Sie sind auf Seite 1von 9

American Journal of Botany 94(9): 1506–1514. 2007.

A GENERIC 3D FINITE ELEMENT MODEL OF TREE ANCHORAGE


INTEGRATING SOIL MECHANICS AND REAL ROOT SYSTEM
ARCHITECTURE1

LIONEL X. DUPUY,2,3,6 THIERRY FOURCAUD,4 PATRICK LAC,2 AND ALEXIA STOKES5


2
Université Bordeaux 1, Mixed Research Unit: Unité Sciences du Bois et des Biopolymères (UMR US2B), Talence, F-33405
France; 3Institut pour le Développement Forestier (IDF), Paris, F-75007 France; 4Centre de Coopération International en
Recherche Agronomique pour le Développement (CIRAD), Mixed Research Unit: Botanique et Bioinformatique de
l’Architecture des Plantes (UMR AMAP), TA A-51/PS2, Boulevard de la Lironde, 34398 Montpellier Cedex 5, France; and
5
Institute National de la Recherche Agronomique (INRA), Mixed Research Unit: Botanique et Bioinformatique de l’Architecture
des Plantes (UMR AMAP), TA A-51/PS2, Boulevard de la Lironde, 34398 Montpellier Cedex 5, France

Understanding the mechanism of tree anchorage in a forest is a priority because of the increase in wind storms in recent years
and their projected recurrence as a consequence of global warming. To characterize anchorage mechanisms during tree uprooting,
we developed a generic finite element model where real three-dimensional (3D) root system architectures were represented in a 3D
soil. The model was used to simulate tree overturning during wind loading, and results compared with real data from two poplar
species (Populus trichocarpa and P. deltoides). These trees were winched sideways until failure, and uprooting force and root
architecture measured. The uprooting force was higher for P. deltoides than P. trichocarpa, probably due to its higher root volume
and thicker lateral roots. Results from the model showed that soil type influences failure modes. In frictional soils, e.g., sandy soils,
plastic failure of the soil occurred mainly on the windward side of the tree. In cohesive soils, e.g., clay soils, a more symmetrical
slip surface was formed. Root systems were more resistant to uprooting in cohesive soil than in frictional soil. Applications of this
generic model include virtual uprooting experiments, where each component of anchorage can be tested individually.

Key words: biomechanics; poplar; Populus; root anchorage; tree stability; windthrow.

In recent years, attention has focused on the consequences of mechanistic tree anchorage models based on a description of
climate change on the composition and sustainability of forest root architecture have not yet been linked to field data, but
ecosystems around the world. One likely consequence is an would be a major step toward developing a powerful tool to
increase in the number and intensity of wind storms (Bloom- predict tree uprooting during storms.
field, 2000), which would cause major economic damage and Vulnerability to uprooting has been traditionally investigated
disturbance to temperate forests. Although much experimental at the population level, by correlating forest damage to certain
research has been carried out on tree anchorage and stand characteristics, i.e., tree species, silviculture, soil type, and
biomechanics, particularly since the storms Lothar and Martin wind speed (Putz et al., 1983; Ruel, 2000; Cucchi and Bert,
hit Europe in 1999 (Cucchi et al., 2004; Achim et al., 2005; 2003; Mayer et al., 2005). However, this empirical analysis of
Nicoll et al., 2005; Stokes et al., 2005; Brüchert and Gardiner, windthrow is incomplete because root system characteristics
2006; Peltola, 2006), mechanistic numerical models of such and other variables influencing tree stability (Khuder et al.,
phenomena are lacking (Blackwell et al., 1990; Dupuy et al., 2007) are not always taken into account. Experimental studies
2005a, b; Fourcaud et al., 2007). The use of such models is whereby trees are winched until failure and the force necessary
three-fold: they can be used to investigate the contribution of to uproot or cause stem breakage is measured are useful in that
different plant or soil parameters to tree anchorage; virtual they provide detailed data at the tree level (Fraser, 1962; Stokes,
experiments can be carried out, thus reducing destructive and 1999; Cucchi et al., 2004; Nicoll et al., 2005; Stokes et al., 2005;
Peltola, 2006). Nevertheless, even if root system architecture is
expensive fieldwork; and finally, the models can be incorpo-
measured, data are often incomplete because roots are usually
rated into decision support systems for forest managers
damaged during the tests (Stokes et al., 2007). The role of root
(Gardiner et al., 2000; Cucchi et al., 2005). Although the system architecture in tree anchorage has been alluded to in
initial development of such models is time consuming, once studies of root system morphology, but without the appropriate
calibrated, they could be adapted fairly easily to different experimental data, it is difficult to quantify the influence of each
species in a variety of situations. To our knowledge, parameter during the uprooting process (Nicoll and Ray, 1996;
Mickovski and Ennos, 2003; Danjon et al., 2005; Di Iorio et al.,
1
Manuscript received 5 January 2007; revision accepted 10 July 2007. 2005; Soethe et al., 2006; Khuder et al., 2007).
The authors thank M. Maitrot and C. Drenou (IDF) for help in the While root architecture strongly affects tree anchorage, soil
measurements of root architecture and F. Charnet (IDF) for data on soil physical and mechanical properties are major determinants of
properties at Chinon. Funding was provided by a French governmental
uprooting. Because the root–soil plate of a tree is a composite
CIFRE bursary, the poplar ECOFOR project (no 2002.06), and the E.U.
project Eco-Slopes (QLK5-2001-00289). Authors are also grateful to the structure, soil type and its interaction with the root system will
two anonymous reviewers for very constructive advice that helped contribute to determining the overall anchorage capacity of a
improve this manuscript. tree (Moore, 2000). When soil is waterlogged, both shear
6
Author for correspondence (e-mail: lxd20@cam.ac.uk), present strength and the root–soil bond are reduced, and therefore, the
address: Department of Plant Sciences, Downing Street, Cambridge entire system fails more easily during wind loading (Coutts,
CB2 3EA, UK, phone: þ44 01223-7-66545 1986). In dry soils or soils with a high shear strength, trees tend
1506
September 2007] D UPUY ET AL .—3D TREE ANCHORAGE MODEL 1507

