads for buildings', (2002)

[29]

EN 1991-1-2 'Eurocode 1: Actions on structures. Part 1-2: General actions. Actions on structures
exposed to fire', (2002)

[30]

EN 1991-1-5 'Eurocode 1: Actions on structures. Part 1-5: General actions -Thermal actions',
(2003)

[31]

EN 1992-1-2 'Eurocode 2: Design of concrete structures. Part 1-2: General rules - Structural fire
design', (2004)

[32]

EN 1996-1-2 'Eurocode 6. Design of masonry structures. General rules. Structural fire design',
(2005)

[33]

EN 1998-1 'Eurocode 8: Design of structures for earthquake resistance', (2004)

[34]

EN 301 'Adhesives, phenolic and aminoplastic for load-bearing timber structures - Classification
and performance requirements'(2001)

[35]

EN 314-1 'Plywood. Bonding quality. Part 1: Test methods', (2004)

Factors Influencing Bond Performance

77

COST E34 - WG1: Bonding on Site

[36]

Marra, A. 'Technology of wood bonding. Principles in Practice' Van Nostrand Reinhold, New York,
1992

[37]

Pocius, A.V. 'Adhesion and Adhesives Technology. An Introduction. 2nd Ed' Hanser Publishers,
Munich, 2002

[38]

Adams, R.D. and Wake, W.C. 'Structural Adhesive Joints in Engineering' Elsevier Applied Science
Publishers, 1984

[39]

Minford, J.D. 'Treatise on Adhesion and Adhesives' Marcel Dekker, Inc., 1991

[40]

Davis, G. 'The performance of adhesive systems for structural timbers', International Journal of
Adhesion & Adhesives 17 (1997) 247-255

[41]

Hernndez, R.E. 'Effect of two wood surfacing methods on the gluing properties of sugar maple
and white spruce', Forest Products Journal 44 (7/8) (1994) 63-66

[42]

Singh, A.P. et al. 'The effect of planing on the microscopic structure of Pinus radiate wood cells in
relation to penetration of PVA glue', Holz als Roh- und Werkstoff 60 (2002) 333-341

[43]

Rabiej, R.J., Ramrattan, S.N. and Droll, W.J. 'Glueline shear strength of laser cut wood', Forest
Products Journal 44 (7/8) (1994) 63-66

[44]

Murmanis, L., River, B.H. and Stewart, H. 'Microscopy of abrasive-planed and knife-planed
surfaces in wood-adhesive bonds', Wood and Fiber Science 15 (2) (1983) 102-115

[45]

Sernek, M. et al. 'Comparative analysis of inactivated wood surface', Holzforschung 58 (2004) 2231

[46]

Nussbaum, R.M. 'Natural surface inactivation of Scots pine and Norway spruce evaluated by
contact angle measurements', Holz als Roh- und Werkstoff 57 (1999) 419-424

[47]

Nussbaum, R.M. 'The critical time limit to avoid natural inactivation of spruce surfaces intended for
painting or gluing.' Holz als Roh- und Werkstoff 53 (1995) 384

[48]

Miroy, F., Eymard, P. and Pizzi, A. 'Wood hardening by methoxymethyl melamine', Holz als Rohund Werkstoff 53 (4) (1995) 276

[49]

Dimri, M.P. et al. 'Effect of preservative treatment of veneers on the glue bond in treated plywood',
Journal of Timber Development Association of India 39 (2) (1993) 18-22

[50]

Herzog, B. 'The effect of creosote and copper naphthenate preservative systems on the adhesive
bondlines of FRP-glulam composite beams', Forest Products Journal 54 (10) (2004) 82-90

[51]

Kilmer, R.W. et al. 'Laminating creosote-treated hardwoods', Wood and Fiber Science 30 (2)
(1998) 175-184

[52]

Lisperguer, J.H. and Becker, P.H. 'Strength and durability of phenol-resorcinol-formaldehyde bonds
to CCA-treated radiata pine wood', Forest Products Journal 55 (12) (2005) 113-116

[53]

Manbeck, H.B. and et al. 'Creosote treatment effect on hardwood glulam beam properties', Wood
and Fiber Science 23 (3) (1995) 239-249

[54]

Sellers, T. and Miller, G.D. 'Evaluation of three adhesive systems for CCA-treated lumber', Forest
Products Journal 47 (10) (1997) 73-76

[55]

Triantafillou, T.C. 'Shear reinforcement of wood using FRP materials', Journal for Materials in Civil
Engineering 9 (2) (1997) 65-69

[56]

Vick, C.B. 'Coupling Agent Improves Durability of PRF Bonds to CCA-Treated Southern Pine',
Forest Products Journal 45 (3) (1995) 78-84

[57]

Vick, C.B. 'Enhanced adhesion of melamine-urea and melamine adhesives to CCA-treated


southern pine lumber', Forest Products Journal 47 (7-8) (1997) 83-87

[58]

Vick, C.B. and Christiansen, A.W. 'Cure of phenol-formaldehyde adhesive in the presence of CCAtreated wood by differential scanning calorimetry', Wood and Fiber Science 25 (1) (1993) 77-86

[59]

Vick, C.B., Christiansen, A.W. and Okkonen, E.A. 'Reactivity of Hydroxymethylated Resorcinol
Coupling Agent as it Affects Durability of Epoxy Bonds to Douglas-Fir', Wood and Fiber Science 30
(3) (1998) 312-322

78

Factors Influencing Bond Performance

COST E34 - WG1: Bonding on Site

[60]

Vick, C.B. and Kuster, T.A. 'Mechanical Interlocking of Adhesive Bonds to CCA-Treated Southern
Pine. A Scanning Electron Microscopic stud', Wood and Fiber Science 24 (1) (1992) 36-46

[61]

Tascioglu, C., Goodell, B. and Lopez-Anido, R. 'Bond durability characterization of preservative


treated wood and E-glass/phenolic composite interfaces', Composites Science and Technology 63
(7) (2003) 979-991

[62]

Vick, C.B. et al. 'Structural bonding of acetylated Scandinavian softwoods for exterior lumber
laminates', International Journal of Adhesion & Adhesives 3 (1993) 139-149

[63]

Broughton, J.G. and Hutchinson, A.R. 'Overview of Structural Adhesive Bonding, Adhesive
Systems for Timber Structures Performance Classification & Code ', 2003

[64]

Adams, R.D., Comyn, J. and Wake, W.C. 'Structural Adhesive Joints in Engineering' Chapman &
Hall, London, 1997

[65]

TC 193 WI 'Adhesives, for on site assembling or restoration or timber structures - Comparative


evaluation of the shear strength of adhesive joints and solid wood.CEN/TC 193, Draft Standard',
(2005)

[66]

Broughton, J.G. and Hutchinson, A.R. 'Pull-out behaviour of steel rods bonded into timber',
Materials and Structures 34 (236) (2001) 100-109

[67]

Broughton, J.G. and Gardelle, V. 'An analytical method for the determination of end-of-hole
profiling in bonded-in rod connections for structural timber'. Adhesion 05, Oxford, UK, 2005, pp.
206-209

[68]

TR57 'Strengthening concrete structures using fibre composite materials: acceptance, inspection
and monitoring', The Concrete Society, UK, 2003

[69]

Gustafsson, P.J. 'Analysis of generalized Volkersen joints in terms of non linear fracture
mechanics'. CIB W18-A, 1987, paper 20-18-2

[70]

Gustafsson, P.J., Serrano, E., Aicher, S. and Johansson, C.J. (2001) 'A strength design equation
for glued-in rods'. In: Aicher, S.Reinhardt, H.W. (Eds) Joints in Timber Structures. RILEM
Publications, Paris, 323-332.

