You are on page 1of 6

Engineering Structures 80 (2014) 147152

Contents lists available at ScienceDirect

Engineering Structures
journal homepage: www.elsevier.com/locate/engstruct

Timber beams loaded perpendicular to grain by multiple connections


A.J.M. Leijten , J.C.M. (Dennis) Schoenmakers
Faculty of Architecture, Unit of Structural Design, Eindhoven University of Technology, P.O.Box 513, 5600MB Eindhoven, The Netherlands

a r t i c l e

i n f o

Article history:
Received 5 November 2013
Revised 26 August 2014
Accepted 27 August 2014
Available online 18 September 2014
Keywords:
Timber
Beams
Connections
Splitting failure
Perpendicular to grain

a b s t r a c t
In timber structures, beams which are loaded perpendicular to grain by connections along the span may
fail by splitting, resulting in the collapse of the beam. In the past, empirical, semi-empirical and fracture
mechanical models have been developed aiming at predicting the splitting failure load. Recent experiments with two and three connections show the empirical models to be very non-conservative in predicting the effect of multiple connections. Fracture mechanical models that are better equipped to estimate
the failure load but the result dependents on the assumed crack growth direction. Test results with three
connections cannot be explained by any of the models yet. Some structural design codes provisions are
unsuitable and can result in very non-conservative predictions while others result in a conservative estimation of the splitting failure load.
2014 Elsevier Ltd. All rights reserved.

1. Introduction

2. Predictive models for single and two connections

A bottom rail loaded by studs in a timber frame is a well-known


example where studs expose the rail to compression stresses perpendicular to grain. If one of the studs is loaded in the opposite
direction, however, a connection is necessary in order to transmit
the tensile force. This also applies for heavy timber structures
where a connection is applied to transmit loads to the beam that
spans the distance between two supports. The strength of the connection can be calculated using the established models given in
design equations of the Building Design Codes. Dependent on the
position over the depth of the beam, however, the connection is
able to force splitting failure along the grain of the beam. In that
case it is the splitting that is governing and not the strength of
the connection. If, for instance, screws or glued in rods are used
for the connection, Fig. 1 and if the connection itself is strong
enough, splitting will occur at the location where the total connection load is transferred to the beam. This location is known as the
largest loaded edge distance which, in the particular case of Fig. 1,
is equal to the penetration or insertion depth of the fasteners. It
will be of no surprise to nd that this distance is an important
parameter in the models that aim to predict the splitting failure
load.

Jensen et al. [1] gave an excellent overview of the models


currently available in literature to predict the splitting resistance
of beams loaded by single connections. Models that predict
the splitting resistance for more than one connection are rare.
Some models do claim to estimate the splitting capacity of
multiple connections and they can be divided into two categories,
(semi-)empirical and fracture mechanical models.
One frequently referenced empirical model that aims at predicting the splitting strength using an analytical approach is Ehlbeck
et al. [2]. After some simplications this semi-empirical model
was introduced in the German design standard DIN 1052 [3]. It is
calibrated using experimental results of dowel-type fastener connections, positioned mainly at mid span of a simply supported
beam. The model consists of one strength parameter and many
geometrical parameters, Eq. (1). The main assumption of the model
is that the tensile strength perpendicular to grain, ft,90 is the governing material parameter. To account for the inuence of the
number of fasteners, the connection area and type of fasteners a
number of additional parameters were introduced. The empirical
nature of this model becomes apparent when for an innitesimal
small loaded edge distance, he Eq. (1) does not goes down to zero.
The German design standard DIN 1052 stipulates that multiple
connections, spaced more than twice the beam depth, can be considered as independent connections meaning that no interaction
takes place. Taking this point of view, the splitting load of each
connection can be considered separately.

Corresponding author. Tel.: +31 402473928; fax: +31 402450328.


E-mail address: a.j.m.leijten@tue.nl (A.J.M. Leijten).
http://dx.doi.org/10.1016/j.engstruct.2014.08.048
0141-0296/ 2014 Elsevier Ltd. All rights reserved.

