Download as pdf or txt
Download as pdf or txt
You are on page 1of 23

VII International Conference on Computational Plasticity

COMPLAS 2003
E. Oñate and D.R.J. Owen (Eds)
c CIMNE, Barcelona, 2003

MULTI-SURFACE PLASTICITY OF
CLEAR SPRUCE WOOD

Peter Mackenzie-Helnwein, Josef Eberhardsteiner and Herbert A. Mang ?

Institute for Strength of Materials


Vienna University of Technology
Karlsplatz 13, A-1040 Vienna, Austria
e-mail: Peter.Mackenzie-Helnwein@tuwien.ac.at
web page: www.fest.tuwien.ac.at

Key words: spruce wood, orthotropic material, multi-surface plasticity, biaxial loading

Abstract. Biaxial experiments on spruce wood show that consideration of an elliptic fail-
ure surface according to Tsai & Wu, and of an elastic model for stress states within this
envelope, gives an insufficient description of the mechanical behavior. As compression
perpendicular to grain occurs, a nonlinear stress path results from a proportional biax-
ial strain path. Moreover, a phenomenological single-surface model does not permit easy
identification of failure modes and thus renders the description of different post-failure
mechanisms very difficult. Investigation of characteristic samples for various biaxial load-
ing conditions enables the identification of four basic mechanisms covering the behavior
of wood under plane stress conditions.
The experimentally observed mechanical behavior will be described by means of a multi-
surface plasticity model. It consists of four surfaces representing four basic failure modes.
The first is a modified tension cut-off for the description of fiber rupture. The second is
a mixed mode normal tension-shear model by Weihe applied to the perpendicular to grain
direction. The third is an extension of the authors’ prior model for perpendicular to grain
compression, and the fourth surface covers the compressive failure parallel to grain. The
model represents the orthotropic multi-surface elasto-plastic material clear wood.
The aim of this paper is to present and discuss selected experimental data from biaxial
tests with respect to distinct failure modes, and to develop an orthotropic plasticity model
for its mathematical description. Since available experimental data cover only plane stress
in the LR-plane, both orthotropic failure and yield surfaces, respectively, are restricted to
this case.

1
Peter Mackenzie-Helnwein, Josef Eberhardsteiner and Herbert A. Mang

1 INTRODUCTION
From the mechanical point of view, wood exhibits a pronounced orthotropic behavior
with large ratios of mechanical properties such as Young’s modulus or strength between
the respective values parallel and transverse to the grain direction. Furthermore, it is well
known that strength differs in tension and compression.
Experimental data for the strength of spruce wood under biaxial loading have been
given by Spengler [22], Ehlbeck and Hemmer [8], Hemmer [13], and Eberhardsteiner [5].
The work by Spengler [22], and Ehlbeck and Hemmer [8], Hemmer [13] are restricted
to the observation of failure locations in stress space without accompanying deformation
measurement. Their tests cover the characteristics in the LT -plane1 and the interaction
between perpendicular to grain normal stress and shear stress (LR-plane).
Filling the gap on biaxial strength tests in the LR-plane, as well as the measurement of
the stress-strain behavior under biaxial loading in the pre-failure domain, were the design
goals for a novel experiment by Eberhardsteiner [5], Eberhardsteiner et al. [6, 7], and
Gingerl [11]. Approximately 450 individual tests cover the whole set of distinguishable
stress states for an orthotropic material under plane stress conditions. The significant
difference to any previously performed experiment is that both the biaxial stress and the
strain history have been measured for each individual test. This makes these tests an
ideal basis for constitutive modeling.
The most common model for the mathematical description of yield surfaces and fail-
ure surfaces of anisotropic materials is the elliptic failure surface according to Tsai and
Wu [23]. Applying it to the experimentally obtained failure locations for spruce wood,
Eberhardsteiner [5] obtained a reasonably good fit for the failure locations. However,
the shortcoming of a single-surface model according to Tsai and Wu [23] is its inabil-
ity to distinguish between different failure modes. The following approach is based on
the experimental data by Eberhardsteiner [5] but replaces the single-surface model by a
micro-mechanically motivated multi-surface plasticity model.
A first attempt in this direction has been presented by Helnwein et al. [12]. They
introduced a second surface for the description of non-linear phenomena due to radial
compression. This paper will present a multi-surface orthotropic plasticity model for bi-
axially stressed clear wood based on experimental work, micromechanical considerations,
and their macroscopic continuum mechanics interpretation. This strategy enables the easy
identification of a failure mode. Moreover, it will enable a comparably simple modeling
of the respective post-failure behavior.
Section 2 presents a brief overview of the experimental basis and presents test results
relevant for the subsequent development. In Section 3, the experimentally observed mech-
anisms will be discussed and a respective mathematical model will be presented for each
of them. After remarks on the numerical implementation in Section 4, the simulation of
a cross-wise joint will be presented in Section 5.
1
L is the longitudinal or grain direction, R is the radial, and T is the tangential direction of the stem.

2
Peter Mackenzie-Helnwein, Josef Eberhardsteiner and Herbert A. Mang

2 EXPERIMENTAL INVESTIGATION OF SPRUCE WOOD UNDER BI-


AXIAL LOADING
Realistic finite element ultimate load analyses of structural details as well as of shell
structures made of wood require suitable constitutive equations for the prediction of
the deformational behavior of biaxially loaded solid wood. However, there is a lack of
adequate biaxial experimental data, particularly if loading situations are considered where
the principal loading directions do not coincide with the principal material directions.
The first relevant material tests on spruce wood under complex loading conditions
were performed and reported in the mid 1980’s. Spengler [22] investigated loading zones
of beams. The specimen used was a small plate subjected to shear stress under superposed
lateral compression. The results were restricted to the failure locations for this particular
set of load cases. No deformation measurement has been carried out. Ehlbeck and
Hemmer [8], Hemmer [13] used cylindrical pipes subjected to tension, torque, internal
pressure, and combinations of it. These experiments cover strength characteristics in the
LT -plane. No deformation information was reported under combined loading.
Constitutive modeling of the biaxial behavior requires knowledge of both, the stress-
strain relations in the pre-failure domain as well as the failure locations for arbitrary
strain paths. To fulfill both requirements, a comprehensive experimental investigation
of the stiffness and strength behavior of wood under arbitrary two-dimensional loading
was carried out by Eberhardsteiner et al. [6, 7], Gingerl [11]. Based on a cruciform
specimen made out of clear spruce wood without imperfections such as knots or any visible
irregularities (see Fig. 1) and a moisture content of u = 12 %, a series of approximately
450 displacement-driven biaxial strength tests were performed. These tests cover the
whole set of distinguishable stress states for an orthotropic material under plane stress
conditions in the LR-plane.

