Sunteți pe pagina 1din 10

Brittle Failures in Timber Beams Loaded

Perpendicular to Grain by Connections


Jørgen L. Jensen, Ph.D. 1; Pierre Quenneville 2; Ulf Arne Girhammar 3; and Bo Källsner 4
Downloaded from ascelibrary.org by GADJAH MADA UNIVERSITY on 02/23/16. Copyright ASCE. For personal use only; all rights reserved.

Abstract: A state-of-the-art review of simple analytical fracture mechanics models for calculation of the splitting capacity of timber beams
loaded perpendicular to the grain direction by connections is presented. It is shown that most of the already available models are closely
related and appear naturally as special cases of the most general model available. A new model, which is a semiempirical extension of an
existing model based on a beam-on-elastic-foundation theory, is proposed. The so-called van der Put model, which forms the theoretical basis
for the splitting equations used in the European and Canadian timber design codes, appears as a special case of the proposed model. The
treatment of the splitting problem in some major timber design codes is reviewed and discussed based on the theoretical models and new test
results. The approach used in the European timber design code where the maximum shear force on either side of a connection is considered
rather than the total load applied on a connection is not in agreement with the test results presented. While the European and Canadian timber
design codes apply a constant value for a material property related to the splitting performance irrespective of the material considered, the
presented experimental results indicate that the material property for Radiata pine laminated veneer lumber can be close to twice the value for
Douglas fir glulam. The presented test results also show that despite the fact that Douglas fir glulam has a significantly higher mean
perpendicular-to-grain tensile strength than Radiata pine laminated veneer lumber, the splitting failure load of Radiata pine laminated veneer
lumber is nevertheless significantly higher than that of Douglas fir glulam. The latter finding seems to be in disagreement with the German
timber design code, according to which the splitting strength is proportional to the perpendicular-to-grain tensile strength. DOI: 10.1061/
(ASCE)MT.1943-5533.0001275. © 2015 American Society of Civil Engineers.
Author keywords: Bolted connections; Timber structures; Splitting; Edge distance; Fracture mechanics; Design codes.

Introduction Further, a new model is proposed. The model is a semiempirical


modification of an existing so-called quasi-nonlinear fracture me-
Timber is an orthotropic material with high strength in the parallel- chanics model, which is based on a beam-on-elastic-foundation
to-grain direction and low strength in the perpendicular-to-grain theory. With the semiempirical modification, the quasi-nonlinear
direction. A timber beam loaded perpendicular to the beam axis fracture mechanics model is brought to contain the until now most
by a connection is subjected to concentrated tensile stresses advanced linear elastic fracture mechanics model as a special case.
perpendicular to the grain and is at risk of failing in splitting. In the European [EN 1995-1-1:2004 (CEN 2004)] and Canadian
Explicit equations for calculation of the splitting capacity of beams [CSA O86-09 (Canadian Standards Association 2009)] timber de-
loaded perpendicular to the grain by connections have recently sign codes, design against splitting of beams loaded perpendicular
been introduced in some major timber design codes. However, to grain by connections is based on a simple fracture mechanics
required minimum edge distances are still in many design codes model, while an empirical or semiempirical approach is used in
relied on as the sole measure to avoid splitting. the German timber design code [DIN 1052:2008 (DIN 2008)].
The character of splitting failures in timber makes it natural to The approaches used in the design codes are discussed in the
apply fracture mechanics, and a number of simple analytical frac- present paper based on theoretical models and on new test results,
ture mechanics models have been presented in recent years for and some issues are emphasized, which need further attention.
analysis of beams loaded perpendicular to grain by connections. The aim of this paper is three-fold, as follows: (1) to give a state-
In the present paper, a state-of-the-art review of such models is of-the-art review of existing simple analytical models for prediction
given, and it is shown that most of the available models are not of the splitting failure load of beams loaded perpendicular-to-grain
independent, but appear as special cases of the (until now) most by connections, and to show that most of the models are related and
general model available. appear as special cases of the most general model; (2) to propose a
semiempirical modification of an existing model based on a beam-
1
Research Fellow, Dept. of Civil and Environmental Engineering, on-elastic-foundation theory, and to indicate by testing the appro-
Faculty of Engineering, Univ. of Auckland, Auckland 1142, New Zealand priateness of the proposed model; and (3) to discuss the approaches
(corresponding author). E-mail: jljensen2908@yahoo.com used in some major timber design codes for beams loaded
2
Professor, Head of Dept., Dept. of Civil and Environmental Engineering,
perpendicular to grain by connections, and to present experimental
Faculty of Engineering, Univ. of Auckland, Auckland 1142, New Zealand.
3
Professor, Division of Structural and Construction Engineering–
evidence for their inadequacy.
Timber Structures, Luleå Univ. of Technology, 971 87 Luleå, Sweden.
4
Professor, School of Engineering, Linnæus Univ., 352 52 Växjö, Sweden.
Note. This manuscript was submitted on June 23, 2014; approved on State-of-the-Art of Simple Analytical Models
January 13, 2015; published online on March 11, 2015. Discussion period
open until August 11, 2015; separate discussions must be submitted for The simple analytical fracture mechanics models currently avail-
individual papers. This paper is part of the Journal of Materials in Civil able for analysis of splitting of beams loaded perpendicular to grain
Engineering, © ASCE, ISSN 0899-1561/04015026(10)/$25.00. by connections may all be considered to be based on the energy

© ASCE 04015026-1 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., 2015, 27(11): 04015026


All models are mathematical abstractions and therefore approx-
imations to a physical reality. More or less realistic assumptions
may be made in the mathematical models, leading to more or less
good agreement with the physical reality. A number of different
solutions may be derived by means of Eq. (6) by making certain
Fig. 1. Cracked beam loaded perpendicular to grain by a connection assumptions about how to calculate the compliance, CðAÞ. The
choice of assumptions leads to more or less good agreement with
experimental results, but from a modeling point of view the models
are equally acceptable. From a design point of view, certain models
balance approach (Gustafsson 2003). A beam as indicated in Fig. 1 may be more attractive than others. Usually, models applied for
is considered. The beam is loaded perpendicular to the grain by a practical design need to be sufficiently simple, robust, in good
connection. The external load acting on the connection is denoted agreement with test results, and if possible be based on material
Downloaded from ascelibrary.org by GADJAH MADA UNIVERSITY on 02/23/16. Copyright ASCE. For personal use only; all rights reserved.

