Sunteți pe pagina 1din 33

CALTECH ASCI TECHNICAL REPORT 093

caltechASCI/2000.093



Three Dimensional Finite-Element Simulation of the
Dynamic Brazilian Tests on Concrete cylinders
G. Ruiz, M. Ortiz, A. Pandolfi


INTERNATIONAL JOURNAL FOR NUMERICAL METHODS IN ENGINEERING
Int. J. Numer. Meth. Engng 2000; 48:963994
Three-dimensional nite-element simulation of the
dynamic Brazilian tests on concrete cylinders
Gonzalo Ruiz
1
, Michael Ortiz
2;;
and Anna Pandol
3
1
Departamento de Ciencia de Materiales; Universidad Polit ecnica de Madrid; 28040 Madrid; Spain
2
Graduate Aeronautical Laboratories; California Institute of Technology; Pasadena; CA 91125; U.S.A.
3
Dipartimento di Ingegneria Strutturale; Politecnico di Milano; 20133 Milano; Italy
SUMMARY
We investigate the feasibility of using cohesive theories of fracture, in conjunction with the direct simulation
of fracture and fragmentation, in order to describe processes of tensile damage and compressive crushing in
concrete specimens subjected to dynamic loading. We account explicitly for microcracking, the development
of macroscopic cracks and inertia, and the eective dynamic behaviour of the material is predicted as an
outcome of the calculations. The cohesive properties of the material are assumed to be rate-independent and
are therefore determined by static properties such as the static tensile strength. The ability of model to predict
the dynamic behaviour of concrete may be traced to the fact that cohesive theories endow the material with
an intrinsic time scale. The particular conguration contemplated in this study is the Brazilian cylinder test
performed in a Hopkinson bar. Our simulations capture closely the experimentally observed rate sensitivity of
the dynamic strength of concrete in the form of a nearly linear increase in dynamic strength with strain rate.
More generally, our simulations give accurate transmitted loads over a range of strain rates, which attests to
the delity of the model where rate eects are concerned. The model also predicts key features of the fracture
pattern such as the primary lens-shaped cracks parallel to the load plane, as well as the secondary profuse
cracking near the supports. The primary cracks are predicted to be nucleated at the centre of the circular
bases of the cylinder and to subsequently propagate towards the interior, in accordance with experimental
observations. The primary and secondary cracks are responsible for two peaks in the load history, also in
keeping with experiment. The results of the simulations also exhibit a size eect. These results validate the
theory as it bears on mixed-mode fracture and fragmentation processes in concrete over a range of strain
rates. Copyright ? 2000 John Wiley & Sons, Ltd.
KEY WORDS: concrete; fracture; cohesive elements; dynamic strength; mixed mode fracture; size eect;
strain rate eect

Correspondence to: Michael Ortiz, Graduate Aeronautical Laboratories, California Institute of Technology, Firestone
Flight Sciences Laboratory, Pasadena, CA 91125, U.S.A.

E-Mail: ortiz@madrid.caltech.edu
Contract=grant sponsor: Direcci on General de Ense nanza Superior, Ministerio de Educaci on y Cultura, Spain
Contract=grant sponsor: Army Research Oce; contract=grant number: DAAH04-96-1-0056
Received 29 January 1999
Copyright
?
2000 John Wiley & Sons, Ltd. Revised 15 September 1999
964 G. RUIZ, M. ORTIZ AND A. PANDOLFI
1. INTRODUCTION
The dynamic behaviour of brittle materials, including glasses, ceramics, rocks and concrete, sub-
jected to high strain rates often involves the development of complex fracture and fragmentation
patterns. Cracks may nucleate from pre-existing aws or defects which populate the material in
large numbers, or at stress concentrations induced by heterogeneities present in the material on
the microscale. Once nucleated, fracture may proceed in a distributed fashion, as microcracking,
or in the form of a few dominant macrostructural cracks. The paths followed by such cracks are
often tortuous and undergo frequent branching, specially under dynamic conditions. Cracks may
link up with free surfaces or with each other to form fragments. In concrete, the microstructural
length scale is commensurate with the aggregate size and fracture processes often occur on a scale
comparable to the geometrical dimensions of the structure. In addition to the energy required to
fracture the material, other sources of dissipation often operate simultaneously, including plastic
working, viscosity and heat conduction. Finally, microinertia often plays a signicant role in shap-
ing the eective macroscopic behaviour of solids at high rates of deformation (e.g. References
[1, 2]).
Several avenues have been traditionally followed in order to model this complex behaviour. One
avenue is to attempt to bury all dissipative mechanisms into the constitutive relations. However, in
practice the eective behaviour of systems with complex microsctructures can only be characterized
analytically, if at all, by recourse to sweeping simplifying assumptions. One particularly vexing
diculty inherent to constitutive descriptions concerns their inability to endow brittle materials
with a well-dened fracture energy, i.e. a material-characteristic measure of energy dissipation
per unit area of crack surface. Thus, the conventional thermodynamic route to the formulation of
general constitutive relations regards dissipation as an extensive quantity possessing a well-dened
density per unit volume. Attempts to build fracture into such formulations by means of a softening
stressstrain law leads to pathologies such as a spurious scaling of the eective fracture energy
with the discretization size. The formulation of appropriate functional forms of the energy of a
solid which allow for both bulk and fracture-like behaviour is a challenging mathematical problem
which is presently the subject of active research (e.g. References [3, 4]).
Fracture mechanics specically addresses the issue of whether a body under load will remain
intact or new free surface will form. However, the classical theory of fracture mechanics is pred-
icated around the assumption of a pre-existing dominant crack. Under these and other restrictive
assumptions, fracture mechanics successfully predicts when a crack will grow, in what direc-
tion and how fast, and how the results of laboratory experiments can be scaled up to structural
dimensions. However, the reliance on a pre-existing dominant crack entirely forgoes the issue of
nucleation, which is for the most part foreign to classical fracture mechanics. Situations, such as this
arise in fragmentation or crushing, involving many intersecting cracks also fall outside the purview
of classical fracture mechanics. Additionally, the conditions for crack growth are typically expressed
in terms of parameters characterizing the amplitude of autonomous near-tip elds. The applicability
of these criteria is therefore contingent upon the existence and establishment of such near-tip elds.
This in turn severely restricts the scope of the theory, e.g. by requiring that the plastic or process
zone be small relative to any limiting geometrical dimension of the solid such as the ligament size,
conditions which are rarely realized in materials such as concrete. In addition, the structure of the
autonomous near-tip elds depends sensitively on the constitutive behaviour of the material, which
inextricably ties the fracture criteria to the constitution of the material. In dynamic fracture, the frac-
ture criterion is additionally responsible for accounting for the microinertia which accompanies the
Copyright ? 2000 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2000; 48:963994
THREE-DIMENSIONAL FINITE-ELEMENT SIMULATION 965
motion of the crack tip. Another complicating factor is non-proportional and cyclic loading, such
as arises in high-cycle fatigue, which renders important near-tip parameters, such as the J-integral,
inapplicable.
An alternative approach to fracture is based on the use of cohesive models to describe processes
of separation leading to the formation of new free surface [530]. Cohesive theories of fracture
address many of the same issues contemplated by classical fracture mechanics while eectively
overcoming most of the diculties alluded to above. Indeed, cohesive models furnish a complete
theory of fracture, and permit the incorporation into the material description of bona de frac-
ture parameters such as a fracture energy and a spall strength. By focusing specically on the
separation process, a sharp distinction is drawn in cohesive theories between fracture, which is
described by recourse to cohesive laws, and bulk material behaviour, which is described through
an independent set of constitutive relations. The use of cohesive models is therefore not limited
by any consideration of material behaviour, nite kinematics, non-proportional loading, dynamics,
or the geometry of the specimen.
Another appealing aspect of cohesive laws as models of fracture is that they t naturally within
the conventional framework of nite element analysis. However, one important numerical require-
ment pertaining to the use of cohesive theories is that the cohesive zones must be adequately
resolved by the discretization in order to obtain mesh-size-independent results [23]. In materials
for which the characteristic cohesive length is small, this may necessitate substantial mesh rene-
ment near crack tips. In materials such as concrete, by contrast, the characteristic cohesive length
is comparatively large, typically of the order of 10 cm, and the mesh-size requirements are often
less severe.
The feasibility of using cohesive theories of fracture for the direct simulation of fragmentation
processes in brittle materials subjected to impact was established by Camacho and Ortiz [23].
In this approach, individual cracks are tracked as they nucleate, propagate, branch and possibly
link up to form fragments. Clearly, it is incumbent upon the mesh to provide a rich enough set
of possible fracture paths, an issue which may be addressed within the framework of adaptive
meshing. The ensuing granular ow of the comminuted material is also simulated explicitly. In
this manner, the daunting task of developing constitutive relations which account for crushing and
fragmentation is avoided altogether.
The delity of cohesive elements in applications involving dynamic fracture of ductile materials
has recently been investigated by Pandol et al. [30], who have simulated the fragmentation of
expanding aluminium rings. The numerical simulations have been found to be highly predictive
of a number of observed features, including: the number of dominant and arrested necks; the
fragmentation patterns; the dependence of the number of fragments and the fracture strain on the
expansion speed; and the distribution of fragment sizes at xed expansion speed. Pandol et al. [31]
have additionally shown that simulations of dynamic crack propagation based on cohesive models
are highly predictive of a number of observed features, including: the crack growth initiation time;
the trajectory of the propagating crack tip; and the formation of shear lips near the lateral surfaces.
Repetto et al. [32] have applied cohesive models to the simulation of failure waves in glass rods
subjected to impact. Their calculations correctly capture the development and propagation of a
sharp failure wave, its propagation speed and the bursting of the comminuted material following
the passage of the failure wave.
The simulations of dynamic fracture of concrete available in the literature commonly model the
tensile strength as an increasing function of strain rate [3343]. In addition, the fracture energy
is often presumed to be constant and independent of strain rate [33, 44]. These rate-dependency
Copyright ? 2000 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2000; 48:963994
966 G. RUIZ, M. ORTIZ AND A. PANDOLFI
laws are necessarily empirical and endeavor to model the eective or macroscopic behaviour of
concrete under dynamic loading.
In this paper we investigate the feasibility of using cohesive theories of fracture, in conjunction
with the direct simulation of fracture and fragmentation, in order to describe processes of tensile
damage and compressive crushing in concrete specimens subjected to dynamic loading. We account
explicitly for microcracking, the development of macroscopic cracks and inertia, and the eective
dynamic behaviour of the material is predicted as an outcome of the calculations. Indeed, the
cohesive properties of the material are assumed to be rate-independent, and are therefore determined
by static properties such as the static tensile strength. The ability of model to predict key aspects
of the dynamic behaviour of concrete, such as the strain-rate sensitivity of strength may be traced
to the fact that cohesive theories, in addition to building a characteristic length into the material
description, they endow the material with an intrinsic time scale as well [23]. This intrinsic time
scale permits the material to discriminate between slow and fast loading rates and ultimately allows
for the accurate prediction of the dynamic strength of the material as a function of strain rate and
other rate eects.
The particular conguration contemplated in this study is the Brazilian cylinder test performed in
a Hopkinson bar. Dynamic Brazilian tests performed using a split-Hopkinson pressure bar (SHPB)
have been proposed as a convenient means of determining the tensile strength of cohesive materials
[41, 37, 36, 4547]. These tests have yielded a wealth of information on the dynamic behaviour
of concrete. For instance, it is well-established that the eective tensile strength of the material
increases with strain rate [41, 37, 36, 33]. By contrast, the fracture energy is ostensible insensitive
to strain rate [33, 44]. Detailed measurements of the near-tip stress elds attendant to cracks
propagating in mode I and mixed mode [35] have shown that the length of the cohesive or
crack-bridging zone at a crack tip decreases with increasing strain rate.
Dynamic Brazilian tests therefore furnish a convenient yet exacting validation test for cohesive
theories of fracture applied to concrete. Our simulations give accurate transmitted loads over a
range of strain rates, which attests to the delity of the model where rate eects are concerned.
The model also predicts key features of the fracture pattern such as the primary lens-shaped cracks
parallel to the load plane, as well as the secondary profuse cracking near the support. The primary
cracks are predicted to be nucleated at the centre of the circular bases of the cylinder and to
subsequently propagate towards the interior, in accordance with experimental observations. The
primary and secondary cracks are responsible for two peaks in the load history, also in keeping
with experiment. The results of the simulations also exhibit a size eect, i.e. a dependence of the
eective behaviour on the size of the specimen [48]. These results validate the theory as it bears
on mixed-mode fracture and fragmentation processes in concrete over a range of strain rates. Our
simulations also provide useful insights into the interpretability of the dynamic Brazilian test, and
the sensitivity of the measurement to details of the experimental design, issues which have been
addressed extensively in the experimental literature [4952].
The organization of the paper is as follows. A brief outline of the particular form of the cohesive
models and their nite-element implementation employed in calculations is given in Section 2. In
Section 3 we begin by compiling the main parameters which dene the simulations, including
the specimen geometry, material parameters, load and boundary conditions, and the details of the
mesh design. The section closes with a detailed comparison of the results of the calculations
with corresponding experimental observations, including load and energy histories, crack patterns
and estimated crack velocities. In Section 4 we proceed to explore the eect of variations in
selected parameters of the model, including strain rate, specimen size, the width of the bearing
Copyright ? 2000 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2000; 48:963994
THREE-DIMENSIONAL FINITE-ELEMENT SIMULATION 967
Figure 1. Cohesive surface traversing a 3D body. Figure 2. Loadingunloading rule from linearly de-
creasing loading envelop expressed in terms of an
eective opening displacement o and traction t.
strips, specimen size and the ratio [ of modes II to I toughness. Finally, a summary and some
concluding remarks are collected in Section 5.
2. FINITE-ELEMENT MODEL
By way of a general framework, we start by considering a deformable body occupying an initial
conguration B
0
R
3
. The boundary cB
0
of the body is partitioned into a displacement boundary
cB
0,1
and a traction boundary cB
0, 2
. The body undergoes a motion described by a deformation
mapping D: B
0
[0, 1] R
3
, where [0, 1] is the duration of the motion, under the action of body
forces j
0
b and prescribed boundary tractions