( −1) ( 0) ( +1)
Uv
V
1 vU2 V=v̄ UvV 3 X2

DV +vX1
X1 VV
u1 u2 u3 +u
X2 VH

UH
( +1) 2, v u4U(H−1)
v4
DH
U(H0 ) uU5(H0=ū
)

ϕϕ 1, u v5 DH
U(H−1) u(H+16)
U
v6
VH X2

VV X1
X1 DV

( +1) ( 0)
X2 UV UV UV( −1)

Figure 1: Wooden specimen mounted in Figure 2: Applied displacement compo-


the biaxial testing device nents for κ = ū : v̄

The developed testing equipment, described in detail by Eberhardsteiner [5], consists

3
Peter Mackenzie-Helnwein, Josef Eberhardsteiner and Herbert A. Mang

of a biaxial servo-hydraulic testing device for anisotropic materials and of a 3D Electronic


Speckle Pattern Interferometer (3D ESPI) for three-dimensional contactless full-field de-
formation analysis. The latter was used to measure the incremental in-plane deforma-
tions and its distribution in the central region of the specimen. Based on the obtained
displacement field the state of strain and the strain history were determined by numerical
differentiation.
The thickness t of the specimen’s measuring area (140×140 mm) has been chosen to be
4.5 mm for states of biaxial tension and combined tensile and (small) compressive stresses.
In order to prevent buckling of the specimen, biaxial compression tests were performed
on specimens with t = 7.5-9.5 mm. The prescribed displacements were applied by means
of 24 servo-hydraulic devices applied to the 12 loading points as shown in Fig. 1. This
enables individual control of the in-plane position for each of the loading points with an
accuracy of 2-3 µm.
The mechanical parameters of the experimental investigation are: (i) The angle ϕ
between principal loading direction and grain direction (ϕ = 0◦ , 7.5◦ , 15◦ , 30◦ , and 45◦ ).
(ii) The ratio κ defined as the displacement ratio ū : v̄, with ū and v̄ according to Fig. 2.
This ratio is in good correlation to the ratio ε11 /ε22 , with ε11 and ε22 being the axial
components of strain along the horizontal and vertical axis of the specimen, respectively.
For grain angles other than zero this involves a non-zero shear strain γ12 as indicated in
Fig. 2.
Most of the displacement-driven experiments were characterized by a proportional
stepwise loading until fracture was reached. The displacement steps applied at each of
the 12 discrete load application points of the specimen varied from 4 to 10 µm. In addition
to the proportional biaxial strength tests, a small number of tests with loading-unloading-
reloading cycles at different load (displacement) levels was performed.
Fig. 3 contains the collection of all experiments performed at grain angles of ϕ = 0◦
and ϕ = 15◦ , respectively. Similar diagrams for grain angles of ϕ = 7.5◦ , 30◦ , and 45◦
are available in Eberhardsteiner [5]. The axis values are the principal stresses, σ1 and
ϕ = 0◦ ϕ = 15◦
10 10

5 5
σ2 [MPa]

σ2 [MPa]

0 0

-5 -5

-10 -10
-50 -25 0 25 50 75 -40 -20 0 20 40
σ1 [MPa] σ1 [MPa]

Figure 3: Biaxial strength data and evolution of principal stress ratio σ2 /σ1 for grain directions
ϕ = 0◦ and ϕ = 15◦

4
Peter Mackenzie-Helnwein, Josef Eberhardsteiner and Herbert A. Mang

σ2 , which are aligned with the horizontal and the vertical axis of the test specimen.
Each line represents the stress path obtained from a single test under proportionally
increasing prescribed displacements. The symbol at the end of each line indicates material
failure, which has been defined as the initial appearance of an extreme value on any of
the two measured components of stress. This definition of material failure underlies
the assumption that once the measurement area of the specimen is no longer under a
homogeneous state of deformation, the influence of the loading zones becomes significant
with respect to the observed forces at the loading points.
0 75
100
σ1 ϕ = 0◦ σ1 ϕ = 0◦
κ = −4 : +5 κ = +1 : 0
144
-25 0
0 -0.002 0.004
293 -0.002 ε 0.004 ε

σ1
ϕ = 7.5◦
-30 κ = −10 : +3 σ22 [MPa]
-0.006 ε 0.010
0 ϕ = 15◦
266 κ = −3 : −5
σ1
σ11
-25
-0.015 ε 0.030
b

c a

ϕ = 30◦ ϕ = 0◦
κ = 0 : −1 κ = 0 : −1
0 0
426 998
ε11
σ1 σ2 ε22
ε12
-7 -6
-0.015 ε 0.010 -0.015 ε 0.002

Figure 4: Elliptic failure envelopes by Tsai and Wu [23], and characteristic stress-strain curves
and fracture types for different biaxial loading situations

The obtained failure locations reveal an elliptic shape of the failure envelope with
very little scatter in the compressive domain as well in most parts of the mixed tensile-

5
Peter Mackenzie-Helnwein, Josef Eberhardsteiner and Herbert A. Mang

compressive domain. Based on this observation Eberhardsteiner [5] proposed a fit with a
second-order tensor polynomial according to Tsai and Wu [23] as a first approximation of
the actual material behavior. In Fig. 4 three different elliptic failure envelopes are shown.
The curves “a” and “b” are based on strength data obtained for 0◦ ≤ ϕ ≤ 45◦ , whereas
curve “c” is only valid for strength data with respect to grain angle ϕ = 0◦ . In contrast
to curve “a” where all available strength data were considered, curve “b” resulted from
an exclusion of the extreme values of the six obtained biaxial strength data for each test
configuration {ϕ ; κ}.
Analyzing the stress-strain relations for different load ratios reveals four basic modes of
failure:
(i) Brittle tensile failure in fiber direction. This mode has a cascading crack pattern as
shown in the upper right image in Fig. 4. This mode shows a large variation of strength.
The elliptical failure envelopes tend to overestimate tensile strength in the shaded domain
of the stress space. This mode can only be found for small grain angles ϕ.
(ii) Brittle tensile failure perpendicular to grain. A distinct straight crack develops
parallel to grain as shown in the upper left image in Fig. 4. This failure mode can be
found for all grain angles ϕ.
(iii) Ductile compressive behavior perpendicular to grain. A behavior very similar to
saturation type hardening as known from metal plasticity can be observed from the bottom
image of Fig. 4. Compressive loading precisely perpendicular to grain does not cause
strength degradation. On the macroscopic level, the deformation occurs homogeneous.
On the microscopic level, however, it appears localized in single rows of cells.
(iv) Compressive failure in fiber direction. Bands of damaged cells develop. Strength
degradation to 70–80 % of the initial failure stress occurs (see the two diagrams on the
upper left side of Fig. 4). Uniaxial compressive tests show a stress plateau after an initial
strength degradation.
At large compressive deformations, both modes (iii) and (iv) show a phenomenon called
densification, where the strength rises and the material becomes almost rigid. This phe-
nomenon cannot be observed within the deformation range of the biaxial testing device.