P, and a crack with crack area A is assumed to have been initiated. properties obtained from independent standard tests.
The beam depth is denoted h and the loaded edge distance is
denoted he .
The considered beam is a system with one degree of freedom Model 1
and may be regarded as a simple spring. The system is assumed to All the models reviewed in the present paper may be based
be linear elastic. Let k denote the spring stiffness, i.e., P ¼ kδ, on Eq. (6). To illustrate the use of Eq. (6), a new model is derived.
where δ is the deflection of the loading point. The stiffness k is A close-up of the part of the beam where the connection is located
a function of the crack area, A, i.e., k ¼ kðAÞ. The compliance, is considered in Fig. 2. The external load is assumed to act in
C, is defined as C ¼ CðAÞ ¼ 1=kðAÞ, and thus a single point, and the loaded part of the beam with depth he is
assumed to behave like a beam of length L ¼ 2a, where 2a
δðAÞ ¼ PCðAÞ ð1Þ is the crack length, with fully fixed supports. If taking into account
bending and shear deformations, the compliance of the beam is
If the load, P, is below a certain critical value, the crack will not
given by
propagate. When P reaches the critical value, Pu , the crack prop-
agates. The situation where the load has reached exactly the critical
a3 a
value, P ¼ Pu , is considered. If the crack area increases by an CðaÞ ¼ þ ð7Þ
infinitely small amount, dA (the load is assumed to be constant 24EI 2 GAs
during the crack propagation), the compliance increases corre-
spondingly by the amount dCðAÞ, and therefore from Eq. (1) where E = modulus of elasticity (MOE) in the grain direction;
dδðAÞ ¼ Pu dCðAÞ. The increase in deflection means that the exter- G = shear modulus; I = moment of inertia; and As = shear area
nal load, Pu , has moved to a state of lower potential energy. The of the beam with depth he . The crack is assumed to penetrate
change in potential energy of the external load, dEpot;ext , due to the the beam through its entire width, and if a beam with rectangular
increase of the crack area, dA, is thus given by cross section with width b is considered, then I ¼ bh3e =12
and As ¼ bhe =β s , where β s is a shear correction factor
dEpot;ext ðAÞ ¼ −Pu dδðAÞ ¼ −dCðAÞP2u ð2Þ (β s ¼ 6=5 for a rectangular cross section according to ordinary
beam theory). The crack area, A, becomes A ¼ 2ba, and from
Due to the increase in the deflection, dδðAÞ, the increase in elas- Eqs. (6) and (7)
tic strain energy, dEstrain , stored in the cracked beam becomes sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2GGf he
1 1 Pu ¼ 2b ð8Þ
dEstrain ðAÞ ¼ Pu dδðAÞ ¼ dCðAÞP2u ð3Þ 3 GE ðhae Þ2 þ β s
2 2

Fracture energy, Gf , is defined as the work needed to bring a The model given by Eq. (8) is to the knowledge of the writers
unit area of a crack surface from its initial unloaded state to com- new and has not been presented previously. However, as will be
plete separation. If the crack area increases by the amount dA, the shown in a subsequent paragraph, Eq. (8) is just yet another model,
energy dissipated due to fracture, dEdis , becomes which appears as a special case of an existing model (Model 3). The
model is presented as an illustration of the fact that so-called new
dEdis ðAÞ ¼ Gf dA ð4Þ models may be presented, which do not lead to any new insight.
Further, the model may be regarded as a so-called missing link
It is assumed that the fracture process occurs sufficiently slowly between two existing models as illustrated in Fig. 6 (discussed
to neglect kinetic energy. The total energy, Etotal , of the system is in detail in a subsequent paragraph).
constant (Etotal ¼ Epot;ext þ Estrain þ Edis = a constant value), and
thus the change in total energy, dEtotal , due to an infinitely small
extension of the crack area, must be zero

dEpot;ext ðAÞ þ dEstrain ðAÞ þ dEdis ðAÞ ¼ 0 ð5Þ

Inserting Eqs. (2)–(4), in (5), the critical value, Pu , of the load,


which makes the crack propagate, becomes
sffiffiffiffiffiffiffiffiffi
2Gf
Pu ¼ dCðAÞ ð6Þ
Fig. 2. Fully fixed beam model
dA

© ASCE 04015026-2 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., 2015, 27(11): 04015026


Model 2 Model 4
Neglecting bending deformations and only taking shear deforma- If neglecting the bending deformations and only taking into
tions into account (finite G, E → ∞) leads to account the shear deformations (finite G, E → ∞), Eq. (10) is
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi reduced to
Pu ¼ 2b 2GGf he =β s ð9Þ sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2GGf he
Pu ¼ 2b ð11Þ
If assuming β s ¼ 1 in Eq. (9), the solution derived in Larsen and β s ð1 − αÞ
Gustafsson (2001) is obtained. The solution presented in Larsen
and Gustafsson (2001) is thus a special case resulting from the For β s ¼ 6=5, Eq. (11) may also be written as
model shown in Fig. 2 and given by Eq. (8). An even more general sffiffiffiffiffiffiffiffiffiffiffiffi
model, which contains Eq. (8) as a special case, is obtained by he
Pu ¼ 2bC1 ð12aÞ
Downloaded from ascelibrary.org by GADJAH MADA UNIVERSITY on 02/23/16. Copyright ASCE. For personal use only; all rights reserved.

