You are on page 1of 38

Preprint submitted to Nuclear Engineering and Design

Formal publication: https://doi.org/10.1016/j.nucengdes.2019.110480

Framework for Dynamic Analysis of Radioactive Material Transport


Packages under Accident Drop Conditions
Xian-Xing (Lambert) Lia,*, Catherine Wangb, Jim Satob

________________________________________________________
aEngineering Mechanics Department, Ontario Power Generation Inc., 889 Brock Rd,
Pickering, Ontario, Canada L1W 3J2
bNuclear Waste Design Engineering, Ontario Power Generation Inc., 1340 Pickering

Pkwy, Pickering, Ontario, Canada L1V 0C4

Email addresses: lambert.li@opg.com (Xian-Xing (Lambert) Li),


catherine.wang@opg.com (Catherine Wang), jim.sato@opg.com (Jim Sato)

*Corresponding author
________________________________________________________

ABSTRACT

Radioactive material transport packages are designed to withstand a 9-meter accident


drop onto an essentially rigid surface in accordance with the IAEA regulations SSR-6.
This study presents a framework for dynamic analysis of the transport packages under
accident drop and impact conditions. To evaluate structural behavior and performance
of the transport packages with large deformation occurring in a very short duration,
material models incorporated in the dynamic analysis framework are established with
four key aspects: 1) true stress and strain are used to describe characteristics of large
deformation of materials; 2) effect of strain rate is included to reveal dynamic response
of materials to high-energy impact loads; 3) material constitutive models are established
to capture realistic mechanical behavior of materials under highly triaxial stress states;
and 4) damage and fracture criteria of materials are incorporated to limit the plastic
deformation to material fracture and to safeguard the containment boundary of the dry
storage container (DSC) that contains radioactive materials. Critical material and model
parameters and properties associated with the material constitutive models, true stress-
strain relationships, strain rate effect, and damage and fracture criteria are appropriately
determined based on experimental results or theoretically derived based on overserved
material behavior in elementary stress states. An example of the dynamic analysis of a
transport package is presented to predict the maximum true plastic strain, strain rate,
and damage condition of the materials. The proposed analysis framework is capable of
addressing accident drop and impact conditions of radioactive material transport
packages and DSCs to meet the nuclear licensing requirements.

1
Keywords: Transport package, True stress, True strain, Triaxiality, Strain rate,
Constitutive model

________________________________________________________

1. Introduction

A licensed transport package carrying radiative materials is designed to withstand a 9-


meter drop onto an essentially rigid surface in the most damaging orientations in a
hypothetical transportation accident scenario in accordance with Paragraph 727(a) of
the IAEA Safety Regulations SSR-6 (2012 edition). A transport package in general
consists of an outer packaging using two impact limiters constructed from steel shells
encasing high-density crushable foam and an inner packaging using a dry storage
container (DSC) with a cask and lid comprised of inner and outer steel liners that
surround reinforced heavyweight concrete. The inner chamber of the DSC cask is used
to house steel modules that hold radioactive material bundles. The inner chamber of the
cask is sealed with the lid using welded or bolted connections between the flanges of
the lid and cask. The impact limiters are designed to absorb the majority of the kinetic
energy of the falling transport package through large plastic deformations, which limit
the deceleration gravity loads on the DSC to within the acceptance criteria for the DSC
to act as the radioactive material containment boundary.
For dynamic analysis concerning impact kinetic energy absorption, four key aspects
must have for material models: 1) true stress and strain to quantify the mechanical
behavior of materials under large deformation; 2) effect of strain rate on dynamic
response to high-energy impacts; 3) material constitutive models and plastic
deformation evolution rules under highly triaxial stress states; and 4) stress triaxiality-
dependent damage and fracture criteria of materials are incorporated to limit the plastic
deformation material fracture and to safeguard the containment boundary of the DSC
that contains radioactive materials.
For reasonable material constitutive models, they should have four unique
characteristics: 1) Constitutive relations should not be affected by arbitrary coordinate
transformations. Material constitutive models are established using principal stresses
and/or stress tensor invariants. 2) Material constitutive models are also established
based on phenomenological approach. Yield surface functions can be reduced to
simpler yield conditions with elementary stress states such as uniaxial compression and
tension, pure shear, equibiaxial compression and tension, etc., so that the material
model parameters can be properly calibrated based on the observed characteristics
from elementary experiments. 3) Material hardening rules are defined with the
dependence of a yield surface function on internal hardening variables. The hardening
behavior in triaxial stress state can be correlated to the elementary hardening/softening
behavior under the uniaxial stress state. 4) Given the material hardening rules, plastic

2
flow rules are established to relate the plastic defamation increments to stresses and
perhaps other quantities via the plastic potential function. Material parameters in the
plastic potential function should be calibrated based on the overserved material
behavior in elementary stress states.
In conventional designs under normal design loads, design inputs to a structural
system are generally external loads or forces. The output of a structural analysis is
generally the stresses that are checked against permissible stress criteria. Since the
plastic deformation is small up to the permissible stress, the engineering stress and
strain are generally used because the difference between the engineering and true
stresses or strains is negligibly small.
In large structural impact scenarios, kinetic energy is normally the input to a structural
system. Large deformation in some parts of the system is required to transfer the
kinetic energy into high strain energy. The acceptance criteria for a containment
boundary component such as a DSC liner is the permissible fracture strain to prevent
leakage of radioactive materials. In this case, true stress and strain should be used in
the dynamic analysis to quantify large deformation. For steel materials, the true stress is
increasingly higher than the engineering stress with increasing strain and the true
fracture strain is larger than the engineering fracture strain.
Static characteristics of materials are much different from the dynamic characteristics,
especially when materials are subjected to short duration impact with high strain rates.
Based on the investigation of a transport package under the 9-meter drop scenario
presented in this paper, the maximum strain rate of the impact limiter steel shell can
reach approximately 300/s, and that of the impact limiter polyurethane foam 500/s.
Consequently, the yield strengths of the impact limiter steel and foam can increase by
more than 50%. This increase is considered as very significant.
There is no steadfast rule to judge whether or not using engineering stress-strain
relationship and ignoring strain rate effect could lead to conservative result. On the one
hand, ignoring strain rate effect reserves conservatism of the material strengths. On the
other hand, ignoring the strain rate effect softens and weakens the actual mechanical
behavior of the materials that are directly in contact with targets, which overestimate of
the deformation and strain energy of the energy absorbent materials such as impact
limiters and target objects, and consequently underestimate the impact load transferred
to the more critical DSC. Reduction of strain energy in some parts of a structural
assembly increases the strain energy in the remaining parts due to the conservation of
energy. Using engineering stress-strain relationship and ignoring strain rate effect can
significantly alter the distribution of the strain energy and, in turn, result in unrealistic
strain and stress of the materials. To reveal the realistic effect of accident drop and
impact scenarios on the safety of the DSC as a containment boundary, the true stress
and strain with realistic strain rate effect should be incorporated in the dynamic analysis.