3 MICROMECHANICALLY MOTIVATED MULTI-SURFACE PLASTIC-


ITY MODEL FOR WOOD UNDER BIAXIAL LOADING
The work presented by Eberhardsteiner [5] covers the orthotropic elastic behavior as
well as challenging identification of the onset of failure under biaxial loading. The single-
surface failure envelope suggested in that work has been extended by a hardening type
plasticity model for the description of early compaction perpendicular to grain by Hel-
nwein et al. [12] and Mackenzie-Helnwein et al. [16].
This paper extends the latter approach to a micro-mechanically motivated multi-surface
plasticity model. Moreover, the originally suggested single-surface model according to
Tsai and Wu [23] will be dropped.

6
Peter Mackenzie-Helnwein, Josef Eberhardsteiner and Herbert A. Mang

The following subsections will address the elastic domain and each of the considered
failure and yield criteria.

3.1 Orthotropic invariants for plane stress


For the mathematical description of orthotropic tensor functions needed for the subse-
quent development, one has to identify a complete set of linearly independent orthotropic
invariants. These are constructed as simultaneous invariants of a tensor, e.g., σ and the
structural tensors [2, 20, 26]. ML := AL ⊗ AL , MR := AR ⊗ AR , and MT := AT ⊗ AT ,
where AL , AR , and AT are unit normal vectors locally aligned with the L (grain), R
(radial), and T (tangential) directions of the tree as shown in Fig. 5.

A3 = A T
A 2 = AR

A1 = A L

Figure 5: Material orientation and definition of the reference frame aligned to the growth di-
rections of the tree

Restricting the problem of interest to plane stress in the LR-plane, one suitable set of
invariants is

σL = tr σML , σR = tr σMR , τLR 2 = tr σMR σML . (1)

Note that the shear component appears squared, i.e., its sign is meaningless under or-
thotropic symmetry transformations.
For the described orientation of the material, a set of strain invariants identical to
those given in (1) can be found for the strain tensor. Using engineering strain for shear
deformations, we obtain

εL = tr εML , εR = tr εMR , γLR 2 = 4 tr εMR εML . (2)

There exists another strain invariant εT = tr εMT . Utilizing the plane stress condition
σT = τRT = τT L = 0, it can be eliminated from the list of arguments of orthotropic
constitutive functions.
Having a system of independent invariants, any scalar valued orthotropic function can
be expressed as a polynomial of these invariants [2, 4, 20, 21, 26]. This is utilized in the
mathematical description of the Helmholtz free energy as

ψ = ψ̃(ε) = ψ̂(εL , εR , γLR 2 ) . (3)

7
Peter Mackenzie-Helnwein, Josef Eberhardsteiner and Herbert A. Mang

Furthermore, it appears with any type of yield function or failure condition fi (σ) in stress
space. Thus any of these conditions can be expressed as

fi = fi (σL , σR , τLR 2 ) ≤ 0 . (4)

Equations (3) and (4) state that the only relevant experimental coordinates are actually
the invariants given in (1) and (2), respectively.

3.2 Linear elastic orthotropic behavior


As can be seen from Fig. 4, wood possesses linear elastic behavior both initially and
at unloading, and both in the tensile and in the compressive regime. The standard
hyperelastic orthotropic formulation based on a Helmholtz free energy function ψ is used.
Using the orthotropic strain invariants εL , εR , and γLR = 2εLR of (2), it can be expressed
as
EL e 2 ER e 2 νLR ER e e GLR e 2
ψ(εe , α) = (εL ) + (εR ) + εL εR + γ + H(α) , (5)
2∆ 2∆ ∆ 2 LR
where εe = ε − εP is the elastic part of the total strain ε, εP is the plastic strain, and
EL , ER , νLR , and GLR are four independent material parameters with ∆ = 1 − νRL νLR =
2
1−νLR ER /EL . H(α) is a hardening potential used for the description of densification with
α being a suitable set of strain-type hardening variables (see Subsection 3.6). Evolution
equations for the plastic strain will be given in Section 4.1.
Hyperelastic behavior defines the stress as
∂ψ
σ= = C : (ε − εP ) . (6)
∂ε
The linear elastic stiffness tensor C in (6) follows from (5) and C = ∂ 2 ψ/∂ε ⊗ ∂ε.

3.3 Tensile failure in grain direction


Tensile failure in grain direction of softwood is characterized by a cascading crack
pattern (see Fig. 4). The reason for this phenomenon is a similar tensile strength of the
fibers and the inter-fiber bond (shear) strength. Thus an initially perpendicular to grain
crack (fiber rupture) typically turns to a shear crack along the cell interfaces (parallel to
grain). Shear cracks are mostly located along the boundary between late-wood and early-
wood. The crack evolves along the cell boundaries until a weaker fiber is reached and
the failure mode once again changes to fiber rupture and a perpendicular to grain crack
orientation. Detailed discussions and micrographs of this crack pattern can be found in
Sippola and Koponen [19].
The combination of fiber rupture and inter-fiber shear failure and thus the ability of
the crack pattern to follow the weaker cells in the cell compound wood does not directly
permit a simple macroscopic description based on fiber strength. It rather yields an
explanation for the observed scatter of tensile strength in the macroscopic experiment

8
Peter Mackenzie-Helnwein, Josef Eberhardsteiner and Herbert A. Mang

due to the statistical variation of fiber strength over the sample size and the importance
of the weakest cell for the macroscopic strength.
The equal priority of shear failure and fiber rupture, however, yields a suitable con-
stitutive assumption for tensile failure parallel to grain as follows: “Tensile failure under
uniaxial states parallel and near parallel to grain occurs at the same level.” Based on the
current uniaxial strength βt,L and the above constitutive assumption, this failure criterion
can be expressed as

σL σR2 (τLR )2 βt,L


ft,L (σ, βt,L ) = 0
+ 0 2
+ 0 2
− 0 ≤ 0, (7)
βt,L (βt,L ) (βt,L ) βt,L
0
where βt,L is the initial tensile strength of the fibers. This condition defines a parabolic
surface with the (−σL )-axis as its axis. Its initial configuration is shown in Fig. 7.
Due to the smeared annual ring structure, the described crack pattern cannot, however,
be reproduced at the macroscopic level. One rather has to think about a crack zone in
which the cascading crack pattern evolves. The use of a discontinuity surface for its
macroscopic description should be viewed as a mathematical limit state rather than an
attempt to predict the microscopic crack pattern. The actual width of this zone is defined
by an additional material parameter `c , often referred to as characteristic length.
The strain-softening behavior is modeled by a single-mode softening function
0
βt,L = βt,L exp (−ξL ) (8)

and the non-associative softening rule for the strain-like parameter ξL as


0
βt,L `c
ξ˙ = I
γ̇ , (9)
Gf,L

where GIf,L is the mode I fracture toughness of the fibers, and γ̇ is the consistency param-
eter (Lubliner [14]).