t applied over cB
0, 2
. The attendant deformation
gradients are denoted F and the rst PiolaKirchho stress tensor P (cf. e.g. Reference [53]).
In addition, the solid contains a collection of cohesive cracks. The locus of these cracks on the
undeformed conguration is denoted S
0
(Figure 1).
Under these conditions, the weak form of linear momentum balance, or virtual work expression,
takes the following form:

B
0
[j
0
(b D) W P
0
W] dJ
0

S
0
t <W= dS
0
+

cB
0 , 2

t W dS
0
= 0 (1)
where a superposed dot denotes the material time derivative,
0
is the material gradient, W is
an arbitrary virtual displacement satisfying homogeneous boundary conditions on cB
0,1
, t are the
cohesive traction over S
0
, and < = denotes the jump across an oriented surface.
From (1) is evident that the presence of a cohesive surface results in the addition of a new
term to the virtual work expression. In order to complete the denition of the problem, a set of
constitutive relations for the cohesive tractions t must be provided, in addition to the conventional
Copyright ? 2000 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2000; 48:963994
968 G. RUIZ, M. ORTIZ AND A. PANDOLFI
constitutive relations describing the bulk behaviour of the material. To this end, we postulate the
existence of a free energy density per unit undeformed area over S
0
of the general form
[=[(T, 0, q; n) (2)
where
T =<D= (3)
are the opening displacements over the cohesive surface, 0 is the local temperature, q is some
suitable collection of internal variables describing the current state of decohesion of the surface, and
n is the unit normal to the cohesive surface in the deformed conguration. The explicit dependence
of [ on n is required to allow for dierences in cohesive behaviour for opening and sliding. By
recourse to the Coleman and Noll method (e.g. References [54, 55]) it is possible to show that
the cohesive law takes the form
t =
c[
cT
(4)
The potential structure of the cohesive law is a consequence of the rst and second laws of
thermodynamics. The evolution of the internal variables q is governed by a set of kinetic relations
of the general form
q =f(T, 0, q) (5)
A more general class of free energies which allows for surface anisotropy and nite opening
displacements have been considered by Ortiz and Pandol [28].
Following Camacho and Ortiz [23] and others [27, 28], we consider a simple class of mixed-
mode cohesive laws accounting for tensionshear coupling obtained by the introduction of an
eective opening displacement:
o =

[
2
o
2
S
+ o
2
n
(6)
where
o
n
=T n (7)
is the normal opening displacement and
o
S
=|T
S
| =|T o
n
n| (8)
is the magnitude of the sliding displacement. The parameter [ assigns dierent weights to the
sliding and normal opening displacements.
Assuming that the free energy potential [ depends on T only through the eective opening
displacement o, i.e.
[=[(o, 0, q) (9)
Copyright ? 2000 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2000; 48:963994
THREE-DIMENSIONAL FINITE-ELEMENT SIMULATION 969
the cohesive law (4) reduces to
t =
t
o
([
2
T
S
+ o
n
n) (10)
where
t =
c[
co
(o, 0, q) (11)
is a scalar eective traction, which expression follows from (6) and (10):
t =

[
2
|t
S
|
2
+ t
2
n
(12)
where t
S
and t
n
are the shear and the normal traction, respectively. From this relation, we observe
that [ denes the ratio between the shear and the normal critical tractions. In brittle materials, this
ratio may be estimated by imposing lateral connement on specimens subjected to high-strain-rate
axial compression [56, 57]. In addition, the parameter [ roughly denes the ratio of K
IIc
to K
Ic
of
the material.
Upon closure, the cohesive surfaces are subjected to the contact unilateral constraint, including
friction. We regard contact and friction as independent phenomena to be modelled outside the
cohesive law. Friction may signicantly increase the sliding resistance in closed cohesive sur-
faces. In particular, the presence of friction may result in a steadyor even increasingfrictional
resistance while the normal cohesive strength simultaneously weakens.
Figure 2 depicts a particular type of irreversible cohesive laws envisioned here; o
c
is the tensile
strength and o
c
the critical opening displacement. Irreversibility manifests itself upon unloading.
Therefore, an appropriate choice of internal variable is the maximum attained eective opening
displacement o
max
. Loading is then characterized by the conditions: o =o
max
and

o0. Conversely,
we shall say that the cohesive surface undergoes unloading when it does not undergo loading. We
assume the existence of a loading envelop dening a relation between t and o under conditions of
loading. A simple and convenient relation is furnished by the linearly decreasing envelop shown
in Figure 2. Following Camacho and Ortiz [23] we shall assume unloading to the origin, Figure 2,
giving
t =
t
max
o
max
o if o o
max
or

o 0 (13)
For the present model, the kinetic relations (5) reduce to a straightforward computation of o
max
.
It is a well-known fact [58, 58] that cohesive theories introduce a well-dened length scale
into the material description and, in consequence, are sensitive to the size of the specimen. The
characteristic length of the material is

c
=
EG
c
[
2
ts
(14)
where G
c
is the fracture energy and [
ts
the static tensile strength. Camacho and Ortiz [23] have
noted that in conjunction with inertia cohesive models introduce a characteristic time as well.
This characteristic or intrinsic time is
t
c
=
jco
c
2[
ts
(15)
Copyright ? 2000 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2000; 48:963994
970 G. RUIZ, M. ORTIZ AND A. PANDOLFI
Figure 3. Geometry of cohesive element. The surfaces S