3
Lo Frano et al (2014) performed a numerical analysis of a type IP-2 package under a
9-meter flat end drop scenario. The mechanical behavior of the packaging (steel shells
encasing concrete) was assumed to be elastic-perfectly plastic, and that of the steel
materials of the container a piece wise linear stress-strain curve without description of
the data used. No details of strain rate effect and fracture criteria were provided. Qiao et
al (2012) presented a finite element model for evaluation of the performance of
transport and storage casks. The analysis model was validated only by comparing the
computed local decelerations with the recorded deceleration from a test. However, no
details of stress-strain relationships, strain rate effect, material constitutive models, and
fracture criteria were provided. Sprung et al (2000) developed an in-house finite element
program PRONTO-3D for nonlinear transient analysis of generic spent fuel casks.
Idealized Ramberg-Osgood stress-strain relationship and bi-linear elastic-plastic
material model are used to represent mechanical behavior of steels and energy
absorbent materials. No details of strain rate effect and triaxial constitutive models were
provided. The EPRI reports 1009929 and 1011817 (EPRI, 2005a and 2005b)
incorporated true stress and strain. Fracture criteria were provided, but not associated
with stress triaxiality. No detail of strain rate effect was provided.
In general, materials used in a radioactive material container and transport package
include materials of the impact limiter steel shell, crushable foam, impact limiter
connection components, DSC steel liners/flanges, weld and/or bolt metals, concrete,
reinforcing steel, fuel modules, and fuel bundles. For these materials, establishments of
material constitutive models, uniaxial true stress-strain relationships, effect of strain
rate, and fracture and damage criteria incorporated in the dynamic analysis framework
are discussed, respectively, in the following sections.
In this paper, stress and modulus of elasticity are in the metric unit, MPa, and density
in kg/m3. Compressive stresses assume negative values, unless noted otherwise.
Stress and strain refers to true stress and strain, respectively, unless indicated
otherwise.

2. True stress and true strain


An engineering stress-strain relationship from a standard uniaxial test is established
using the initial cross-sectional area and gauge length of the specimen in general. Such
determined stress and strain are termed as engineering stress and strain. When the
specimen is subjected to large deformation, the cross-sectional area and gauge length
change significantly. In this case, the stress and strain should be determined using
current cross-sectional area and current reference length, which are termed as true
stress and strain. In a high-velocity impact scenario involving large deformation, true
stress-strain relationships are required. The engineering stress and strain can be
converted into the true stress and strain using the following equations:
𝑠
𝜎 = (1−𝜐(𝑒)𝑒)2 and 𝜀 = 𝑙𝑛(1 + 𝑒) (1)

4
where 𝜎 and 𝜀 are the true stress and strain, respectively; 𝑠 and 𝑒 are the engineering
stress and strain, respectively; and 𝜐(𝑒) is the Poisson’s ratio which is a function of
strain. Eqs. (1) are valid for brittle materials under tension and compression, ductile
materials under tension up to the peak engineering stress, and ductile materials under
compression.
For a uniaxial tensile test of a specimen made of ductile materials, initiation of cross-
section necking occurs at the peak tensile load or peak engineering stress. Until the
neck forms, the deformation is essentially uniform throughout the gauge length of the
specimen. After necking, all subsequent deformation takes place in the necking region,
and the engineering strain varies with the change of the gauge length of the specimen.
In the plastic-flow regime following yield, ductile materials flow with negligible change in
volume. The true strain after necking is then measured as
𝐴
𝜀 = 𝑙𝑛 ( 𝐴0 ) (2)

where 𝐴0 and 𝐴 are the initial and current minimum cross-sectional area in the necking
region.
After necking occurs, the specimen necking increases with increasing extension. The
state of stress in the necking region always becomes increasingly non-uniaxial. The true
stress is lower than the average value over the minimum cross-sectional area due to the
triaxial stress state in the necking region. The true stress can be calculated based on
the geometry of the necking as discussed by Yan et al (2018). Alternatively, an
empirical form is proposed based on the work by Chen (2010) to correct the average
stress after the initiation of necking for the true stress:

𝐹 𝐴 1+𝑒
𝑢
𝜎 = 𝛾𝑛 (𝐴) (1 + √𝐴 ) = 𝑠 ∙ 𝛾𝑛 exp(𝜀) (1 + √exp(𝜀) ) (3)
𝑝

where 𝐹 = applied tension; 𝐴𝑝 = 𝐴0 /(1 + 𝑒𝑢 ) is the cross-sectional area prior to the


onset of necking; 𝑒𝑢 the engineering strain at the peak engineering stress; and 𝛾𝑛 =
0.50 for mild steels (Chen, 2010) or 𝛾𝑛 = 0.57 for stainless steels to fit the experimental
result by Blandford et al (2007).
The plastic strain rate is related to the total strain rate through

𝐸𝑇
𝜀̇𝑝𝑙 = (1 − ) 𝜀̇ (4)
𝐸

in which 𝐸 is the static modulus of elasticity and 𝐸𝑇 the static tangent modulus. This
equation is used to convert experimental data that uses the total strain rate into analysis
inputs in terms of the plastic strain rate as required in analysis software.

3. Constitutive model for rigid polyurethane foams


3.1 Yield surface function and plastic flow potential of polyurethane foams
High-density rigid polyurethane foams have found use in transportation of radioactive
materials due to their excellent impact energy-absorbing capability. One of the greatest
attractions of the high-density polyurethane foams is the nearly isotropic crush
properties in the directions of perpendicular and parallel to the foam rise.

5
The most promising crushable foam material model is the model with volumetric
hardening which assumes that the evolution of the yield surface is controlled by the
volumetric compacting. The yield surface function adopted in ABAQUS (2017) is
defined in the meridional plane as

𝑝𝑙 𝑝𝑙 2
𝐹(𝜎𝑖𝑗 ) = √𝑞(𝜎𝑖𝑗 )2 + 𝛼 2 (𝑝(𝜎𝑖𝑗 ) − 𝑝0 (𝜀𝑣𝑜𝑙 )) − 𝐵(𝜀𝑣𝑜𝑙 )=0 (5)

where
1
𝑝(𝜎𝑖𝑗 ) = − 3 𝑡𝑟𝑎𝑐𝑒(𝜎𝑖𝑗 ) , the hydrostatic pressure being the function of the stress tensor 𝜎𝑖𝑗 ;
3
𝑞(𝜎𝑖𝑗 ) = √2 𝑆𝑖𝑗 𝑆𝑖𝑗 , the Mises stress being the function of the deviatoric stress tensor 𝑆𝑖𝑗 =
𝜎𝑖𝑗 + 𝛿𝑖𝑗 𝑝(𝜎𝑖𝑗 );
𝑝𝑙
𝑝𝑙 𝑝𝑐 (𝜀𝑣𝑜𝑙 )−𝑝𝑡
𝑝0 (𝜀𝑣𝑜𝑙 ) = , the centre of the yield surface on the p-axis;
2
𝑝𝑙
𝑝𝑐 (𝜖𝑣𝑜𝑙 )
is the evolved hydrostatic yield stress under compaction as a function of the
𝑝𝑙
volumetric compacting plastic strain 𝜀𝑣𝑜𝑙 ;
𝑝𝑡 is the yield stress in hydrostatic tension;
𝑝𝑙
𝑝𝑙 𝑝𝑐 (𝜀𝑣𝑜𝑙 )+𝑝𝑡
𝐵(𝜀𝑣𝑜𝑙 )=𝛼 , a yield surface evolution parameter with the hardening variable
2
being the volumetric plastic strain;
3𝑘𝑐
𝛼= )(3−𝑘 )
, a material constant that defines the shape of the yield surface;
√(3𝑘𝑡 +𝑘𝑐 𝑐
𝑘𝑐 = 𝜎𝑐0 /𝑝𝑐0 and 𝑘𝑡 = 𝜎𝑡0 /𝑝𝑐0;
𝜎𝑐0 is the initial uniaxial compressive strength;
𝜎𝑡0 the initial uniaxial tensile strength; and
𝜎𝑝0 the initial hydrostatic compressive strength.
The plastic flow is non-associative with the plastic flow potential being defined as