3.4 Fiber compressive behavior


Compressive strength of softwood in fiber direction can be caused either by brittle
failure of the cell caps or by cell-wall buckling (Gibson and Ashby [10]). Combined loading
with σL and τRL mainly effects the cell-wall buckling. Considering the variability in the
microstructure regarding cell sizes and cell-wall thickness and chemistry, a combination
of both modes will occur within a characteristic macroscopic reference volume.
Contrary to the tensile strength prediction in fiber direction, the interaction with both
the shear stress and the lateral normal stress has to be considered. In order to avoid
unwanted interference with the tensile strength criterion the yield surface is assumed to
be parabolic with its axis parallel to the σL -axis. So far, the micromechanics of this failure
mode is not sufficiently understood in the case of combined loading. On the other hand,

9
Peter Mackenzie-Helnwein, Josef Eberhardsteiner and Herbert A. Mang

the failure envelope according to Tsai and Wu [23] performs well in this regime. Thus
the parameters for the parabolic yield surface are chosen such as to define a second-order
approximation to the single-surface failure envelope proposed by Eberhardsteiner [5] at
the location of min σL . Its mathematical representation can be given as

σL σR σL2 (τLR )2
fc,L (σ, q) = − 0
+ a 0
+ b 0 2
+ c 0 2
+q −Y ≤ 0, (10)
βc,L βc,L (βc,L ) (βc,L )
0
where a, b > 0, and c > 0 are dimensionless parameters, and βc,L is the initial compressive
strength in fiber direction.
Strength degradation and densification cause the yield surface to move along the σL -
axis while simultaneously slightly changing its curvature. Cell cap failure and inelastic cell
wall buckling causes strength degradation and occurs in a localized manner which requires
control by the fracture toughness GIf c,L and a characteristic length `c (crushing band
width). Densification eventually causes a homogeneous deformation on the macroscopic
scale, thus should depend neither on the fracture toughness nor on the characteristic
length.
The experimentally observed mode of failure shows minor non-linearity in the pre-
failure domain and softening beyond the stress maximum, as can be seen on the upper
left diagrams of Fig. 4. The softening is not as strong as that observed in the tensile
regimes. Assuming approximatively linear elastic behavior in the pre-failure domain, one
can extract a softening behavior as shown in Fig. 6.

βc,R

β0c,L −q
Y1

Y0
strength degradation densification
αc,L
αL,d αL,∞
Figure 6: Strength degradation and densification in fiber direction

The mathematical description of the post-failure behavior consists of two parts: the
initial strength degradation (strain softening) and the subsequent densification.
Strength degradation is modeled by the strength function Y (αL ) as
1
Y (αL ) = 0
[Y0 + Y1 exp(−kc,L αL )] , (11)
βc,L

10
Peter Mackenzie-Helnwein, Josef Eberhardsteiner and Herbert A. Mang

0
where Y0 and Y1 = βc,L − Y0 are strength parameters, and
0
βc,L `c
kc,L = I
. (12)
Gf c,L

The evolution equation for the equivalent strain variable αL is given by the associative
flow rule
∂fF C
α̇L = γ̇ = γ̇ . (13)
∂q
Densification causes an equilibrium state of internal stresses in the cell structure. Build-
ing up these internal stresses does not cause additional dissipation though the effective
compressive strength increases. Thus the increase of strength is modeled by the hardening
stress q as
HL,d hαL − αL,d i2
q(αL ) = − 0 , (14)
βc,L αL,∞ − αL
where HL,d is a hardening stiffness parameter, αL,d and αL,∞ mark the onset of densi-
fication and the ultimate densification strain, respectively, and hxi := (x + |x|)/2. The
dissipative part due to cell cap failure and inelastic cell wall buckling is marked by the
yellow area in Fig. 6.

3.5 Radial tension and inter-fiber shear failure


The failure mode observed in this regime is characterized by a distinct crack parallel
to grain and mixed mode I/mode II crack development. If an already developed crack is
subjected to compressive normal stress, i.e., σR < 0, friction enables the transfer of shear
stress τRL over the crack. This frictional shear stress is smaller than the initial shear
strength of the uncracked material.
Initially developed for frictional materials like soil, the model of Weihe [24, 25] is best
suited for the description of such a failure mechanism. It has been successfully applied
to the microscopic modeling of crack development in the RT -plane of spruce wood by
Lucena-Simon et al. [15].
Specialization of this model for a crack aligned with a grain boundary yields

(τRL )2
fW (σ) = −(σR − βt,R − a)2 + + a2 ≤ 0 , (15)
tan2 φ
where βt,R is the current tensile strength in the R-direction, a is a stress-type shape
parameter, and tan φ is the friction coefficient of the crack surface of a fully developed
crack.
This yield condition defines a hyperbolic surface in the orthotropic stress space. Its
geometric representation consists of two mirrored surfaces where only that one satisfying
σR ≤ βt,R is of physical relevance.

11
Peter Mackenzie-Helnwein, Josef Eberhardsteiner and Herbert A. Mang

The stiffness degradation along the crack is controlled by the fracture toughnesses
(energy release rates) for mode I and mode II, namely GIf,R and GIIf,R , respectively. The
dissipated energy for each failure mode is described by two normalized state variables ξ I
and ξII with the respective evolution equations
`c `c
ξ˙I = I
βt,R ε̇PR = I βt,R hMR : ri γ̇ (16)
Gf,R Gf,R
`c
ξ˙II = P
(|τRL | − h−σR tan φi) |γ̇RL |
GII
f,R
`c p
= I
(|τRL | − h−σR tan φi) 4 tr(MR rML r) γ̇ , (17)
Gf,R

where r is the plastic flow direction and hxi := (x + |x|)/2. The mathematical description
of r will be discussed in Section 4.
The strength parameters in (15) evolve as
0
βt,R = (1 − κn ) βt,R , a = (1 − κa ) a0 , (18)

where
κn = min(1, ξI + ξII ) , κa = ξII , (19)
0
and βt,R is the initial radial tensile strength, and a0 the initial parameter value for a.