148

A.J.M. Leijten, J.C.M. (Dennis) Schoenmakers / Engineering Structures 80 (2014) 147152

F 90

s
GGc ha
b
0:61  a

where
F90 the maximum shear force on one side of the connection.
a the ratio between the largest loaded edge distance, a and
beam depth, h.
b thickness of the beam.
G shear modulus.
Gc fracture energy per unit crack surface.

Fig. 1. Part of timber beam loaded perpendicular to grain by screws or glued in


rods.

F 90 ks kr 6:5

18he
h

!
0:8

t ef h f t;90

where
F90 is the splitting capacity.
ks a factor taking account of the number of fasteners in a row.
kr a factor taking account of the number of fasteners in line with
the connection force.
he edge distance of the most distant fastener from the loaded
edge.
h beam depth.
ft,90 tensile strength perpendicular to grain.
tef effective width, dependent of the penetration depth of the
fastener.
The rst model based on linear elastic fracture mechanics for
this application was formulated by Van der Put [4] and later
extended to connections by Van der Put and Leijten [5] based on
the compliance method as described by Hellan [6] given by Eq.
(2). In Fig. 2 the starting point of the model is shown representing
the symmetrical half of a simply supported beam with a connection at mid span. The crack opening is visualised by the separation
of beam 1 into beam 2 and 3. The load that forces the crack opening
is shown at mid span. Basically, it does not consider how the load is
applied but studies the conditions for unstable crack growth
outside the joint area. For this reason the parameters that
describe the fastener type and the connection width and pattern
are absent.

The model assumes shear, fracture mechanics mode II, as the


driving mechanism for failure. In reality it probably is a mixed
Mode I and II mechanism. The ratio between both might depend
on the way in which the crack is initiated as it could depend on
the number of fasteners and the slenderness of the dowel-type fasteners, among others, Van der Put and Leijten [5]. In this respect it
is of interest to quote the experience of Jensen et al. [1]. There are
large differences in splitting at the characteristic level for different
wood species and wood products. Comparison of the test results of
the Douglas r glulam beams and the Radiata pine LVL beams
show the same thing: A 100% difference in splitting capacity may
be observed. A difference of that magnitude ought to be reected
in timber design codes in order not to unfairly discriminate against
competitive wood species and products.
For all these reasons the parameter GGc in Eq. (2) is changed
from a parameter based on the determination of both material
parameters G and Gc to a so-called apparent fracture parameter
based on calibration to test results. Eq. (2) is introduced in Eurocode 5 [7] together with a limiting shear force criterion. The latter
stipulates that the shear force at the support should not exceed the
maximum shear force caused by a single mid-span connection.
This implies that for a simply supported beam with one connection
close to the support, only 50% of the splitting capacity is allowed
compared to a single mid-span connection. Test by [1] with Radiata
Pine LVL beams and by Schoenmakers [15] with beams of Spruce
show that the failure load is not affected by the position along
the grain. This shear provision seems, therefore, to result in a conservative approach. This shear force restriction also limits the total
load, in case of multiple connections, in the same way. An elaborate comparison between the models of Eq. (1) and (2) is presented
by Aicher and Finn [8]. Another fracture mechanical model, where
deformations are modelled as deections of a beam on elastic
foundation for the dowel-type fasteners close to the loaded edge,
is presented by Jensen [9] No predictions are provided, however,
for multiple connections.
Ballerini and Rizzi [10] proposed a model based on a parametric
numerical study of the splitting phenomenon and is semi-empirical in nature, trying to combine both previous models. Eq. (2) is
taken as a starting point. Additional parameters, as shown in Eqs.
35, are added to account for inuences of the connection location
and area.

Fig. 2. Model by Van de Put [4].