to be better anchored but the stem breaks during loading,


whereas trees in waterlogged soils or soils with a lower shear
strength are usually uprooted (Coutts, 1986; Moore, 2000). In
simulations of uprooting experiments of simplified model trees,
soil internal friction angle modified the position of the root
system rotation axis during overturning (Dupuy et al., 2005b).
Such an adjustment to the mechanics of the composite root–soil
system in turn strongly influences the force required to cause
failure of this system. However, although the models by Dupuy
et al. (2005a, b) and others (Ennos, 1990; Niklas et al., 2002;
Fourcaud et al., 2007) explicitly consider soil and root system
properties, the models do not use real root system architectures
in uprooting simulations. Hence, testing models that can
accurately predict overturning behavior of trees in different
types of soils is a priority.
The development of automatic and generic numerical
methods that use accurate descriptions of plant morphology
for biomechanical analyses can be extremely useful, and such Fig. 1. Diagrams of pulling tests for trees in the field. (a) The trees
were winched sideways using a motorized winch. The winch was attached
models have already been applied to both real and simulated to the base of an anchoring tree at the longest possible distance to the
architectures subjected to mechanical loading in different winched tree to obtain a small deviation from pure horizontal forces. The
environments (Fourcaud et al., 2003a, b; Sellier et al., 2006). height of the cable attachment was only 6.0 m so as to induce root failure
We focused on the use of finite element modelling (FEM) to instead of stem breakage. (b) The finite element case study was formed of
simulate the uprooting mechanism of realistic root architecture a block of soil 2.0 m deep and 4.4 m wide. The underside face was built in
in different soil types. and lateral faces were blocked horizontally. A displacement was imposed
The validation of FEM models with real data is, however, on the rigid trunk at a height of 6.0 m.
particularly difficult as mechanical testing often greatly
damages the root system. Therefore, trees selected for complete (69% sand, 21% clay, and 10% loam) and lacked a hard pan in the top 1.0 m (F.
uprooting were different from trees used for root system Charnet, Institut pour le Développement Forestier, personal communication).
measurements. We also chose two poplar cultivars (Populus During the investigation, the ground water table was 1.2 m deep. Soil was
trichocarpa Torr. Gray and P. deltoides Batr. Marsh) to sampled at a depth of 0.30 m from the root–soil plate of trees winched
illustrate interspecific variations in root morphological proper- sideways. The wet soil mass was measured immediately, and samples were
then oven-dried for 5 d at 1058C to calculate soil moisture content (g water/100
ties and anchorage (e.g., branching density, rooting depth) g soil). In both plantations, soil moisture was 21.7 6 1.4% (Raspalje) and 21.7
(Harrington and DeBell, 1996). 6 0.9% (Beaupré) and thus was not thought to influence anchorage between
Winching tests were carried out in the field, and the force trees.
required to cause failure was recorded. Root system architec-
ture was then measured, and these data were fed into the Static bending tests—The method used for uprooting trees was similar to
model, which calculated the force necessary to uproot the trees that used by Coutts (1983), Moore (2000), Mickovski and Ennos (2003), Cucchi
in two types of soil. Results were compared with experimental et al. (2004), and Nicoll et al. (2005). Two trees of each cultivar were winched
data to better understand the major parameters influencing tree sideways until anchorage failure, which occurred when the tree was deflected
;108 from the vertical, forming an 808 angle with the soil. One other tree for each
uprooting during a wind storm. species was winched sideways almost to the point of failure, i.e., the tree was
deflected 58 at the stem base, then allowed to return to the vertical. Trees that
were bent 58 were not damaged and were used for measuring root system
MATERIALS AND METHODS architecture, which required intact root systems. A motorised winch (type Hit-
Trac 16B, Habegger, Switzerland), with a 16 kN maximal strength capacity, was
Tree material and site—To provide data from real trees, which could be used to winch trees sideways. The winch was attached to the base of an anchoring
used to test the 3D model developed later, we performed static bending tests tree at the longest possible distance to the winched tree to obtain a small deviation
and measured root architecture of forest trees. We chose poplars (Populus sp.) from pure horizontal forces (Fig. 1). The cable was attached to the winched tree at
because poplar plantations suffered severe losses during the 1999 storms in only 6.0 m so as to induce root failure instead of stem breakage. The tension
Europe (Bergonzini and Laroussinie, 2000). The poplar trees selected for this applied to the winched tree was measured by a load cell (type K25H-20kN
study were planted as live pole cuttings, each pole being several meters long Scaime S.A., France), attached between the winch and the anchoring tree, and
and pushed into the ground at a depth of 1 m. Adventitious lateral roots then was recorded every second using a data logger (type Almemo 2290–8, Ahlborn,
grow from this pole, which can be considered analogous to a central tap root. Germany). An inclinometer was attached to the tree stem at a height of 6.0 m to
Therefore, root systems of plantation-grown poplars are unusual but simpler to measure the stem deflection angle during the winching test. Deflection angles
describe than seed-grown, tap-rooted conifers such as maritime pine (Pinus were recorded every second with a second data logger, which was synchronized
pinaster Ait., Danjon et al., 2005). Because these data were to be fed into the and simultaneously activated with the load cell data logger. Trees were winched
3D root and soil model in a first attempt to calibrate the model with real data, along the direction of the westerly prevailing wind.
we preferred to use simple systems, thus reducing computation time.
Six 7-yr-old trees were used, three each of two North American poplar Root systems—Root systems of trees that were winched but not uprooted
cultivars: ‘Beaupré’ (Populus trichocarpa Torr. and Gray) and ‘Raspalje’ (stem deflection of 58) were then excavated. Those trees that had been uprooted
(Populus deltoides Batr. ex Marsh). Stem diameter at breast height (DBH) was could not be used to measure roots because the root systems were damaged. A
0.24 6 0.01 m (mean 6 SE) for Beaupré and 0.28 6 0.01 m for Raspalje. mechanical shovel was used to cut a circular trench with a radius of 2.0 m
Trees were planted at a density of 200 stemsha1 in two neighboring stands around the trunk, and the root system was lifted out of the soil. An air spade
located in northwest France (Moulin de Bariteau, Chinon, 0814 0 E, 47810 0 N). (Rizzo and Gross, 2000) was then used to remove the soil and clean the root
The study area was flat with an elevation of 37 m. Mean annual precipitation is system. Root system architecture was measured manually in the field, and
820 mmyr1, and the prevailing wind direction is northwest. Mean wind speed during these measurements, each root system was discretized into small
ranges from 4.5–5.5 ms1 (Météo-France). The soil consisted of a clayey sand sections. Following a recursive pathway along the central tap root and each
1508 A MERICAN J OURNAL OF B OTANY [Vol. 94