3.6 Radial compressive behavior


All biaxial tests possessing compressive stress in radial direction show behavior similar
to the one shown at the bottom of Fig. 4. The figure shows the radial stress over both
the longitudinal and the radial component of strain, respectively, for one characteristic
test. As can be seen from the unloading-reloading cycles, the material possesses an elastic
domain in which it shows linear behavior. Moreover, in the investigated range of small
strains, no significant stiffness degradation occurs. The non-linearity in the stress-strain
curves results from inelastic deformations which evolve during the loading process. Based
on a micromechanical investigation, this plastic phenomenon can be explained by crushing
of whole rows of cells within the wooden structure. The first cells to be crushed are
the large cells growing in spring (early-wood). With increasing stress the smaller cells
are crushed row by row toward the late-wood. The process slows down and eventually
stops once the pore space has been reduced to an ultimate minimum (Gibson and Ashby
[10]). This last phase is called densification and is characterized by a steep increase of
the compressive strength. Densification occurs at large compressive strain and will be
addressed in this work with significant simplifications.
The most characteristic features are the linear elastic behavior in the unloading-
reloading cycles and a hardening type plastic behavior above a certain value of σR . Hel-
nwein et al. [12] presented a plasticity model for this phenomenon. The present approach

12
Peter Mackenzie-Helnwein, Josef Eberhardsteiner and Herbert A. Mang

contains a minor variation to the original formulation for the description of densification.
The respective yield condition now reads as

fRC (σ, q) = −σR − c σL + µ τLR 2 + q − Y ≤ 0 . (20)

Therein, c and µ are material parameters, q is the hardening stress for the description of
densification, and Y is a strength parameter which depends on the hardening parameter
αR . This yield function describes a parabolic surface as shown in Fig. 7.
Modeling densification follows the same line of arguments as given in Section 3.4. The
hardening stress q used for the macroscopic description of densification is expressed as

hαR − αR,d i2
q(αR ) = −HR,d , (21)
αR,∞ − αR

where HR,d is a hardening stiffness parameter, and αR,d and αR,∞ mark the onset of
densification and the ultimate densification strain, respectively.
The strength function Y (αR ) describes the initial shape of the stress plateau observed
in any radial compression test. It can be described by means of a Prony series. In the
present context the three-parameter form

Y (αR ) = Y0 + Y1 [1 − exp(−kαR )] (22)

is used and considering the variability of wood yields satisfactory results. Other functional
forms representing the actual cell structure of an annual growth ring may be used instead.
Accompanying uniaxial compression tests motivate the introduction of a flow potential
as follows:
gRC (σ, q) = −σR − c̄ σL + µ τLR 2 + q − Y , (23)
where 0 ≤ c̄ < c allows a correction of the plastic flow component in fiber (L) direction.
The evolution equation for the equivalent strain variable αR is obtained on the basis
of associative plasticity as
∂gRC ∂fRC
α̇R = γ̇ = γ̇ = γ̇ . (24)
∂q ∂q
3.7 Visualization of the multi-surface wood model
In the Subsections 3.3–3.6 a series of four mechanisms for the plane stress static re-
sponse of clear wood have been discussed. In order to identify the active domains for each
of these mechanisms one has to analyze which condition occurs first for each individual
proportional stress path.
Similar to the Heigh-Westergard stress space for isotropic problems, one may visualize
plane stress orthotropic problems in a three-dimensional orthotropic stress space. The
coordinate axes of such a space are defined by the orthotropic stress invariants given in

13
Peter Mackenzie-Helnwein, Josef Eberhardsteiner and Herbert A. Mang

τRL σL
Fiber tension τRL Fiber compression
−σR
σR
Helnwein et al. Weihe & Kröplin
(radial compression) σL (radial tension/shear)

Figure 7: Initial state of the multi-surface wood model; shown in the plane stress orthotropic
stress space

(1). For easier interpretation of the obtained geometric picture, the invariant (τRL )2 is
replaced by τRL leading to symmetric surfaces with respect to the plane τRL = 0.
Fig. 7 illustrates the initial state of the four yield conditions (resp. failure envelopes).
It is clearly visible that the radial compression surface and the radial tension-shear surface
dominate the behavior. The fiber tension surface is only relevant for a very limited domain,
having no effect for grain angles ϕ larger than approximately 10◦ . The fiber compressive
surface covers a larger domain than the fiber tension surface but its importance still seems
limited.

4 NUMERICAL IMPLEMENTATION
The numerical implementation of the constitutive model is based on the elastic predic-
tor – plastic corrector algorithm and the implicit Euler backward algorithm as described
in Simo and Hughes [18]. Even though that algorithm is solely applied to isotropic elasto-
plastic material in Simo and Hughes [18] it can be applied to small strain orthotropic
plasticity without changes in the procedure. The only remarkable difference is a con-
sequence of the special form of the flow rules of each of the four models, namely their
linearity with respect to the stress tensor. This reduces the number of variables to be
involved in an iterative procedure for the solution of the nonlinear consistency condition
to the incremental consistency parameters of the set of active yield and failure conditions.

4.1 Generic description of orthotropic failure envelopes, yield conditions and


flow rules
All four of the presented models can be written in a generic form. It is based on the
second-order tensor polynomial according to Tsai and Wu [23] as follows:

f (σ, q) = a : σ + σ : b : σ + q − Y
= aRR σR + aLL σL + bRRRR σR2 + bLLLL σL2 (25)
+2 bRRLL σR σL + 4 bRLRL (τRL )2 + q − Y = 0 ,

where a and b are constant orthotropic tensors of second and fourth order, respectively, q
is a hardening stress, and Y is a strength parameter. The latter has to be chosen as Y = 1
to obtain the classical formulation of Tsai and Wu [23]. For plane stress orthotropy, only

14
Peter Mackenzie-Helnwein, Josef Eberhardsteiner and Herbert A. Mang

the six independent material parameters aLL , aRR , bLLLL , bRRRR , bRRLL , and bRLRL are
required to define these tensors.
While the failure criterion according to Tsai and Wu [23] is typically considered to
describe a closed elliptic surface, it may as well describe open parabolic and hyperbolic
surfaces. An elliptic surface is best suited to single-surface models. For the present
multi-surface model, a combination of parabolic and hyperbolic surfaces is be used.
Comparison of (25) and (1) yields the description of the parameter tensors a and b as
a = aLL ML + aRR MR (26a)
and
b = bLLLL ML ⊗ ML + bRRRR MR ⊗ MR +
+ bRRLL (MR ⊗ ML + ML ⊗ MR ) + bRLRL M , (26b)
where M = (AR ⊗ AL + AL ⊗ AR ) ⊗ (AR ⊗ AL + AL ⊗ AR ). Equation (25) shows that
the tensorial criterion is a complete second order polynomial of the orthotropic stress
invariants given in (1).

4.2 Generic orthotropic evolution law for plastic strain


The evolution law for the plastic strain εP can either be described on the basis of the
representation theory of tensor valued functions [3, 4, 26] or by means of a flow potential
formulation [14, 17, 18]. It has been proven by Betten [3] that the flow potential theory
is only first-order accurate due to its incomplete representation of a second-order tensor
valued function by the gradiant of a scalar valued function. The flow potential theory,
2
however, simplifies the task by introducing a scalar flow potential g(σL , σR , τLR ) rather
than identifying parameters for a possible complete but non-linear flow rule. Based on
the flow potential, the plastic flow rule is obtained as
 
P ∂g ∂g ∂g ∂g
ε̇ = γ̇ = γ̇ ML + MR + 2 2 M =: γ̇ r , (27)
∂σ ∂σL ∂σR ∂τLR
with γ̇ as the consistency parameter to be computed, and r as the plastic flow direction.
Constructing g(σ, q) similar to f (σ, q) according to (25) by replacing the parameter
tensors a and b by two independent parameter tensors a∗ and b∗ , respectively, yields the
plastic flow direction as
∂g(σ, q)
r= = a∗ + 2b∗ : σ . (28)
∂σ
Setting g(σ, q) = f (σ, q) yields to the associative flow rule with a∗ = a and b∗ = b.