assuming that the beam in Fig. 2 instead of being fully fixed 1 − hhe
has rotational and vertical translational springs attached at the ends.
The equations will, however, not be derived in this paper since it is rffiffiffiffiffiffiffiffiffiffiffiffiffi
not immediately obvious what stiffness the springs should be as- 5
C1 ¼ GGf ð12bÞ
signed. A model which may be regarded as being a special case of 3
such a spring model and which automatically includes appropriate
stiffness of the springs will be reviewed in a subsequent paragraph.
Eqs. (12a) and (12b) constitute the so-called van der Put or van der
Put/Leijten model first presented in van der Put and Leijten (2000),
Model 3 which now forms the basis of design against splitting in the European
[EN 1995-1-1:2004 (CEN 2004)] and Canadian [CSA O86-09
Fig. 3 shows a different structural system from which the compli- (Canadian Standards Association 2009)] timber design codes.
ance may be determined. The cracked beam is modeled as a beam The models presented in Larsen and Gustafsson (2001) and van
structure consisting of four horizontal beam elements and two ver- der Put and Leijten (2000) [Eqs. (9), (12a) and (12b), respectively]
tical beam elements, all rigidly connected. The uncracked parts of and the new model given by Eq. (8) are all special cases resulting
the beam are modeled using horizontal beam elements with depth from Eq. (10).
h, while the cracked part is modeled by means of two horizontal
beam elements with depths he and (h − he ). The vertical beam el-
ements are assumed to be infinitely stiff. All connections between Model 5
the beam elements are assumed to be fully rigid. In Jensen et al. (2003) and Jensen (2005a, b, c), models based on a
The compliance of the structure shown in Fig. 3 may be ex- Timoshenko beam on a Winkler foundation are presented. Fig. 4
pressed analytically, and according to Jensen (2005a) follows from schematically illustrates the model for a simply supported beam.
Eq. (6) An infinitely thin fictitious fracture layer is assumed along the po-
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi tential crack path. The part of the beam below the fracture layer is
2GGf he treated as a beam with depth he on an elastic foundation, and the
Pu ¼ 2b ð10Þ
3 E ðhe Þ ð1 − α3 Þ þ β s ð1 − αÞ
G a 2
part above the fracture layer (h–he ) is assumed to be infinitely stiff.
The fracture layer, which provides the foundation stiffness, is as-
where α ¼ he =h. sumed to be linear elastic and ideal brittle, and it is assigned the
By comparison of Eqs. (8) and (10), Eq. (8) appears as a special properties of the wood, i.e., the perpendicular to grain tensile
case of Eq. (10), namely for α → 0. This is in accordance not only strength, ft , and Mode I fracture energy, Gf . The stiffness, K,
with the mathematics, but can also easily be realized from the phys- of the foundation may thus be given in terms of the material proper-
ics; α → 0 is in accordance with assuming h → ∞ for a finite ties as K ¼ f 2t =ð2Gf Þ.
value of he , i.e., all beams except the beam with depth he are as- For a single load acting far from the end of a beam, e.g., as
sumed to be infinitely stiff. The beam with depth he and length 2a shown in Fig. 4, the solution may for small crack lengths
thus becomes fixed at the ends exactly like the beam considered (a → 0) according to Jensen (2005c) be given as
in Fig. 2.
The model given by Eq. (10) is built on the assumption that a Pu ¼ γPu;LEFM ð13aÞ
point load is placed at midspan of a simply supported beam and the pffiffiffiffiffi
crack is assumed to propagate symmetrically to the left and right Pu;LEFM ¼ 2bC1 he ð13bÞ
from the point load. In Schoenmakers (2010) an extended version
of the model given by Eq. (10) is given in which asymmetric crack rffiffiffiffiffiffiffiffiffiffiffiffiffi
propagation is considered. The approach does, however, not lead to 5
C1 ¼ GGf ð13cÞ
a simple explicit solution. 3

Fig. 3. Beam structure for modeling the compliance of a cracked beam Fig. 4. Beam-on-elastic-foundation model

© ASCE 04015026-3 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., 2015, 27(11): 04015026


pffiffiffiffiffiffiffiffiffiffiffiffiffi
2ζ þ 1 illustrated in Fig. 3, rigidly connected, which is an idealization that
γ¼ ð13dÞ in general leads to an unsafe solution. Gustafsson (1988) intro-
ζþ1
duced an additional contribution to the compliance in a beam model
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi for analysis of end-notched beams in order to account for the abrupt
C G1 change in cross-sectional depth.
ζ¼ 1 10 ð13eÞ Since the failure load according to Eqs. (13a)–(13e) is given as a
ft E he
factor, γ, multiplied by a linear elastic fracture mechanics solution,
Pu;LEFM , and this linear elastic fracture mechanics solution results
where a subscript in Eqs. (13a) and (13b) refers to linear elastic from Eqs. (12a) and (12b) for h → ∞, it may naturally be sug-
fracture mechanics (LEFM). gested to use the linear elastic fracture mechanics solution given
The solution Pu ¼ Pu;LEFM is obtained from Eq. (9) for β s ¼ by Eqs. (12a) and (12b) in Eqs. (13a)–(13e) as a substitution
6=5 and from Eqs. (12a) and (12b) for h → ∞. The solution Pu ¼ for Pu;LEFM
Downloaded from ascelibrary.org by GADJAH MADA UNIVERSITY on 02/23/16. Copyright ASCE. For personal use only; all rights reserved.

Pu;LEFM is obtained from Eqs. (13a)–(13e) if assuming f t → ∞,


or if assuming E → ∞. Pu ¼ γPu;LEFM ð15aÞ
In Jensen (2005d), the model given by Eqs. (13a)–(13e) is
further extended to not only considering the loaded edge distance, sffiffiffiffiffiffiffiffiffiffiffiffi
he , but also the influence of the fastener’s distance from the beam he
Pu;LEFM ¼ 2bC1 ð15bÞ
end, le . An approximate solution is given by 1 − hhe
0 pffiffiffiffiffiffiffiffi
1 1
þ Pbf t le
rffiffiffiffiffiffiffiffiffiffiffiffiffi
2 2ζþ1 u;LEFM
Pu ¼ Pu;LEFM · min@ pffiffiffiffiffiffiffiffi A ð14Þ 5
2ζþ1 C1 ¼ GGf ð15cÞ
ζþ1
3

pffiffiffiffiffiffiffiffiffiffiffiffiffi
where Pu;LEFM is given in Eqs. (13a)–(13e). A typo in Jensen 2ζ þ 1
(2005d) has been corrected in Eq. (14). γ¼ ð15dÞ
ζþ1
In, for example, Jensen (2005b) it has been shown that if the
foundation stiffness is chosen as K ¼ f 2t =ð2Gf Þ, which is the case
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
for Eqs. (13a)–(13e) and (14), the same failure load is obtained
C G1
from the beam on elastic foundation whether using (1) a conven- ζ¼ 1 10 ð15eÞ
tional stress failure criterion or (2) the compliance method of ft E he
fracture mechanics as given by Eq. (6). For other choices of foun-
dation stiffness, e.g., if including the elastic perpendicular to grain
strain of the beams, the two approaches do not lead to the same The γ factor is limited to the range [0, 1] and may be regarded as
failure load. an effectiveness factor, which is a function of the material proper-
The perpendicular to grain tensile strength appears in ties E, G, ft , and Gf and of the loaded edge distance, he . For h →
Eqs. (13a)–(13e) and (14), and the failure load is in general not ∞ Eqs. (15a)–(15e) reduce to Eqs. (13a)–(13e), and for ft → ∞ or
proportional to the square root of the fracture energy. The solutions G=E → 0 Eqs. (15a)–(15e) reduce to Eqs. (12a) and (12b). Fig. 5
do thus not possess the characteristics of linear elastic fracture shows γ as a function of ζ (which is typically limited to the lower
mechanics, but on the other hand cannot be characterized as non- part of the range [0, 6]).
linear fracture mechanics either. The solutions belong to a class Also Eq. (14) may semiempirically be generalized by introduc-
of fracture mechanics solutions, which have been termed quasi- ing the expression for Pu;LEFM as given in Eqs. (15a)–(15e). The
nonlinear fracture mechanics solutions (Serrano and Gustafsson generalized Eq. (14) takes into account the effects of edge and
2006). The LEFM solutions appear as special cases from the quasi- end distances, and contains the models presented in van der
nonlinear solutions. The claim made by van der Put (2013a, b) that Put and Leijten (2000) and Larsen and Gustafsson (2001) as
the quasi-nonlinear fracture mechanics models are not fracture special cases.
mechanics models therefore seems irrelevant. Fig. 6 shows schematically how the reviewed Models 1–6 are
The model considered in Fig. 2 and given by Eq. (8) may be related.
generalized by introducing rotational and vertical translational
springs at the supports instead of the fixed end boundary condi-
tions. However, for such a spring model to be relevant, appropriate
spring stiffness must be assigned. The beam on elastic foundation
model may be regarded as a special case of such a spring model,
where the appropriate stiffness of the springs is automatically
given.