TABLE 1. Mechanical properties of soils and roots used in the analysis.


The soils were modelled using Mohr–Coulomb yield criteria, and the
roots were modelled using the von Mises yield criterion. Refer to list
of symbols at end of Discussion for definitions.
Property Frictional soil Cohesive soil Root

E (MPa) 20 20 2000
m 0.3 0.49 0.3
c (kPa) 1.5 20 —
/ (8) 25 0 —
MOR (MPa) — — 15
c (kN/m3) 20 20 0

lateral root, we measured the distance (link) between each root base and
daughter root, along with the diameter at the root section base and halfway
along the link. The depth, branching angle, and azimuth of each link were also
noted, as well as any abrupt changes in azimuthal direction. Data were recorded
and encoded simultaneously in the software ARCHIROOT (Dupuy, 2003,
www.archiroot.org.uk). ARCHIROOT stores data in the format of a directed
Tree Graph (TG) using the AMAP MTG format as described by Danjon et al.
(1999). The data were used to compute the total root biomass and the distribution
of diameter size and to produce 3D images of the root system (Fig. 1).

Finite element modelling—Soil and roots material definition—Modelling


was carried out using the FEM software Abaqus (ABAQUS, Inc., http://www.
abaqus.com/). Soil was modelled as an elastic perfectly plastic material using a
linear Mohr Coulomb model available in the Abaqus software (see Abaqus
user’s manual for details). For frictional material, the model was used with
nonassociated flow, with a dilation angle w ¼ 0 (see Whitlow [1995] for an
introduction to soil mechanics). Root material was considered to be elastic
Fig. 2. Graphs of root architecture data from Beaupré (Populus
linear with a plastic threshold modelled by a von Mises yield criterion trichocarpa) and Rasplje (Populus deltoides) cultivars’ root systems used
(Kachanov [2004]). to simulate uprooting. (a) The Beaupré cultivar’s root system had a high
Results from the 3D mechanical model were then to be compared with the branching density but a low root volume. (b) The Raspalje cultivar’s root
experimental data from the static bending tests. However, information system had fewer but larger roots. XX and ZZ correspond to width and
concerning the mechanical properties of soil at the field site was not available. depth, respectively.
Therefore, we used theoretical values (Whitlow, 1995), which were considered
representative of the site conditions. Because the soil to be modelled comprised mesh of the root system was automatically derived from the architectural data
70% sand particles, we chose the following values: Young’s modulus E ¼ 20 using a computer program specifically dedicated to this task (see flowchart in
MPa, Poisson’s coefficient m ¼ 0.3, density c ¼ 20k N/m3, and friction angle / Fig. 2). This program gave a generic character to the meshing process, so that
¼ 258 (Table 1). A low cohesion value c ¼ 1.5 kPa was also used to account for the program is applicable to any branching structure encoded with an AMAP
the presence of about 20% clay in the soil (Table 1). MTG format.
Anchorage simulations were then also performed using a model of The initial discretization was based on the nodes defined in the TG files, i.e.,
undrained pure clay-like soil so that the root–soil mechanical response could resulting from the field measurements. The automatic mesh generation
be analyzed in different conditions, where the cohesion c ¼ 20 kPa, m ¼ 0.49, consisted of sequentially reading the pathway traced in the TG file and
and / ¼ 08. The remaining mechanical parameters were kept the same as in the creating a separate list of nodes (NODE) as well as a list of beam elements
preliminary soil model (Table 2). (BEAM). The format of these elements must be consistent with the B21
The wood material of roots was characterized by a Young’s Modulus of 2.0 element type chosen in this study. It should be noted that this section of the
GPa, and the yield stress (rr) associated with the von Mises criterion equalled program can be changed easily if another type of element or FEM software is
15 MPa. used. Each line of NODE contains a node number and its spatial coordinates,
whereas each line of BEAM contains an element number and the number of the
Geometry and meshing—The root architecture data, which were recorded in two associated nodes delimiting the extremities of the element.
the TG files, provided the topology and geometry of the whole root structure. The soil domain consisted of a parallelepipedic block discretized with 20
These data are the list of consecutive nodes, the type of connection (link) these nodded brick elements, i.e., the 3D continuum element C3D20 in the ABAQUS
nodes had with the antecedent node of the same branching order, and the elements library.
associated geometrical attributes.
In this FEM model, root systems were discretized using 2-node linear beam Root–soil interaction—The following stage consisted of creating the
elements B21 available in the structural element library of ABAQUS. The FEM physical links between the soil and roots. Different techniques exist to model
explicit surface to surface interaction through the use of constraints
TABLE 2. Lowest and highest values of c, /, w, and MOR used to analyze implemented with Lagrange multipliers or interface elements. Such methods
are computationally extensive and may sometimes lead to convergence
the sensitivity of root and sand-like soil parameters on resistance to
difficulties. However, the hypothesis of a rigid root–soil interaction was
uprooting. Refer to list of symbols at end of Discussion for
supported by field observations that most of the tree roots remained embedded
definitions. in the mass of soil after uprooting, suggesting that most of the soil lifted during
Parameter Low High
uprooting was trapped in the network of roots and that little relative
displacement between soil and roots occurred.
c (kPa) 1.5 2 Therefore, a node to node interaction between soil and roots was
/ (8) 20 25 implemented using kinematic coupling (see the ABAQUS theory manual).
w (8) 0 5 Let Xr denote the position of a root node (master node) and Xs be the coupled
MOR (MPa) 12.7 15.0 soil position (slave) in reference configuration. The initial reference
configuration is expressed as the translation vector N:
September 2007] D UPUY ET AL .—3D TREE ANCHORAGE MODEL 1509