4.3 Time integration – return map algorithm


Time integration of the plastic flow rule for multiple active yield and failure conditions
by means of the Euler backward algorithm yields
X
εPn+1 = εPn + γi,n+1 ri,n+1 (29)
i∈Jact

15
Peter Mackenzie-Helnwein, Josef Eberhardsteiner and Herbert A. Mang

where Jact is the set of active conditions, γi,n+1 is the incremental consistency parameter
of the i-th surface, and ri,n+1 the plastic flow direction of the respective surface. The
latter follows from (28) as
ri,n+1 = a∗i + b∗i : σ n+1 . (30)
Substituting (30) into (29) and the obtained result into (6) yields
" #−1 !
X X
σ n+1 = C−1 + γi,n+1 2b∗i : εn+1 − εPn − γi,n+1 a∗i . (31)
i∈Jact i∈Jact
| {z }
=:Ξn+1

The hardening parameters αi,n+1 and softening parameters ξi,n+1 are updated as

αi,n+1 = αi,n + Ki γi,n+1 and ξi,n+1 = ξi,n + Ki γi,n+1 , (32)

where Ki is a scalar factor dependent on the respective hardening rule. The softening
rules (16) and (17) directly involve the stress at tn+1 and thus require an extended set
of variables being involved in solving the consistency conditions at tn+1 . By introducing
explicit integration for ξI and ξII the solution procedure could be simplified.
The elastic predictor or trial step is defined by the previous equations with γi,n+1 = 0
for all surfaces. If any of the yield resp. failure condition
trial
fi,n+1 = fi (σ trial trial
n+1 , qi,n+1 ) < 0 (33)

is violated a plastic corrector step is required. The set of active surfaces is initiated as
trial
Jact = {i|fi,n+1 ≥ 0} . (34)

Then for all i ∈ Jact the consistency parameter γi,n+1 6= 0 is introduced as a free variable.
The values of these variables are the solution of the consistency conditions

fi,n+1 = fi (σ n+1 (γj,n+1), qi,n+1 (γj,n+1 )) = 0 ∀ i, j ∈ Jact . (35)

For some configurations one may obtain consistency parameters γi,n+1 < 0 and thus vio-
lating the Kuhn-Tucker loading/unloading conditions [18]. In such a case the respective
surface has to be removed from Jact and the modified set of consistency conditions re-
solved. The updated solution for the stress σ n+1 , the plastic strain εPn+1 and the relevant
hardening and softening parameters are obtained by back-substitution of γi,n+1 into (29)
to (32).
The consistent tangent operator can be derived exactly as shown in chapter 5 of Simo
and Hughes [18]. It is obtained as
dσ n+1 X X ij
= Ξn+1 − gn+1 Ξn+1 : ri,n+1 ⊗ rj,n+1 : Ξn+1 , (36)
dεn+1 i∈J j∈Jact act

16
Peter Mackenzie-Helnwein, Josef Eberhardsteiner and Herbert A. Mang

where
ij
[gn+1 ] = [gij,n+1 ]−1 = [ ri,n+1 : Ξn+1 : rj,n+1 + Hij,n+1 ]−1 . (37)
The brackets symbolize a matrix of the given components. Hij,n+1 are scalar hardening
functions derived from the respective hardening and softening laws.

5 NUMERICAL SIMULATION: CROSS-WISE JOINT OF RODS


The presented material model shall be used for the analysis of a cross-wise joint of two
wooden rods as shown in Fig. 8. The top and bottom part of the joint are 100×100 mm

L
R

2 (` = 200 mm)
100×100 mm

L 1
R

L
R

100×100 mm

Figure 8: Shape of the analyzed configuration, grain orientation (L), loading conditions, and
finite element discretization

spruce columns with vertically oriented grain direction. The short horizontal rod is made
out of the same material but has its grain angle oriented in the horizontal direction. The
vertical column is subjected to axial loading. This load has to be transmitted through
the horizontal rod which, due to the weakness of wood perpendicular to grain, eventually
suffers crushing and fiber cracking. Such a loading condition may occur either single sided
or double sided in wooden trusses. It can also serve as a symmetrized model of a wooden
bearing, though the horizontal part would be made of hardwood rather than softwood.
The latter would require a different set of parameters but the basic concept of the model
still applies.
Only a quarter of the structure needs to be analyzed due to the double symmetry of the
problem. Plane stress conditions are assumed to be appropriate. The actual connection

17
Peter Mackenzie-Helnwein, Josef Eberhardsteiner and Herbert A. Mang

Table 1: Material parameters for the multi-surface plasticity model


0
fiber tension βt,L = 65.0 MPa GIf,L = 1650 J/m2
0
fiber comp. βc,L = 49.9 MPa Y1 = 10.0 MPa GIf c,L = 50000 J/m2
αL,d = 0.50 αL,∞ = 0.75 HL,d = 15.0 MPa
radial tension φ = 25◦ GIf,R = 300 J/m2 GIIf,R = 300 J/m
2

radial comp. Y0 = 4.00 MPa Y1 = 1.80 MPa k = 30.0


αR,d = 0.10 αR,∞ = 0.75 HR,d = 2.00 MPa
model aLL aRR bLLLL bRRRR bLLRR bLRLL
fiber tension 0.015385 0.000000 0.000000 0.00023669 0.000000 0.00005917
fiber comp. -0.020000 0.234533 0.000000 0.179664 0.000000 0.002052
radial tension 0.000000 0.444183 0.000000 -0.005997 0.000000 0.006895
radial comp. -0.025000 -1.000000 0.000000 0.000000 0.000000 0.026025