[71]

Beevers, A. and Bowditch, M.R. 'Tensile butt or lap shear? - the testing question'. Proc. seminar on
the Durability of Adhesive Joints, Institute of Materials, 1995,

[72]

Bond, A.E., G.C., E., Jones, C.M. and Jones, G.D. 'Finite element analysis of dry and wet butt and
lap joints. Report No 12', 1996

[73]

Kinloch, A.J. 'Adhesion and Adhesives Science and Technology' Chapman and Hall, Ltd., 1987

[74]

Anon. 'Evaluation of published durability data. Report No 2.' DTI-MTS Project 3, 1994

[75]

Taylor, A.C. and Kinloch, A.J. 'The durability performance of structural epoxy adhesives'. 5th
international conference Structural Adhesives in Engineering, Bristol, UK, 1998, pp. 53-58

[76]

Weitzenbck, J.R., Echtermeyer, A.T. and B., H. 'Selection of adhesives and pre-treatments for
structural bonding of aluminium in ships proc'. Int. Conf. On Structural Adhesives in Engineering V,
Bristol, UK., 1998,

Factors Influencing Bond Performance

79

COST E34 - WG1: Bonding on Site

DURABILITY OF HISTORIC STRUCTURES REPAIRED WITH ADHESIVES


Pizzo B.

9.1

Overview

The practice of using adhesives for on site bonding on new structures is quite recent, whereas there is a
relatively long experience of the repairing of ancient wooden structures. Adhesives for on site bonding
were therefore firstly used for interventions on old timber structures. In general, an ever increasing trend
has developed during the last thirty years in using epoxy adhesives for wood consolidation (initial
examples are given in [1-4] but this list is certainly not exhaustive), mainly thanks to the excellent
adaptive capability of the products to the various proposed design solutions and operational conditions,
together with good mechanical performances.

It is possible to use epoxy resins as:

adhesives for gluing of bars, rods or plates, metallic or polymer based, both for the repair and for
reinforcing purposes. In this typology can be also included the transversal bars used to close or
tie longitudinal drying fissures and the use as a binder for mortar applied as gap fillers (see
below);

consolidating resins for impregnation. In such case low viscosity products are used, which can
be injected from specifically drilled holes or directly from the longitudinal shrinkage fissures;

gap fillers for damaged zones or cracks, eventually by simultaneously using some pieces of
wood or other materials.

In any case, while the intervention practice is well established some concerns exist about the long-term
durability of joints. In fact, wood is a hygroscopic and technologically highly complex material because it
varies in size with changing ambient humidity and to different extents depending on the considered
anatomical direction. Additionally, it can be degraded under given conditions by biological agents, mainly
fungi and insects, and some decay has been occasionally observed even for surfaces in contact with
metallic elements, probably because of condensation effects. On the other hand epoxies, both for
chemical composition and physical-mechanical behaviour, differ appreciably from wood, and even if the
short-term performance of bonds is reliable this is not a guarantee for the long-term durability of the joints.
This last factor is also strongly influenced by the effective execution of the joint (geometry, wood species,
gluing procedures, etc.) and therefore it is possible that the same product applied in a different manner
behaves differently during the service life. It appears therefore useful to set up some laboratory
techniques in order to predict the long-term behaviour of the adhesive joints through accelerated ageing
tests [5], although this approach could give information limited to the experimental conditions.
Conversely, a series of examinations on interventions carried out in the past by using adhesives enables
to consider the effective behaviour of such products (and even techniques) when subjected to real service
conditions and therefore to obtain a greater reliability of the evaluations.

9.2

Examination of past interventions

Due to the great variety of procedures of intervention using epoxy adhesives renders it is impracticable to
verify all possible techniques. Moreover, it is certainly difficult to confirm, after a long time, the response
to the original project requirements of the elements or the structures subjected to the intervention. The
most common approach is based on visual inspection; it gives only qualitative hints but allows obtaining
useful information both on the specific intervention typologies and generally on the long-term efficacy to
increase the reliability of the given structure. Moreover, in the case of complex reconstructions making
use of adhesives, it is highly interesting to carry out two evaluation steps, firstly referring to the
intervention typologies and to the use of specific techniques (structural type evaluation), and secondly
relating to the specific durability of the interface between wood and adhesive, that is in the final analysis
to the long-term duration of bonded joints.

80

Durability Of Historic Structures Repaired With Adhesives

COST E34 - WG1: Bonding on Site

9.2.1

Structural type evaluations

Bertolini [6] reports on inspections of about thirty buildings in which epoxy adhesives have been used in
interventions realised approximately fifteen years ago. The examined consolidation techniques can be
grouped in two classes: rehabilitation of beam-heads and consolidation of truss-joints. Concerning the
first group, in which the wooden decayed part is substituted by epoxy polymer concrete connected to the
sound part of the beams with bars, the author observes that "in some cases the contact surfaces between
wood and resin was detached", even though "for such cases the mechanical efficiency seems not
reduced". Moreover, "the loss of adherence between surfaces is caused by twisting or shrinkage, mainly
in presence of elements with a high slope of grain". Concerning the second group, the common typology
is the stiffening of the truss-joints by blocking them with bars connected to each other and drowned in
epoxy-polymer concrete. In such cases more important durability problems have been noticed, and in fact
failures "along some of the elements" have been observed. The author concludes that "the major
problems of these techniques are related to the blocking of the joints which changes the structural
scheme of trusses and the cyclic movement of the wood".
The inspections made by Ceccotti and Marradi [7] are related to interventions carried out seven years ago
in eleven buildings. All examined typologies refer to the restoration of trusses and consider the use of
glulam prostheses for the substitution of decayed beam-heads, the stiffening of joints and the filling of
drying fissures in the wood. The observations give evidence that in the majority of cases no deterioration
and no time dependent loss of functionality has been developed. Concerning the use of glulam
prostheses, the authors observe "a detachment and fracture of the resin or wood due to shrinkage
causing separation of the restored part from the existing wooden element", a problem already evidenced
by [6]. In addition, the same authors observe a loosening of the applied metallic wires and attribute such
occurrences, however inconsequential, to the creep of the wood.
9.2.2

Evaluations regarding the durability of the interface

Rizzo et al. [8] have examined specific interventions dating back ten to twenty-five years. In all the cases
the filling of wood fissures and cracks (Figure 9.1 and 9.2) has been executed, accompanied in one case
by the complete internal injection of wood by pouring the resin through the same cracks (Figure 9.2). In at
least two of the examined buildings the wood cracks have been nailed with transversal bars (Figure 9.3)
probably, as assumed by authors, "in order to try a stitching or blocking of the cracks", mainly when the
slope of grain was particularly anomalous.
The surveys highlighted several problems in the adhesion between wood and epoxy systems: the
movement of wood after the intervention caused alterations at the interface resin-wood, evidenced by the
opening of the fillings (Figure 9.1 and 9.2), even in the cases of internal injections. The authors noticed
that "the new fissures are not wide, but the separation between the two materials is clear". In some cases
the new fissures were within the area of the injected material: the interface between wood and resin
remained intact, but the resin showed cohesive failure (Figure 9.4). Conversely, in one of the examined
buildings it is pointed out that trusses are in a well preserved condition: neither detachments between
wood and resin nor any kind of failures have been noticed.

Figure 9.1: Filling of shrinkage cracks with a thixotropic


product. Hygroscopic movements of timber
induce a clear separation at the wood
/adhesive interface (darker line in the photo).

Durability Of Historic Structures Repaired With Adhesives

Figure 9.2: Filling of drying fissures carried out by


pouring a low viscosity product through them.
Even in this case the interface fails due to the
shrinkage of wood after the intervention.