TABLE 3. Summary of experimental and numerical pulling tests for


Beaupré (Populus trichocarpa) and Raspalje (Populus deltoides)
cultivars.
Beaupré Raspalje

Angle Displacement Force Angle Displacement Force


Test (8) (m) (kN) (8) (m) (kN)

Quick test 2 0.2 6.3 4.1 0.4 9.5


Complete test 1 5.9 0.6 8.3 10.5 1.1 12.3
Complete test 2 11 1.1 9.2 8.7 0.9 15.1
Simulation sand 9.6 1 10.5 9.6 1 15.9
Simulation clay 9.6 1 19.4 9.6 1 31.3

were created, with an increasing number of soil elements, i.e., 256 (8 3 8 3 4)


elements, 720 (12 3 12 3 5) elements, 1372 (14 3 14 3 7) elements, and 2976
(18 3 18 3 9) elements. These different meshes were used to analyze two root
systems in the frictional soil. The evolution of the reaction force was related to
the number of elements. Calculation time (CPU) was monitored to find an
optimal number of elements for a reasonable CPU.

Sensitivity analysis of material properties—Because we did not possess any


quantitative data on soil and root mechanical properties at the field site for the
static bending tests on poplar trees, a sensitivity analysis was conducted to
study the influence of small variations in frictional soil properties (c, /, w) and
root mechanical parameters. Each parameter of the initial configuration was
varied successively by 15% (Table 2). The sensitivity S of a parameter a was
then quantified as the percentage increase in anchorage resistance for an
increase in 1% of the specific parameter:
Fig. 3. Flowchart of the mesh generator program. BEAM and NODE DF=F
are lists defining nodes and beam elements, respectively, in the FEM S¼ ð3Þ
Da=a0
format of the input file. The program reads the MTG file incrementally and
finds at each line i the branching order o(i) and the spatial coordinates
needed to write a new beam element (,i. and ,i, j. in the BEAM and
NODE list). N(o) represents the number of the anterior node of the oth
branching order. The operator  indicates the insertion of a new line ,i.
RESULTS
or ,i, j. in the NODE or BEAM list.
Static bending tests—Of the four trees winched sideways
N¼ Xs  Xr ð1Þ
until failure, the stem of two of each cultivar broke at heights of
2 m; the remaining two trees, which were deflected to 5 rather
In the current configuration, the root node position xr and the soil node position than 108, failed through uprooting. The mean force that caused
xs are linked with the following relationship: failure was 8.75 kN in Beaupré poplars and 13.5 kN in
xs ¼ xr þ Cr  N; ð2Þ Raspalje (Table 3). Because we were simulating uprooting
r
behavior in these trees, we could not compare results from the
where C denotes the rotation matrix associated to the root node rotation. The model with data from trees that had failed in the stem;
constrained degrees of freedom are expressed as a function of the master node
degrees of freedom. This reduces the total number of degrees of freedom in the
therefore, only two trees could be used to compare observed
tangent stiffness matrix and is therefore more computationally efficient than and predicted uprooting results. For the two trees, which were
explicit surface to surface interaction methods. winched sideways but not to failure, only the initial part of the
Soil–root node pairs were connected using the ABAQUS command *TIE. force displacement curve could be compared with results from
the model.
Boundary conditions and bending test simulation—The reconstructed root
system model was placed at the center of a parallelepipedic domain of soil, 4.4 Description of root systems—The root system of the
3 4.4 m wide and 2.0 m deep. Boundary conditions were defined to reproduce Beaupré cultivar consisted of a main vertical root 1.11 m deep
the action of the surrounding infinite earth: translation on lateral faces was fixed with a basal diameter of 0.31 m. This central root bore
horizontally, and the underside face was built in (Fig. 3). A rigid beam 6 m long
was fixed to the top of the root system, representing the tree trunk beneath the numerous secondary laterals with an average density of 35
crown. To simulate uprooting, lateral displacements of 1.0 m were applied laterals/m and a mean diameter of 32.0 6 0.02 mm. The root
incrementally on the rigid trunk up to a height of 6.0 m. The reaction force was system had a total volume of 0.12 m3 and had 115 structural
recorded at this point, and the force–displacement curve then determined. The roots (branching order ,5, 90% were smaller than 50 mm in
maximum force at the end of the lateral displacement was used to compare the diameter), with a total root length of 94 m (Fig. 1a).
soil–root anchorage efficiency. The field of equivalent plastic strain was The root system of the Raspalje cultivar had fewer but larger
monitored in the vertical plane in the direction of pulling. The plastic area
around the roots was also visualized in the complete root system, and the
roots than the Beaupré. One massive central vertical root with a
motion of the stump was monitored for each increment. These variables were length of 0.95 m and a basal diameter of 0.34 m bore secondary
used to understand the different mechanisms occurring during uprooting. laterals with an average density of 24 laterals/m and a mean
diameter of 0.55 6 0.04 m. The root system had a total volume
Mesh size sensitivity analysis—A preliminary analysis of the influence of of 0.21 m3 and had 82 structural roots (branching order ,5)
mesh size on the calculation was also carried out. Four different soil meshes with a total root length of 82.0 m (Fig. 1b).
1510 A MERICAN J OURNAL OF B OTANY [Vol. 94