Model 6
The beam on elastic foundation models are by nature limited to
situations for which the loaded edge distance is small as compared
with the total beam depth. The same limitation applies to the so-
lutions given in Eqs. (8) and (9) and illustrated in Fig. 2. The model
illustrated in Fig. 3 does in principle not possess any limitation
Fig. 5. Effectiveness factor, γ, as a function of ζ
to the loaded edge distance. However, the beam elements are, as

© ASCE 04015026-4 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., 2015, 27(11): 04015026


Downloaded from ascelibrary.org by GADJAH MADA UNIVERSITY on 02/23/16. Copyright ASCE. For personal use only; all rights reserved.

Fig. 7. Definition of some parameters used in Eq. (17)


Fig. 6. Relations between the reviewed and proposed models

beam loaded at quarter-span, and 2 times the resistance of a can-


Design Approaches tilever beam. The length of the beams is of no influence.
The same expression for Rk as given in Eqs. (16a)–(16d) is used
Specified minimum edge distances have traditionally been relied on in the Canadian timber design code [CSA O86-09 (Canadian Stan-
in most timber design codes as a sufficient guard against splitting of dards Association 2009)], but in the Canadian timber design code it
beam loaded perpendicular to the grain by connections. In some is required that the total connection design load must be less than
design codes, an additional check of the shear stresses in the beam the corresponding design value of the resistance.
has been required at the location of the connection. It is, however, The German timber design code [DIN 1052:2008 (DIN 2008)]
only recently that explicit equations for design against splitting is based on an empirical approach and the fact that a connection in
have been introduced in major timber design codes. reality usually does not load the beam in a single point is accounted
In the European [EN 1995-1-1:2004 (CEN 2004)] and Canadian for. Eqs. (17a)–(17d) shows the requirements according to DIN
[CSA O86-09 (Canadian Standards Association 2009)] timber design 1052:2008 (DIN 2008). Definitions of related parameters are given
codes, it has been decided to base the design against splitting on the in Fig. 7. In Eqs. (17a)–(17d), f t denotes the perpendicular-to-grain
simple linear elastic fracture mechanics model given by Eqs. (12) and tensile strength of the timber, and bef is an effective beam width,
(12b). While it in the Canadian timber design code [CSA O86-09 which depends on the slenderness of the fasteners [for further
(Canadian Standards Association 2009)] has been decided to details refer to DIN 1052:2008 (DIN 2008)]
consider the total load on the connection as in Eqs. (12) and
(12b), it has in the European timber design code [EN 1995-1- R ¼ μ N μ m R0 ð17aÞ
1:2004 (CEN 2004)] been decided to consider the shear force instead
of the total connection load. In the European timber design code [EN   2 
h
1995-1-1:2004 (CEN 2004)], it is required that Eqs. (16a)–(16d) be R0 ¼ 6.5 þ 18 e ðbef hÞ0.8 f t ð17bÞ
fulfilled. Subscripts d and k denote design and characteristic values, h
respectively; V 1;d and V 2;d are the design shear forces on the two
 
sides of the connection; and R denotes resistance 1
μN ¼ max ð17cÞ
0.7 þ 1.4 lhN
V d ≤ Rd ð16aÞ

m
V d ¼ maxfV 1;d ; V 2;d g ð16bÞ μm ¼ P m h1 2
ð17dÞ
i¼1 ð hi Þ
sffiffiffiffiffiffiffiffiffiffiffiffi
he In Ballerini and Rizzi (2007) another semiempirical model is
Rk ¼ bC1 ð16cÞ presented, which attempts to take into account the effect of multiple
1− he
h fasteners. The model, which is given by Eqs. (18a)–(18e), utilizes
an empirically modified version of Eqs. (12a) and (12b) for the
C1 ¼ 14 N=mm1.5 ð16dÞ failure load caused by a single fastener. The effect of multiple fas-
teners is taken into account by multiplying the failure load caused
by a single fastener by some modification factors, μN and μm , in the
Eqs. (12a) and (12b) have been derived by considering a simply same way as in Eqs. (17a)–(17d)
supported beam loaded at midspan, and is in principle only valid for
that special situation. However, in practical design situations, en- R ¼ μ N μ m R0 ð18aÞ
gineers need to be able to design for other support conditions and
other locations of the connections as well, and by considering the sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
he
shear force as done in Eqs. (16a)–(16d), all situations can in prin- R0 ¼ 2bC1 ðcharacteristic C1 ¼ 9 N=mm1.5 Þ ð18bÞ
ciple be handled by the designer in a very simple way. 1 − ðhhe Þ3
For a simply supported beam loaded at midspan, Eqs. (12a),
(12b), and (16a)–(16d) give the same failure load for C1 ¼  
14 N=mm1.5 . Eqs. (16a)–(16d) predicts that a beam loaded at mid- 1 þ 0.75 lhN
μN ¼ min ð18cÞ
span has a splitting resistance which is 1.5 times that of the same 2

© ASCE 04015026-5 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., 2015, 27(11): 04015026