81

COST E34 - WG1: Bonding on Site

Figure 9.3: Transversal bar used for the sewing of a


shrinkage crack. In this case the
particle-size of the fillers is quite coarse.

Figure 9.4: Filling of a drying fissure with an


adhesive containing fillers with a quite
coarse particle size. The crack-opening
movement of wood induced a cohesive
failure in the adhesive.

Concerning the use of transversal bars, the authors remark "detachments of little influence in
correspondence of the shrinkage fissures" just for the repaired elements, but in one of the buildings "the
extent of such detachments reach a few centimetres in beams showing a significant grain deviation".
Even for both cases in which such technique has been used, authors noticed a deviation from the normal
plane of the two edges of a shrinkage fissure so that both the edges form a sort of lip.
Finally, it is interesting to evidence one of the precautions considered at design level: in one building
showing little detachments "a proper ventilation of the space above the rafters has been assured by
leaving an air space connected to the exterior by means of tubes able to assure a chimney effect".
Similar inspections have been carried out in the United Kingdom on buildings related to a period of 3-7
years after the interventions [9]. Only in one case, where a beam has been externally completely coated
with epoxy products, the conditions were judged totally unacceptable. In the other cases (some of them in
conditions of exposure to the weather) no detachments were observed, neither of fillings nor of glued-in
bars and plates. However, because of the substantially different climatic conditions, these results can not
be directly related to the Mediterranean countries.

9.3

Considerations based on the inspections

All the adhesion problems observed during the surveys were related to different dimensional changes of
wood compared to the adhesives: in fact, unlike the adhesives, wood changes in size according to its
moisture content, mostly in the transverse (tangential and radial) directions. The moisture content of wood
changes in timber elements, particularly in roof structures, during the succession of the seasons and
therefore some stresses can be induced to the interface.
In the examples reported by Rizzo et al. [8] the epoxy adhesives fail to follow the wood deformations and
hence at a certain point, because of the developing tensile stresses orthogonal-to-the-grain, they break in
their weakest points or layers within the bulk adhesive or at the interface to the wood.
The bonding defects observed by Bertolini [6] (who reports even of failures in some timber elements) can
be explained by the blocking of truss-joints, which practically transforms in a rigid joint, which could be
considered to act like a hinge. This implies a qualitative and quantitative modification of the distribution of
stresses, due to the different mechanical system, and therefore the development of new internal stresses
when the wood tries to change in size in consequence of the hygro-thermal variations of the environment
(see for example [10]). This can induce effects not predictable a priori, and ,in the worst cases, leads to
82

Durability Of Historic Structures Repaired With Adhesives

COST E34 - WG1: Bonding on Site

the observed defects. On the other hand, in such cases the problem is not directly related to the use of
epoxy adhesives but to the use of an incorrect intervention technique rather than to the use of the epoxy
adhesive.
The separation of the prostheses from the sound part of the elements has been attributed by Bertolini [6].
to the elevated slope of grain of the beams. Not only this implies to a higher rate of size variation with
changing ambient parameters in the transversal directions but also in the longitudinal one, thus
intensifying the stresses even more. A manifestation of the same occurrence, even if related to different
techniques (stitching of shrinkage fissures), could be the formation of the sort of lips observed by
Rizzo et al. [8].

9.4

Summary and recommendations

In most of the examinations carried out on repaired timber structures in historic monuments (even those
specifically devoted to the evaluation of the durability of the interface) a relatively small amount of data
has been collected specifically concerning the epoxy adhesives applicability for repair work, whereas the
emphasis has been given on the procedures of application and on the execution of interventions.
Therefore, it has been possible to get information of general validity regarding for example the design
criteria but not concerning the use of specific types of adhesive. Some reasons for this lack of knowledge
are:

often the information regarding the applied product is simply missing: the basic chemical nature
and the producer are unknown;

sometimes the producers are known but they modify the chemical composition of the adhesive
components during the years, either for technical or safety reasons, even if the commercial
name remains unchanged;

in some circumstances, the products can be modified on site for example by adding organic or
inorganic fillers, and this can substantially alter their characteristics.

Finally, the influence of the application procedures of such products on the long-term behaviour has to be
considered. It is therefore necessary to further develop experimental laboratory tests for a more rapid and
reliable assessment of the adhesives used in the repair work [11-13].
Based on the inspections reported above the following recommendations are formulated in order to
enhance the durability of bonding on site repair works:

Do not limit the dimensional variations of wood. Obviously, the situation is facilitated when the
structural elements are ventilated considering that the cubature of buildings confers to the
system of high inertia and helps to reduce the effects of ambient moisture exchanges.

Take into account the specific characteristics of the materials during the designing phase of the
intervention.

The execution of joint-blocking (stiffening) interventions (a typical example is that related to the
tie-rafter joint in trusses) has generally to be considered with caution. In fact these operations,
unless they are specifically designed, may change the original mechanical system: for example,
for trusses a hybrid system can be formed combining elements of hinged structures and of rigidly
or semi-rigidly joined frames.

The stitching of drying fissures or cracks carried out by inserting bars perpendicular to the
beam axis is often useless and can potentially induce the development of anomalous stresses.
When this type of intervention is really necessary it is important to consider some expedients
enabling the wood to change in size with ambient moisture fluctuations, thus avoiding the
uncontrollable modification of the stresses distribution.

An injection of large and deep drying fissures or cracks with adhesive is sometimes carried out
to restore a solution of continuity in the wooden tissue, thus improving the shear strength or the
moment of inertia of the element. In such cases one has to consider that very likely, as
evidenced by practically all the inspections, new fissure can originate at the wood-adhesive
interface.

The interface gluing (usually end grain to end grain), as between prostheses and sound part of a
beam, should be avoided or at least it should not be considered as contributing to the overall

Durability Of Historic Structures Repaired With Adhesives

83

COST E34 - WG1: Bonding on Site

mechanical resistance of the element, unless some portions of the longitudinal surfaces are
involved in the adhesive bonding.

Consider systems suited to prevent future mechanical degradation or decay. Care has to be
taken in order to avoid any accidental build-up of dampness for a long period of time, by
facilitating the moisture exchanges with the environment. Some good solutions are:

enabling the ventilation of the heads of the beams, mainly for the elements inserted in the
external walls;
avoiding the direct contact of wood with the masonry, for example by interposing a polymeric
or cork sheet;
facilitating, where possible, the aeration of the upper surfaces of the beams to prevent or at
least to hamper the effects related to the formation or permanence of condensation.
All these considerations, based on visual inspections, ignore the evaluations on the influence of the
adhesives as such on the long term behaviour of joints. Specific laboratory tests evidenced that epoxy
adhesives, even if they are marketed for wood applications, do not have the same behaviour regarding
accelerated ageing tests: whereas some of them have a poor performance, others exhibit high strength
values, sometimes comparable to those of solid wood [12, 13]. Therefore, the application of adhesives
with high mechanical compatibility with wood allows greater assurance of long-term duration of the
interventions.

On the contrary, some practices relatively diffused have to be considered with the greatest caution. For
example, altering the inert quantity or distribution by on site adding fillers is highly imprudent, unless
explicitly authorized by the producer in the technical information sheet of the adhesive. In the same way,
altering the mixing ratio is very dangerous because it can induce a dramatic decrease of the final
mechanical characteristics of the cured product.
Finally great attention has to be paid to the treatment of the gluing surfaces: basically, they must be as
clean as possible and free from dust, oils and fatty residuals.

9.5

Research needs

A comprehensive evaluation of long-term durability of joints in which bonding on site has been carried out
should try to correlate:

the environmental conditions, from the moment of the bonding to the service life of elements
(temperature, relative humidity, etc.)

the bonding conditions (preparation of surfaces, time after mixing, etc.)

the quality of adhesives (type and suitability for use with wood, rheological and mechanical
characteristics, adding of fillers at the moment of the application, etc.).