Fig. 4. Influence of the soil mesh on the uprooting simulations. The


circles (*) and crosses (x) indicate the maximum reaction force for
Beaupré and Raspalje, respectively, for a given mesh size. The solid line
(—) indicates the calculation time (CPU) required for a given mesh size. Fig. 5. Simulated fields of plastic strain generated in the frictional soil
(a) and in the root system (b) during the uprooting of the Raspalje cultivar.
In (a), permanent deformations were greater on the windward side of the
Mesh size sensitivity analysis—The study of mesh size tree. In (b), failure of root elements appeared in the region of soil plastic
influence on the numerical results demonstrated that the failure.
calculated reaction force opposed to a given stem displacement
decreased with the size of the soil elements (Fig. 4). The
calculated reaction force vs. mesh size did not stabilize to a experiment than in the simulation with the frictional soil (Table
constant value with the range of element size used in this 3). Simulations with the cohesive soil resulted in an uprooting
sensitivity analysis (Fig. 4); however, the absolute difference resistance 70% higher than that measured in the field.
between the mechanical responses of the two root systems The influence of small variations in soil and root properties,
remained constant. On the other hand, tthe calculation time i.e., c, /, w, and MOR, showed that the cohesion c was the
(CPU) increased exponentially with the number of soil most influential parameter with regard to uprooting, with a
elements; therefore, a mesh with 1372 elements was finally sensitivity of 0.5 (Fig. 8). The properties / and MOR had a
used for the comparative analysis. This choice provided a good similar effect on uprooting resistance, with a sensitivity of less
compromise between the accuracy of the numerical results and than 0.3 (Fig. 8), while w had the least influence on uprooting
the calculation time required for the simulations. resistance (Fig. 8).

Static bending test simulations—The displacement applied


at the top of the rigid stem induced a bending moment at the DISCUSSION
center of the bole. The roots on the ‘‘leeward’’ side (the side
held in compression during bending) tended to plunge Understanding how wind forces are transferred to the soil via
downward, which locally compressed the soil. On the the root system and what determines the anchorage resistance
‘‘windward’’ side (the side held in tension during bending), of trees is crucial for predicting windthrow in forests. To date,
roots were pulled upward from the soil, decreasing soil our knowledge and ability to predict tree uprooting has been
compression on this side of the tree. The failure areas were limited by our ability to understand the nature of the root–soil
highlighted by the field of equivalent plastic strain (Fig. 5) and interaction in the context of real root system architectures. Root
took the form of a slip surface. In cohesive soil, the slip morphologies are usually very complex and their variability
surfaces were symmetric and circular (Fig. 6a), whereas in poorly understood. Also, by current experimental approaches
frictional soil, permanent deformations expanded further on the we are unable to measure the essential mechanical variables
windward side than on the leeward side (Fig. 6b). The plot of (e.g., strains, stresses, and displacement fields) in both roots
the displacement field of a strip of soil in the plane of pulling and soils during in situ tree pulling tests. Finite element and
(centered on the stem) showed that the kinematics of uprooting other numerical methods used in physics and engineering are
in these trees followed a rather regular rotation around an axis potentially very powerful tools for obtaining quantitative in-
at half the rooting depth (Fig. 6). depth information on such mechanisms, and the model
The differences in anchorage resistance between the two trees presented here was a first attempt in using such a method to
were similar between both the experimental tests and computer understand the mechanical properties of root–soil composites.
simulations (Fig. 7a, b). In the field, the Raspalje cultivar was In this paper we primarily described the development and
30% more resistant than the Beaupré. Simulations in the testing of the finite element method in tree anchorage studies.
frictional soil produced a difference of 40% between the two The developed model allowed us to simulate the uprooting of
cultivars, and a difference of 50% uprooting resistance resulted two different poplar cultivars while accounting for an accurate
when the simulations were carried out in cohesive soil. When geometrical and topological description of their two distinct
comparing the simulations to experimental data, we found that root system architectures. To compare the model’s predictions
20% less force was required to uproot the poplars in the with the anchorage of real trees, destructive winching tests
September 2007] D UPUY ET AL .—3D TREE ANCHORAGE MODEL 1511

Fig. 8. Results of the sensitivity analysis describing how properties


affected the force required to uproot trees. (a) Soil cohesion, c, had the
most influence on uprooting. (b) Friction angle, /, and (c) root strength
MOR, rr, had a similar influence, with sensitivities of 0.27 and 0.26,
respectively. (d) Dilation, w, had the least influence on resistance to
uprooting.