and S
+
coincide in the reference
conguration of the solid.
where j is the mass density and c the longitudinal wave speed. Owing to this intrinsic time scale,
the material behaves dierently when subjected to fast and slow loading rates. This sensitivity
to the rate of loading confers to the cohesive models the ability to reproduce subtle features of
the dynamic behaviour of brittle solids, such as crack-growth initiation times and propagation
speeds [31]; and the dependence of the pattern of fracture and fragmentation on strain rate [30].
In addition, the calculations presented subsequently demonstrate the ability of cohesive theories to
account for the dynamic strength of brittle solids, i.e. the dependence of the dynamic strength on
strain rate.
Another appealing aspect of cohesive laws as models of fracture is that they t naturally within
the conventional framework of nite-element analysis. We follow Camacho and Ortiz [23] and
adaptively create new surfaces as required by the cohesive model by duplicating nodes along
previously coherent element boundaries. The introduction of cohesive surfaces may result in drastic
changes in the topology of the model [29]. The nodes are subsequently released in accordance
with a tensionshear cohesive law. The particular class of cohesive elements used in calculations
has been developed by Ortiz and Pandol [28] and consists of two six-node triangles endowed
with quadratic displacement interpolation (Figure 3).
Inserting the displacement interpolation into the virtual work expression (1) leads to a system
of semi-discrete equations of motion of the form
M x + f
int
(x) =f
ext
(t) (16)
where x is the array of nodal co-ordinates, M is the mass matrix, f
ext
is the external force array,
and f
int
is the internal force array. In calculations we use the second-order accurate central dier-
ence algorithm to discretize (16) in time [5961]. Despite the fact that the time step is bounded
by stability [60], explicit integration is particularly attractive in three-dimensional calculations,
where implicit schemes lead to system matrices which often exceed the available in-core storage
capacity. Yet another advantage of explicit algorithms is that they are ideally suited for concurrent
computing [62].
Copyright ? 2000 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2000; 48:963994
THREE-DIMENSIONAL FINITE-ELEMENT SIMULATION 971
3. APPLICATION OF COHESIVE MODEL TO THE PREDICTION OF
THE DYNAMIC BEHAVIOUR OF CONCRETE
We proceed to simulate the dynamic behaviour of concrete using the cohesive model just described.
The particular conguration contemplated in the study is the Brazilian cylinder test performed in a
Hopkinson bar. The simulations give accurate transmitted loads over a range of strain rates, which
attests to the delity of the model where rate eects are concerned. The model also predicts key
features of the fracture pattern such as the primary lens-shaped cracks parallel to the load plane,
as well as the secondary profuse cracking near the support. The primary cracks are predicted to be
nucleated at the centre of the circular bases of the cylinder and to subsequently propagate towards
the interior, in accordance with experimental observations. The primary and secondary cracks are
responsible for two peaks in the load history, also in keeping with experiment.
These results validate the theory as it bears on mixed-mode fracture and fragmentation processes
in concrete over a range of strain rates. It is particularly noteworthy that neither the bulk material
behaviour, which remains essentially elastic throughout the calculations reported here, nor the
cohesive law are themselves rate dependent. As noted earlier, cohesive laws, in conjunction with
inertia, endow the material with a characteristic or intrinsic time scale, an attribute which ultimately
accounts for the ability of cohesive theories to capture rate eects accurately.
3.1. Experimental set-up
Dynamic Brazilian tests performed using a split-Hopkinson pressure bar (SHPB) have been pro-
posed as a convenient means of determining the tensile strength of cohesive materials
[41, 37, 36, 4547]. The SHPB consists of an incident bar and a transmitter bar, with a short
specimen placed between them. In addition, a striker bar impacts the incident bar and produces
a longitudinal compressive pulse which propagates toward the specimen (Figure 4(a)). The spec-
imen is typically in the form of a cylinderalthough other geometries can be used, including
cubes and prismsand it is placed transversely to the bars, i.e. such that the axis of the specimen
is perpendicular to the axis of the bars. The specimen is connected to the bars by two bearing
strips (Figure 4(b)). The amplitude and duration of the pulse depends on the material properties,
velocity and length of the striker. The pulse is partially reected back into the incident bar, and
partially transmitted through the specimen. As is well known [63], under quasistatic conditions the
diametral loading of a cylindrical specimen generates tension perpendicular to the loading plane
which eventually causes the specimen to split. This quasistatic solution is often taken as a basis
for interpreting the dynamic experiments as well [51].
The bars are designed to remain elastic throughout the test, and by virtue of their slenderness,
one-dimensional stress-wave theory applies to the bars to a good approximation. The strain cor-
responding to the incident, reected and transmitted pulses may then be measured using gauges
glued to the surface of the bars, and these signals, recorded by means of an oscilloscope and
a data acquisition system, may be used to calculate the corresponding stress pulses (Figure
5). In addition, the dynamic splitting tensile stress, [
td
, may be inferred from the following
equation:
[
td
=
2P
max
WD
(17)
where P
max
is the maximum load transmitted through the cylinder and W and D are the width and
diameter of the cylinder, respectively. P
max
is in turn calculated from the maximum transmitted
Copyright ? 2000 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2000; 48:963994
972 G. RUIZ, M. ORTIZ AND A. PANDOLFI
Figure 4. Experimental set-up (a), and details of the specimen (b).
Figure 5. Typical stress pulses (adapted from Reference [41]). Figure 6. Simplied inci-
dent load pulse used in
the analysis.
stress, o
max
, with the result
P
max
=R
2
o
max
(18)
where R is the SHPB radius. The strain rate follows as
c =
[
td
Et
(19)
where E is the elastic modulus of the specimen material and t is the time delay between the start
and the peak of the transmitted pulse.
Copyright ? 2000 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2000; 48:963994
THREE-DIMENSIONAL FINITE-ELEMENT SIMULATION 973
Table I. Concrete parameters.
Density (in Reference [40]) 2405 kg}m
3
Elastic modulus 37.9 GPa
Acoustic velocity 3620 m}s
Static compressive strength
Cylinder 152 305 mm 48.3 MPa
Cylinder 51 51 mm 57.1 MPa
Static splitting tensile strength
Cylinder 51 51 mm 3.86 MPa
Static direct tensile strength
Cylinder 51 51 mm, square notch 4.53 MPa
Fracture energy 66.2 N}m
Characteristic length 122 mm
3.2. Specimen geometry and material parameters
We specically aim to simulate the experiments reported by Hughes et al. [37], Tedesco et al.
[41] and Ross et al. [36]. The specimens in these experiments were concrete cylinders of 50.8 mm
in diameter and height obtained by coring from a concrete block. The maximum aggregate size
was 8.5 mm, or
1
6
of the specimen diameter. The corresponding material parameters are listed in
Table I. All these parameters were measured through independent tests with the exception of the
fracture energy, which we have estimated following the recommendations set forth in the Model
Code [64].
It is important to bear in mind the uncertainties which are inevitably inherent to the material
properties listed in Table I. Thus, for instance, the measured compressive strength depends on
both the shape and size of the specimen, ranging from 48.3 MPa, corresponding to the standard
specimen [65], to 57.1 MPa, when the 5151 mm cylinder is used. This variation is in agreement
with the expected size eect in concrete, namely, an increasing strength with decreasing specimen
size and increasing stiness of the loading device. In the calculations of interest here, however,
the response of the specimen is dominated by tensile cracking and the compressive strength of the
material is seldom reached. Under these conditions, the bulk behaviour may simply be modelled
as elastic to a good approximation (see, e.g., Reference [48, Section7.1.6]). As a slightand fairly
inconsequentialimprovement, we assume that the material obeys J
2
-plasticity upon the attainment
of its compressive strength. Details of the specic implementation of J
2
-plasticity employed in
calculations may be found elsewhere [66].
The tensile strength is highly dependent on the testing technique and conventions adopted for
its measurement. The static direct tensile strength of 4.53 MPa listed in Table I was measured
by Hughes et al. [37], Tedesco et al. [41] and Ross et al. [36] using cylinders similar to those
employed in the dynamic splitting tests. The specimens included a circumferential square notch
3.2 mm deep around their mid-section. In addition, the same authors performed static Brazilian tests
on specimens of the same geometry and measured a tensile strength of 3.86 MPa. However, it has
been shown by Rocco [52] that the static Brazilian test yields estimates of the tensile strength which
are strongly dependent on the shape and size of the specimen, and on the boundary conditions,
Copyright ? 2000 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2000; 48:963994
974 G. RUIZ, M. ORTIZ AND A. PANDOLFI
and are, therefore, unreliable. In view of these uncertainties, we have employed the tensile strength
of 4.53 MPa measured by direct tensile testing on notched specimens.
The material is assumed to obey the linear irreversible cohesive law shown in Figure 2 with the
measured tensile strength taken as the cohesive strength of the material. According to this law, the
material remains coherent up to the attainment of its tensile strength. This criterion is veried at
all interfaces between tetrahedral elements throughout the calculations. When the criterion is met,
a cohesive element is inserted at the interface, which subsequently opens in accordance with the
linearly decreasing cohesive law. This approach is particularly well suited to explicit dynamics as
it does not aect the critical time step.
There is experimental evidence [6769] which suggests that the intrinsic fracture toughness of
concrete, i.e. the critical stress intensity factor required to advance a semi-innite crack within its
plane in the absence of kinking, is much larger in pure mode II that in pure mode I owing to the
interlocking of aggregate particles. This in turn suggests adopting a large value of the coupling
parameter [ in (6), since, as remarked earlier, [ gives the ratio of modes II to I fracture toughness.
Based on a suggestion by Gustafsson and Hillerborg [70] on the relative strengths of concrete in
tension and shear, in calculations we take [ =10. Under the assumptions just stated, a semi-innite
crack subjected to mixed-mode loading will tend to kink at an angle roughly corresponding to the
maximum circumferential stress in its K-eld. Indeed, the maximum circumferential stress criterion
is known to lead to accurate predictions of crack paths in concrete [71, 72].
It should be carefully noted that, under the conditions envisioned in this study, and more gener-
ally when dealing with the fracture of concrete, conventional concepts from linear-elastic fracture
mechanics must be applied with caution or not at all. Thus, from (14) and the properties listed