Unfortunately, most of this information is often missing, not only for the past interventions but also, in too
many cases, for the more recent ones. On the contrary, information on the techniques of execution of the
interventions are often available, sometimes even in detail.
In fact, some of this data is difficult to collect. For example, monitoring of structures during their service
life is expensive and needs a great maintenance effort. However, it could be useful to improve the
reliability of on site repair work by

84

further developing experimental laboratory or on site approaches for a rapid and reliable
assessment of the bonding performance of on site adhesives (both from a chemical and a
mechanical point of view)

provision of a record-book, specific for each construction and intervention in which a series of
basic information is reported, related to the original situation, all executed repair/reinforcing
works and to the maintenance plan for future surveys

directly monitoring selected key parameters at well defined and significant sites (for example
moisture content, temperature, deflection) after the realisation of complex interventions by
economic and highly durable devices and data-logging systems.

Durability Of Historic Structures Repaired With Adhesives

COST E34 - WG1: Bonding on Site

9.6

References

[1]

Amato, R.O., Cigni, G., Chigi, P., Perroneo, U. and Rocchi, P. (eds) New consolidation techniques
of wooden beams. Original in Italian: Nuove tecniche di consolidamento di travi in legno. Roma,
Kappa.1981

[2]

Sanders, P.H., Emkin, L.Z. and Avent, R.R. Epoxy repair of timber roof trusses, Journal of the
Construction Division 103 (3) (1978)

[3]

Stumes, P. Testing the efficiency of wood epoxy reinforcement systems , APT Bulletin VII (3)
(1975)

[4]

Stumes, P. The W.E.R.-System Manual: structural rehabilitation of deteriorated timber,


Association for Preservation Technology Pubs, 1979

[5]

v.d. Kuilen, J.W. and Cruz, H. (2003) Test Methods and Prediction of Performance. In: Dunky, M.,
Pizzi, A.van Leemput, M. (eds) Cost Action E13. State of the art report - Vol 1. Luxembourg, 155157.

[6]

Bertolini Cestari, C. (1994) Techniques of consolidation and their durability: intervention problems
and designing aspects. Original in Italian: Tecniche di consolidamento e loro durabilit: problemi di
intervento e aspetti progettuali. In: (eds) Timber, a structural material from the past to the future.
48th General Workshop RILEM, Trento, Italy.

[7]

Ceccotti, A. and Marradi, P. (1993) New technologies in the interventions of rehabilitation of


ancient wooden trusses: materials and methods. Original in Italian: Nuove tecnologie negli
interventi di recupero delle antiche capriate di legno: materiali e metodi. In: (eds) Acts of the
International Workshop Il recupero degli edifici antichi: manualistica e nuove tecnologie. Clean,
Napoli, Italy.

[8]

Rizzo, G., Pizzo, B., Berti, S., Lavisci, P. and Lauriola, M. Consolidation of wooden structures by
using epoxy adhesives. Part 1: Investigations on durability of interventions. Original in Italian:
Consolidamento con adesivi epossidici delle strutture lignee. Parte 1: Indagine sulla durabilit
degli interventi, Arkos 1 (2002)

[9]

Mettem, C.J., Page, A.V. and Davis, G. Long-term performance of resin bonded systems for
structural timbers: Case studies of repairs in service, TRADA, High Wycombe, UK, 1995

[10]

Borri, A. and Vetturini, R. On the possible mechanical modifications related to the consolidation of
joints of ancient wooden trusses. Original in Italian: Sulle possibili modificazioni statiche
conseguenti al consolidamento dei nodi di antiche capriate lignee, Bollettino degli Ingegneri della
Toscana, Firenze 7-8 (1994)

[11]

Lavisci, P., Berti, S., Pizzo, B., Triboulot, P. and Zanuttini, R. A shear test for structural adhesives
used in the consolidation of old timber, Holz als Roh- und Werkstoff 59 (1-2) (2001) 145-152

[12]

Pizzo, B., Lavisci, P., Misani, C. and Triboulot, P. The compatibility of structural adhesives with
wood, Holz als Roh- und Werkstoff 61 (4) (2003) 288-290

[13]

Pizzo, B., Lavisci, P., Misani, C., Triboulot, P. and Macchioni, N. Measuring the shear strength
ratio of glued joints within the same specimen, Holz als Roh- und Werkstoff 61 (4) (2003) 273-280

Durability Of Historic Structures Repaired With Adhesives

85

COST E34 - WG1: Bonding on Site

10 QUALITY CONTROL ON SITE


Smedley D., Cruz H., Paula R.

10.1

Introduction

Refurbishment, repair or upgrading interventions generally involve the following principal procedures:

An in depth survey of the building components and an assessment of damage, decay and the
possibility of overloaded structural components, and any other items that may require an
intervention

Prior to commencing any work and based on the above assessment provision of all necessary
structural propping and confirmation that adequate access to the fabric of the building is
available

The careful removal of any decayed structural timber

Where necessary the installation of suitable formwork

The measurements of those timber splice replacements which will need to be installed

The use of pre-calculated data and suitable dimensioned drawings to carry out the drilling of
holes and slots in the existing sound timber and where necessary the new pre-prepared
additional component splices

The carrying out of all abrading and cleaning of all surfaces to be bonded

The supply of the pre-measured, pre-cut metallic or FRP components, either rods or plates, and
checking the size, configuration and fit

Supply the pre-calculated volumes of the appropriate adhesives, grouts or mortars, mix and
apply in the correct sequence for fitting the new components

The carrying out of all necessary site test procedures and make the necessary samples, all in
order to validate the quality of the work

Clean the site and remove, after a suitable cure period, all support scaffolding, staging and
access equipment

In order to ensure a complete and satisfactory intervention programme it is necessary to carry out the
appropriate quality control procedures at each and every stage of the intervention.
It is essential that the full complement of appropriate tools and equipment are available to carry out the
work.
It must be ensured that the correct bonding adhesives and grouts suitable for the specific operation are
available.
The use of these materials must be controlled by a competent supervisor and workforce, and their actions
complemented by adequate data and instructions from the material supplier.
Quality Plan: The work should be planned in advance with the aid of a pre-prepared, sequential work
sheet and all necessary dimensioned sketches/drawings. The engineers calculations should be available
in order to check the accuracy of items such as length and diameter of cut, bond length, etc.
The work should be carried out by experienced and qualified (Non Vocational Qualification - NVQ or
similar) operatives.
The described procedure for quality control on site was established within the scope of the European
CRAFT Project LICONS [1].