9
𝐺(𝜎𝑖𝑗 ) = √𝑞(𝜎𝑖𝑗 )2 + 2 𝑝(𝜎𝑖𝑗 )2 (6)

In the process of plastic deformation, the tensile strength, 𝑝𝑡 , is assumed to be


𝑝𝑙
constant. By contrast, the compressive strength, 𝑝𝑐 (𝜀𝑣𝑜𝑙 ) , evolves as a result of
3𝑘𝑐
compaction of the material. Based on 𝛼 = , the hardening rule of the
√(3𝑘𝑡 +𝑘𝑐 )(3−𝑘𝑐 )
compressive strength with volumetric hardening is
𝑝𝑙 𝑝𝑙 1 1 𝑝
𝑝𝑙 𝜎𝑐 (𝜀𝑣𝑜𝑙 )[𝜎𝑐 (𝜀𝑣𝑜𝑙 )( 2 + )+ 𝑡 ]
𝛼 9 3
𝑝𝑐 (𝜀𝑣𝑜𝑙 )= 𝑝𝑙 (7)
𝜎𝑐 (𝜀 )
𝑣𝑜𝑙
𝑝𝑡 +
3

along with the fact that


𝑝𝑙 𝑝𝑙
𝜀𝑣𝑜𝑙 = (1 − 2𝜐𝑓𝑐 )𝜀𝑎𝑥𝑖𝑎𝑙 (8)

6
in the uniaxial compression, where 𝜐𝑓𝑐 is the plastic Poisson’s ratio of crushable foams.
It is noted that 𝜐𝑓𝑐 is conventionally assumed to be zero for crushable foams with the
volumetric hardening model. With a zero plastic Poisson’s ratio, Eq. (8) reduces to the
𝑝𝑙 𝑝𝑙
equation 𝜀𝑣𝑜𝑙 = 𝜀𝑎𝑥𝑖𝑎𝑙 , which may have been used in the built-in constitutive models in
analysis software. To use this built-in model for foams of non-zero plastic Poisson’s
ratio, the hardening rule should be specified as a compressive stress - volumetric plastic
𝑝𝑙
strain relationship, 𝜎𝑐 (𝜀𝑣𝑜𝑙 ) , which can be obtained from modifying the uniaxial
𝑝𝑙
relationship 𝜎𝑐 (𝜀𝑎𝑥𝑖𝑎𝑙 ) by scaling the axial plastic strain using Eq. (8).

3.2 Calibration of material parameters in the yield surface function


To determine the parameter, 𝛼, of the yield surface function, the initial hydrostatic
compressive strength, 𝑝𝑐0 , needs to be given so as to calculate the constants 𝑘𝑐 and
𝑘𝑡 . However, this strength is not available in general unless specifically tested as such.
Instead, the initial shear strength, 𝜏0 , may be provided by foam vendors. If the initial
uniaxial compressive strength, 𝜎𝑐0 , initial uniaxial tensile strength, 𝜎𝑡0 , and initial shear
strength, 𝜏0 , are given, the yield surface function of Eq. (5) can be reduced to three
corresponding simpler equations for the three individual elementary cases. Solving
these three equations, the initial hydrostatic strength, 𝑝𝑐0 , can be obtained as follow:

𝜏0 2 [𝜎𝑡0 2 +𝜎𝑐0 2 +6(𝜎𝑡0 −𝜎𝑐0 )𝑝0𝑖 ]


𝑝𝑐0 = 𝑝0𝑖 + √𝑝0𝑖 2 + (9)
3(6𝜏0 2 −𝜎𝑡0 2 −𝜎𝑐0 2 )

where

1 (3𝜏0 2 −𝜎𝑡0 2 )𝜎𝑐0 2 −(3𝜏0 2 −𝜎𝑐0 2 )𝜎𝑡0 2


𝑝0𝑖 = (6) (3𝜏0 2 −𝜎𝑡0 2 )𝜎𝑐0 +(3𝜏0 2 −𝜎𝑐0 2 )𝜎𝑡0
(10)

𝑝𝑙
is the initial value of 𝑝0 (𝜀𝑣𝑜𝑙 ), i.e., the initial center of the yield surface. It is noted that
the uniaxial compressive strength, 𝜎𝑐0 , is a positive value.

3.3 Uniaxial compressive stress-strain relationship of polyurethane foams


As per Liu and Subhash (2004), the engineering stress-strain curves fitting
experimental data of specimens in uniaxial compression can be expressed with a
simple general form:

exp(𝛿𝑒)−1
𝑠 = 𝑎 exp(𝛽𝑒)+𝑏 + exp(𝑐)(𝑒𝑥𝑝(𝛾𝑒) − 1) (11)

where 𝑠 and 𝑒 are uniaxial engineering compressive stress and strain, respectively; a, b,
c, δ, β and γ are model parameters related to foam density. If the uniaxial compressive
stress-strain relationship is established using engineering strain as the common
representation of the laboratory test results, it should be converted into true stress-strain
relationship using Eqs. (1) for large deformation analysis.

7
3.4 Compressive Poisson’s ratio of polyurethane foams
To convert an engineering stress-strain relationship to a true stress-strain relationship,
the Poisson’s ratio used in Eqs. (1) should be given. For low-density foams, Kraus et al
(2013) proposed a general form of the Poisson’s ratio as
𝜈𝑓0
𝑓𝑜𝑟 𝑒<0
1+exp(𝛽𝑣 (𝜀𝑜𝑐 −𝑒))
𝜈𝑘 (𝑒) = { 𝜈 (12)
𝜈𝑓0 + 1+exp(𝛼𝑓𝑒(𝜀 −𝑒)) 𝑓𝑜𝑟 𝑒≥0
𝑣 𝑜𝑡

where 𝑒 is the engineering strain. Under large uniaxial compressive strain, the Poisson’s
ratio approaches to zero. Kraus et al (2013) proposed that 𝜈𝑓0 = 0.3, 𝜈𝑓𝑒 = 0.2, 𝛼𝑣 = 20,
𝛽𝑣 = 14, 𝜀𝑜𝑡 = 0.25, and 𝜀𝑜𝑐 = −0.28 to fit the experimental results of the low-density
conventional foams by Choi and Lakes (1992).
For high-density foams, the Poisson’s ratio may not approach to zero under high
compressive strain. For example, the experimental result of the Poisson’s ratio of the
LAST-A-FOAM® FR-3718 provided by General Plastics Manufacturing Company is
shown in Fig. 2. The density of the foam is 𝜌 = 288 𝑘𝑔/𝑚3 . When the foam is densified,
the Poisson’s ratio reaches the plastic Poisson’s ratio. For high-density foams, the
Poisson’s ratio is proposed to take the form