in Table I one computes a characteristic length for concrete of 122 mm, which is larger than
the diameter of the specimen, or 50.8 mm. The conditions of the analysis correspond to a fully
yielded situation and linear-elastic fracture mechanics, which fundamentally rests on the small-
scale yielding condition, does not apply. Cohesive theories of fracture, by way of contrast, are not
so constrained, and can be applied to situations such as envisioned here in which the size of the
process zone is comparable to the ligament length or any other limiting geometrical dimension.
Further discussions of this and related issues may be found in References [73, 20, 48].
The simulations of dynamic fracture of concrete available in the literature commonly model the
tensile strength as an increasing function of strain rate [3343]. In addition, the fracture energy
is often presumed to be constant and independent of strain rate [33, 44]. These rate-dependency
laws are necessarily empirical and endeavor to model the eective or macroscopic behaviour of
concrete under dynamic loading. By way of contrast, we account explicitly for microcracking, the
development of macroscopic cracks and inertia, and the eective dynamic behaviour of the material
is predicted as an outcome of the calculations. In particular, we stress again that the cohesive law,
despite being rate-insensitive, endows the material with an intrinsic time scale which discriminates
between slow and fast loading rates and ultimately allows for the accurate prediction of the dynamic
strength of the material as a function of strain rate and other rate eects.
3.3. Load and boundary conditions
We consider several loading cases covering a range of load levels and strain rates and correspond-
ing to the tests reported by Hughes et al. [37], Tedesco et al. [41] and Ross et al. [36]. In the
actual tests, the width of the steel bearing strips connecting the specimen to the bars equalled
1
8
of the diameter of the cylinder. The load applied to the specimenand the load transmitted by
Copyright ? 2000 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2000; 48:963994
THREE-DIMENSIONAL FINITE-ELEMENT SIMULATION 975
Table II. Parameters for the incident load pulses.
Rise time Duration Stress level Related velocity
Load case no. t
r
(s) t
d
(s) o
i
(MPa) t (m}s)
1 in Reference [41]1 66 100 60.2 1.5
2 in Reference [41]2 72 100 72.8 1.8
1 in Reference [37]3 80 100 79.4 2.0
2 in Reference [37]4 85 100 122.5 3.1
3 in Reference [37]5 41 100 184.5 4.7
3 in Reference [41]6 48 100 264.3 6.7
itmay be readily obtained from the strain recorded by the gages glued to the bars. The loading
pulse can be idealized as a linearly rising ramp followed by a plateau (Figure 6).
In calculations it is convenient to simply prescribe displacements at the contact between the
striker bar and the specimen. To this end, the prescribed velocity at the contact may be estimated
as [74]
t =
o
i
jc
(20)
where j and c are the one-dimensional density and wave speed of the incident bar, respectively, and
o
i
is the incident stress. Equation (20) simply amounts to requiring that the motion of the specimen
matches the incident wave prole. The rise time, duration, and stress and velocity levels for the
loading cases under consideration are collected in Table II. By way of contrast, the contact between
the specimen and the transmitter bar is assumed to be rigid, leading to constrained displacements
over the width of the transmitter bearing strip. This simplication is warranted by the observation
that the transmitted pulses measured in the tests were very weak, roughly one order of magnitude
lower than the incident pulses.
3.4. Computational mesh
Figure 7 shows the mesh used in the calculations. All the surfaces and the interior of the specimen
are meshed automatically by the advancing front method [75] using 10-node quadratic tetrahedra.
It should be carefully noted that the midplaneor load planeof the specimen, which is a likely
crack path is not built into the mesh. Instead, the load plane is crossed by tetrahedra, (Figure
7(b)), which forces some degree of roughness on any macroscopic crack contained within it. Post-
mortem examination of actual concrete specimens tends to reveal a similar degree of roughness
in the form of protruding aggregate particles. The computational mesh comprises 8378 nodes and
5669 tetrahedra, (Figure 7(a)), and is designed so as to be ne and nearly uniform on and in the
vicinity of the load plane, with a size commensurate with half the maximum aggregate size, and
to gradually coarsen away from the load plane up to the maximum aggregate size.
It is important to note that the mesh size roughly ranges from
1
15
to
1
30
of the characteristic co-
hesive length (14) of the material and may, therefore, be expected to yield objective and mesh-size
insensitive results [23]. In order to verify this point, we have compared test solutions obtained from
three dierent meshes: the computational mesh just described and two increasingly ner meshes.
The results of the three tests are collected in Figure 8. The maximum transmitted load obtained
Copyright ? 2000 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2000; 48:963994
976 G. RUIZ, M. ORTIZ AND A. PANDOLFI
Figure 7. External appearance of the mesh (a), split specimen by the load plane (b) and
by the middle circular cross-section (c).
Figure 8. A comparison of numerical results corresponding to three meshes of varying degrees of re-
nement, illustrating the sensitivityor lack of it thereofof the various phases of the solution to mesh
size. (a) Histories of transmitted load and expended cohesive energy, and (b) asymptotic behaviour of the
cohesive energy consumption for the same mesh sizes.
Copyright ? 2000 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2000; 48:963994
THREE-DIMENSIONAL FINITE-ELEMENT SIMULATION 977
Figure 9. Comparison of experimental and numerical rate-sensitivity curves and dynamic strength of concrete.
from the intermediate and ne meshes diers from that obtained from the coarser computational
mesh by 2 per cent and +0.6 per cent, respectively, which we take to be acceptable error. In
addition, in all cases the cohesive energy consumption curves dier negligibly up to the peak load,
which provides an indication of the good accuracy of the coarse solution.
It is interesting to note that the post-peak solution, which is characterized by extensive micro-
cracking and macroscopic crack propagation and is therefore inherently unstable, is highly sensitive
to small changes in the parameters of the model, including the details of the mesh design. Thus,
because of the vast number of possible paths which the system can choose from, corresponding
to variations in local microcracking and fragmentation patterns, the solution inevitably exhibits a
marked random or stochastic character in the unstable region following the peak load. Indeed,
short-term and local uctuations in the solution exhibit broad variation even between meshes pos-
sessing the same element-size distribution. Under these highly stochastic conditions, convergence
of the solution can only reasonably be understood in terms of aggregate or average macroscopic
quantities such as energy, dissipation, microcrack densities, and other similar features. For instance,
in calculations the accumulated cohesive energy expenditure remains within a narrow band and
converges to a mesh-independent asymptote independent of mesh size (Figure 8(b)), which attests
to the requisite convergence in the mean.
3.5. Simulation results
Selected results of the calculations and comparisons with experimental data are shown in
Figures 911. The main features of these results, as regards dynamic strength load histories, crack
patterns, and crack velocity are next discussed in turn.
3.5.1. Dynamic strength and rate sensitivity. In keeping with the experimental procedure, we
identify the dynamic strength of the material with the peak transmitted load. Our computation of
the strain rate is based on the time interval between the start and the peak of the transmitted
load wave. In order to avoid spurious eects due to the bearing strip-cylinder contact, we have
estimated the experimental starting time of the load wave as the intersect of the tangent to the
linear ramp-up with the time axis.
Copyright ? 2000 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2000; 48:963994
978 G. RUIZ, M. ORTIZ AND A. PANDOLFI
Figure 10. Comparison of experimental and numerical transmitted load and energy histories for loading
case 2. (b) Experimental crack-pattern sequence [41].
Figure 9 compares the predicted and observed ratio of static to dynamic strengths for all loading
cases under consideration, and the dependence of the dynamic strength on strain rate. In addition,
the curve inset in Figure 9(a) represents a linear t to the experimentally observed strain-rate
dependence of the dynamic strength. As may be seen from Figure 9(a), the calculations capture
well the overall rate sensitivity of the material, which takes the form of a steady rise in dynamic
strength with increasing strain rate. It is interesting to note that the simulations corresponding to
loading cases 13 yield comparable results (Figure 9(a)), which may be due to the compensating
eect of a simulataneous increase in rise time and impact velocity in the load pulse (cf. Table II).
The accuracy in the calculation of the peak transmitted load and, by extension, of the dynamic
strength, Equation (17), is equally satisfactory, with the exception of loading case 1 (Figure
9(b)). Interestingly, the match between the numerical and experimental maximum transmitted loads
improves at high strain rate, with the error decreasing below 10 per cent for load cases 5 and 6.
This trend may be attributed to the fact that the eective opening displacements of the cracks at
the peak load tend to be larger at low strain rates, with the result that the computed peak load is
sensitively dependent on the shape of the cohesive law. Contrarily, at high strain rates the opening
displacements are comparatively smaller and the peak load correlates more closely with the tensile
strength. This suggests that the delity of the predicted peak loads at low strain rates may be
improved by a more realistic choice of cohesive law.
Again, it bears emphasis that these features of the dynamic behaviour of concrete are predicted
byand not built intothe theory. Indeed, the sole strength parameter of the theory is the static
strength of concrete. In addition, by itself, the cohesive law of the material is rate-independent. The
good qualitative and quantitative agreement between theory and experiment just reported suggests
that microinertia associated with microcracking and dynamic fracture is the dominant mechanism
underlying the eective rate sensitivity of concrete.
It is evident in Figure 9 that the experimental scatter is substantial, specially where load cases 1
and 6 are concerned. Thus, while the loading conditions corresponding to load case 1 are ostensibly
similar to those corresponding to load case 2 (cf. Table II), the strain rate and peak load reported
for the former are considerably lower that those reported for the latter. In particular, the dynamic
Copyright ? 2000 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2000; 48:963994
THREE-DIMENSIONAL FINITE-ELEMENT SIMULATION 979
Figure 11. Development of fracture patterns on the surface and in the interior of the specimen. Displace-
ments are magnied by a factor of 100 to aid visualization. The damage eld represents the fraction of
expended fracture energy to total fracture energy per unit surface.
strength reported for load case 1 is below the static strength, which constitutes a clear anomaly.
In addition, the strain rate reported for load case 6 is abnormally low in comparison to that for
load case 5, which corresponds to a o
i