10.2

Quality control of materials

All materials and products should be accompanied by the manufacturers/suppliers Product Data Sheet,
which should include all the information required for the practical use of the product and in particular the

86

Quality Control On Site

COST E34 - WG1: Bonding on Site

surface preparation required for specific adherends (the indigenous site components and the repair
components introduced to the site).
It is mandatory that a users general safety guide plus a health and safety data sheet (COSHH) are
supplied for each and every site introduced product and these must be available to all operatives and
supervisors prior to the use of the product.
10.2.1 Solid timber splice
The solid timber splices should be of the same species of the timber to be repaired, or compatible in
terms of its mechanical properties (by using, for instance, wood engineered products (WEPs)), durability
and colour. However, when higher durability is required, other timber species of higher natural durability
may be used [2]. Alternatively, low durability timber can be used provided that it is treated with the correct
preservative products [3], which are compatible with the appropriate hazard (use) class [4]. The choice of
timber replacement type and specification remains the choice of the overseeing engineer.
Timber or wood engineered products (WEPs) supplied in whole or in part for the preparation and
installation of timber resin splices (TRSs) should be sound and of suitable quality, free of defects and with
a moisture content (MC) of up to but not exceeding 20% (preferably, between 14% and 16%). However,
irrespective of the above, it is essential that they should match the MC of the parent timber whenever any
bonding process is involved, with a variance of no more than 1.5% from the in situ parent timber
component.
10.2.2 Epoxy adhesives and grouts
These are all characterized by the following.
The products should be supplied in separate, clearly labelled, closed packages in pre-weighed
proportions (stoechiometric quantities), supplied by the manufacturer in batch coded containers.
The pot-life characteristics of individually manufactured products in terms of polymerization (cure) and
exothermic rate will depend upon room temperature, material temperature and volume of installed
material. This data and associated information regarding these specific characteristics are all available on
each manufacturers product data sheets and should be clearly understood by supervisors and operatives
prior to the use of any individual product.
10.2.3 Metallic rods and plates
The metallic reinforcement elements should be of stainless or galvanized steel, or steel adequately
protected against corrosion. In the case of galvanized steel, a separate adhesion promoting primer will be
necessary.
10.2.4 FRP rods and plates
The rods and plates should have good corrosion resistance unaffected by wood tannin and both solvent
and aqueous based timber preservatives and treatments.
The rods and plates (secondary adherends) should be compatible with the bonding products (adhesives,
grouts or mortars).

10.3

Quality control of tools and equipment

It is necessary to be aware of all the criteria regarding the various materials, their characteristics, the tools
and equipment and the best and most appropriate way to handle them for individual and specific
applications.
Tools and equipment used for timber cutting, drilling slots, surface cleaning, mixing and application of
adhesives and grouts shall be suitable for each individual project and kept clean and in good condition.
Site work should be conducted by well informed, trained and certificated operatives, under the
supervision of the Site Manager, all in order to ensure compliance with the specifications of the Quality
Plan. All other procedures regarding deployment and the handling of materials, equipment and tools shall
take into consideration both the Quality Plan and the Health and Safety on site regulations.

Quality Control On Site

87

COST E34 - WG1: Bonding on Site

10.3.1 Timber cutting and drilling slots


The cutting on site and removal of timber components and slot cutting of timber beams should be carried
out using a chainsaw or modified chainsaw and purpose made jig designed specifically to conform to on
site safety regulations.
All work should be expedited with due regard to current statutory Health and Safety regulations. The
electric chainsaw equipment should have a power of at least 1600 watts in order to avoid damage to the
timber. The chain and bar should be in acceptable service condition to ensure the regularity of the cutting
surface. Extreme temperatures produced by ineffective or blunt chains should be avoided as this process
may cause scorching (oxidation) of the timber therefore possibly and subsequently compromising the
timber adhesive bond.
The drilling of holes and the formation of slots will be carried out with the appropriate electric drill, router
and drills all in order to prevent scorching or oxidation of the timber surface.
The cutting of medium and long slots will be carried out with chain saw equipment of standard or modified
type, approved by the supervising engineer.
10.3.2 Surface cleaning
Cleaning of the timber elements should be carried out with an industrial vacuum cleaner or aspirator.
10.3.3 Mixing and application
All ancillary equipment to be used for preparation, mixing and installation of the adhesive or grout should
be that recommended or specified by the material manufacturer and/or supplier of the adhesives and
grouts. In the case of surface application or installation into slots or drilled holes, the correct/approved
injection guns, cartridges, tubing and, if necessary, hand or pumping equipment, should be used as
described by the adhesive manufacturer.
The adhesive products should be applied by spatula to the surfaces to be bonded. Suitable injection
guns with cartridges should be used for the injection into slots and holes. Suitable extension tubes should
be available and used to ensure that deep holes and/or slots are completely filled.
10.3.4 Tool maintenance
All equipment should be cleaned with suitable paper tissue following its use with any adhesive or grout.
This process of cleaning must be carried out prior to cure of the epoxy components. Solvents of any type
should not be introduced into the proximity of the workplace.
In the event that equipment has to be cleaned with solvent it should be removed and cleaned in an area
outside the perimeter of the works. In the event that epoxy materials cure on the surface of tools or
equipment, then mechanical means of removal and cleaning will be required to be carried out outside the
perimeter of the works.
Sufficient back up or supplementary mechanical equipment should be available at all times.

10.4

Quality control on site

10.4.1 Contract Preparation


Prior to commencement of any site work a suitable temporary and approved support and access system
should be installed within the repair area in order to comply with local current Health and Safety
regulations.
Any existing timbers that are to be repaired should be clearly identified and the section size(s) to be
removed agreed with the supervising engineer before proceeding with any cutting operation.
10.4.2 Removal of decayed timber
When a portion or section of a timber element is to be removed then the element should be opened up
from the appropriate face and the decayed timber removed without compromising the sound timber.
In the case of removal of decayed sections, the extremity of the timber to be removed should be cut in
shape and in accordance with the design. The cut line could be vertical or at an angle of 45, depending

88

Quality Control On Site

COST E34 - WG1: Bonding on Site

on the engineers instructions. It is essential that all decayed timber should be removed. In the event that
there is further decay than the design anticipated, then the decay should be removed and the engineer
informed in writing of the new situation.
If future insect or fungal attack is anticipated, suitable preserving treatment should be applied to the
remaining timber - all in accordance with recommendations from the timber preservative manufacture.
This work should only be carried out following complete installation and cure of the new repair in order
that any bond line is not compromised by the potential use of preservation fluids.
10.4.3 Drilling and slotting
In the case of certain hardwoods, e.g. greenheart, it may be necessary to ream any drilled holes or slots
in order to enhance adhesion of the epoxy adhesive by increasing surface area of bond. Cleaning of the
drilled holes, slots and the surfaces to be bonded should be carried out with an industrial vacuum cleaner
or aspirator in order to remove all sawdust or other deleterious substances that may compromise the
bonding process.
The drilling of holes and formation of slots, and the cleaning operation should all be carried out
immediately prior to the application of the adhesive and the installation of a secondary adherend - ideally
within 2 hours.
If surface preservative treatment is specified this should not be brought to the work area until the bonding
process is completed and cured.
10.4.4 Cleaning of bonded surfaces
All mixing and application of the specified adhesive products and grouts should follow the protocol
provided in the Product Data Sheet and Health and Safety information provided by the product
manufacturer and any supplementary information specific to the contract.
Opening of component packages should be carried out immediately prior to mixing, application and
installation. Mixed components should be used within the pot-life indicated by the manufacturer.
It is essential that mixing and final placement takes place in sufficient time to avoid stresses caused by
movement taking place after the cure regime has commenced.
The polymerization of the products depends upon the following factors: volume of mix, initial
temperatures of components prior to mixing and ambient temperature.
If necessary, the components should be pre-placed at a room temperature suitable to their ideal mixing
and application conditions. Prior to their application, it may be necessary to create and/or maintain the
conditions specified by the manufacturer to ensure the ideal curing conditions of the materials. For
instance, suitable temporary thermal foam boards may be used to insulate and contain the temperature
during cold weather conditions.
The injection and adhesive filling of drilled anchor holes should be carried out with suitable injection
extension tubes, all in order that the adhesive completely fills the drilled holes commencing from the rear
of all holes.
Slots should also be filled or injected from the blind base of the slot or cut, all in order to minimize air
pockets which may subsequently compromise the bond characteristics of the joint rod or plate.
An ambient site temperature record should be noted during each and every application and recorded on
the job sheet, as part of the Quality Assurance (QA) system and in order to determine at what stage any
temporary propping or shuttering may be removed.
Removal of any support structure should only take place when it is certain that the adhesive/grout
systems are fully cured.
10.4.5 Mixing
Thixotropic epoxy adhesive (2-component)
The total contents of the smaller container (hardener) are added to that of the larger container (base
resin). The two components are mixed thoroughly, preferably with a purpose provided hand mixing knife
or, in the case of larger component packs, the use of a mechanical drill and paddle. The injection
application cartridges are fitted using the filler.