𝑒 𝑚𝑐 𝜈 −𝜈
𝜈𝑓𝑐 + (𝜈𝑓0 − 𝜈𝑓𝑐 )exp[(𝑒 ) ln(𝜈 𝑐𝑖 −𝜈𝑓𝑐 )] 𝑓𝑜𝑟 𝑒<0
𝑐𝑖 𝑓0 𝑓𝑐
𝜈𝑓 (𝑒) = { 𝑚𝑡 𝜈 −𝜈
(13)
𝑒
𝜈𝑓𝑡 + (𝜈𝑓0 − 𝜈𝑓𝑡 )exp[(𝑒 ) ln(𝜈 𝑡𝑖 −𝜈𝑓𝑡 )] 𝑓𝑜𝑟 𝑒≥0
𝑡𝑖 𝑓0 𝑓𝑡

where (𝑒𝑐𝑖 , 𝜈𝑐𝑖 ) and (𝑒𝑡𝑖 , 𝜈𝑡𝑖 ) are the inflection points on the compression and tension
branches of the Poisson’s ratio - engineering strain relation curve, respectively; 𝑚𝑐 and
𝑚𝑡 the material parameters controlling the curvature of the compression and tension
braches, respectively; 𝜈𝑓𝑐 and 𝜈𝑓𝑡 the plastic compressive and tensile Poisson’s ratios at
high strain, respectively; and 𝜈𝑓0 the elastic Poisson’s ratio. For the LAST-A-FOAM® FR-
3718, the material parameters are taken as 𝑒𝑐𝑖 = 0.12, 𝜈𝑐𝑖 = 0.30 , 𝜈𝑓0 = 0.34, 𝜈𝑓𝑐 =
0.24, and 𝑚𝑐 = 3 to fit the experimental results as shown in Fig. 1.
0.36 Test
0.34 This paper

0.32
Poisson's ratio

0.3

0.28

0.26

0.24

0.22

0.2
0 0.1 0.2 0.3 0.4
Engineering strain
Fig. 1. Poisson’s ratio of polyurethane foam LAST-A-FOAM® FR-3718

8
3.5 True stress-strain relationship of polyurethane foams

For the LAST-A-FOAM® FR-3718, the true stress-strain relationship can be obtained
by converting the engineering stress-strain using Eqs. (1) with the proposed Poisson’s
ratio in Eq. (13). The engineering and true stress vs plastic strain relationships are
depicted in Fig. 2. It can be found from the figure that the foam in the densification
regime with true stress-plastic strain relationship behaves softer than that indicated by
the engineering stress-plastic strain relationship.
In Fig. 3, the relationship of the true stress vs the volumetric plastic strain or the
modified axial plastic strain multiplied by (1 − 2𝜐𝑓𝑐 ) in accordance with Eq. (8) is also
shown. This relationship is required to be used as the hardening rule of the compressive
strength of the foams with volumetric hardening.
35 Engineering stress-axial plastic strain (Test)
30 True stress-axial plastic strain
True stress-volumetric plastic strain
25

20
Stress (MPa)

15

10

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 1.1 1.2 1.3
Plastic Strain
Fig. 2. Stress-strain relationships of polyurethane foam LAST-A-FOAM® FR-3718

3.6 Effect of strain rate on dynamic stress of polyurethane foams


As train rate increase, crushable foams show an increase in the yield stress. The
dynamic increase factor (DIF) of the uniaxial compressive stress may be defined as
𝑝𝑙 𝑝𝑙
𝑝𝑙 ̃𝑐 (𝜀𝑎𝑥𝑖𝑎𝑙 ,𝜀̇ 𝑎𝑥𝑖𝑎𝑙 )
𝜎
𝑟 (𝜀̇ ) = 𝑝𝑙 (14)
𝜎𝑐 (𝜀𝑎𝑥𝑖𝑎𝑙 )

𝑝𝑙
It is noted that 𝜀̇ = 𝜀̇𝑎𝑥𝑖𝑎𝑙
𝑝𝑙
for the uniaxial compression and tension.
𝑝𝑙 𝑝𝑙
The strain-rate-dependent size of the yield surface, 𝐵̃ (𝜀 , 𝜀̇ ) , can be calibrated by

𝑝𝑙 𝑝𝑙
𝑝𝑙 𝐵̃(𝜀 ,𝜀̇ ) 𝑝𝑙 3𝑘 +𝑟(𝜀 ̇ 𝑝𝑙 )[𝑘 +𝑘 (3−𝑘 )]
𝑅 (𝜀̇ ) = 𝑝𝑙 = 1 + (𝑟(𝜀̇ ) − 1) 𝑡 𝑐 𝑡
𝑝𝑙
𝑐
(15)
𝐵(𝜀 ) (1+𝑘𝑡 )(3𝑘𝑡 +𝑟(𝜀̇ )𝑘𝑐 )

9
In this paper, the DIF of the yield stress in the plateau regime of the uniaxial
compression curve is proposed in the following form similar to that presented by Song
et al (2009):
𝑝𝑙
𝑝𝑙 𝑎𝑥𝑖𝑎𝑙 𝜀̇
𝑟(𝜀̇𝑎𝑥𝑖𝑎𝑙 ) = 1 + 𝑟𝑐 𝑙𝑜𝑔( 10−3
) ≥ 1.0 (16)

For polyurethane foams of the high-density approximately 𝜌 = 300 𝑘𝑔/𝑚3 , the model
parameter in Eq. (16) can be taken as 𝑟𝑐 = 0.12 to fit the experimental results of the
plateau stress by Kasparek et al (2012), Lu (2014), Song et al (2009), and Maji et al
(1995) (perpendicular to rise), as illustrated in Fig. 3. It was demonstrated by Kasparek
et al (2011) that the DIF of the yield stress in the stress plateau regime can be applied
to the densification regime of compaction.
1.9

1.8 Kasparek et al (2012)


Lu (2014)
1.7
Song et al (2009)
1.6 Maji et al (1995)
DIF of stress

This paper
1.5

1.4

1.3

1.2

1.1

1
0.001 0.01 0.1 1 10 100 1000 10000
Strain Rate
Fig. 3. Dynamic increase factor of stress of polyurethane foam of 𝜌 = 300 𝑘𝑔/𝑚3

4. Concrete damaged plasticity model


The concrete damaged plasticity model is one of most promising concrete models
used for simulation of concrete failure. This model is proposed by Lubliner et al (1989)
for monotonic loading and later modified by Lee and Fenves (1998) for dynamic and
cyclic loading. The constitutive model not only captures the concrete elasticity and
plasticity including hardening and softening, but also the damage-dependent elastic
stiffness degradations and irreversible strains in the hardening and softening branches
of the stress-strain curves. Therefore, it provides the ability to model the mechanical
behavior of reinforced concrete structures subjected to impacts loads.
The concrete damaged plasticity model requires a full definition of material-specific
parameters that include: 1) uniaxial compressive stress-strain relationship; 2) uniaxial
tensile stress-strain relationship that involves fracture energy; 3) compressive damage
index evolution in terms of inelastic strain; 4) tensile damage index evolution in terms of

10
crack strain; 5) dilation angle in meridional plane; 6) ratio of equibiaxial compressive
strength to uniaxial compressive strength; and 7) ratio of the second stress invariant on
the tensile meridian to that on the compressive meridian.