A.J.M. Leijten, J.C.M. (Dennis) Schoenmakers / Engineering Structures 80 (2014) 147152

F 90

s
he
f f
2b  9 
1  a3 w r

f r 1 1:75

lr l1
6 2:0
h

nhm =1000
1 nhm =1000

f w 1 0:75

where
l1 the distance between clusters of fasteners (connections).
lr connection width parallel to grain.
hm the difference between the smallest and largest loaded edge
distance.
n number of fastener rows parallel to grain.
The authors of [10] claim that when two connections are separated by more than 1.6 times the beam depth, the splitting strength
is doubled compared to a single connection. The test result, on which
this conclusion is mainly based, refers to Yasumura et al. [11] and the
test setup is given in Fig. 3. In [10] it is stated that the distance
between two connections, l1 is twice the distance between the centre of the connection and the end of the beam for one end of a simply
supported or cantilevered beams. This assumes that tensile stresses
perpendicular to grain govern the splitting and that, at a certain end
grain distance, the splitting cracks are no longer affected by the end
distance. It seems that for this approach, the test setup of Fig. 3 is
mirrored and the beams are virtually nger jointed to obtain a beam
with two single fastener connections. This approach is questionable.
The test data used in [10] for two clusters of fasteners that are less
than 1.6 times the beam depth apart are given in [2,12]. This data
showed the splitting strength to be lower than 1.7 times the single
connection strength.
In [10] the systematic experimental research, by Kasim and
Quenneville [13] that used two connections with increasing spacing, is mentioned but these results are apparently not used for
the evaluation. In [13] simply supported glued laminated timber
beams with depth of 304 mm with two connections were tested.
The spacing along the grain of the connections (with two fasteners
each) incrementally increased up to 2.3 times the beam depth
whilst maintaining the symmetry of the test setup. The aim was
to determine an upper limit for connection spacing beyond which
the connection could be assumed to behave independently. The
result is presented in Fig. 4 and shows that, with increasing spacing, the splitting strength of each connection does not exceed
about 75% of the single connection splitting strength. An upper
bound was reached after 1.4 times the beam depth. In Fig. 4 both
the mean and the 5% lower curves are given. As yet no models
are available to explain this behaviour.

Fig. 3. Test set up by Yasumura et al. [10].

149

Fig. 4. Increasing distance between two connections, Kasim and Quenneville [13].

Two researchers tried solving this multiple connection problem.


Jensen [14] presented a schematization similar to the single connection model by Van der Put and Leijten [5] with two symmetrically placed connections, Fig. 5, where b is the ratio of the crack
length and beam depth, b = k/h.
The model could not, however, explain the experimental results
by Kasim and Quenneville [13]. The splitting failure load appeared
to be either the same load as for a single mid span connection in
which case the cracks merged between the connections, or twice
the failure load of a single mid span connection. In the latter case,
the cracks did not merge between the joints before failure. Another
attempt was made by Schoenmakers [15] who applied the same
compliance method, Fig. 6 but made one very important difference.
The model was set up in such a way that while the rate of crack
growth could differ on both sides of each connection the condition
for crack growth could be different.
When a (dominant) crack grows, other cracks usually grow
simultaneously but possibly at different rates. Plausible situations
were investigated and evaluated. The model results are presented
in Fig. 7.
The critical failure load of one mid span connection is taken as a
reference (100%), top (dotted) curve. In that case, the crack growth
is symmetrical and the resistance drops with increasing crack
length. Fig. 7(b) shows the symmetrical half of a simply supported
beam loaded by two connections. The beam is exposed to a critical
splitting load F introduced by the connection, resulting in a crack
on either side of the connection. The crack length left of the connection has a length k4 while k3 is the crack length on the other
side of the connection. An incremental symmetrical crack growth
is shown extending the crack in both directions. To consider different crack growths on either side of the connection, the parameter
xc is introduced. This parameter represents the ratio of the length
of two growing cracks. For instance xc = k4/k3 is ratio of the crack
length k4 and k3. In case of symmetrical crack growth, k4 = k3 and
so xc = 1. The result of the model is now represented by a curve
which starts at 100% but drops more rapidly than the curve associated with a single mid span connection. This means that, for this
case, the splitting strength of two connections is double the single
connection which agrees with Jensens model [14]. Curves for other
crack growth ratios are also given in Fig. 7(a). For these cases this
reveals that the ultimate load lies somewhere between 0.71 and
0.85 of the single connection splitting strength. The lowest curve
associated with a dominant crack growth towards the support
(while the crack growth towards mid span is very small), the critical load per connection will become 0.71 times the single mid
span critical load. This is in agreement with the test results of
[13] (See Fig. 4).