Quality Control On Site

89

COST E34 - WG1: Bonding on Site

Structural Timber grout (3-component)


The liquid contents of the hardener (curing agent) container should be added to the liquid contents of the
base container and thoroughly mixed and then poured into the outer plastic mixing pail. The contents of
the powder filler bag are added and the liquid and powder mixed with the paddle and mechanical drill
assembly to form a flowable grout.
Any fissures or cracks in the timber should be sealed with a two part solvent free rapid curing epoxy putty.
The mixed grout may then be poured or pumped in situ in such a way that there is a continuous head of
flowable grout mix to completely fill any specified voids in the timber component.
10.4.6 Installation of secondary adherends
The number and correct dimension of rods or plates and their position for insertion should be installed in
accordance with the engineers dimensioned design sketches and instructions.
The glass fibre rods/plates, except those with a polyurethane matrix, should be abraded with the correct
grade of carborundum paper and subsequently wiped with acetone immediately prior to installation.
The rods/plates should be installed subsequent to this treatment and before the surface is compromised
by any contamination.
If necessary, abraded and solvent wiped components may be stored by wrapping in clingfilm to prevent
contamination or in the case of grit blasted steel rust oxidation.
Secondary adherend components should only be unwrapped immediately prior to use. This treatment is
particularly successful for protecting grit-blasted steel plates, which may, depending on the particular site
humidity conditions, be subject to premature flash rusting.
The rods/plates should be installed immediately after the injection of the thixotropic epoxy adhesive. The
holes or slots should be filled to the pre-calculated amount of adhesive which, when added to by the
insertion of the volume effect of the rod or plate, ensures that the hole or slot will be completely filled and
the secondary adherend totally encapsulated in the adhesive. It is essential to have available a slight
excess of adhesive to guarantee that the drilled hole or slot is full. Any excess may be removed.
The reinforcement elements (rods/plates) should then be anchored in sound timber.
10.4.7 Manufacture of the solid timber splice (TRS)
The solid timber splice (TRS) should be manufactured in accordance with the project specification
information or in accordance with any subsequent instructions from the supervising engineer. However, it
is necessary to know in advance the maximum length of the new piece of timber to be installed. After
manufacture, it will be impossible to extend the length of the pre-made unit. If there is any doubt about the
eventual amount of existing timber to be removed, then the pre-made TRS unit should be made in excess
of the required size. It may then be cut or profiled to size in situ.
If the TRS is pre-treated with a solvent based preservative, it should be ensured that all volatiles have
been dispersed prior to use with any adhesive. For water based preservative treated timber the timber
must be sufficiently dry prior to use with any adhesive. As a guide, the time in days before adhesives may
be used on the substrate is equal to the manufacturers stated re-entry period for both aqueous and
emulsion based preservatives. This is normally the equivalent of 7 days at 20 degrees C and 50% RH.
The foregoing relates to the timber treatment which occurs prior to commencement of the bonding on site
assembly.
If surface preservative in situ treatment is specified, the preservative should be applied only after the
splice is assembled and the adhesives have cured. Normally this takes 7 days assuming the temperature
is not less than 20 degrees C.
When necessary, the rods/plates should be fixed according to the description in item 10.4.6 of this
chapter.
10.4.8 Quality control for generic repair systems
The following works should be completed, in the correct sequence and following the principles indicated
in the corresponding paragraphs.

90

Quality Control On Site

Inject thixotropic epoxy adhesive into holes/slots (4.4, 4.5)

Place rods / plates in slots (4.6)

Fit formwork and seal it to the beam with bonding paste (4.5)

Seal any cracks in existing timber with bonding paste, if necessary (4.4, 4.5)

Pour epoxy timber grout (4.4, 4.5)

Remove formwork after curing of grout

BEAM REINFORCEMENT
Modified flitch (reinforcement of beams
with rods and epoxy grout by the top
slot method)
Reinforcement of beams with rods and
thixotropic epoxy adhesive on bottom
slot of beam
Reinforcement of beams with plates
and thixotropic epoxy adhesive
Repair of beams above decorative
ceilings (side slot)
CORRECTION OF FISSURES IN
SOLID TIMBER / DELAMINATION
OF GLULAM BEAMS
TRUSS CONSOLIDATION
MODIFIED TENON
BONDED-IN DOWEL JOINTS
BONDED-IN PLATE JOINTS

Quality Control On Site

Pour epoxy timber grout into slot (4.4, 4.5)

Place rods (4.6)

Fit solid timber splice and wedge to level with existing timber (4.5)

Place rods / plates in slots (4.6)

Drill holes at the end of beam (4.3)

Inject thixotropic epoxy adhesive into holes/slots (4.4, 4.5)

Cut off decayed end (4.2)

Drill holes on element to be repaired / elements to be connected (4.3)

Drill slot on the beam (4.3)

Manufacture of solid timber splice and drilling up slot (4.7)

Manufacture of solid timber splice and fixing of rods (4.7)

Preparatory works (4.1)


DECAYED BEAM END REPAIR
Beam end repair using epoxy timber
grout with removable formwork
Beam end repair using epoxy timber
grout with permanent formwork
Beam end repair using solid timber
splice with rods and epoxy timber
grout on top slot of the beam
Beam end repair using solid timber
splice with rods and thixotropic epoxy
adhesive on side slots of the beam
Repair of timber structures using solid
timber splice with rods and epoxy
timber grout on top slot of the splice
Repair of timber structures using solid
timber splice with thixotropic epoxy
adhesive on side slots of the splice
Repair of timber structures using solid
timber splice with epoxy timber grout
on top slot of the beam and solid
timber splice
Repair of timber structures using solid
timber splice with rods and thixotropic
epoxy adhesive on side slots of beam
and solid timber splice

Apply bonding paste to bonding surfaces of solid timber splice and beam (4.4, 4.5)

COST E34 - WG1: Bonding on Site

91

COST E34 - WG1: Bonding on Site

10.4.9 Health and safety


All works should be executed in accordance with current Health and Safety regulations and with local
regulations and the contractors pre-submitted Health and Safety Plan (HSP).
Health and Safety procedures provided by the product manufacturer must be observed.
When handling any epoxy resin materials, operatives must wear appropriate goggles, rubber gloves and
uniform (smock, overall).
The work place must be well ventilated. Smoking and eating are not allowed within the work or mixing
area.

10.5

Quality plan

10.5.1 Responsibilities and records


The Job Supervisor, under the supervision of the Site Manager, should ensure compliance with the
specifications of any Quality Plan (QP) and the training/informing of the site staff, under his/her authority,
regarding the procedures to be deployed and the handling of materials, equipment and tools.
The Job Supervisor (or his/her designated substitute) shall ensure that there is adequate supporting
documentation and the regular submission to the Site Manager of completed records (as indicated in this
QP), materials test certificates, delivery slips, etc. All records should be dated and shall clearly identify the
Job site, responsible person and/or Job Supervisor.
He/she is also responsible for requesting any clarification, information and resources as required.
10.5.2 Reception of materials
The arrival and removal of all materials on the work site should be recorded. An example of a typical
record sheet is given below.
Form A Record of materials
Designation

(*)

Code(*) Lot(**) Validity(*)

Entry

Date Quant.