between the different parts acts as frictional contact. From practical observations it seems
a reasonable simplification to assume sticky conditions throughout the simulation. This
yields a minor error at the end of unloading but is sufficient for most of the simulated
load history.
The material parameters chosen for this analysis are given in Table 1. Some of these
parameters have been taken from the biaxial tests by Eberhardsteiner [5], some from
papers by Lucena-Simon et al. [15], Adalian and Morlier [1], Gibson and Ashby [10], and
[9]. These parameters originate from various material samples and do not represent the
best parameter fit for the present model. Further studies are in progress and the reader
is encouraged to look for better parameter sets in upcoming publications.
At load levels P < 40 kN the joint responds solely elastic. For the shown discretization,
this is the first instance where plastic deformations occur. Except for the small domain
underneath the edge of the vertical column, this stress state is elastic. There, however,
a stress concentration causes the radial component of stress, σR , to exceed the threshold
value Y0 = 4.00 MPa, and thus activating the radial compression surface. Fig. 9 shows the
distribution of both the horizontal and the vertical normal component of stress at P =
40 kN. Due to the horizontal and vertical grain orientation of both rods, respectively, the
stress invariants σL and σR are identical either to the vertical or the horizontal components
of stress σ11 (stress 1) and σ22 (stress 2). For all Figs. 9–11 the deformed configuration
is shown at a scaling factor of 1.
Fig. 10 shows the joint after increasing the load level to P = 60 kN. The vertical
column impressed the horizontal rod like a molding die. The radial stress σR underneath
the contact surface with the column became more homogeneous than that of the elastic
solution. A radial stress σR > Y0 + Y1 = 5.80 MPa shows that the material undergoes
densification. The zone of densification initiated underneath the edge of the column
and extends with increasing load level towards the symmetry plane, which has not been

18
Peter Mackenzie-Helnwein, Josef Eberhardsteiner and Herbert A. Mang

_________________
STRESS 1 _________________
STRESS 2

-5.03E+00 -7.95E+00
-4.43E+00 -7.26E+00
-3.83E+00 -6.57E+00
-3.23E+00 -5.89E+00
-2.63E+00 -5.20E+00
-2.03E+00 -4.51E+00
-1.43E+00 -3.83E+00
-8.32E-01 -3.14E+00
-2.32E-01 -2.45E+00
3.67E-01 -1.77E+00
9.67E-01 -1.08E+00
1.57E+00 -3.93E-01
2.17E+00 2.93E-01

Time = 4.00E+00 Time = 4.00E+00

σL in the horizontal rod [MPa] σR in the horizontal rod [MPa]


σR in the vertical rod [MPa] σL in the vertical rod [MPa]
Figure 9: Stress state at a total load of P = 40 kN

_________________
STRESS 1 _________________
STRESS 2

-9.93E+00 -7.57E+00
-8.04E+00 -6.79E+00
-6.14E+00 -6.01E+00
-4.24E+00 -5.23E+00
-2.34E+00 -4.44E+00
-4.39E-01 -3.66E+00
1.46E+00 -2.88E+00
3.36E+00 -2.10E+00
5.26E+00 -1.32E+00
7.16E+00 -5.39E-01
9.06E+00 2.42E-01
1.10E+01 1.02E+00
1.29E+01 1.80E+00

Time = 6.00E+00 Time = 6.00E+00

σL in the horizontal rod [MPa] σR in the horizontal rod [MPa]


σR in the vertical rod [MPa] σL in the vertical rod [MPa]
Figure 10: Stress state at a total load of P = 60 kN

reached in Fig. 10.


At the upper surface of the horizontal rod right beside the edge of the column, a fiber
tensile failure evolves. Though the respective failure surface is being activated, a better
coverage of the physical phenomenon in this area would require a finer discretization
around the stress concentration. Moreover, due to the large deformations in that area, a
geometrically non-linear formulation would be desirable.
Increasing the load level to P = 100 kN yields a significant increase of lateral defor-

19
Peter Mackenzie-Helnwein, Josef Eberhardsteiner and Herbert A. Mang

_________________
STRESS 1 _________________
STRESS 2

-8.28E+01 -6.40E+01
-7.39E+01 -5.84E+01
-6.50E+01 -5.29E+01
-5.61E+01 -4.74E+01
-4.72E+01 -4.18E+01
-3.83E+01 -3.63E+01
-2.94E+01 -3.08E+01
-2.05E+01 -2.52E+01
-1.16E+01 -1.97E+01
-2.74E+00 -1.42E+01
6.16E+00 -8.62E+00
1.51E+01 -3.08E+00
2.40E+01 2.45E+00

Time = 1.00E+01 Time = 1.00E+01

σL in the horizontal rod [MPa] σR in the horizontal rod [MPa]


σR in the vertical rod [MPa] σL in the vertical rod [MPa]
Figure 11: Stress state at a total load of P = 100 kN

mation as shown in Fig. 11. Two thirds of the horizontal rod are in an extended state of
densification. The previously mentioned crack could not develop further. The latter is a
faulty consequence of the linear kinematic description. The general tendency corresponds
well with practical experience, but the numerically obtained results must no longer be
viewed as reliable due to the insufficient description of the obtained large deformations
using linearized kinematic relations.
Another important result is the vertical stiffness of the joint, which is shown in Fig. 12.
Furthermore, the dominant failure mechanisms can be identified from the active yield
conditions and entered in the diagram. In the present case, radial compression causing
crushing of cells and accompanying large lateral plastic deformations are found as the
dominating mode. As the plastic deformation continues, the crushed cell structure gets
compacted (densification) and further load can be applied. The observed fiber cracking
20
thickness change (symm)
thickness change (edge)
0
total force [kN]

−20

unloading initial loading


−40
cell crushing
densification
−60

−80

−100
0.65 0.7 0.75 0.8 0.85 0.9 0.95 1
relative thickness

Figure 12: Stiffness of the joint; micromechanical failure mechanisms

20
Peter Mackenzie-Helnwein, Josef Eberhardsteiner and Herbert A. Mang

on the surface of the horizontal rod is of minor significance for the vertical stiffness. A
homogeneous state of radial stress would produce a distinct plateau at σR = Y0 + Y1 =
5.80 MPa (P = 58.0 kN). The given configuration shows inhomogeneous distributions
of the radial stress (see Figs. 9–11) and consequently shows a different response. The
absence of the plateau force can be explained by the local densification in the horizontal
rod underneath the edge of the vertical column.

6 CONCLUSIONS
Within this paper a new constitutive model for the simulation of clear spruce wood
under biaxial states of plane stress has been presented. It is based on observations from
a comprehensive test series for the experimental identification of the failure envelope for
clear spruce wood under arbitrary biaxial states of plane stress. Employing this and
both microscopic observations and their macroscopic translation led to a multi-surface
plasticity model.
It consists of four surfaces representing four basic failure modes. The first is a modified
tension cut-off for the description of fiber rupture. The second is a mixed mode radial
tension-shear model of Weihe applied to the perpendicular to grain direction. The third
is an extension of the model of Helnwein et al. for perpendicular to grain compression,
and the fourth surface covers the compressive failure parallel to grain.
The introduced model can be used to describe failure locations in stress space just as can
be done by the single-surface elliptic failure envelope of Tsai and Wu, which has been fitted
by Eberhardsteiner to the biaxial test data. Additionally, the multi-surface model enables
the identification of the relevant microscopic failure modes. The separated description of
four modes easily enables the modeling of their respective post-failure behavior.
Each of the four parts of the model can be described by means of a generic condition
(yield conditions and failure envelopes) which enables their reasonably simple numerical
implementation. The treatment of multi-surface softening plasticity requires some more
work regarding stability at larger time steps. The computations also pointed out the di-
rection of future enhancements, namely the need for a geometrically non-linear framework
for the proper description of the obtained large strains.