However, significant intraspecific variability occurred during


the pulling tests, and the model predictions can only provide
indications about the differences in anchorage between
cultivars. More trees need to be tested to validate our model,
and more elaborated poro-viscoplastic models could be
developed to incorporate the effect of pore water in the
mechanisms of uprooting.
Fig. 6. Displacement fields at soil nodes in a vertical plane along the The properties of the field soil (clayey sand) used are
direction of pulling for the Beaupré cultivar in (a) clay-like soil and in (b) between those of sand and clay, and yet the model simulations
sand-like soil. predicted much higher resistance than was measured in the
field. Results from the simulations seem, therefore, to have
were conducted on four additional trees. The use of cultivars overestimated the actual uprooting resistance. Apart from the
from the poplar plantation at Moulin de Bariteau aimed at fact that the trees that were winched to failure in the field were
reducing the morphological variability among species (Har- not those for measured for root architecture, the difference in
rington and DeBell, 1996), while the Beaupré and Raspalje uprooting resistance between simulated and real anchorage
cultivars were chosen for their morphological differences. tests could be explained in several ways. The rigid node-to-
node interaction used in the model does not enable consider-
ation of sliding or opening of the root–soil surface interaction
during failure. Studies on the influence of interface properties
have also shown that roughness of the interface on foundation
piles can account for 15–50% of the variation in resistance,
depending on the type of loading and pile configuration (Chen
and Martin, 2002; Jeong et al., 2004). Therefore, the rigid root–
soil adhesion used in our model can be expected to artificially
increase the uprooting resistance. Finally, the material model
used for roots and soils is probably not suitable for large
deformations. Highly anisotropic and fragile models, which
can involve failure of individual root elements, will more
realistically represent root breakage.
Although the generic 3D uprooting model developed in this
Fig. 7. Simulated curves of force vs. displacement (with a mesh size of study does not yet include all the elements necessary to
1372) for (a) the Beaupré and (b) the Raspalje cultivars. Experimental accurately predict uprooting resistance, it can nevertheless
pulling tests are represented with a dashed line, the simulation in frictional provide useful information on how root system structures
soil is represented with circles, and the simulation in cohesive soil is interact within a soil. In the simulations of uprooting in clay
represented with squares. soil, failure in the soil was symmetrical on both the windward
1512 A MERICAN J OURNAL OF B OTANY [Vol. 94

and leeward sides of the root system. In analogous studies on considering the efficiency of landslide stabilizing piles, but pile
laterally loaded piles on cohesive clay soils (Yang and Jeremic, failure is concerned with soil cohesion in compression in the
2002), deformations were also symmetrical. However, in the direction of pulling. In our model, soil failure was predominant
same study on a frictional soil, Yang and Jeremic (2002) found on the windward side of the tree, where roots and soil were
that the plastic deformation not only occurred close to the soil held in tension during overturning.
surface but also expanded laterally further away from the pile. Changing root strength had little effect on overturning
In our model of uprooting in a frictional soil, plastic resistance in the model. To our knowledge, no quantitative
deformation was greater on the windward than on the leeward studies of the influence of root strength on anchorage
side of the root system. Although this phenomenon occurs in efficiency of trees has been done, although indicators of its
real situations of tree overturning (Coutts, 1983, 1986; Crook importance, particularly in shallowly rooted forest trees, are
and Ennos, 1996; Cucchi and Bert, 2003), it has not been available (Stokes and Mattheck, 1996).
explained comprehensively. We show that soil mobilized on With regard to the technical aspects of the generic 3D model
the windward side of the tree was subjected to low or negative described here, CPU time was high, and the problem of
isostatic pressure when lifted up by roots. Therefore, failure convergence does not yet enable models to be compared
occurred and uprooting was induced. readily to real data. Nevertheless, such a numerical tool is very
The geometrical structure of the root systems also affected promising and can be applied to any root architecture from
soil behavior during overturning in our simulations. The two which topology and geometry have been accurately described.
root systems, which were from two different cultivars, were Recording the structure with an AMAP MTG format also
distinctly different from each other with regard to the number allows the topological data of the root system to be read
and diameter of first order lateral roots borne on the central tap automatically and incorporated in the FEM anchorage model.
root. The Raspalje cultivar possessed few but large lateral The numerical approach developed in our study can also be
roots, whereas the Beaupré had twice as many smaller lateral used to explore in detail the role of each of the structural
roots and also more total roots (115 structural roots in Beaupré, elements involved in tree uprooting. In particular, how any
85 structural root in the Raspalje root system). Because the given root interacts within the root–soil matrix can be
stiffness of a root is a function of its diameter to the fourth determined by isolating and removing that root from the
power, numerous small roots will decrease anchorage rigidity model, or by adding new roots (Dupuy et al., 2005a; Fourcaud
but increase resistance in tension (Stokes et al., 1995). The et al., 2007). Such virtual experiments are easier to perform
‘‘group effect’’ of large lateral roots near the tree trunk will than complex field experiments, whereby root systems are
cause plastic failure in the soil to occur further from the tree physically manipulated (Khuder et al., 2007).
axis, thus augmenting resistance to failure (Coutts, 1983;
Dupuy et al., 2005a). The simulated failure zone was larger
with the Raspalje than with the Beaupré cultivar. This LIST OF SYMBOLS
difference in the size of plastic failure in the soil is probably
due to the effect of the large lateral roots in the Raspalje tree. BEAM: Structure recording beam definition in FEM input file
Clonal, morphological differences affect the anchorage NODE: Structure recording node definition in FEM input file
efficiency of poplar trees in a plantation (Harrington and i: line number in Multi Tree Graph (MTG) file
DeBell, 1996). Uprooting resistance was higher in the Raspalje
cultivar than the Beaupré in both the numerical and real o(i): Branching order of the root at the ith line in the MTG
uprooting tests. This increase in anchorage efficiency can be file
explained by the high root volume and number of large lateral N(o): Number of the anterior node of branching order o
roots in the Raspalje tree. The resistance of lateral roots held in , .: A line in the NODE or BEAM list
tension during uprooting is a major component of anchorage : Addition of a new line in the NODE or BEAM list
(Coutts, 1983, 1986; Stokes et al., 1995; Crook and Ennos, Xs: Soil nodal coordinates of the sth soil node
1996) because of the high tensile resistance of wood (Genet et Xr: Root nodal coordinates of the rth root node
al., 2005). Numerous thin lateral roots, even if they are not
large, will augment uprooting resistance. However, our model þ: Symbol indicating the initiation of a new branch in
does not yet include the anisotropic material properties of root the MTG file
wood, and including such properties is a priority for future ,: Symbol indicating a new portion of root in the MTG
modelling studies. file
Although root architecture has long been considered a major MTG: Multiscale Tree Graph. Architectural data file for the
component of root anchorage (Coutts, 1983, 1986; Danjon et coding of plant structure
al., 2005; Dupuy et al., 2005a, b; Khuder et al., 2007), we c: Soil cohesion
showed that soil cohesion c was also an important factor
affecting uprooting resistance. The weight of the system alone /: Soil friction angle
was not sufficient to produce enough isostatic pressure to w: Soil dilation angle
prevent soil failure according to the Mohr–Coulomb theory. m: Poisson coefficient
Additionally, the soil that is lifted on the windward side of the c: Volumetric weight
tree during uprooting is subject to low or negative isostatic
pressure, thus increasing sensitivity to variations in soil shear
strength, which was triggered by the soil cohesion c. Soil LITERATURE CITED
friction angle also influenced uprooting resistance, but dilation ACHIM, A., J.-C. RUEL, AND B. GARDINER. 2005. Effect of precommercial
angle less so. Chen and Martin (2002) showed that these two thinning on the resistance of balsam fir to windthrow. Canadian
parameters were more important than soil cohesion when Journal of Forest Research 35: 1844–1853.
September 2007] D UPUY ET AL .—3D TREE ANCHORAGE MODEL 1513