Exit
Delivery
Delivery
Date Quant.
Note
Note

Stock

Unit.
price

As indicated by the supplier


Suppliers lot or sequential number attributed in job site

(**)

The Job Supervisor should ensure that all materials received on site are subject to inspection upon
reception: verification of delivery note and goods bills delivered to ensure that the correct products have
been delivered to meet project specifications. These checks should include integrity of packaging, where
applicable, as well as recording the materials expiry date. All information should be recorded in the
project day book.
10.5.3 Inspections and tests
Additional tests and/or inspections should be carried out and recorded. Examples of suitable forms are
given as forms B to F below.
Samples are collected/completed in accordance with the specific procedures of each test. Upon
collection/execution, samples are identified by their works number, date of collection/preparation, name of
material and, where applicable, batch number and/or sample number. Those additional retained samples
shall be delivered to the Depot or Job Supervisor.
Where relevant, records (charts) indicating the locations where samples have been collected should be
documented and, where applicable, the environmental conditions should be recorded (temperature,
humidity, etc.).
The analysis of test results is the responsibility of the Site Manager unless a different person has been
indicated in the current QP.
92

Quality Control On Site

COST E34 - WG1: Bonding on Site

Form B Record of Inspections and Tests:


Epoxy adhesives and grouts, metal rods/plates, FRP rods/plates
Control, inspection or test and
specifications to be verified
RECEPTION OF MATERIALS
Verification of test
bulletins/certificates of
conformance.

DMM

(1)

n.a.

Supporting
documentation

Frequency of
(2)
inspection

Sample
characteristics(3)

Resp.(4)

Project
BoS specification

100%

n.n.

S.M.

In tables B to F:
(1)
Measurement and Monitoring Devices.
(2)
Quantity of inspections by quantity of work carried out.
(3)
Quantity, type, dimensions, etc., comprising the sample to be collected.
(4)
S.M. Site Manager; J. S. Job Supervisor; Op. Operative - BoS Bonding on Site
Form C Record of Inspections and Tests: Solid timber splice fabrication
Control, inspection or test and
specifications to be verified
SOLID TIMBER SPLICE FABRICATION
Visual Inspection: verification of timber
characteristics.
The solid timber splice and the piece to be
repaired shall be of the same type of timber
or, should this not prove possible, they
shall have the same mechanical, durability
and colour characteristics. The use of
timber with raised natural durability can be
justified, even if it is different from the
existing timber.

DMM

(1)

Supporting
documentation

Frequency
Sample
(4)
of
(3) Resp.
(2) characteristics
inspection

n.a.

BoS
Specification
Project

100%

n.a.

S.M.

n.a.

EN Standards

100%

n.a.

S.M.

Verification of moisture content: moisture


content of solid timber splice shall be
between 14 and 16%.

moisture
meter

Project
BoS spec.

100%

n.a.

S.M.

Verification of geometry of the piece:


tolerance: 5mm in cross-section, 10mm
in length.

tape
measure

Project

100%

n.a.

J.S.

Project

100%

n.a.

J.S.

Project
BoS spec.

100%

n.a.

J.S.

Visual inspection: verification of anomalies


and flaws in the timber.
Timber shall be free, as much as possible,
of anomalies and flaws.

Verification of location, diameter and depth


of holes/slots: tolerance: 4mm in location,
+4mm in diameter and +5mm in depth.
Verification of rods/plates characteristics.

Quality Control On Site

n.a.

93

COST E34 - WG1: Bonding on Site

Form D Record of Inspections and Tests: Workshop


Control, inspection or test and specifications to
be verified

DMM(1)

PREPARATION OF STRUCTURAL ELEMENTS TO BE REPAIRED


Verification of propping.
n.a.

Supporting
documentation

Frequency
Sample
(4)
of
(3) Resp.
(2) characteristics
inspection

Project

100%

n.a.

S.M.

Verification of moisture content: timber shall


have a moisture content below 20%.

moisture
meter

Project
BoS spec.

100%

n.a.

S.M.

Ensuring holes, slots and gluing surfaces are


clean.

n.a.

Project
BoS spec.

100%

n.a.

S.M.

tape
measure

Project

100%

n.a.

J.S..

Project
BoS spec.

100%

n.a.

J.S..

Project
BoS spec.

100%

n.a.

J.S..

DIMENSIONAL VERIFICATIONS
Verification of cutting of degraded timber areas:
tolerance of +10mm in length.
Verification of location, diameter and depth of
holes/slots: tolerance: 4mm in location, +4 mm
in diameter and +5mm in depth.
Verification of rods/plates characteristics.

n.a.

Form E Record of Inspections and Tests: Mixing and application of adhesives


Frequency
Sample
(4)
of
(3) Resp.
(2) characteristics
inspection

Control, inspection or test and specifications to


be verified

DMM(1)

Supporting
documentation

VERIFICATION OF FINAL WORK


Visual inspection: the surface of the repaired
element shall be free of flaws and deformations.

n.a.

100%

n.a.

W.E.

Tape
measure

100%

n.a.

W.E.

Verification of the geometrical compatibility


between the solid timber splice and the structural
element.

Form F Record of Inspections and Tests: Verification of final work


Control, inspection or test and specifications to
be verified
VERIFICATION OF FINAL WORK
Visual inspection: the surface of the repaired
element shall be free of flaws and deformations.
Verification of the geometrical compatibility
between the solid timber splice and the structural
element.

10.6

DMM

(1)

Supporting
documentation

Frequency
Sample
(4)
of
(3) Resp.
(2) characteristics
inspection

n.a.

100%

n.a.

W.E.

Tape
measure

100%

n.a.

W.E.

Certification of operatives

10.6.1 Introduction
Certificated operators and supervisors are necessary in order to ensure that this type of specialist activity
and its critical procedures are carried out and recorded with due diligence.
Certification procedures will depend on previous background. Many operatives are already trained and
practising carpenters and therefore the work of bonding on site is an added discipline. There is currently
no acknowledged course for this type of training in Europe. However from Dec 2007 there will be in the
UK an NVQ (Non-Vocational Qualification) for in situ timber engineering repair. This is in the early stages
of being validated by the UKL NVQ Certification Board. In the interim in the UK there are one or two
94

Quality Control On Site

COST E34 - WG1: Bonding on Site

companies who are offering training, in particular for operatives intending to use chain saws and modified
chain saws on site.
The certification foreseen hereafter follows the NVQ principles. It is valid for three years and may only be
renewed for the same period only by repeating the practical exam.
10.6.2 Training
The training and study period is organised by the applicant himself/herself under the guidance of a works
engineer who uses this document as a guide and reference coupled with his/her own experience. The
Supervisor shall provide the applicant with documentation, which the latter should use to consolidate
his/her theoretical knowledge.
10.6.3 Certification procedure
The certification procedure is divided into two parts:

Theoretical examination, oral and/or written test, relating to the relevant procedures (weighted
40%)

Practical demonstration of relevant on site procedures (weighted 60%)

It is recommended that the examinee should at least attain an overall 70 percent grade to pass, with not
less than a 60 percent grade in any single part of the examination. If failure in one section of the test
occurs the candidate may retake the failed portion on two further occasions. If failure still occurs the
candidate will be required to repeat both sections of the test.
10.6.4 Theoretical examination
The theoretical examination may consist of 20 (twenty) questions of a true /false nature, intended to
assess knowledge about materials, procedures, equipment, safety and quality control issues. Allow 100
points are allowed for the practical examination.
10.6.5 Practical examination
The applicant should demonstrate his/her ability to correctly and successfully carry out the working
procedures adapted to the job that is chosen at random by the examiner. This includes: using equipment
and materials compatible with the system chosen; checking the condition of the equipment; ensuring the
proper use of safety and protection equipment. The necessary materials (including a suitable real size
piece timber), equipment and tools will be provided to the applicant.
The examiner will carefully observe the preparation and execution of the test piece repair, after which
he/she will check that in principle the applicant has followed the standard recommended method.
Where applicable, the effect of low/high temperatures and volume of the material on epoxy adhesives and
grouts pot life and cure time should be borne in mind.
10.6.6 Evaluation
100 points are allowed for the practical examination. Points are deducted for unsatisfactory sections of
repair as noted below. All items should be evaluated and graded.
Ref.
1
2
3
4
5

Designation
Lack of accuracy in mixing of adhesives and grouts
Holes or slots deficiently filled or other flaws in the application or injection of
epoxy adhesives and grouts
Equipment dirty or not correctly maintained
Work rate too slow
Failure to observe safety procedures

Points to be
deducted
45
40
5
5
5

10.6.7 Inspection and Testing


During the practical examination, the standard on site inspection and tests applicable to the type of repair
in question should be observed. Test results will help the evaluation of items 1 and 2 in the table above.