roughly 50 per cent lower than that of load case 6, t
r
being
nearly identical in both cases.
These anomalies, and the overall scatter of the experimental data, attest to the diculties inherent
to the testing of highly heterogeneous materials such as concrete. For instance, nominally identical
loading conditions may result in sharply diering local conditions at the supports, leading to
noticeable dierences in the overall response of the specimens. These experimental uncertainties,
as well as the uncertainties in the computed solution alluded to above, must be borne in mind in
making meaningful comparisons between the numerical predictions and experimental data.
Copyright ? 2000 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2000; 48:963994
980 G. RUIZ, M. ORTIZ AND A. PANDOLFI
Figure 11. Continued.
3.5.2. Load and energy history curves. A comparison between computed and experimental load
histories for loading case 2 is shown in Figure 10(a). The two histories match nearly exactly during
the early stages of loading. The discrepancy in the peak loads, of the order of 17 per cent, and in
the slope of the ramp portion of the record are well within the experimental scatter. Indeed, local
eects near the bearing strips are notoriously dicult to control experimentally, which detracts
considerably from the reproducibility of the ramp slope. In view of the good match between the
experimental and computed strain rates for loading case 5, we may reasonably surmise that the
experimental load history is likewise better captured in that case. Regrettably, that history was not
reported in the original publications.
It is interesting to note that the time required for elastic pressure waves to travel through the
diameter of the cylinder is of the order of 14 s. In view of the smoothness of the rising part
of the transmitted load history, it would appear that the number of reverberations which take
Copyright ? 2000 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2000; 48:963994
THREE-DIMENSIONAL FINITE-ELEMENT SIMULATION 981
Figure 11. Continued.
place during the ramp-up time of the loading pulse is sucient to establish a rather uniform state
of deformation within the specimen. It follows as a corollary that the waviness of the softening
part of the load history curve is due to fracture processes rather than wave eects. Furthermore,
the absence of signicant wave eects justies the use of the test data for purposes of inferring
constitutive properties of the material.
The horizontal dash line inset in Figure 10(a) represents the maximum load attained during the
static Brazilian test and is shown for reference. As may be seen from the gure, the dynamic peak
load is of the order of 2.71 times the static maximum load, which attests to the importance of the
dynamic eects under the conditions of the test.
Figure 10(a) also depicts the consumption of cohesive energy and the kinetic energy as a
function of time for the same loading case. The various stages of cracking observed experimentally
Copyright ? 2000 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2000; 48:963994
982 G. RUIZ, M. ORTIZ AND A. PANDOLFI
are also shown in Figure 10(b) for comparison. As is evident from Figure 10(a), the kinetic
energy initially builds up to the initiation time of the main crack contained within the mid-
plane of the specimen. The kinetic energy subsequently increases slowly during the growth of the
main crack. The peak load is attained when the main crack extends clear through the specimen,
Figures 10(a) and 10(b), point 7. It therefore follows that the dynamic growth of the main crack is
stable in the sense of requiring an increasing applied load to proceed. By way of contrast, following
the peak load, a sharp upturn in the kinetic energy occurs while the applied load simultaneously
decreases. This phase of the energy-history record is accompanied by a marked increase in the
expenditure of cohesive energy (Figure 10(a)). These features are indicative of profuse and unstable
microcrackinghence the high cohesive energy expenditureresulting in widespread dynamic
fragmentationhence the loss of load-carrying capacity and increase in kinetic energy. Eventually,
the specimen loses all load-bearing capacity, the fragmentation process arrests and the expended
cohesive energy saturates asymptotically to a constant value (Figure 8(b)).
The horizontal dash line shown for reference in Figure 10(a) represents the fracture energy
expended in the formation of a single planar crack cutting through the mid-plane of the specimen.
As may be observed, the actual fracture energy expended is greatly in excess of that value, which
is indicative of a far more complex and intricate crack pattern. It is also interesting to note that
no energy is dissipated through plastic work at any time during the calculations. Hence, cohesive
fracture accounts for the totality of energy dissipation.
3.5.3. Crack pattern. The predicted sequence of crack patterns also follows closely the experi-
mental patterns observed by means of a high-speed camera (Figure 10(b)) [41]. Figure 11 depicts
a sequence of snapshots of the deformed specimens at intervals of 10 s, showing the distribution
of cracks. It should be carefully noted that, in this plot, displacements have been magnied by
a factor of 100 in order to aid visualization. It may also be recalled that the peak load occurs
roughly at 70 s, corresponding to snapshot (Figure 11(f)). Also shown in the gures are level
contours of damage, dened as the fraction of expended fracture energy to total fracture energy per
unit surface, or critical energy release rate. Thus, a damage density of zero denotes an uncracked
surface, whereas a damage density of one is indicative of a fully cracked or free surface. It bears
emphasis that this damage eld is dened on any internal surface of the body, i.e. it represents a
density per unit areaas opposed to a density per unit volume. In Figure 11 we have chosen to
represent the extent of damage on the mid-plane, or load plane, of the specimen.
Remarkably, both the experimental observations and the numerical solution clearly exhibit a
main crack on the mid-plane of the specimen which initiates near the centre of the cylinder and
subsequently propagates towards the bearing strips, eventually causing the specimen to split into
two main fragments (cf. Figure 10(b)). The observed initiation time is roughly 30 s, which is
in fair agreement with the results of the calculations (Figures 11(b) and 11(c)). Furthermore,
the simulation also captures some early localized cracking in the loading area (Figures 10(b)
and 11(b)). Evidently, the numerical simulation provides a wealth of information regarding the
three-dimensional character of the evolving fracture patterns which is inaccessible to experimental
diagnostics, which are for most part limited to the observation of the surface of the specimen. In
this regard it is interesting to note that, as expected, the initiation and growth of the main crack
is far from being uniform through the width of the specimen. Indeed, our simulations suggest that
lenticular cracks initiate from the surface of the specimen, i.e. the ends of the cylinder (Figures
11(c) and (d)) and subsequently propagate inward within the mid-plane with increasing load.
Eventually, the surface cracks coalesce and form a single through-crack (Figures 11(d) 11(f)).
Copyright ? 2000 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2000; 48:963994
THREE-DIMENSIONAL FINITE-ELEMENT SIMULATION 983
The development of the main crack on the mid-plane of the specimen is by no means a clean
process. Thus, the main crack is in constant competition with incipient cracks initiating on adjacent
parallel planes, e.g. Figure 11(d). However, these competing cracks are quickly shut out by the
unloading set forth by the dominant crack and their growth is stunted. A similar process of trial
cracks has been observed by Xu and Needleman to accompany the dynamic growth of cracks [17].
The softening regime is likewise far from monotonic and often exhibits multiple peaks, which is
indicative of a haltingrather than smoothcrack propagation. Thus, as noted earlier, the rst peak
load coincides with the formation of a well-developed through-crack. This event generates relief
waves which temporarily halt the splitting process. The specimen subsequently reloads, inducing
further crack growth and a new peak load is attained (Figures 10(a) and 12).
Figures 11(e)11(f) further reveal the development of profuse cracking at the impact point on the
surface of the specimen, in keeping with observation (Figure 10). As noted earlier, this generalized
microcracking, which often leads to fragmentation, occasions a sharp upturn in energy dissipation.
The ability to account for such complex fracture patterns with relative ease is a remarkable feature
of cohesive theories of fracture.
3.5.4. Crack velocity. Tedesco et al. [41] also reported crack speeds estimated from high-speed
camera pictures. These rough crack-speed estimates range from 10 to 50 per cent of the Rayleigh
wave speed, which equals 3620 m}s for the concrete used in the tests. An identical procedure
applied to loading case 5 returns a crack speed of 1320 m}s, or 36 per cent the Rayleigh wave
velocity, which is within the experimental range.
4. PARAMETRIC STUDY
The preceding comparisons with experiment may be regarded as validation of the cohesive theory
of fracture adopted in the calculations as a model of the dynamic behaviour of concrete. In
particular, the predictions of the theory have been found to be in good agreement with several
salient features of the experimental record, including the crack pattern and the transmitted load
histories measured by Tedesco et al. [41] in a split Hopkinson-bar Brazilian test conguration,
and the attendant eective rate sensitivity of the dynamic strength of concrete. Now that the
delity of the model has been established, we proceed to apply the model predictively with a
view to exploring the eect of variations in selected parameters of the model, including strain
rate, specimen size, the width of the bearing strips, the ratio [ of modes II to I toughness and
specimen geometry. This parametric study reveals useful insights into the interpretability of the
dynamic Brazilian test and the role of material parameters in shaping the dynamic behaviour of
concrete.
We dene a base case with respect to which we introduce systematic variations in parameters
of interest. The specimen geometry in the base case is as previously considered. However, in
order to speed up calculations, we conne the analysis to a thin slice of the cylinder of thickness
D}12, and constrain the normal displacement on the lateral surfaces of the slice. Thus, whereas
the calculations are carried out in three dimensions, the conditions of the analysis approximate a
plane-strain constraint such as may be expected to be in force at the centre of the cylinder. The
material properties remain as listed in Table I. The loading pulse at the end of the incident bar is
idealized as a linearly rising ramp in velocity of a duration of t
i
=50 js, followed by a constant
plateau at 2 m}s maintained up to a time t
d
=100 s (Figure 6). The boundary conditions remain
Copyright ? 2000 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2000; 48:963994
984 G. RUIZ, M. ORTIZ AND A. PANDOLFI
as described in Section 3.3. The mesh design conforms to the pattern described in Section 3.4.
However, owing to the restriction of the analysis to a single slice of the cylinder the computational
mesh is conveniently reduced to 1231 nodes and 630 tetrahedra only.
Finally, throughout this section results are presented in dimensionless form in order to facilitate
comparisons between dierent cases. To this end, we dened the normalized transmitted load P as
P