BERGONZINI, J. C., AND O. LAROUSSINIE. 2000. Les écosystèmes forestiers GARDINER, B. A., H. PELTOLA, AND S. KELLOMAKI. 2000. Comparison of two
dans les tempêtes. ECOFOR, Ministère de ’Agriculture et de la Pêche, models for predicting the critical wind speeds required to damage
Paris, France. coniferous trees. Ecological Modelling 129: 1–23.
BLACKWELL, P. G., K. RENNOLLS, AND M. P. COUTTS. 1990. A root GENET, M., A. STOKES, F. SALIN, S. B. MICKOVSKI, T. FOURCAUD, J.-F.
anchorage model for shallowly rooted Sitka spruce. Forestry 63: 73– DUMAIL, AND R. VAN BEEK. 2005. The influence of cellulose content on
92. tensile strength in tree roots. Plant and Soil 278: 1–9.
BLOOMFIELD, J. 2000. The potential impacts of global warming on HARRINGTON, C. A., AND D. S. DEBELL. 1996. Above- and below-ground
America’s forests. Critical findings for forest lands from the First characteristics associated with wind toppling in a young Populus
National Assessment of the potential consequences of climate plantation. Trees—Structure and Function 11: 109–118.
variability and change. Website www.climatehotmap.org/impacts/ JEONG, S., J. LEE, AND C. J. LEE. 2004. Slip effect at the pile–soil interface
forests.html [accessed 1 January 2007]. on dragload. Computer and Geotechnics 31: 115–126.
BRÜCHERT, F., AND B. GARDINER. 2006. The effect of wind exposure on the KACHANOV, L, M. 2004. Fundamentals of the theory of plasticity. Dover
tree aerial architecture and biomechanics of Sitka spruce (Picea Publications Inc., Mineola, New York, USA.
sitchensis, Pinaceae). American Journal of Botany 93: 1512–1521. KHUDER, H., A. STOKES, F. DANJON, AND K. GOUSKOU. 2007. Is it possible to
CHEN, C.-Y., AND G. R. MARTIN. 2002. Soil-structure interaction for manipulate root anchorage in young trees? Plant and Soil 294: 87–
landslide stabilizing piles. Computer and Geotechnics 29: 363–386. 102.
COUTTS, M. P. 1983. Root architecture and tree stability. Plant and Soil 71: MAYER, P., P. BRANG, M. DOBBERTIN, D. HALLENBARTER, J.-P. RENAUD, L.
171–188. WALTHERT, AND S. ZIMMERMANN. 2005. Forest storm damage is more
COUTTS, M. P. 1986. Components of tree stability in Sitka spruce on peaty frequent on acidic soils. Annals of Forest Science 62: 303–311.
gley soil. Forestry 59: 173–197. MICKOVSKI, S. B., AND A. R. ENNOS. 2003. Anchorage and asymmetry in
CROOK, M. J., AND A. R. ENNOS. 1996. The anchorage mechanics of deep the root system of Pinus peuce. Silva Fennica 37: 161–173.
rooted larch, Larix europea 3 L. japonica. Journal of Experimental MOORE, J. R. 2000. Differences in maximum resistive bending moments of
Botany 47: 1509–1517. Pinus radiata trees grown on a range of soil types. Forest Ecology
CUCCHI, V., AND D. BERT. 2003. Wind-firmness in Pinus pinaster Aı̈t. and Managment 135: 63–71.
stands in southwest France: influence of stand density, fertilisation NICOLL, B. C., A. ACHIM, S. MOCHAN, AND B. A. GARDINER. 2005. Does
and breeding in two experimental stands damaged during the 1999 steep terrain influence tree stability? A field investigation. Canadian
storm. Annals of Forest Science 60: 209–226. Journal of Forest Research 35: 2360–2367.
CUCCHI, V., C. MEREDIEU, S. A. F. DE COLIGNY, J. SUAREZ, AND B. GARDINER. NICOLL, B. C., AND D. RAY. 1996. Adaptive growth of tree root systems in
2005. Modelling the windthrow risk for simulated stands of maritime response to wind action and site conditions. Tree Physiology 16: 891–
pine. Forest Ecology and Management 213: 184–196. 898.
CUCCHI, V., C. MEREDIEU, A. STOKES, S. BERTHIER, D. BERT, M. NAJAR, A. NIKLAS, K. J., F. MOLINA-FREANER, C. TINOCO-OJANGUREN, AND D. J.
DENIS, AND R. LASTENNET. 2004. Root anchorage of inner and edge PAOLILLO JR. 2002. The biomechanics of Pachycerus pringlei root
trees in stands of maritime pine (Pinus pinaster Ait.) growing in systems. American Journal of Botany 89: 12–21.
different podzolic soil conditions. Trees—Structure and Function 18: PELTOLA, H. M. 2006. Mechanical stability of trees under static loads.
460–466. American Journal of Botany 93: 1501–1511.
DANJON, F., D. BERT, C. GODIN, AND P. TRICHET. 1999. Structural root PUTZ, F. E., P. COLEY, D. K. LU, A. MONTALVO, AND A. AIELLO. 1983.
architecture of 5-year-old Pinus pinaster measured by 3D digitising Uprooting and snapping of trees: structural determinants and