Quality Control On Site

95

COST E34 - WG1: Bonding on Site

Following the mixing of the epoxy adhesives


Samples should be taken and tests carried out according to Draft Standards [5-7].
Following the mixing of epoxy grouts used in the reconstitution of timber component sections: the test
piece should be drilled or sawn in a position and in such a way that the examiner can detect application
flaws, such as areas without grout, areas where the grout has improperly polymerised and any such other
defects.

10.7

References

[1]

Anon. 'Low Intrusion Conservation Systems for Timber Structures.' 2006, Website:
http://www.licons.org/

[2]

EN 350-2 'Durability of wood and wood-based products. Natural durability of solid wood. Part 2:
Guide to natural durability and treatability of selected wood species of importance in Europe.'
(1994)

[3]

EN 460 'Durability of wood and wood-based products. Natural durability of solid wood. Guide to the
durability requirements for wood to be used in use classes.' (1994)

[4]

EN 335-2 'Durability of wood and wood-based products. Definition of use classes. Part 2:
Application to solid wood', (2006)

[5]

CEN 'TC 193/SC1/WG11. Adhesives for on site assembling or restoration of timber structures. On
site acceptance testing: Part 1: Sampling and measurement of the adhesives cure schedule. Doc.
N20.' (2003)

[6]

CEN 'TC 193/SC1/WG11. Adhesives for on site assembling or restoration of timber structures. On
site acceptance testing: Part 2: Verification of the shear strength of an adhesive joint. Doc. N21.'
(2003)

[7]

CEN 'TC 193/SC1/WG11. Adhesives for on site assembling or restoration of timber structures. On
site acceptance testing: Part 3: Verification of the adhesive bond strength using tensile proofloading. Doc. N22.' (2003)

96

Quality Control On Site

COST E34 - WG1: Bonding on Site

11 SUMMARY OF ACHIEVEMENTS
Richter K., Cruz H.
A SWOT analysis was performed by the working group at the beginning of the WG activities to establish a
common understanding of the positive and critical aspects of Bonding on Site (figure 11.1). It shows not
only the potentials and challenges but also the critical aspects which need to be resolved in order for
further successful development of the on site bonding processes.
Strengths
Conservation of historic structures
Allows larger spans
Possibility of solving static problems
Permits change of use of existing buildings
Cost-efficiency
Combination of different materials
Low intrusion
Less time consuming than current methods
Superior stiffness

Weaknesses
Lack of knowledge on durability and long
term performance
Lack of accepted standards
Lack of accepted calculation methods
Variability of materials and properties
Sensitivity to temperature and fire
Brittleness/incompatibility
with
primary
adherend
Insufficient universally accepted quality
control methods

Opportunities
New markets
Enhances existing markets
Ductility
Adaptability of formulation
Reversibility
Higher temperature resistance than current
formulations

Threats
Individual job site variability
Difficulty in controlling ambient conditions
Efficiency and acceptability of existing
mechanical repair methods
Requirement for addressing potential health
and safety exposure issues
Operative/installer lack of knowledge
compromising safety of the structure
New technology suspicion syndrome by
potential users

Figure 11.1: SWOT analysis of Bonding on Site (BoS)


It was the objective of the working group to compile and assess as much qualitative and quantitative
knowledge on structural BoS processes as possible in order to reduce the weaknesses and to counter the
perceived threats of on site bonding. The state of the art presented in previous chapters is based on a
literature review, on individual relevant experiences of the authors, plus focused expert discussions aimed
at providing practical proposals for solutions for a modus operandi from the varying existing procedures.
Regarding obstacles to acceptance of BoS, most progress has been achieved in recent years in quality
control processes. The implementation of the systematic quality assurance procedures as they are
described in Chapter 10 will certainly contribute to a better overall quality of on site repair works and shall,
in the mid and long term, increase the reliability of future repair, refurbishment and upgrading of
interventions. The information presented in Chapter 10, as well as in the other chapters, is also a sound
basis for practical teaching and training of the craftsmen and operators, and will serve for certification
procedures.
The concerns related to intrinsic properties of the adhesives used in the repair or bonding work have not
been addressed directly in the course of the Cost E 34 Action. However, design methods taking into
account brittleness, temperature and fire sensitivity are currently being investigated both in scientific
studies as well in the associated chemical industry, so adhesive systems with performance related
ductility and better thermal stability are realistic. In this context we refer to the information summarized by
working group 3. Although temperature is a vital component of the durability assessment procedure of
adhesive joints, excessive temperature durations are not generally considered realistic and should be
used with special caution for accelerated durability evaluation. This is one of the conclusions of Chapter
8, where the variety of factors influencing bond performance is discussed and a considerably important
Summary of achievements

97

COST E34 - WG1: Bonding on Site

contribution to the existing knowledge on durability and long term performance of adhesive bonds is
given. A more practical focus on durability is presented in Chapters 7 and 9, including proposals for an
enhancement of durability in repair work of both contemporary and historic structures.
A most important aspect is the lack of accepted design and calculation methods for on site bonding
processes. This shortcoming could not be solved within the timeframe of this cooperation. Nevertheless,
some initial suggestions for calculation methods are presented and discussed, focusing on the
strengthening of timber beams (Chapter 3), including pre-stressing (Chapter 4) and the procedure for
glued-in-rods (Chapter 5). These proposed calculation methods should be regarded as a contribution to
the ongoing expert discussion which needs to be intensified in the relevant committees and expert
groups, e.g. CIB W18. Through such discussions a second important prerequisite for better accepted
processing conditions will be triggered: the definition of technical standards for on site bonding. Today,
first drafts for three standards have been proposed to assess the adhesives for use in timber structure
rehabilitation. It can be expected that it will take years to upgrade these drafts to harmonize with
European standards. As well as the agreement on technical questions a decision needs to be made on
whether adhesives used on site are covered by the CEN Mandate M 127-EN and, if so, how to enable
CE marking to be undertaken on the basis of these standards.
Despite the progress made in recent years, many open questions have yet to be answered in order to
foster the development of on site bonding. Cooperation in research and development between academic
and industrial partners is essential as well as an exchange of ideas and requirements with practitioners.
The compilation of research requirements at the end of each chapter will help to focus on the relevant
topics. Besides many engineering and technical problems, an improved knowledge of the economical and
environmental considerations regarding BoS is identified as relevant for the promotion of these
procedures.
We further hope that the information summarized in this core document will be used to disseminate
knowledge to industry, research, society and practitioners. Such knowledge transfer will create more
practical applications of on site bonding procedures, and the communication of site experiences to
research laboratories will trigger further innovative techniques aimed at making BoS as acceptable and
reliable as factory controlled structural bonding.

98

Summary of achievements

Você também pode gostar