ACKNOWLEDGEMENTS
This work was granted by the APART-program of the Austrian Academy of Sciences,
the Austrian Science Fund, and the Federal Ministry of Education, Science and Culture,
which is gratefully acknowledged.

REFERENCES
[1] C. Adalian and P. Morlier. “Wood Model” for the dynamic behaviour of wood in
multiaxial compression. Holz als Roh- und Werkstoff, 60(6):433–439, 2002.

21
Peter Mackenzie-Helnwein, Josef Eberhardsteiner and Herbert A. Mang

[2] J. Betten. Applications of Tensor Functions in Solid Mechanics, volume 292 of CISM
courses and lectures, chapter 11–14. Springer, 1987.
[3] J. Betten. Kontinuumsmechanik – Elasto-, Plasto- und Kriechmechanik. Springer,
1993.
[4] J.P. Boehler. Applications of Tensor Functions in Solid Mechanics, volume 292 of
CISM courses and lectures, chapter 1–7. Springer, 1987.
[5] J. Eberhardsteiner. Mechanisches Verhalten von Fichtenholz – Experimentelle Bes-
timmung der biaxialen Festigkeitseigenschaften. Springer, Vienna, 2002.
[6] J. Eberhardsteiner, M. Gingerl, and L. Ondris. Experimental investigation of the
strength of solid wood under biaxial loading oblique to the grain direction. In
P. Morlier and G. Valentin, editors, COST Action E8: Workshop Damage in Wood,
Bordeaux, France, May 27–28, 1999.
[7] J. Eberhardsteiner, F. Pulay, and H.A. Mang. Zur Frage der Lasteinleitung
bei experimentellen Festigkeitsuntersuchungen von zweiaxial beanspruchtem Holz.
Österreichische Ingenieur- und Architekten-Zeitschrift, 136(6):265–272, 1991.
[8] J. Ehlbeck and K. Hemmer. Erfassung, systematische Auswertung und Ermittlung
von Grundlagen über das Zusammenwirken von Längs, Quer- und Schubspannungen
bei fehlerfreiem und fehlerbehaftetem Nadelholz. Technical report, Stuttgart: IRB
Verlag, 1986.
[9] C. Faessel, P. Navi, and M. Jirasek. 2D anisotropic damage model for wood in
tension. In P. Morlier and G. Valentin, editors, COST Action E8: Workshop Damage
in Wood, pages 21–37, Bordeaux, France, May 27–28, 1999.
[10] L.J. Gibson and F. Ashby. Cellular solids, Structure and properties – Second edition.
Cambridge University Press, 1997.
[11] M. Gingerl. Realisierung eines optischen Deformationsmeßsystems zur experi-
mentellen Untersuchung des orthotropen Materialverhaltens von Holz bei biaxialer
Beanspruchung. Österreichischer Kunst- und Kulturverlag, 1998. In German.
[12] P. Helnwein, J. Eberhardsteiner, and A. Hanhijärvi. Constitutive model for the
short-term failure analysis of wood under multiaxial states of stress: Effect of radial
compression. In Proceedings of the First International Conference of the European
Society for Wood Mechanics, Lausanne, Switzerland, April 19–21 2001. Building
Material Laboratory, Materials Science and Engineering Department, Swiss Federal
Institute of Technology Lausanne (EPFL).
[13] K. Hemmer. Versagensarten des Holzes der Weißtanne unter mehrachsiger Bean-
spruchung. Dr.ing.-thesis, Technical University of Karlsruhe, Germany, 1986.
[14] J. Lubliner. Plasticity Theory. Macmillan Publishing Company, New York, 1990.

22
Peter Mackenzie-Helnwein, Josef Eberhardsteiner and Herbert A. Mang

[15] J. Lucena-Simon, B.H. Kröplin, G. Dill-Langer, and S. Aicher. A fictitious crack


approach for the anisotropic degradation of wood. In Proceedings of the Interna-
tional Conference on Wood and Wood Fiber Composites, pages 229–240, University
of Stuttgart, Germany, 13.–15. April 2000. COST Action E8, Otto-Graf-Institute.
[16] P. Mackenzie-Helnwein, J. Eberhardsteiner, and H.A. Mang. Short-term mechani-
cal behavior of biaxially stressed wood: Experimental observations and constitutive
modeling as an orthotropic multi-surface elasto-plastic material. Journal of Plastic-
ity, 2003. In print.
[17] L.E. Malvern. Introduction to the Mechanics of a Continous Medium. Series in
Engineering of the Physical Sciences. Prentice-Hall, 1969.
[18] J.C. Simo and T.J.R. Hughes. Computational Inelasticity. Interdisciplinary Applied
Mathematics. Springer, 1998.
[19] M. Sippola and S. Koponen. Fracture behaviour of clear softwood; Tests and FEM
models. In COST E8, Damage in Wood, pages 27–28, Bordeaux, France, May 27–28
1999. COST E8, Damage in Wood.
[20] A.J.M. Spencer. Theory of Invariants. In A.C. Eringen, editor, Continuum Physics,
volume 1, pages 239–353, New York, 1971. Academic Press.
[21] A.J.M. Spencer. Applications of Tensor Functions in Solid Mechanics, volume 292
of CISM courses and lectures, chapter 8–10. Springer, 1987.
[22] R. Spengler. Festigkeitsverhalten von Brettelementen aus Fichte unter zweiachsiger
Beanspruchung; Ergebnisse aus experimentellen Untersuchungen. Technical report,
Technical University of Munich, Institute for Civil Engineering II (Bauingenieurwe-
sen II), 1986.
[23] S.W. Tsai and E.M. Wu. A general theory of strength for anisotropic materials.
Journal of Composite Materials, 5:58–80, 1971.
[24] S. Weihe. Modelle der fiktiven Rissbildung zur Berechnung der Initiierung und Aus-
breitung von Rissen: ein Ansatz zur Klassifizierung. PhD thesis, University of
Stuttgart, Germany, 1995.
[25] S. Weihe. Failure induced anisotropy in the framework of multi-surface plasticity. In
T.J.R. Owen, E. Oñate, and E. Hinton, editors, Computational Plasticity, Funda-
mentals and Applications, pages 1049–1056, Barcelona, Spain, 1997. CIMNE.
[26] Q.-S. Zheng. Theory of Representations for Tensor Functions – Unified Invariant
Approach to Constitutive Equations. Applied Mechanics Review, 47(11):545–587,
1994.

23

You might also like