χ grade of LVL). The shear modulus of Radiata pine LVL is usually
μm ¼ 1 þ 1.75 ð18dÞ
1þχ estimated as G ¼ E=20, which will also be assumed in this paper.
The glulam used had a density of 516 kg=m3 (mean value) at
mlm a moisture content of 12–13%. The dynamic MOE in the grain
χ¼ ð18eÞ direction was measured using a log-grader and found to be E ¼
1,000
12.9 GPa (mean value). Two different grades of glulam were or-
dered, but judged from the density and MOE measurements, only
one grade (the higher) was delivered. The glulam was ordered with
Experimental: Materials and Methods the same grade of timber used for all laminas.
Hot galvanized machine bolts of quality 4.2 and with a diameter
Splitting tests were conducted on laminated veneer lumber (LVL) of d ¼ 16 mm were used in the tests.
specimens made of New Zealand grown Radiata pine (Pinusra-
Downloaded from ascelibrary.org by GADJAH MADA UNIVERSITY on 02/23/16. Copyright ASCE. For personal use only; all rights reserved.

diata) and on glue-laminated timber (glulam) specimens made


Beam Tests
of New Zealand grown Douglas fir (Pseudotsugamenziesii). Fur-
ther, material property tests were conducted to determine the Splitting tests were conducted in three-point bending setups.
perpendicular-to-grain tensile strength of the LVL and the glulam, The tested glulam and LVL specimens are shown in Figs. 8
and the Mode I fracture energy of the LVL. So-called plate speci- and 9, respectively. For both glulam and LVL, tests were conducted
men tests were also conducted on glulam and LVL. on beams loaded at midspan and at quarter-span. All beams had a
The LVL used had a density of 593 kg=m3 (mean value) at a 45 × 300 mm2 cross section. Two different edge distances he were
moisture content of 9%. The dynamic MOE in the grain direction tested [(1) he ¼ 4d, and (2) he ¼ 8dðd ¼ 16 mmÞ]. When testing
was measured using a log-grader and found to be E ¼ 15.9 GPa for edge distances larger than 4d, a single bolt often causes signifi-
(mean value; the manufacturer specifies 13.2 GPa for the supplied cant embedment, in which case two closely spaced bolts were used

Fig. 8. Glulam test specimens

Fig. 9. Laminated veneer lumber test specimens

© ASCE 04015026-6 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., 2015, 27(11): 04015026


Downloaded from ascelibrary.org by GADJAH MADA UNIVERSITY on 02/23/16. Copyright ASCE. For personal use only; all rights reserved.

Fig. 12. Plate specimen

R δ¼δ0
mgδ0 þ δ¼0 Pdδ
Gf ¼ ð19Þ
A
Further, so-called plate specimen tests were conducted on spec-
Fig. 10. Perpendicular-to-grain tensile strength specimen imens as shown in Fig. 12. Specimen dimensions as recommended
by Yasumura (2002) were used. The C1 factor may be determined
from Eq. (20) as per Yasumura (2002), where Pu is the measured
failure load. Thirty-three replicates were tested for the glulam and
to increase the bearing area. Two closely spaced bolts are known 15 replicates were tested for the LVL
to give essentially the same splitting failure load as a single bolt
Pu
[Kasim and Quenneville (2002)]. For the glulam specimens, it C1 ¼ pffiffiffiffiffi ð20Þ
was chosen to try a configuration with two bolts aligned along 2b he
the grain as well as a configuration with two bolts aligned across
the grain for specimens with he ¼ 8d. For the LVL specimens, it
was chosen to use two bolts very closely spaced along the grain for Results
all tests. For the glulam, only a limited number of replicates were
tested; six replicates of D-M-4 and three replicates of D-M-8 H, Beam Tests
D-M-8 V, D-Q-4, and D-Q-8. For the LVL, 10 replicates were
tested of all specimens. The mean failure loads and the coefficient of variation (COV) of the
tested glulam and LVL beam specimens are shown in Table 1.
The COV values are not reliable for beams tested with only three
Material Property Tests replications.
Tests were conducted on specimens as shown in Fig. 10 in agree-
ment with EN 408 (CEN 2003) to determine the perpendicular- Material Property Tests
to-grain tensile strength of the glulam and the LVL. Forty-two The perpendicular-to-grain tensile strength tests as shown in Fig. 10
replicates were tested of the glulam and 15 replicates were tested resulted in a mean tensile strength of ft ¼ 2.3 MPa with a coeffi-
of the LVL. The perpendicular-to-grain tensile strength, ft , is cient of variation of 30% for the glulam, and ft ¼ 1.5 MPa with a
determined as f t ¼ Pu =ð45 × 70 mm2 Þ. coefficient of variation of 12% for the LVL.
Tests for determination of Mode I fracture energy of the LVL The Mode I fracture energy tests as shown in Fig. 11 resulted in
were conducted on specimens as shown in Fig. 11 with dimensions a mean fracture energy of Gf ¼ 0.95 N=mm with a coefficient of
in accordance with Gustafsson (2003). Fifteen replicates were variation of 15% for the LVL. The results are based on 13 speci-
tested. The fracture energy, Gf , is calculated as given by Eq. (19), mens since two out of the tested 15 specimens had to be excluded
where P and δ are the recorded load and displacement of the cross due to improper data recordings. The Mode I fracture energy was
head of the testing machine, respectively; m is 7=8 of the mass of not determined for the glulam.
the specimen; g (9.81 m=s2 ) is the gravity acceleration; δ 0 is the
displacement at which the load returns to zero; and A is the frac-
tured area (in this context 24 × 45 mm2 ). The integral term repre- Table 1. Failure Loads of Glulam and LVL Beams
sents the work done by the testing machine, i.e., the area under the Mean failure Coefficient
measured load–displacement curve Specimen Na load (kN) of variation (%)
D-M-4 6 9.3 11
D-M-8 H 3 16.8 14
D-M-8 V 3 18.5 13
D-Q-4 3 9.7 5
D-Q-8 3 20.0 18
R-M-4 10 20.3 13
R-M-8 10 36.6 9
R-Q-4 10 19.3 8
R-Q-8 10 37.0 10
Fig. 11. Mode I fracture energy test specimen a
N = number of replicates.