4.1 Yield surface function of concrete


The yield surface function of the concrete damage plasticity model adopted in
ABAQUS (2017) is defined in the meridional plane taking the form
1
(𝑞(𝜎𝑖𝑗 ) − 3𝛼𝑝(𝜎𝑖𝑗 ) + 𝛽(𝜀̃ 𝑝𝑙 )〈𝜎̂𝑚𝑎𝑥 〉 − 𝛾〈−𝜎̂𝑚𝑎𝑥 〉) − 𝜎̅𝑐 (𝜀̃𝑐𝑝𝑙 ) = 0 (17)
1−𝛼

where
3(1−𝐾 )
𝛾 = 2𝐾 −1𝑐 , 0.5 ≤ 𝛾 ≤ 1.0 ;
𝑐
𝐾𝑐 is the ratio of the second stress invariant on the tensile meridian to that on the
compressive meridian;
(𝑓 ′ /𝑓′ )−1
𝛼 = 2(𝑓𝑏′ /𝑓𝑐 ′ )−1 , 0 ≤ 𝛼 ≤ 0.5;
𝑏 𝑐
𝑓𝑏′ /𝑓𝑐′ is the ratio of initial equibiaxial compressive strength to initial uniaxial compressive
strength;
𝑝𝑙
̅𝑐 (𝜀̃𝑐 )
𝜎
𝛽(𝜀̃𝑝𝑙 ) = 𝑝𝑙 (1 − 𝛼) − (1 + 𝛼) ;
̅𝑡 (𝜀̃𝑡 )
𝜎
𝑝𝑙
𝜎𝑐 (𝜀̃𝑐 )
𝜎̅𝑐 (𝜀̃𝑐𝑝𝑙 ) = 𝑝𝑙 , the effective compressive stress corresponding to the compressive
1−𝑑𝑐 (𝜀̃𝑐 )

stress 𝜎𝑐 (𝜀̃𝑐𝑝𝑙 ) being determined based on the compressive equivalent plastic strain, 𝜀̃𝑐𝑝𝑙 ;
𝑝𝑙
𝑝𝑙 𝜎𝑡 (𝜀̃𝑡 ) 𝑝𝑙
𝜎̅𝑡 (𝜀̃𝑡 ) = 𝑝𝑙 , the effective tensile stresses corresponding to the tensile stress 𝜎𝑡 (𝜀̃𝑡 )
1−𝑑 (𝜀̃ ) 𝑡 𝑡

being determined based on the tensile equivalent plastic strain, 𝜀̃𝑡𝑝𝑙 ;


The parameters 𝑑𝑐 (𝜀̃𝑐𝑝𝑙 ) and 𝑑𝑡 (𝜀̃𝑡𝑝𝑙 ) are the compressive and tensile damage indices
to characterize the degraded elastic compressive stiffness as (1 − 𝑑𝑐 (𝜀̃𝑐𝑝𝑙 )) 𝐸𝑐 and the
degraded elastic tensile stiffness as (1 − 𝑑𝑡 (𝜀̃𝑡𝑝𝑙 )) 𝐸𝑐 .

Lim et al (2016) calibrated from experimental results the following equations for the
ratio, 𝑓𝑏′ /𝑓𝑐′ , and the parameter, 𝐾𝑐 :

𝑓𝑏′ −0.09
= 1.57𝑓𝑐′ (18)
𝑓𝑐′

−0.025
𝐾𝑐 = 0.71𝑓𝑐′ (19)

4.2 Uniaxial compressive stress-strain relationship of concrete


Based on the experimental study, Aslani and Jowkarmeimandi (2012) proposed a
uniaxial compressive stress-strain relationship as shown in Eq. (20). This relationship is
considered as an engineering stress-strain relationship since it directly fits the
experimental results. The engineering stress-strain relationship can be converted into
11
true stress-strain relationship using Eqs. (1) for the purpose of large deformation
analysis.
𝑒
𝑛( 𝑐 )𝑓𝑐′
𝜀𝑐𝑝
𝑠𝑐 (𝑒𝑐 ) = 𝑒
𝑛 (20)
𝑛−1+( 𝑐 )
𝜀𝑐𝑝

where 𝑠𝑐 and 𝑒𝑐 are uniaxial engineering compressive stress and strain, respectively; 𝜀𝑐𝑝 is
the strain at the peak engineering stress, 𝑓𝑐′ ;

𝑓𝑐′
𝑛1 = (1.02 − 1.17 𝐸 )−0.74 𝑓𝑜𝑟 𝑒𝑐 ≤ 𝜀𝑐𝑝
𝑛={ 𝑐 𝜀𝑐𝑝 (21)
𝑛1 + 𝑎 + 28𝑏 𝑓𝑜𝑟 𝑒𝑐 > 𝜀𝑐𝑝

−911
in which 𝑎 = 3.5(12.4 − 0.0166𝑓𝑐′ )−0.46 and 𝑏 = 0.83exp( ) ; and
𝑓𝑐′

𝐸𝑐 is the initial modulus of elasticity, taking the form


𝛾
𝑐
𝐸𝑐 = (3320√𝑓𝑐′ + 6900)(2300)1.5 (22)

in which 𝛾𝑐 is the density of concrete in kg/m3. The strain at the peak compressive
stress, 𝜀𝑐𝑝 , can be considered as either engineering or true strain due to its small vale.
Based on the study by Samani and Attard (2012) for normal-weight concrete, the strain
at the peak compressive stress is proposed as
0.75
𝑓𝑐′ 𝛾
𝜀𝑐𝑝 = 𝜇𝑎 𝑐
(2300 )𝜑 (23)
𝐸𝑐

where 𝜇𝑎 = 4.26 for crushed aggregate or 𝜇𝑎 = 3.78 for rounded aggregate (Samani
and Attard, 2012); and the exponent, 𝜑 , is to incorporate heavyweight concrete. Based
on the limited experimental results performed in author’s organization for the
heavyweight concrete of density 3700 kg/m3, the exponent may be taken as 𝜑 = 0.3.

4.3 Uniaxial tensile stress-strain relationship of concrete


Cracked plain concrete under tension shows strain softening associated with the
localization of a single crack and a corresponding sharp tensile stress release. One way
to account for crack localization in finite element analysis is to assign softening
properties in the material constitutive law in the smeared cracking framework. This is
done by smearing the crack opening to the average strain over a reference domain of
an element dimension. For a three-point bending of a reinforced concrete beam
modeled using the fracture energy-based approach, Malm (2006) shows that there is a
very small difference between the responses of the models with the fine mesh (5 mm)
and the coarse mesh (40 mm).
Reinhardt et al (1986) experimentally derived a tensile stress-crack opening
relationship or cohesion law as follow in static condition:

12
𝜎𝑡 𝑤 3 𝑤 𝑤
= [1 + (𝑐1 𝑤 ) ] exp (−𝑐2 𝑤 ) − 𝑤 (1 + 𝑐1 3 )exp(−𝑐2 ) (24)
𝑓𝑡 𝑐 𝑐 𝑐

where 𝑐1 = 3 ; 𝑐2 = 6.93 ; 𝜎𝑡 is the uniaxial tensile stress; 𝑓𝑡 the initial uniaxial tensile
strength of concrete; 𝑤 the crack opening displacement; and 𝑤𝑐 the critical crack
opening displacement with free of stress determined by
𝐺𝐹
𝑤𝑐 = 5.136 (25)
𝑓𝑡

in which 𝐺𝐹 is the fracture energy that equals the area under the entire stress-crack
opening curve. By fitting the experimental results, Bazant and Becq-Giraudon (2002)
proposed the fracture energy of concrete as

𝜎 0.46 𝑑 0.22
𝑐𝑢
𝐺𝐹 = 𝛼𝑎 (0.051) 𝑎
(1 + 11.27) (𝑤/𝑐)−0.30 (26)

where 𝑑𝑎 is maximum aggregate size, mm; 𝑤/𝑐 the water cement ratio by weight; and
𝛼𝑎 = 1.0 for rounded aggregate or 𝛼𝑎 = 1.44 for crushed aggregate.
In lieu of an exponential cohesion law by Eq (24), a linear descending cohesion law
with a constant slope (de Oliveira e Sousa and Gettu, 2006) may be assumed:

𝜎𝑡 𝑤
𝑓𝑡
=1−𝑤 (27)
𝑒𝑐

where the equivalent critical crack opening displacement is determined by


2𝐺𝐹
𝑤𝑒𝑐 = (28)
𝑓𝑡

Based on the experimental results of the dynamic stress-crack strain or deformation


relationship by Cadoni et al (2013) and Weerheijm and Vegt (2010), a linear cohesion
law may be more appropriate for concrete cracking response under high strain rates.