=
2P
WD[
ts
(21)
With this convention, P

gives the ratio [


td
}[
ts
at peak load, and P

=1 when the transmitted load


is equal to the theoretical maximum static load. We choose to normalize the cumulative cohesive
energy W
coh
by the fracture energy required to cleave the cylinder along its load plane into two
equal fragments with the result:
W

coh
=
W
coh
WDG
c
(22)
Finally, we dened the normalized time as
t

=
t
t
c

c
D
(23)
where t
c
and
c
are the characteristic time and length of the material, respectively (Equations (15)
and (14)). The factor
c
}D is introduced in (23) in order to bring into coincidence the initial
stages of the P

curves for specimens of dierent sizes.


4.1. Inuence of impact velocity
In order to gain additional insight into the role of impact velocity in the dynamic Brazilian test, we
proceed to apply incident pulses of amplitudes t =2, 4, 10, 20, 40 and 100 m}s. All the remaining
parameters of the model, including the ramp-up time t
r
and the duration time t
d
, are held constant as
in the previously dened base case. It should be carefully noted that the highest impact velocities
considered here are well beyond the range of impact velocities in Tedesco et al. experiments
(cf. Table II).
Figure 12 collects the computed load-history curves plotted in terms of normalized variables.
As may be seen from the gure, the rst load peak increases monotonically at low impact speeds,
or, alternatively, at low strain rates, and marks the maximum load transmitted by the specimen. In
particular, the rst load peak is consistently well above the second load peak, and the secondary
load peaks occur during the softening part of the load-history curve. By contrast, the rst load
peak increases less rapidly at high impact speeds and is of comparable magnitude toand in
some cases exceeded bythe second load peak. As a consequence, for high impact speeds the
ascending part of the load-history curve tends to exhibit multiple peaks. Our calculations show
that incident velocities t =10100 m}s cause considerable local crushing, in the form of extensive
microcracking and fragmentation, near the bearing strips which impart the loads to the specimen.
This observation suggests that the increased fracture and fragmentation near the bearing strips
accounts for the change in the peak load structure at high impact speeds.
The transition from monotonic load histories to load histories with multiple peaks prior to the
maximum load is also noteworthy. Indeed, our calculations reveal the existence of a threshold
impact velocity, or, equivalently, a threshold nominal strain rate, beyond which a suciently
Copyright ? 2000 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2000; 48:963994
THREE-DIMENSIONAL FINITE-ELEMENT SIMULATION 985
Figure 12. Normalized load-history curves for dierent prescribed velocities.
uniform state of deformation is not attained prior to the onset of cracking. The existence of this
threshold has been noted by Wu and Gorham [76]. The presence of multiple peaks in the rising
part of the load-history curve is a telltale sign of strong wave eects and eectively disqualies
the dynamic Brazilian test as a constitutive test at high impact speeds.
4.2. Size eect
Because of its large characteristic length, which in the material under study here equals 122 mm,
the fundamental postulates of linear-elastic fracture mechanics, most notably the requirement of
small-scale yielding, are often violated in practice. When the limiting geometrical dimensions are
comparable to the characteristic length of the material, the structure exhibits transitional behaviour
straddling strength of materials and conventional fracture mechanics. This transitional behaviour
is often summarily referred to a size eect. One pernicious consequence of this size eect is
that conventional fracture tests, which are designed to meet all the requirements of linear-elastic
fracture mechanics, require inordinately large specimens and prohibitively expensive experimental
facilities [77, 78].
In order to investigate the role of specimen size in the dynamic Brazilian test, we have performed
calculations for three cylinder diameters: D=50.8, 100 and 150 mm, or D=0.4
c
, 0.8
c
and
1.2
c
. The smaller diameter coincides with that which has been considered previously, whereas
the remaining diameters bear a ratio of 2 and 3 to the smaller diameter, respectively. In order
to facilitate comparisons, the width of the bearing strips and the prescribed impact velocity are
scaled in direct proportion to the diameter. The prescribed velocities are chosen as 2, 4 and 6 m}s.
In consequence of this scaling, the nominal strain rate and contact pressure are independent of the
specimen size. The minimum mesh size is kept constant in all cases so as to provide an equal
resolution of the cohesive length of the material. The resulting meshes are shown in the inset of
Copyright ? 2000 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2000; 48:963994
986 G. RUIZ, M. ORTIZ AND A. PANDOLFI
Figure 13. Normalized load-history curves for three dierent cylinder diameters, D=50.8, 100 and 150 mm.
Figure 13, which also shows the normalized load and cohesive-energy histories resulting from the
calculations.
By virtue of the time normalization, the initial rising part of the load history is ostensibly
identical in all cases. As may be seen from Figure 13, the peak load increases with decreasing
specimen size, which is consistent with the well-known size eect in fully yielded specimens, i.e.
smaller is stronger [48]. Our results suggest that the peak or failure load becomes insensitive to
the size of the specimen beyond a diameter of the order of one characteristic length, in agreement
with the ndings of Ba zant and Planas [48].
The cohesive energy consumption scales almost exactly with the size of the specimen, as may
be seen from Figure 13. The increase in specic fracture energy with specimen size is indicative
of a higher extent of localized and distributed microcracking, as has been pointed out for static
tests by Guinea [49], Guinea et al. [50] and Planas et al. [79].
The contribution of microcracking to the fracture energy may be estimated as follows. For small
specimens, the zone of distributed microcracking may be expected to engulf the entire specimen,
and the fracture energy consumed by microcracking may be estimated as
W

coh
G
c
W
D
2
4
1
!
(24)
The normalized fracture energy then follows as
W
coh
WDG
c
1 +

4
D
!
(25)
where ! is the fragment size, which may be taken to be commensurate with the aggregate size,
and term 1 in (25) accounts for the fracture energy required to split the specimen along the load
plane, which equals WDG
c
. For large specimens, the zone of distributed microcracking may be
expected to be concentrated around the bearing strips. Using Boussinesqs solution for a
Copyright ? 2000 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2000; 48:963994
THREE-DIMENSIONAL FINITE-ELEMENT SIMULATION 987
line load on an elastic half-space, the depth of the microcracking zone may be estimated as
d
P
max
2Wo
c
(26)
and the fracture energy then follows in the form
W
coh
WDG
c
1 +

4

P
max
2Wo
c

2
1
D!
(27)
The maximum applied load may in turn be estimated from (20) as
P
max
o
i
A=jctA (28)
where A is the cross-sectional area of the incident bar. The cross-over between the fracture energies
(25) and (27) occurs for a diameter
D
c