150

A.J.M. Leijten, J.C.M. (Dennis) Schoenmakers / Engineering Structures 80 (2014) 147152

Fig. 5. Model for two connections by Jensen [14].

Fig. 6. Modelling the symmetrical half of the beam, Schoenmakers [15].

Fig. 7. Critical load per connection as function of the crack length. (a) Comparison to the critical load corresponding to the beam with a single mid span connection. (b) Three
cases of dominant crack propagation direction, Schoenmakers [15].

3. Experimental results with multiple connections


Apart from the theoretical model development [15] performed
many tests with one and two connections to verify the two connection model. Recent tests with three equally spaced connections
along the span were performed to assess if the trend found with
two connections would continue, Fig. 8. These tests using sawn
timber beams came from the same batch as used by [15], and were

assigned as strength class C24 (characteristic bending strength of


24 N/mm2). The mean density was 455 kg/m3 and the moisture
content 12.9%. The glued laminated beams tested had a mean density of 450 kg/m3 and moisture content of 12.7%. In Table 1 the test
series are grouped according to the type of fasteners, the dimensions of the beams and other parameters as indicated in col. (2)
to (9). Nailed connections had 5 rows of 5 nails, equalling 25 nails
in a square pattern. For the sawn wood beams, four closely spaced

151

A.J.M. Leijten, J.C.M. (Dennis) Schoenmakers / Engineering Structures 80 (2014) 147152

When two or three connections are tested simultaneously, the


weakest will always fail rst and distort comparison of the
mean value between test series. Therefore, the average values
of the fracture parameter of these test series were adjusted
using established statistical procedures assuming a standard
normal distribution, Douwen et al. [16]. This results in a rise
of the mean fracture parameter of approximately 10%.
Having taken these factors into account, the corrected apparent
fracture parameter is given by col. (11) Table 1. Column (12) shows
the mean of the test series grouped by fastener type and the number of connections, Fig. 10.
For the nailed connections, there is a distinct difference in
strength between tests with one and two connections. The
strength ratio 10.16/13.35 = 0.76 is close to the prediction in [15]
with a lower bound prediction of 0.71. For connections with dowels, the situation is altered because no signicant difference is
found between the corrected fracture parameter of one and two
connections, i.e. 11.97 and 11.68 respectively. However, three connections apparently have a very signicant effect with a drop in
strength to 8.73/11.97 = 0.73. No model is yet able to explain this
behaviour. However, Schoenmakers model might be a good candidate when extended to three connections.
The consequences of these test results are considerable if one
understands that in the semi-empirical model of [2] connections
are considered as separate connections when spaced more than
twice the beam depth. In Fig. 11 the expected total load on a
45  220 mm beam with a single connection, two or three connections is given based on the outcome of the experimental research.
The mean values for beams with two connections represent the
connections with nails and the other one for dowels. Assuming that
the models used for both design standards DIN1052 [3] and Eurocode 5 [7] accurately predict the mean splitting capacity for one
connection, the graph shows a distinct difference for the total load
prediction using multiple connections. The DIN 1052 [3] appears to
be highly non-conservative as the test data is well below the
expectation, while the Eurocode 5 [3] is conservative. This conservative estimation is caused by the limiting shear force criterion.
The Ballerini and Rizzi model [10] has a spacing parameter for
multiple connections, Eq. (4). For the tests with two connections

Fig. 8. Test with three connections, Series 16 and 17, Table 1.