4.4 Compressive damage index of concrete


Under impact loads, stress in concrete experiences ramping up and down
corresponding to the impact loading and unloading. The unloading stiffness of concrete
decreases with progress in unloading from the start point (𝜀𝑐 , 𝜎𝑐 ) at the envelope stress-
strain curve. Aslani and Jowkarmeimandi (2012) proposed the equivalent unloading
tangent modulus as
|𝜎𝑐 |
+0.57𝜀𝑐𝑝
𝐸𝑐
𝐸𝑐𝑢𝑛 = 𝐸𝑐 (|𝜀 ) (29)
𝑐 |+0.57𝜀𝑐𝑝

13
𝑝𝑙
By definition the compressive damage index, 𝑑𝑐 (𝜀̃𝑐 ) , indicating the degradation of
the unloading stiffness as illustrated in Fig. 4, can be expressed as 𝑑𝑐 (𝜀̃𝑐𝑝𝑙 ) = 1 −
𝐸𝑐𝑢𝑛 /𝐸𝑐 . With substitution of Eq. (26) into this expression, the compressive damage
index can be derived as follow in terms of inelastic compressive strain 𝜀̃𝑐𝑖𝑛 = 𝜀𝑐 − 𝜎𝑐 /𝐸𝑐 :

|𝜀̃𝑐𝑖𝑛 |
𝑑𝑐 (𝜀̃𝑐𝑖𝑛 ) = 𝜎 ≤ 0.8 (30)
|𝜀̃𝑐𝑖𝑛 + 𝑐 |+0.57𝜀𝑐𝑝
𝐸𝑐

where a limit of 0.8 is imposed to maintain a minimum stiffness of 0.2𝐸𝑐 . The inelastic
compressive strain can be converted to the equivalent plastic strain using 𝜀̃𝑐𝑝𝑙 = 𝜀̃𝑐𝑖𝑛 −
𝑑𝑐 (𝜀̃𝑐𝑖𝑛 ) 𝜎𝑐
. Eq. (30) is required as input to the concrete damage plasticity model.
1−𝑑𝑐 (𝜀̃𝑐𝑖𝑛 ) 𝐸𝑐

4.5 Tensile damage index of concrete


The unloading stiffness of concrete decreases with progress in unloading from the
start point (𝜀𝑡 , 𝜎𝑡 ) at the envelope stress-strain curve. The unloading stiffness of
concrete is proposed to take the form
𝐸𝑐
𝐸𝑡𝑢𝑛 = 𝐸 𝜀 (31)
1−𝜆+𝜆 𝑐 𝑡
𝜎𝑡

where 𝜆 is a parameter controlling the stiffness degradation. 𝜆 = 0 corresponds to no


degradation, 𝜆 = 1 corresponds to full degradation. Based on the experimental results
by Reinhardt (1984), the degradation parameter can be taken as 𝜆 = 0.2 for concrete
under uniaxial tension.
The tensile damage index, 𝑑𝑡 (𝜀̃𝑡𝑝𝑙 ) , indicating the degradation of the unloading
stiffness as illustrated in Fig. 4, can be expressed as 𝑑𝑡 (𝜀̃𝑡𝑝𝑙 ) = 1 − 𝐸𝑡𝑢𝑛 /𝐸𝑐 at any
unloading start point (𝜀𝑡 , 𝜎𝑡 ). With substitution of Eq. (31) into this expression, the
damage index can be derived as follow in terms of the crack strain 𝜀̃𝑡𝑐𝑟 = 𝜀𝑡 − 𝜎𝑡 /𝐸𝑐 :

𝜆𝐸𝑐 𝜀̃𝑡𝑐𝑟
𝑑𝑡 (𝜀̃𝑡𝑐𝑟 ) = 𝜆𝐸 ̃ 𝑡𝑐𝑟 +𝜎𝑡
≤ 0.8 (32)
𝑐𝜀

where the damage index is limited to 0.8 to maintain a minimum stiffness of 0.2𝐸𝑐 . The
𝑑 (𝜀̃ 𝑖𝑛 ) 𝜎
crack strain can be converted to the equivalent plastic strain via 𝜀̃𝑡𝑝𝑙 = 𝜀̃𝑡𝑖𝑛 − 1−𝑑𝑡 (𝜀𝑐̃ 𝑖𝑛) 𝐸𝑡 .
𝑡 𝑐 𝑐
Eq. (32) is required as input to the concrete damage plasticity model.

4.6 Uniaxial tensile strength of concrete


There are three types of tests to measure the strength of concrete in tension: direct
tension, flexure, and splitting tension. It has been well established (Arioglu et al, 2006)
that the simplest and the most reliable method, which generally provides a lower
coefficient of variation, is the splitting tensile test. The splitting strength can be used to
estimate direct tensile strength by multiplying by a conversion factor, 𝛼𝑠𝑝 , that can be

14
taken as 1.0 as given in the CEB-FIB Model Code 2010 (FIB, 2012), i.e., the splitting
tensile strength is approximately equal to the direct tensile strength.
For concrete without accelerating admixtures and made with crushable aggregates,
the tensile strength in MPa is provided by Arioglu et al (2006) by fitting experimental
results:
0.630
𝑓𝑡 = 𝑓𝑠𝑝 = 0.387𝑓𝑐′ (33)

4.7 Plastic flow potential of concrete


The concrete damaged plasticity model assumes non-associated plastic flow. The
flow potential takes the Drucker-Prager hyperbolic function:

2
𝐺(𝜎𝑖𝑗 ) = √(𝜖𝑓𝑡 tan 𝜓)2 + 𝑞(𝜎𝑖𝑗 ) − 𝑝(𝜎𝑖𝑗 ) tan 𝜓 (34)

where 𝜓 is the dilation angle measured in the meridional plane at high hydrostatic
pressure. 𝜖 is the eccentricity that defines the rate at which the hyperbolic plastic
potential function approaches the asymptote.
Malm (2006) and Ferrotto et al (2018) studied the effect of the dilation angle in the
range between 10𝑜 and 56.3𝑜 on the flexural response of reinforced concrete
specimens. The results indicate that the dilation angle strongly influences the flexural
strength and ductility of the specimens. The larger the dilation angle, the more ductile of
the specimen is, and the higher strength the specimen has. For normal strength
concrete, different authors in the past assumed different values of the dilation angle
between 13° and 49° independently. Lee and Fenves (1998) specified it as 31° for the
verification of the model; Nana et al (2017) suggested 37° based on good agreement
with the shear test results; Jankowiak and Lodygowski (2005) calibrated it to be 38°;
and Malm (2009) showed that a dilation angle between 30° and 40° for normal strength
concrete can have the best agreement with the experimental response.
In this paper, the dilation angle, 𝜓 , is proposed to be

6(𝜈𝑐0 −𝜈𝑐𝑒 )
𝜓 = 𝑡𝑎𝑛−1 [3𝐸𝑐𝜀𝑐𝑝 ] (35)
+2(𝜈𝑐0 −𝜈𝑐𝑒 )−3
𝑓′𝑐

where 𝜀𝑐𝑝 is the strain at the peak compressive engineering stress as given by Eq. (23);
𝜈𝑐𝑒 the elastic (initial) Poisson’s ratio; and 𝜈𝑐0 the Poisson’s ration at the peak
compressive engineering stress. The dilation angle may be limited to the concrete
material angle, 56.3𝑜 (Malm, 2006), but this cap value corresponds to a very high
compressive strength (greater than 140 MPa) which may not be justified due to scarcity
of the experimental results at this high strength.
Derivation of the plastic flow potential with respect to the hydrostatic pressure results
in the inclination of the plastic flow potential curve:

𝑑𝑞 𝑓 2
= tan(𝜓)√1 + ( 𝑞𝑡 𝜖tan(𝜓)) (36)
𝑑𝑝

15
Jiang and Wu (2012) derived a relationship between this inclination and plastic
volumetric and shear strain increments for the case of axial compression with uniform
lateral confinement. Since the inclination of the asymptote is constant, this relationship
can be extended as follow for the total inelastic strains under high hydrostatic pressure:
𝑖𝑛 𝑖𝑛 𝑖𝑛
𝑑𝑞 3𝑑𝜀𝑣𝑜𝑙 3 ∫ 𝑑𝜀𝑣𝑜𝑙 3𝜀𝑣𝑜𝑙
=− 𝑖𝑛 −𝜀 𝑖𝑛 ) =− 𝑖𝑛 −𝜀 𝑖𝑛 ) = 𝑖𝑛 𝑖𝑛 ) (37)
𝑑𝑝 2𝑑(𝜀𝑐𝑎 𝑐𝑙 2 ∫ 𝑑(𝜀𝑐𝑎 𝑐𝑙 2(𝜀𝑐𝑙 −𝜀𝑐𝑎

𝑖𝑛 𝑖𝑛 𝑖𝑛
where 𝜀𝑣𝑜𝑙 , 𝜀𝑐𝑎 , and 𝜀𝑐𝑙 are the inelastic volumetric, axial, and lateral strains,
respectively. To calibrate the dilation angle, the peak engineering stress state under
uniaxial compression is considered since the hydrostatic pressure under uniaxial
compression reaches the maximum as intended for the determination of the dilation
angle. Let the Poisson’s ratio at the peak engineering stress (−𝑓𝑐′ ) under the uniaxial
𝑖𝑛
compression be 𝜈𝑐0 , the inelastic volumetric strain is then 𝜀𝑣𝑜𝑙 = 2(𝜈𝑐0 − 𝜈𝑐𝑒 )𝑓𝑐′ /𝐸𝑐 . In
𝑖𝑛 𝑖𝑛 𝑖𝑛
addition, 𝜀𝑐𝑎 = −𝜀𝑐𝑝 + 𝑓𝑐′ /𝐸𝑐 , 𝜀𝑐𝑙 = (𝜀𝑣𝑜𝑙 𝑖𝑛
− 𝜀𝑐𝑎 )/2 . Substituting these values into Eq.
(37) and neglecting the second term in the square root of Eq. (36) due to small value of
𝑓𝑡 /𝑓𝑐′ result in Eq. (35).
Fig. 4 shows the variation of the dilation angle with the compressive strength in the
range between 20 MPa and 100 MPa for concrete made with crushed aggregate. The
dilation angle falls in the range between 29𝑜 and 42𝑜 for normal strength of concrete
(20-60 MPa) and increases with the increasing the compressive strength.

Fig. 4 Concrete dilation angle increases with increasing compressive strength

4.8 Poisson’s ratio of concrete in compression


Poisson’s ratio is used to convert engineering stress-strain relationships into true
stress-strain relationships by Eqs. (1). For concrete under uniaxial compression, Van
Mier (1986), Imran and Pantazopoulou (1996), Candappa et al (2001), Grassl et al
(2002), Binici (2005), and Poinard et al (2010) indicate that the volumetric strain at the
peak engineering stress under uniaxial compression is approximately zero, i.e., the
Poisson’s ratio at this stress state is approximately 𝜈𝑐0 = 0.5. Based on the study by
Binici (2005), the following Poisson’s ratio for concrete under uniaxial compression is
proposed:

16
𝜈𝑐𝑒 𝑓𝑜𝑟 |𝑒𝑐 | ≤ 𝜀𝑒𝑙
𝑛
𝜈𝑐 (|𝑒𝑐 |) = { |𝑒 |−𝜀 𝜈𝑐𝑝 −𝜈𝑐0 (38)
𝜈𝑐𝑝 − (𝜈𝑐𝑝 − 𝜈𝑐𝑒 ) exp [(𝜀 𝑐 −𝜀𝑒𝑙) ln(𝜈 )] 𝑓𝑜𝑟 |𝑒𝑐 | > 𝜀𝑒𝑙
𝑐𝑝 𝑒𝑙 𝑐𝑝 −𝜈𝑐𝑒

2
where 𝑒𝑐 is the uniaxial engineering compressive strain; 𝜈𝑐𝑒 = 8 × 10−6 𝑓𝑐′ + 0.0002𝑓𝑐′ +
0.138, the elastic (initial) Poisson’s ratio (Candappa et al, 2001); 𝜈𝑐0 = 0.5, the Poisson’s
ratio of concrete at the peak engineering stress; 𝜈𝑐𝑝 is the ultimate Poisson’s ratio at
large plastic strains to be determined; 𝜀𝑒𝑙 = 0.4𝑓𝑐′ /𝐸𝑐 , the strain limit to the elastic regime
under uniaxial compression. 𝑛 is a material constant being taken as 𝑛 = 3 to
approximately fit the experimental results of Candappa et al (2001), Binici (2005), and
Samani and Attard (2014) for small to moderate compressive deformation.
The strain-related Poisson’s ratio is proposed to be expressed as 𝜈𝑐 = 𝜈𝑐𝑒 (1 − 𝜀 𝑝 /𝜀) +
𝜈𝑐𝑝 (𝜀 𝑝 /𝜀) , where 𝜈𝑐𝑒 and 𝜈𝑐𝑝 are the elastic and plastic Poisson’s ratios, respectively.
As the strain increases, the stress decreases and the contribution of the elastic
Poisson’s ratio to the Poisson’s ratio is reduced. The plastic Poisson’s ratio is the
ultimate value when the elastic strain is negligibly small.
The plastic Poisson’s ratio is related to the yielding mechanism. Concrete is a
composite material consisting of a heterogeneous matrix of mortar and aggregates.
Under uniaxial compression of a concrete specimen, plasticity occurs when meso-
cracks are formed in the direction of the load. The extent of plasticity is increased as the
meso-cracks start to propagate and finally coalesce to form macro-cracks. Then the
specimen disintegrates into many parts separated by the macro-cracks, which are
considered as individual separated structures. Additional meso-cracks are formed in the
separated parts as the concrete exhibits a strain softening behavior. Lateral expansion
measured at the exterior surface of the specimen is very sensitive to the macro-crack
width, or the clear distance of the separated parts. Direct measurement of the lateral
expansion at the exterior surface may not be objective, because 1) the separated parts
may be subjected to non-uniform compression owing to non-uniform damage to and
different profiles of the individual parts, which lead to their different structural behavior
and cause different macro-crack widths; 2) the direct measurement at the exterior
surface is also affected by loading and restraining conditions at the ends of the
specimen; and 3) the individual structural behavior and interaction of the separated
parts should be captured by structural analysis, not by material constitutive model.
Therefore, the material behavior should be limited within the matrix of the individual
parts with meso-cracks and the lateral expansion of the specimen is to exclude the
macro-crack width. The plastic Poisson’s ratio represents the lateral plastic dislocation
of grains in the matrix and may be assumed to be constant during the material
softening.
The plastic Poisson’s ratio of concrete, 𝜈𝑐𝑝 , is then determined using the plastic
deformation at the peak engineering stress of concrete prior to the onset of macro
cracks. It is the ratio of the inelastic lateral strain increment to the inelastic axial strain
𝑑𝜀 𝑖𝑛
𝑐𝑙
increment, i.e., 𝜈𝑐𝑝 = − 𝑑𝜀𝑖𝑛 . Using Eqs. (36) and (37), the plastic Poisson’s ratio of
𝑐𝑎
3+2𝑡𝑎𝑛(𝜓)
concrete can be obtained as 𝜈𝑐𝑝 = 6−2𝑡𝑎𝑛(𝜓) . Based on the proposed dilation angle in
Eq. (35), the plastic Poisson’s ratio is given by