P
max
2Wo
c

jctA
2Wo
c
(29)
It follows from these estimates that, provided that the impact velocity t and the specimen width W
are scaled in proportion to the diameter D, the fracture energy decreases for DD
c
as D
1
, and
the eect of microcracking becomes increasingly localized. Under these conditions, the extended
fracture energy tends asymptotically to the nominal value WDG
c
for large D. For small specimens,
DD
c
, microcracking extends to the entire specimen and the fracture energy increases linearly
with D.
For the calculations shown in Figure 13 one has D
c
360 mm. Thus, the specimen diameters
may be expected to be below the transitional regime in which the eect of microcracking saturates.
This accounts for the variation in fracture energy between the three specimens, and the substantial
excess fracture energy over that which is strictly required to cleave the specimens (Figure 13). It
can hardly be overemphasized the relative ease with which the eects are accounted for within a
cohesive fracture framework.
4.3. Inuence of the width of the bearing strips
As already noted, local eects near the bearing strips can have a marked inuence on the overall
response of the specimen and are notoriously dicult to control experimentally. In order to gage
the sensitivity of the test results to these eects, we have carried out calculations for two bearing
strips of widths b =D}8 and D}12. The rst bearing-strip width is as in the previously dened
base case, whereas the second smaller bearing strip is reduced in size by a factor of
2
3
.
Figure 14 shows the resulting load-history curves for the two bearing-strip sizes. Also shown
inset is the slight local modication to the mesh required to accommodate the narrower strip. As
may be seen from the gure, the maximum load diminishes with the size of the bearing strip,
an eect reported by Rocco et al. [80] and by Rocco [52] for the static test. For the geometries
considered in the calculations, the reduction in peak load is of the order of 16 per cent. This
result strongly suggests that the early stages of cracking are governed by the stress P}(Wb) under
the bearing strips. It should be carefully noted that a direct application of the static formula (17)
yields two dierent values of the dynamic strength of the material, which cautions against the
indiscriminate use of such formulae. The cohesive energy consumption is nearly identical in both
cases (Figure 14).
Copyright ? 2000 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2000; 48:963994
988 G. RUIZ, M. ORTIZ AND A. PANDOLFI
Figure 14. Normalized load-history curves for two dierent bearing-strip widths b =D}8 and D}12.
4.4. Eect of [
As noted earlier, the parameter [ determines the ratio of modes II to I fracture toughness. Following
Reference [70], in calculations we have adopted the value [ =10 in order to account for the
toughening eect of the aggregate interlocking mechanism under mode II loading. It should be
carefully noted that for a large (small) value of [ the critical condition for the initiation of
a cohesive crack is ostensively independent of the magnitude of the shear (normal) tractions,
and the ensuing sliding (opening) displacements are constrained to remain close to zero at all
times. In the limit of [ ([ 0), the initiation criterion becomes completely independent
of the shear (normal) traction and the sliding (normal) displacements are constrained to vanish
identically.
In order to ascertain the sensitivity of the results to the choice of [, we have compared the
cases [ =0.1, 1, 10, and 100. Figure 15 shows the results of the simulations in the form of load-
history and fracture energy curves, and deformed meshes and damage patterns at peak load. It
is concluded from these gures that, as expected, the choice of [ has a marked inuence on
the predictions of the model. Thus, for the small [ =0.1, the specimen fails along slip lines and
transmits a small peak load (Figure 15). The deformed mesh and the damage pattern at the peak
load exhibit profuse distributed shear cracking. The cohesive energy consumption is very small up
to the peak load, for large sliding displacements in the cracks are necessary to reach the eective
critical opening displacement. The value [ =1 gives the same weight to normal and shear stresses
for crack initiation and opening, and thus represents a transition between predominantly shear
cracking and predominantly normal opening, e.g. due to the aggregate interlocking constraint.
Figure 15 shows that, for [ =1, the specimen undergoes considerable local crushing near the
incident bearing strip. In addition, a through-crack develops in which some degree of sliding is
evident. Both mechanisms dissipate cohesive energy before the maximum load is attained. The
extent of distributed cracking is smaller than in the case of [ =0.1. The results for [ =10 and
100 dier only slightly, which demonstrates that a choice of [ =10 approximates closely the case
of perfect aggregate interlocking. The load-history and the fracture energy curves are identical
in both cases up to the peak transmitted load, while the deformed meshes and damage patterns
Copyright ? 2000 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2000; 48:963994
THREE-DIMENSIONAL FINITE-ELEMENT SIMULATION 989
Figure 15. Normalized load-history curves and deformed meshes for [ =0.1, 1, 10 and 100, representing
dierent ratios of modes II to I toughness.
reveal the predominance of a single through-crack devoid of any appreciable sliding. For [ =10,
some degree of local crushing remains in the loading area which is almost entirely suppressed for
[ =100. These results demonstrate the ability of the cohesive model to reproduce dierent types
of mixed-mode fracture behaviours.
4.5. Inuence of the geometry of the specimen
Finally, we investigate the eect and shape of the specimen. To this end, we consider an additional
specimen of square cross-section and dimensions such that the cross-section through the load plane
is identical to that of the base specimen. The mesh design near the load plane is likewise identical
in both cases. The details of the specimen geometry and mesh design are shown in Figure 16.
The resulting load and fracture energy histories are also shown in the gure. It follows from these
results that the shape of the specimen has a modest eect on the peak load. The higher peak
load corresponds to the square cross-section and is within 10 per cent of the peak load for the
circular cross-section. The post-peak response of the square specimen is accompanied by increased
distributed cracking near the bearing strips, which results in a higher fracture energy consumption.
Copyright ? 2000 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2000; 48:963994
990 G. RUIZ, M. ORTIZ AND A. PANDOLFI
Figure 16. Normalized load-history curves for two specimen shapes of identical
cross-section through the load plane.
5. SUMMARY AND CONCLUSIONS
We have investigated the feasibility of using cohesive theories of fracture, in conjunction with the
direct simulation of fracture and fragmentation, in order to describe processes of tensile damage
and compressive crushing in concrete specimens subjected to dynamic loading. The particular
conguration contemplated in this study is the Brazilian cylinder test performed in a Hopkinson
bar [41, 37, 36], which furnishes a demanding validation test of the theory. Our approach accounts
explicitly for microcracking, the development of macroscopic cracks and inertia. The eective
dynamic behaviour of the material is predicted as an outcome of the calculations. In particular,
our simulations capture closely the experimentally observed rate sensitivity of the dynamic strength
of concrete, i.e. the nearly linear increase in dynamic strength with strain rate [41, 37, 36, 33].
More generally, our simulations give accurate transmitted loads over a range of strain rates, which
attests to the delity of the model where rate eects are concerned. The model also predicts key
features of the fracture pattern such as the primary lens-shaped cracks parallel to the load plane,
as well as the secondary profuse cracking near the support. These results validate the theory as
it bears on mixed-mode fracture and fragmentation processes in concrete over a range of strain
rates.
We have assumed that the cohesive properties of the material are rate independent and therefore
determined by static properties such as the static tensile strength. However, we have noted that
cohesive theories, in addition to building a characteristic length into the material description, endow
the material with an intrinsic time scale as well [23]. This intrinsic time scale accounts for the
ability of model to predict key aspects of the dynamic behaviour of concrete, such as the strain-
rate sensitivity of strength. Our results suggest, therefore, that most of the strain rate sensitivity
of concrete is attributable to the microinertia attendant to dynamic microcracking and fracture.
We have additionally carried out a parametric study with a view to ascertaining the sensitivity
of the dynamic Brazilian test to details of the experimental set-up such as the size and geometry
of the specimen and the size of the bearing strips. The results of the simulations also exhibit the
Copyright ? 2000 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2000; 48:963994
THREE-DIMENSIONAL FINITE-ELEMENT SIMULATION 991
expected size eect, i.e. smaller specimens are stronger [48]. Our results suggest that the peak or
failure load becomes insensitive to the size of the specimen beyond a diameter of the order of
one characteristic length, in agreement with the ndings of Ba zant and Planas [48]. We have also
found that the cohesive energy expenditure is considerably larger in the dynamic test than in the
static case, which reects the dissipation due to microcracking. An additional insight provided by
the simulations is that the load-history curve exhibits multiple peaks within its rising part at high
strain rates. Yet another nding is that the peak load is sensitively dependent on the size of the
bearing strips through which the loading is imparted to the specimen.
These observations, taken together, should caution experimentalists against a simplistic inter-
pretation of the dynamic Brazilian test, specially as regards the inference of intrinsic material
properties. Similar warnings have been voiced by Rodrguez et al. [45] and Johnstone and Ruiz
[46] in the context of SHPB tests on ceramics.
ACKNOWLEDGEMENTS
Gonzalo Ruiz gratefully acknowledges the nancial support for his stay at the California Institute of Technol-
ogy provided by the Direcci on General de Ense nanza Superior, Ministerio de Educaci on y Cultura, Spain.
Anna Pandol and Michael Ortiz are grateful for support from the Department of Energy through Caltechs
ASCI Center of Excellence for Simulating Dynamic Response of Materials. Michael Ortiz also wishes to
gratefully acknowledge the support of the Army Research Oce through grant DAAH04-96-1-0056. We are
indebted to Dr. Ra ul A. Radovitzky for his assistance in the development of the meshes used throughout this
research.
REFERENCES
1. Ravichandar K, Knauss WG. An experimental investigation into dynamic fracture, Part 1. Crack initiation and arrest.
International Journal of Fracture 1984; 25(4):247262.
2. Liu C, Knauss WG, Rosakis AJ. Loading rates and the dynamic initiation toughness in brittle solids. International
Journal of Fracture 2000; to appear.
3. Braides A, ChiadoPiat V. Integral representation results for functionals dened on sbv. SISSA ISAS 1994; 198(M):
131.
4. Fonseca I, Francfort GA. Relaxation in BV versus quasiconvexication in W
1,
; a model for the interaction between
fracture and damage. Calculus of Variations 1995; 3:407446.
5. Dugdale DS. Yielding of steel sheets containing slits. Journal of the Mechanics and Physics of Solids 1960;
8:100104.
6. Barrenblatt GI. The mathematical theory of equilibrium of cracks in brittle fracture. Advances in Applied Mechanics
1962; 7:55129.
7. Rice JR. Mathematical analysis in the mechanics of fracture. In Fracture, Liebowitz H (ed.). Academic Press:
New York, 1968; 191311.
8. Hillerborg A, Modeer M, Petersson PE, Needleman A. Analysis of crack formation and crack growth in concrete by
means of fracture mechanics and nite elements. Cement Concrete Research 1976; 6:773782.