© ASCE 04015026-7 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., 2015, 27(11): 04015026


The plate specimen tests as shown in Fig. 12 resulted in a mean Table 3. Theoretical to Experimental Failure Load Ratios
value of C1 ¼ 22.7 N=mm1.5 with a coefficient of variation of 8% Specimen he (mm) Na Eqs. (12) and (12b) Eqs. (15a)–(15e)
for the LVL, and C1 ¼ 12.7 N=mm1.5 with a coefficient of varia-
D-M-4 64 6 1.11 —
tion 17% for the glulam.
D-M-8 H 128 3 1.02 —
D-M-8 V 128 3 0.92 —
D-Q-4 64 3 1.06 —
Discussion D-Q-8 128 3 0.85 —
R-M-4 64 10 0.90 1.04
In the European [EN 1995-1-1:2004 (CEN 2004)] and R-M-8 128 10 0.83 1.05
Canadian [CSA O86-09 (Canadian Standards Association 2009)] R-Q-4 64 10 0.95 1.10
timber design codes [Eqs. (16a)–(16d)], a constant characteristic R-Q-8 128 10 0.83 1.03
value C1 ¼ 14 N=mm1.5 is assumed irrespective of the wood a
N = number of replicates.
Downloaded from ascelibrary.org by GADJAH MADA UNIVERSITY on 02/23/16. Copyright ASCE. For personal use only; all rights reserved.

species and product. The plate specimen tests resulted in mean C1


values of 22.7 and 12.7 N=mm1.5 for Radiata pine LVL and Douglas
fir glulam, respectively. These values strongly indicate that a charac- obtained mean values for the glulam and LVL beams are shown in
teristic value of C1 ¼ 14 N=mm1.5 , as used in the European [EN Table 3. In Eqs. (12a) and (12b), the mean values of the C1 factors
1995-1-1:2004 (CEN 2004)] and Canadian [CSA O86-09 (Canadian obtained from the plate specimen tests and Eq. (20) have been used.
Standards Association 2009)] timber design codes, may be signifi- In Eqs. (15a)–(15e), the measured values of the perpendicular-to-
cantly nonconservative for certain wood species and products and grain tensile strength, ft ¼ 1.5 MPa, and the fracture energy,
very conservative for other species and products. The C1 values used Gf ¼ 0.95 N=mm, have been used. No theoretical values are given
in design codes should reflect the actual differences in wood species for the Douglas fir glulam beams using Eqs. (15a)–(15e) since the
and products in order to ensure fair competition. fracture energy of the Douglas fir glulam was not measured.
The large difference in splitting capacity between Radiata pine Eqs. (12a) and (12b) is in principle only valid for beams loaded
LVL and Douglas fir glulam as observed in the plate specimen tests at midspan, but since the results in Table 2 indicate that the location
is consistently also observed in the beam tests in Table 1. of the connection is of no influence, Eqs. (12a) and (12b) is also
The splitting capacity is according to Eqs. (17a)–(17d) in the applied to the beams loaded at quarter-span.
German timber design code [DIN 1052:2008 (DIN 2008)] assumed Table 3 indicates that Eqs. (12a), (12b), and (15a)–(15e) pro-
to be proportional to the perpendicular-to-grain tensile strength. duce acceptable results for design purposes. Eqs. (15a)–(15e)
The material property tests showed, however, that Douglas fir produce results in better agreement with experiments than
glulam has a higher perpendicular-to-grain tensile strength than Eqs. (12a) and (12b), but slightly on the nonconservative side.
Radiata pine LVL, but the latter performs nevertheless much better The predicted to experimental mean failure load ratios for the
in terms of splitting capacity as seen from the plate specimen tests LVL beams using Eqs. (8)–(10), (12a), (12b), (13a)–(13e), and
as well as from the beam test results in Table 1. (15a)–(15e) are given in Table 4. The ratios are given for predicted
The mean COV is 10% for the tested LVL beams presented in failure loads calculated by using two different estimates of the frac-
Table 1. Experiments on wood involving perpendicular-to-grain ture energy. Superscripts indicate values estimated from the plate
loading typically exhibit COVs between 15 and 25%. For splitting specimen (Fig. 12) tests, i.e., Eq. (20), and the definition of C1 as
of beams due to loading perpendicular-to-grain, it is difficult to find given in Eq. (12), and values measured directly by means of the test
reliable data since most test series reported only contain two to specimens (Fig. 11). The fracture energy estimated from the plate
three replications. The probably most comprehensive collection of specimen tests and determined by Eqs. (20), (12a), and (12b)
data for splitting of beams loaded perpendicular to grain by con- is closely related to the model given by Eqs. (10) or (12a) and
nections is found in Schoenmakers (2010). (12b); they should therefore in principle only be used in Eqs. (10),
Experimentally obtained ratios of the failure loads of the beams (12a), and (12b).
loaded at midspan to those loaded at quarter-span for the two tested According to Table 4, identical failure loads are predicted by
edge distances are given in Table 2. In the European timber design means of Eqs. (10), (12a), and (12b), and the predictions are in
code [EN 1995-1-1:2004 (CEN 2004)], the maximum shear force is fair agreement with the experimental results if the fracture energy
considered as shown in Eqs. (16a)–(16d), leading to a ratio of 1.5. is estimated from the plate specimen tests and use of Eqs. (20),
As seen from Table 2, the experimentally obtained ratios are very (12a), and (12b). However, the agreement between the models
close to 1.0, which means that the location of a connection in the and the experiments is not convincing for Eqs. (10), (12a), and
span of a simply supported beam is of no influence, and the ap- (12b) if using the directly measured fracture energy.
proach used in the European timber design code [EN 1995-1- Eqs. (13a)–(13e) gives predictions in fair agreement with the
1:2004 (CEN 2004)], i.e., considering the maximum shear force, experiments if using the directly measured fracture energy (spec-
is not justified. imens shown in Fig. 11). As should be expected, Eqs. (13a)–(13e)
Ratios of the theoretically predicted failure loads determined by produces good results for small edge distances (specimens R-M-4
means of Eqs. (12a), (12b), and (15a)–(15e) to the experimentally and R-Q-4), while the results are less good for large edge distances
(specimens R-M-8 and R-Q-8). Eqs. (13a)–(13e) lead to poor
agreement with experiments if using the fracture energy as esti-
Table 2. Midspan to Quarter-Span Failure Load Ratios mated from the plate specimens. Eqs. (15a)–(15e) are in good
Theoretical as per agreement with the experimental results if using the directly mea-
Edge distance Experimental EC5 (CEN 2004) sured fracture energy. The proposed semiempirical modification of
Eqs. (13a)–(13e) in order to take into account the effect of the total
Douglas Radiata
he (mm) fir glulam pine LVL Eqs. (16a)–(16d) beam depth has in this context led to a significant improvement of
the predictions for large edge distances.
64 0.96 1.06 1.5 Table 4 also shows that the models given by Eqs. (8) and (9)
128 0.92 0.99 1.5
show a different trend than Eqs. (10), (12a), and (12b); namely,