(4d) 12 mm diameter dowels were used, set in a square pattern. All


beams failed in a brittle splitting mode.
In addition [14,15] also tested cantilevered beams with connections both at the end and half way along the cantilever length but
these results were omitted here. Series 16 and 17 beams comprised
10 tests with three connections equally spaced at two times the
beam depth, 2h along the span. All three connections were loaded
by separate hydraulic actuators, each having a load cell to check for
any differences, which, however, proved to be insignicant. Crack
initiation and growth direction were studied with special transducers, mounted at close distance on either side of each connection. In addition, a high speed camera was used to observe the
crack growth visually. In 70% of the tests the crack initiation
started at the connections nearest to the support but a dominant
crack growth direction was difcult to determine, Fig. 9. In 30%
of the tests, however, a symmetric crack growth could be determined. In 50% of the cases, a leading crack direction could not be
established.
The number of connections along the span is given in col. (5),
Table 1. The critical load, Fcrit is the load per connection, col. (10).
To allow comparison between test series using different cross-sections, distance from the support, number and type of fasteners, the
mean apparent fracture parameter (GGc)0.5 was calculated per test
series with Eq. (2).
This fracture parameter was adjusted for the following reasons:
From evaluation of the total data base of test results by [15] a
10% higher value was found with glued laminated beams compared to sawn timber beams. This takes 10% off for test series 1,
2 and 9 and 10.

Table 1
Test results of beam with multiple connections.

(1)
Test
series

(2)
Cross
section

(3)
No of
test

(4)
Span

No

bh
(mm2)

2l
(mm)

(5)
Num of
con

(6)
Distance from
support

(7)
Num of
fast

(8)
Loaded edge
distance

(9)
Diam,

(mm)

he/h
(%)

d
(mm)

(10)
Mean max
load per
connection
Fcrit
(kN)

(11)
Adjusted
calibration
(GGc)0.5
mean
(N/mm1,5]

(12)

mean

1
2
3
4
5a
6
7
8

45  300
45  300
45  220
45  220
45  220
45  220
45  220
45  220

5
5
5
5
5
5
5
5

2600
2600
1600
1600
1600
1400
1200
1000

1
1
1
1
1
1
1
1

650
900
800
800
800
700
600
500

25
25
25
20
20
15
15
15

47
47
47
47
47
47
47
47

4
4
4
4
4
4
4
4

27.78
29.41
22.79
22.54
22.52
21.82
20.11
19.43

13.19
13.97
14.04
13.89
13.88
13.45
12.39
11.97

13.35

9
10

45  300
45  300

5
5

2600
2600

2
2

900
650

25
25

47
47

4
4

19.68
19.83

10.12
10.20

10.16

11
12
13

45  220
45  220
45  220

3
5
5

1400
1200
1000

1
1
1

700
600
500

4
4
4

44
44
44

12
12
12

17.91
18.76
18.17

11.72
12.28
11.90

11.97

14
15

45  220
45  220

5
5

1600
1600

2
2

400
200

4
4

44
44

12
12

16.68
17.56

11.48
11.88

11.68

16
17

45  220
40  220

10
10

2000
2000

3
3

440
440

4
4

46
33

12
12

11.68
8.68

8.24
9.23

8.73

One extreem low left out.

152

A.J.M. Leijten, J.C.M. (Dennis) Schoenmakers / Engineering Structures 80 (2014) 147152

this parameter reaches the upper limit value of 2. This means that
the splitting strength is twice the single connection strength which
Fig. 11 shows is too optimistic.
4. Conclusions
A simply supported beam, loaded by multiple connections perpendicular to the grain, can fail by splitting if this appears to be the
weakest link. Models that predict the splitting failure have been
mainly veried for a single connection at mid span. Only the fracture model by Schoenmakers [15] is able to explain the reduced
splitting capacity for two connections as reported by Kasim and
Quenneville [13]. This investigation focused on multiple connections that were spaced twice the beam depth along the span. The
test result shows a distinct non-linear inuence of the number of
connections. For connections spaced this distance, the model by
Ehlbeck et al. [2] predicts a linear increase with the number of connections which leads to very non-conservative results. Also the
design parameter that accounts for the connection spacing of multiple connections by Ballerini and Rizzi [10] in Eq. (4) result in an
non-conservative estimation of the splitting strength. Present
Eurocode 5 [7] provisions based on the model by Van der Put [4]
result in a conservative but safe approach.