9. Rose JH, Ferrante J, Smith JR. Universal binding energy curves for metals and bimetallic interfaces. Physical Review
Letters 1981; 47(9):675678.
10. Carpinteri A. Mechanical Damage and Crack Growth in Concrete. Martinus Nijho: Dordrecht, The Netherlands,
1986.
11. Needleman A. A continuum model for void nucleation by inclusion debonding. Journal of Applied Mechanics 1987;
54:525531.
12. Ortiz M. Microcrack coalescence and macroscopic crack growth initiation in brittle solids. International Journal of
Solids and Structures 1988; 24:231250.
13. Willam K. Simulation issues of distributed and localized failure computations. In Cracking and Damage, Mazars J,
Bazant ZP (eds). Elsevier Science: New York, 1989; 363378.
14. Needleman A. An analysis of decohesion along an imperfect interface. International Journal of Fracture 1990;
42:2140.
15. Needleman A. Micromechanical modeling of interfacial decohesion. Ultramicroscopy 1992; 40:203214.
Copyright ? 2000 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2000; 48:963994
992 G. RUIZ, M. ORTIZ AND A. PANDOLFI
16. Xu XP, Needleman A. Void nucleation by inclusion debonding in a crystal matrix. Modelling and Simulation in
Materials Science and Engineering 1993; 1:111132.
17. Xu XP, Needleman A. Numerical simulations of fast crack growth in brittle solids. Journal of the Mechanics and
Physics of Solids 1994; 42:13971434.
18. Tvergaard V, Hutchinson JW. The inuence of plasticity on mixed-mode interface toughness. Journal of the Mechanics
and Physics of Solids 1993; 41:11191135.
19. Ortiz M, Suresh S. Statistical properties of residual stresses and intergranular fracture in ceramic materials. Journal of
Applied Mechanics 1993; 60:7784.
20. Planas J, Elices M, Guinea GV. Book chapter: cohesive cracks as a solution of a class of nonlocal problems. In
Fracture and Damage in Quasibrittle Structures. Experiment, Modelling and Computer Analysis, Bazant ZP (ed.).
E & FN SPON, 1994.
21. Xu XP, Needleman A. Numerical simulations of dynamic interfacial crack growth allowing for crack growth away
from the bond line. International Journal of Fracture 1995; 74:253275.
22. Ortiz M. Computational micromechanics. Computational Mechanics 1996; 18:321338.
23. Camacho GT, Ortiz M. Computational modelling of impact damage in brittle materials. International Journal of Solids
and Structures 1996; 33(2022):28992938.
24. Tvergaard V, Hutchinson JW. Eect of strain dependent cohesive zone model on predictions of interface crack growth.
Journal de Physique IV 1996; 6:165172.
25. Tvergaard V, Hutchinson JW. Eect of strain dependent cohesive zone model on predictions of crack growth resistance.
International Journal of Solids and Structures 1996; 33:32973308.
26. Xu XP, Needleman A. Numerical simulations of dynamic crack growth along an interface. International Journal of
Fracture 1996; 74:289324.
27. De-Andr es A, P erez JL, Ortiz M. Elastoplastic nite element analysis of three-dimensional fatigue crack growth in
aluminum shafts subjected to axial loading. International Journal of Solids and Structures 1999; 36:22312258.
28. Ortiz M, Pandol A. Finite-deformation irreversible cohesive elements for three-dimensional crack-propagation analysis.
International Journal for Numerical Methods in Engineering 1999; 44:12671282.
29. Pandol A, Ortiz M. Solid modeling aspects of three-dimensional fragmentation. Engineering with Computers 1998;
14:287308.
30. Pandol A, Krysl P, Ortiz M. Finite element simulation of ring expansion and fragmentation the capturing of length
and time scales through cohesive models of fracture. International Journal of Fracture 1999; 95:279297.
31. Pandol A, Guduru PR, Ortiz M, Rosakis AJ. Finite element analysis of experiments of dynamic fracture ini c.300
steel. International Journal of Solids and Structures 2000; to appear.
32. Repetto EA, Radovitzky R, Ortiz M. Finite element simulation of dynamic fracture and fragmentation of glass rods.
Computer Methods in Applied Mechanics and Engineering 2000; to appear.
33. Yon JH, Hawkins NM, Kobayashi AS. Strain-rate sensitivity of concrete mechanical properties. ACI Materials Journal
1992; 89(2):146153.
34. Yu C-T, Kobayashi AS, Hawkins NM. Energy-dissipation mechanisms associated with rapid fracture of concrete.
Experimental Mechanics 1993; 33(3):205211.
35. Du J, Yon JH, Hawkins NM, Arakawa K, Kobayashi AS. Fracture process zone for concrete for dynamic loading.
ACI Materials Journal 1992; 89(3):252258.
36. Ross A, Tedesco JW, Kuennen ST. Eects of strain rate on concrete strength. ACI Materials Journal 1995;
92:3747.
37. Hughes ML, Tedesco JW, Ross A. Numerical analysis of high strain rate splitting-tensile tests. Computers and
Structures 1993; 47:653671.
38. Guo ZK, Kobayashi AS, Hawkins NM. Dynamic mixed mode fracture of concrete. International Journal of Solids
and Structures 1995; 32(17}18):25912607.
39. Reinhardt HW, Weerheijm J. Tensile fracture of concrete at high loading rates taking account of inertia and crack
velocity eects. International Journal of Fracture 1991; 51:3142.
40. Tedesco JW, Ross CA, McGill PB ONeil BP. Numerical Analysis of high strain rate concrete direct tension tests.
Computers and Structures 1991; 40(2):313327.
41. Tedesco JW, Ross CA, Kuennen ST. Experimental and numerical-analysis of high-strain rate splitting tensile tests. ACI
Materials Journal 1993; 90(2):162169.
42. Yon JH, Hawkins NM, Kobayashi AS. Numerical simulation of mode I dynamic fracture concrete. Journal of
Engineering Mechanics ASCE 1991; 117(7):15951610.
43. Yon JH, Hawkins NM, Kobayashi AS. Fracture process zone in dynamically loaded crack-line wedge-loaded, double-
cantilever beam concrete specimens. ACI Materials Journal 1991; 88(5):470479.
44. van Doormaal JCAM, Weerheijm J, Sluys LJ. Experimental and numerical determination of the dynamic fracture energy
of concrete. Journal de Physique IV 1994; C8:501506.
45. Rodrguez J, Navarro C, S anchez-G alvez V. Splitting tests: an alternative to determine the dynamic tensile strength of
ceramic materials. Journal de Physique IV 1997; 4(C8):101106.
46. Johnstone C, Ruiz C. Dynamic testing of ceramics under tensile stress. International Journal of Solids and Structures
1995; 32(17}18):26472656.
Copyright ? 2000 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2000; 48:963994
THREE-DIMENSIONAL FINITE-ELEMENT SIMULATION 993
47. G alvez F, Rodrguez J, S anchez V. Tensile measurements of ceramic materials at high rates of strain. Journal de
Physique IV 1997; C3:151156.
48. Ba zant ZP, Planas J. Fracture and Size Eect in Concrete and Other Quasibrittle Materials. CRC Press:
Boca Raton, FL, 1998.
49. Guinea GV. Medida de la Energa de Fractura del Hormig on. Ph.D Thesis, Departamento de Ciencia de Materials,
Universidad Polit ecnica de Madrid, ETS de Ingenieros de Caminos, Ciudad Universitaria, 28040 Madrid, Spain, 1990.
(Measurement of the fracture energy of concrete in Spanish.)
50. Guinea GV, Planas J, Elices M. Measurement of the fracture energy using three-point bend tests: 1. Inuence of
experimental procedures. Materials and Structures 1992; 25:212218.
51. Field JE, Walley SM, Bourne NK, Huntley JM. Experimental methods at high rates of strain. Journal de Physique IV
1994; C8:322.
52. Rocco CG. Inuencia del Tama no y Mecanismos de Rotura en el Ensayo de Compresi on diametral. Ph.D. Thesis,
Departamento de Ciencia de Materiales, Universidad Polit ecnica de Madrid, ETS de Ingenieros de Caminos, Ciudad
Universitaria, 28040 Madrid, Spain, 1996. (Size-dependence and fracture mechanisms in the diagonal compression
splitting test, in Spanish.)
53. Marsden JE, Hughes TJR. Mathematical foundations of elasticity. Prentice-Hall: Englewood Clis, NJ, 1983.
54. Lubliner J. On the thermodynamic foundations of non-linear solid mechanics. International Journal of Non-Linear
Mechanics 1972; 7:237254.
55. Lubliner J. On the structure of the rate equations of materials with internal variables. Acta Mechanica 1973;
17:109119.
56. Chen G, Ravichandran WN. Dynamic compressive behavior of ceramics under lateral connement. Journal de Physique
IV 1994; 4:177182.
57. Chen G, Ravichandran WN. Static and dynamic compressive behavior of aluminum nitride under moderate connement.
Journal of the American Ceramic Society 1996; 79:579584.
58. Petersson PE. Crack growth and development of fracture zones in plain concrete and similar materials. Technical Report
TVBM-1006, Division of Building Materials, Lund Institute of Technology, University of Lund, Sweden, 1981.
59. Belytschko T. An overview of semidiscretization and time integration procedures. In Computational Methods for
Transient Analysis, Belytschko T, Hughes TJR (eds). North-Holland: Amsterdam, 1983; 165.
60. Hughes TJR. Analysis of transient algorithms with particular reference to stability behavior. In Computational Methods
for Transient Analysis, Belytschko T, Hughes TJR (eds). North-Holland: Amsterdam, 1983; 67155.
61. Hughes TJR. The Finite Element Method: Linear Static and Dynamic Finite Element Analysis. Prentice-Hall:
Englewood Clis, NJ, 1987.
62. Mathur KK, Needleman A, Tvergaard V. Three dimensional analysis of dynamic ductile crack growth in a thin plate.
Journal of the Mechanics and Physics of Solids 1996; 44:439464.
63. Carneiro FLL, Barcellos A. Tensile strength of concrete. RILEM Bulletin 1953; 13:97123.
64. CEB-FIP Model Code 1990, Final Draft. Bulletin DInformation N. 203, 204 and 205, EEP Lausanne.
65. ASTM C 39. Annual Book of ASTM Standards. vol. 04.02. Chapter Standard test method for compressive strength
of cylindrical concrete specimens. ASTM: Philadelphia, 1991; 2024.
66. Cuiti no AM, Ortiz M. A material-independent method for extending stress update algorithms from small-strain plasticity
to nite plasticity with multiplicative kinematics. Engineering Computations 1992; 9:437451.
67. Jia Z, Castro-Montero A, Shah SP. Observation of mixed mode fracture with center notched disk specimens. Concrete
and Cement Research 1996; 26(1):125137.
68. Swartz SE, Taha NM. Crack-propagation and fracture of plain concrete beams subjected to shear and compression.
ACI Structural Journal 1991; 88(2):169177.
69. Subrumaniam KV, Popovics JS, Shah SP. Testing concrete in torsion: Instability analysis and experiments. Journal of
Engineering Mechanics ASCE 1998; 124(11):12581268.
70. Gustafsson PJ, Hillerborg A. Sensitivity in the shear strength of longitudinally reinforced beams to fracture energy of
concrete. ACI Structural Journal 1988; 85(3):286294.
71. Carpinteri A, Valente S, Ferrara G, Melchiorri G. Is mode II fracture energy a real material property? Computers and
Structures 1993; 48:397413.
72. G alvez JC, Cend on D, Planas J, Guinea GV, Elices M. Fracture of concrete under mixed loading. Experimental
results and numerical prediction. Fracture Mechanics of Concrete Structures, Proceedings FRAMCOS-3, D-79104,
AEDIFICATIO Publishers, Freiburg, Germany, 1998.
73. Planas J, Elices M. Nonlinear fracture of cohesive materials. International Journal of Fracture 1991; 3:139157.
74. Nicholas T, Recht RF. Introduction to impact phenomena. In High Velocity Impact Dynamics, Zukas JA (ed.). Wiley:
New York, 1990; 163.
75. Radovitzky R, Ortiz M. Tetrahedral mesh generation based on node insertion in crystal lattice arrangements and
advancing-front-Delaunay triangulation. Computer Methods in Applied Mechanics and Engineering 2000; in press.
76. Wu XJ, Gorhan DA. Stress equilibrium in the split Hopkinson pressure bar test. Journal de Physique IV 1997;
7(C3):9196.
77. Albertini C, Cadoni E, Labibes K. Impact fracture process and mechanical properties of plain concrete by means of
an Hopkinson bar bundle. Journal de Physique IV 1997; C3:915920.
Copyright ? 2000 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2000; 48:963994
994 G. RUIZ, M. ORTIZ AND A. PANDOLFI
78. Fandrich RG, Clout JMF, Bourgeois FS. The CSIRO Hopkinson bar facility for large diameter particle breakage.
Minerals Engineering 1998; 11(9):861869.
79. Planas J, Elices M, Guinea GV. Measurement of the fracture energy using three-point bend tests: 2. Inuence of bulk
energy dissipation. Materials and Structures 1992; 25:305312.
80. Rocco CG, Guinea GV, Planas J, Elices M. The eect of the boundary conditions on the cylinder splitting strength.
In Fracture Mechanics of Concrete Structures, Wittmann FH (ed.). Aedicatio Publishers: Freiburg, Germany, 1995;
7584.
Copyright ? 2000 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2000; 48:963994

S-ar putea să vă placă și