and analysed with AMAPmod. Plant and Soil 217: 49–63. ecological consequences. Canadian Journal of Forest Research 13:
DANJON, F., T. FOURCAUD, AND D. BERT. 2005. Root architecture and 1011–1020.
windfirmness of mature Pinus pinaster Aı̈t. New Phytologist 168: RIZZO, D. M., AND R. GROSS. 2000. Distribution of Armillaria on pear root
261–269. systems and a comparison of root excavation techniques. In A. Stokes
DI IORIO, A., B. LASSERRE, G. S. SCIPPA, AND D. CHIATANTE. 2005. Root [ed.], The supporting roots of trees and woody plants: form, function
system architecture of Quercus pubescens trees growing on different and physiology, vol. 87, 305–311. Kluwer, Dordrecht, Netherlands.
sloping conditions. Annals of Botany 95: 351–361. RUEL, J.-C. 2000. Factors influencing windthrow in balsam fir forests:
DUPUY, L. 2003. Modélisation de l’ancrage racinaire des arbres forestiers, from landscape studies to individual tree studies. Forest Ecology and
Ph.D. thesis, Bordeaux I, Bordeaux, France. Management 135: 169–178.
DUPUY, L., T. FOURCAUD, AND A. STOKES. 2005a. A numerical investigation SELLIER, D., T. FOURCAUD, AND P. LAC. 2006. A finite element model for
into factors affecting the anchorage of roots in tension. European investigating effects of aerial architecture on tree oscillations. Tree
Journal of Soil Science 56: 319–327. Physiology 26: 799–806.
DUPUY, L., T. FOURCAUD, AND A. STOKES. 2005b. A numerical investigation SOETHE, N., J. LEHMANN, AND C. ENGELS. 2006. Root morphology and
into the influence of soil type and root architecture on tree anchorage. anchorage of six native tree species from a tropical montane forest
Plant and Soil 278: 119–134. and an elfin forest in Ecuador. Plant and Soil 279: 173–185.
ENNOS, A. R. 1990. The anchorage of leek seedlings: the effect of root STOKES, A. 1999. Strain distribution during anchorage failure of Pinus
length and soil strength. Annals of Botany 65: 409–416. pinaster Ait. at different ages and tree growth response to wind-
FOURCAUD, T., L. DUPUY, D. SELLIER, P. ANCELIN, AND P. LAC. 2003b. induced root movement. Plant and Soil 217: 17–27.
Application of plant architectural models to biomechanics. In B. Hu STOKES, A., M. ABDGHANI, F. SALIN, F. DANJON, H. JEANNIN, S. BERTHIER,
and M. Jaeger [eds.], Proceedings of international symposium on A. KOKUTSE, AND H. FROCHOT. 2007. Root morphology and strain
plant growth modeling, simulation, visualization and their applica- distribution during tree failure on mountain slopes. In A. Stokes, I.
tions, Beijing, China 2003, 462–479 Tsinghua University press, Spanos, J. E. Norris, and L. H. Cammeraat [eds.], Eco- and ground
Beijing, China. bio-engineering: the use of vegetation to improve slope stability.
FOURCAUD, T., J.-N. JI, Z.-Q. ZHANG, AND A. STOKES. 2007. Understanding Developments in plant and soil sciences, vol. 103, 165–173. Springer,
the impact of root morphology on overturning mechanisms: a Dordrecht, Netherlands.
modelling approach. Annals of Botany: in press. STOKES, A., A. H. FITTER, AND M. P. COUTTS. 1995. Responses of young
FOURCAUD, T., P. LAC, P. CASTÉRA, AND P. DE REFFYE. 2003a. Numerical trees to wind and shading: effects on root architecture. Journal of
modelling of shape regulation and growth stresses in trees. II. Experimental Botany 46: 1139–1146.
Implementation in the AMAPpara software. Trees—Structure and STOKES, A., AND C. MATTHECK. 1996. Variation of wood strength in tree
Function 17: 31–39. roots. Journal of Experimental Botany 47: 693–699.
FRASER, A. I. 1962. The soil and roots as factors in tree stability. Forestry STOKES, A., F. SALIN, A. D. KOKUTSE, S. BERTHIER, H. JEANNIN, S. MOCHAN,
23: 90–95. L. DORREN, N. K. KOKUTSE, AND T. FOURCAUD. 2005. Mechanical
1514 A MERICAN J OURNAL OF B OTANY [Vol. 94

resistance of different tree species to rockfall in the French Alps. YANG, Z., AND B. JEREMIC. 2002. Numerical analysis of pile behaviour
Plant and Soil 278: 107–117. under lateral loads in layered elastic–plastic soils. International
WHITLOW, R. 1995. Basic soil mechanics. Longman Scientific & Journal for Numerical and Analytical Methods in Geomechanics 26:
Technical, Harlow, Essex, UK. 1385–1406.

Das könnte Ihnen auch gefallen