© ASCE 04015026-8 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., 2015, 27(11): 04015026


Table 4. Theoretical to Experimental Failure Load Ratios for LVL Beams
Specimen Eq. (8) Eq. (9) Eq. (10) Eqs. (12) and (12b) Eqs. (13a)–(13e) Eqs. (15a)–(15e)
R-M-4a 0.68 0.68 0.90 0.90 0.66 0.74
R-M-8a 0.54 0.54 0.83 0.83 0.55 0.73
R-Q-4a 0.72 0.72 0.95 0.95 0.69 0.78
R-Q-8a 0.53 0.53 0.83 0.83 0.55 0.72
R-M-4b 1.06 1.06 1.41 1.41 0.92 1.04
R-M-8b 0.84 0.84 1.30 1.30 0.79 1.05
R-Q-4b 1.12 1.12 1.49 1.49 0.97 1.10
R-Q-8b 0.83 0.83 1.29 1.29 0.78 1.03
a
Fracture energy determined by means of plate specimen (Fig. 12) and Eq. (20).
b
Fracture energy determined directly by specimen shown in Fig. 11.
Downloaded from ascelibrary.org by GADJAH MADA UNIVERSITY on 02/23/16. Copyright ASCE. For personal use only; all rights reserved.

better agreement with the experiments if using the directly mea- order to take into account the total beam depth results in a signifi-
sured fracture energy as compared with using the fracture energy cant improvement of the predictions for larger edge distances as
determined by means of the plate specimens. This may be said to be compared with Eq. (14).
surprising since Eqs. (8) and (9) are just special cases of Eqs. (10), Hindman et al. (2010) and Patel and Hindman (2012) present
(12a), and (12b). experimental results of splitting tests on various wood products
Eqs. (10), (12a), and (12b) produce fair predictions as long as for comparison of Eqs. (12a) and (12b) with Eqs. (13a)–(13e).
the C1 factor is determined from the plate specimen tests and The conclusion is that Eqs. (13a)–(13e) perform better than
Eq. (20). This should be expected since Eq. (20) stems from Eqs. (12a) and (12b).
Eqs. (12a) and (12b). The fact that Eqs. (10), (12a), and (12b) Judged from the limited results available in Tables 3 and 4 and
do not lead to good agreement with experiments if using the di- from the results presented in Jensen et al. (2012), the proposed
rectly measured value of the fracture energy seems to suggest that semiempirically generalized model as given by Eqs. (15a)–(15e)
the linear elastic fracture mechanics model, on which Eqs. (10), seems to produce convincing and significantly improved predic-
(12a), and (12b) are based, may contain some deficiencies. tions as compared with Eqs. (13a)–(13e). Since the results pre-
The experimental failure loads of the tested LVL beams are in sented in Hindman et al. (2010) and Patel and Hindman (2012)
Fig. 13 compared with the predicted failure loads as given by suggest that Eqs. (13a)–(13e) perform better than Eqs. (12a)
Eqs. (12a), (12b), (13a)–(13e), and (15a)–(15e). The predictions and (12b), Eqs. (15a)–(15e) thus seems to produce better results
given by Eqs. (12a) and (12b) are based on the fracture energy de- than do Eqs. (12) and (12b). The model should, however, be tested
termined by means of plate specimens as shown in Fig. 12 while the on other wood species and products before conclusive statements
predictions given by Eqs. (13a)–(13e), and (15a)–(15e) are based are made.
on the fracture energy test specimens shown in Fig. 11. The data
presented in Fig. 13 suggest that Eqs. (15a)–(15e) lead to better
predictions than Eqs. (12a), (12b), and (13a)–(13e), and that the Conclusions
semiempirical modification of Eqs. (13a)–(13e) successfully brings Currently available analytical fracture mechanics models for pre-
the predictions given by Eqs. (15a)–(15e) in agreement with the diction of the splitting failure load of beams loaded perpendicular
experimental results for larger (8d) edge distances. to grain by connections were reviewed. It was shown that most
Jensen et al. (2012) present experimental results for splitting test of the available models are not independent despite having been
for connections consisting of single bolts loading a Radiata pine derived independently by various researchers, but are related
LVL beam perpendicular to the grain near the beam end. Very good and appear as special cases of the most general models. A new
agreement between experimental results and the generalized semiempirical generalization was proposed, which unifies two
version of Eq. (14) is reported. The findings in Jensen et al. classes of models, namely (1) the linear elastic fracture mechanics
(2012) suggest that the semiempirical modification introduced in models, and (2) the so-called quasi-nonlinear fracture mechanics
models.
Splitting tests were conducted on beams made of Douglas fir
glulam and Radiata pine LVL. The beams were loaded by bolted
connections located either at midspan or at quarter-span. The
Radiata pine LVL beams showed roughly twice the splitting capac-
ity as compared with the Douglas fir glulam beams. The tests also
showed that the location of the connection in the span of the beam
does not affect the splitting failure load; the same failure load was
obtained for beams loaded at midspan and at quarter-span for given
edge distances.
Perpendicular-to-grain tensile strength tests were also con-
ducted for the Douglas fir glulam and the Radiata pine LVL used
for the beam tests. The tests resulted in a lower mean value of the
perpendicular-to-grain tensile strength for the Radiata pine LVL
than for the Douglas fir glulam.
Theoretically predicted splitting failure loads using the pro-
Fig. 13. Experimental failure loads for LVL beams compared with
posed new semiempirically modified quasi-nonlinear fracture me-
Eqs. (12a), (12b), (13a)–(13e), and (15a)–(15e)
chanics model were compared with the experimentally determined

© ASCE 04015026-9 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., 2015, 27(11): 04015026


ones, and very good agreement was obtained. The experimental Gustafsson, P. J. (2003). “Fracture perpendicular to grain—Structural
data available are, however, too limited to conclude whether the applications.” Timber engineering, S. Landersson and H. J. Larsen,
good agreement is a coincidence. Further testing is needed to ex- eds., Wiley, Chichester, U.K.
plore the capacity of the proposed model. Hindman, D. P., Finkenbinder, D. E., Loferski, J. R., and Line, P. (2010).
In the European timber design code, a concept of considering “Strength of sawn lumber and wood composite connections loaded
the maximum shear force on either side of a connection rather than perpendicular to grain: Fracture mechanics equations.” J. Mater. Civ.
Eng., 10.1061/(ASCE)MT.1943-5533.0000141, 1226–1234.
considering the total load on the connection is applied. According
Jensen, J. L. (2005a). “Quasi-non-linear fracture mechanics analysis
to this concept, the splitting capacity of a beam loaded at midspan is
of splitting failure in moment-resisting dowel joints.” J. Wood Sci.,
50% higher than the capacity of the same beam loaded at quarter- 51(6), 583–588.
span. This, however, is in disagreement with experimental results, Jensen, J. L. (2005b). “Quasi-non-linear fracture mechanics analysis of
which unanimously show that the capacity is the same for beams splitting failure in simply supported beams loaded perpendicular to
loaded at midspan and at quarter-span. grain by dowel joints.” J. Wood Sci., 51(6), 577–582.
Downloaded from ascelibrary.org by GADJAH MADA UNIVERSITY on 02/23/16. Copyright ASCE. For personal use only; all rights reserved.