Fig. 9. Location of crack initiation.

Apparent (adusted) fracture parameter

16

mean

14

Nails

12

Dowels

References

10
8
6
4
2
0

Number of conecons
Fig. 10. Inuence of multiple connections.

Fig. 11. Code predictions and test results for a 45  220 mm beam.

[1] Jensen JL, Girhammar UA, Quenneville P, Kllsner. Splitting of beams loaded
perpendicular to grain by connections simple fracture mechanics models. In:
Proceedings of the world conference of timber engineering, Auckland, New
Zealand; July 2012
[2] Ehlbeck J, Grlacher R, Werner H. Determination of perpendicular-to-grain
tensile stresses in joints with dowel-type-fasteners. In: Proceedings of CIBW18/paper 22-7-2, 1989, Berlin, Germany; 1989.
[3] DIN 1052: 2004-08. Design of timber structures general rules and rules for
buildings. Beuth Verlag GmbH, 10772 Berlin. Germany.
[4] Van der Put TACM. Tension perpendicular to grain at notches and joints. In:
Proceedings of CIB-W18/paper 23-10-1, Lisbon, Portugal; 1990.
[5] Van der Put TACM, Leijten AJM. Evaluation of perpendicular to grain failure of
beams caused by concentrated loads of joints. In: Proc. of CIB-W18/paper 337-7, Delft, The Netherlands; 2000.
[6] Hellan K. Introduction to fracture mechanics. MacGraw-Hill Book Company;
1985.
[7] Eurocode 5. EN 1995-1-1, European committee for standardization (CEN):
design of timber structures. Brussels, Belgium, 2004.
[8] Aicher S, Finn R. Joints in timber structures loaded perpendicular to grain,
comparison of design approaches. Otto-Graf-J 2005;16:28599.
[9] Jensen JL. Quasi non-linear fracture mechanics analysis of the splitting failure
of single dowel joints loaded perpendicular to grain. J Wood Sci
2005;51:55965.
[10] Ballerini M, Rizzi M. Numerical analyses for the prediction of the splitting
strength of beams loaded perpendicular-to-grain by dowel-type connections.
Mater Struct 2007;40:13949.
[11] Yasumura M, Murota T, Sakai H. Ultimate properties of bolted joints in gluedlaminated timber. In: Proc. CIB-W18/paper 20-7-3, Dublin, Ireland; 1987.
[12] Reske RG, Mohammed M, Quenneville P. Inuence of joint conguration
parameters on strength of perpendicular-to-grain bolted timber connections.
In: Proc. of the 6th world timber engineering conference, WCTE 2000,
Whistler, BC Canada; August 2000.
[13] Kasim M, Quenneville P. Effect of row spacing on the capacity of bolted timber
connections loaded perpendicular-to-grain. In: Proc. of CIB-W18/ paper 35-76, Kyoto, Japan; 2002.
[14] Jensen. Splitting strength of beams loaded by connections. In: Proc. CIB-W18/
paper 36-7-8, Portland, Colorado, USA; 2003.
[15] Schoenmakers D. Fracture and failure mechanisms in timber loaded
perpendicular to grain by mechanical connections. PhD thesis, Eindhoven
University of Technology, Eindhoven, The Netherlands; 2010.
[16] Douwen AA, Kuipers J, Loof HW. Corrections to the mean value and the
standard deviation from test series with symmetrical test specimens. Stevin
report 4-82-9/oe-5, May TU-Delft; 1982.

You might also like