It is in the German timber design code assumed that the splitting Jensen, J. L. (2005c). “Quasi-non-linear fracture mechanics analysis of
capacity of a timber beam is proportional to the perpendicular-to- the splitting failure of single dowel joints loaded perpendicular to
grain tensile strength of the wood, and that the perpendicular- grain.” J. Wood Sci., 51(6), 559–565.
to-grain tensile strength is the only strength parameter affecting the Jensen, J. L. (2005d). “Splitting strength of beams loaded perpendicular to
splitting capacity. However, the research reported in this paper grain by dowel joints.” J. Wood Sci., 51(5), 480–485.
show that the mean splitting failure capacity of Radiata pine Jensen, J. L., Gustafsson, P. J., and Larsen, H. J. (2003). “A tension fracture
LVL was approximately twice that of Douglas fir glulam despite model for joints with rods or dowels loaded perpendicular to grain.”
the fact that the measured mean perpendicular-to-grain tensile Proc., Int. Council for Research and Innovation in Building and Con-
strength of the Radiata pine LVL was lower than that of the Douglas struction, Rotterdam, Netherlands.
fir. This indicates that fracture energy may play an important role. Jensen, J. L., Quenneville, P., Girhammar, U. A., and Källsner, B. (2012).
The proposed semiempirical generalization of a quasi-nonlinear “Splitting of timber beams loaded perpendicular to grain by
fracture mechanics model, which predicts the influence of both connections—Combined effect of edge and end distance.” Constr.
Build. Mater., 35, 289–293.
the perpendicular-to-grain tensile strength and the fracture energy,
Kasim, M., and Quenneville, P. (2002). “Effect of row spacing on the
may be a key to understanding the splitting behavior of different
capacity of bolted timber connections loaded perpendicular-to-grain.”
wood species and products. Proc., Int. Council for Research and Innovation in Building and Con-
struction, Rotterdam, Netherlands.
Larsen, H. J., and Gustafsson, P. J. (2001). “Dowel joints loaded
Acknowledgments
perpendicular to grain.” Proc., Int. Council for Research and Innovation
Funding of the research reported in this paper was provided by the in Building and Construction, Rotterdam, Netherlands.
Patel, M. C., and Hindman, D. P. (2012). “Comparison of single- and
New Zealand Structural Timber Innovation Company (STIC), by
two-bolted LVL perpendicular-to-grain connections. II: Fracture mod-
the County Administrative Board in Norrbotten, the Regional
els.” J. Mater. Civ. Eng., 10.1061/(ASCE)MT.1943-5533.0000304,
Council of Västerbotten (Sweden), and by the European Union’s
339–346.
Structural Funds (the Regional Fund). The funding is greatly Schoenmakers, J. C. M. (2010). “Fracture and failure mechanisms in timber
appreciated. The work done by undergraduate students Rui Li, loaded perpendicular to grain by mechanical connections.” Ph.D. thesis,
Nelson Veerasingam, Rayan Hoshino, and HarithBarakat, who Univ. of Technology, Eindhoven, Netherlands.
did much of the experimental aspect of the research reported in this Serrano, E., and Gustafsson, P. J. (2006). “Fracture mechanics in timber
paper during their final-year projects at the University of Auckland, engineering–Strength analyses of components and joints.” Mater.
is acknowledged and appreciated. Struct., 40(1), 87–96.
van der Put, T. A. C. M. (2013a). “Discussion of ‘Strength of sawn lumber
and wood composite dowel connections loaded perpendicular to
References grain. I: NDS design equations’ by Daniel P. Hindman, David E.
Finkenbinder, Joseph R. Loferski, and Philip Line.” J. Mater. Civ.
Ballerini, M., and Rizzi, M. (2007). “Numerical analysis for the prediction
Eng., 10.1061/(ASCE)MT.1943-5533.0000355, 1143–1147.
of the splitting strength of beams loaded perpendicular-to-grain by
van der Put, T. A. C. M. (2013b). “Discussion of ‘Strength of sawn
dowel-type connections.” Mater Struct., 40(1), 139–149.
lumber and wood composite dowel connections loaded perpen-
Canadian Standards Association. (2009). “Engineering design in wood.”
dicular to grain. II: Fracture mechanics equations’ by Daniel P.
CSA O86-09, Toronto.
CEN (Comité Européen de Normalisation). (2003). “Timber structures— Hindman, David E. Finkenbinder, Joseph R. Loferski, and Philip
Structural timber and glue laminated timber—Determination of some Line.” J. Mater. Civ. Eng., 10.1061/(ASCE)MT.1943-5533.0000317,
physical and mechanical properties.” EN 408, Brussels, Belgium. 1147–1150.
CEN (Comité Européen de Normalisation). (2004). “Design of timber van der Put, T. A. C. M., and Leijten, A. J. M. (2000). “Evaluation of
structures.” EN 1995-1-1:2004, Brussels, Belgium. perpendicular to grain failure of beams caused by concentrated loads
DIN (Deutsches Institut für Normung). (2008). “Draft. Analysis and design of joints.” Proc., Int. Council for Research and Innovation in Building
of timber structures—Basic rules for analysis and design of building and Construction, Rotterdam, Netherlands.
structures.” DIN 1052:2008, Berlin. Yasumura, M. (2002). “Determination of fracture parameter for dowel-type
Gustafsson, P. J. (1988). “A study of strength of notched beams.” Proc., Int. joints loaded perpendicular to wooden grain and its application.” Proc.,
Council for Research and Innovation in Building and Construction, Int. Council for Research and Innovation in Building and Construction,
Rotterdam, Netherlands. Rotterdam, Netherlands.

© ASCE 04015026-10 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., 2015, 27(11): 04015026

S-ar putea să vă placă și