17
𝐸𝑐 𝜀𝑐𝑝 −2𝜈𝑐𝑒 𝑓𝑐′
𝜈𝑐𝑝 = (39)
2(𝐸𝑐 𝜀𝑐𝑝 −𝑓𝑐′ )

For normal strength concrete made with crushed aggregates, the plastic Poisson’s ratio
is approximately 1.0.
For the 40 MPa concrete, the proposed Poisson’s ratio is displaced in Fig. 5. This
figure also shows the derived Poisson’s ratios based on 1) the lateral strain model by
Samani and Attard (2014) which fits experimental results of the specimens of standard
geometry with presence of macro-cracks; 2) the Poisson’s ratio model by Candappa et
al (2001) which fits the experimental results under moderate inelastic lateral expansion;
3) the Poisson’s ratio model by Binici (2005) who calibrated the model such that the
curve crosses through the Poisson ratio (0.5) at the peak engineering stress and fits the
experimental results with presence of macro-cracks. The plots show good agreement of
the proposed Poisson’s ratio model with the experimental results in small to moderate
compressive strains up to 0.0025. At high compressive strains, the models with macro-
cracks included are not appropriate as discussed previously. Ferretti (2004) also
indicates that lateral strain measurements acquired on specimen surface are not real
lateral expansion of concrete material due to macro-cracks propagating through the
specimen. Therefore, the results in high strain range proposed by Binici (2005),
Candappa et al (2001) and Samani and Attard (2014) may have overestimated the
realistic lateral strain of the concrete due to presence of macro-cracks.
3
Samani and Attard (2014)
Binici (2005)
3
Candappa et al (2001)
This paper
2
Poisson's ratio

0
0 0.001 0.002 0.003 0.004 0.005 0.006 0.007 0.008 0.009 0.01
Engineering strain
Fig. 5 Comparison of Poisson’s ratio models

4.9 Effect of strain rate on the response of concrete


As strain rate increases, concrete shows an increase in compressive and tensile
strengths as well as modulus of elasticity. Eq. (19) for uniaxial compressive stress-strain
relationship of concrete and Eq. (23) or (26) for uniaxial tensile stress-crack

18
displacement relationship can be used for the corresponding dynamic relationships with
the updated materials properties due to effect of strain rates.

4.9.1 Concrete compressive strength


The dynamic increase factor (DIF) of the uniaxial compressive strength in terms of
uniaxial strain rate, 𝜀̇𝑐 , presented by Malvar and Ross (1998) and adopted in CSA S850
(2012), is given by

𝜀̇
(𝜀̇ 𝑐 )1.026𝛼𝑠 ≥ 1.0 𝑓𝑜𝑟 𝜀̇𝑐 ≤ 30𝑠 −1
𝑓̃𝑐′ (𝜀̇ 𝑐 )
𝑟𝑐 (𝜀̇𝑐 ) = = { 𝑐0 (40)
𝑓𝑐′ 𝜀̇
𝛾𝑠 (𝜀̇ 𝑐 )1/3 𝑓𝑜𝑟 𝜀̇𝑐 > 30𝑠 −1
𝑐0

where 𝜀̇𝑐0 = 30 × 10−6 𝑠 −1 , the static strain rate; 𝛼𝑠 = 1/(5 + 9𝑓𝑐′ /𝑓𝑐𝑜 ) ; 𝑙𝑜𝑔(𝛾𝑠 ) =
6.15𝛼𝑠 − 2 ; and 𝑓𝑐𝑜 = 10 𝑀𝑃𝑎. Since 𝐸𝑇 (𝜀𝑐 ) = 0 at the peak stress, the DIF in terms of
𝑝𝑙 𝑝𝑙
the plastic strain rate, 𝜀̇𝑐 , is 𝑟 𝑝𝑙 (𝜀̇𝑐 ) = 𝑟(𝜀̇𝑐 ) as per Eq. (4).

4.9.2 Concrete tensile strength


The DIF of the uniaxial tensile strength in terms of uniaxial strain rate, 𝜀̇𝑡 , presented
by Malvar and Ross (1998) and adopted in CSA S850 (2012), is given by

𝜀̇
(𝜀̇ 𝑡 )𝛿𝑠 ≥ 1.0 𝑓𝑜𝑟 𝜀̇𝑡 ≤ 1𝑠 −1
𝑓̃𝑡 (𝜀̇ 𝑡 )
𝑟𝑡 (𝜀̇𝑡 ) = = { 𝑡0 𝜀̇ (41)
𝑓𝑡
𝛽𝑠 (𝜀̇ 𝑡 )1/3 𝑓𝑜𝑟 𝜀̇𝑡 > 1𝑠 −1
𝑡0

where 𝜀̇𝑡0 = 10−6 𝑠 −1, the static strain rate; 𝛿𝑠 = 1/(1 + 8𝑓𝑐′ /𝑓𝑐𝑜 ) ; and 𝑙𝑜𝑔(𝛽𝑠 ) = 6𝛿𝑠 − 2.
Since 𝐸𝑇 (𝜀𝑐 ) = 0 at the peak stress, the DIF in terms of the plastic strain rate, 𝜀̇𝑡𝑝𝑙 , is
𝑟 𝑝𝑙 (𝜀̇𝑡𝑝𝑙 ) = 𝑟(𝜀̇𝑡 ) as per Eq. (4).

4.9.3 Modulus of elasticity of concrete


The effect of uniaxial strain rate, 𝜀̇𝑢 , on the modulus of elasticity may be estimated
using the equation adopted in CEB-FIP Model Code 2010 (FIB, 2012):

𝐸̃𝑐 (𝜀̇ 𝑢 ) 𝜀̇
𝑟𝐸 (𝜀̇𝑢 ) = = (𝜀̇ 𝑢 )0.026 ≥ 1.0 (42)
𝐸𝑐 𝑐0

The DIF of the modulus of elasticity cannot be presented in terms of the plastic strain
rate since the modulus is a quantity measured in the elastic regime. It is noted that there
may be no built-in method in a finite element analysis software to incorporate the strain-
rate-dependent modulus of elasticity. Then a user-defined subroutine may be needed to
be implemented in the software for the modulus to account for the strain rate effect.

4.9.4 Strain at peak compressive engineering stress of concrete


The effect of strain rate on the strain at the peak compressive engineering stress may
be estimated using the equation adopted in CEB-FIP Model Code 2010 (FIB, 2012):

19
𝜀̃𝑐𝑝 (𝜀̇ 𝑐 ) 𝜀̇
𝑟𝜀𝑝 (𝜀̇𝑐 ) = = (𝜀̇ 𝑐 )0.02 ≥ 1.0 (43)
𝜀𝑐𝑝 𝑐0

The DIF in terms of the plastic strain rate, 𝜀̇𝑐𝑝𝑙 , is assumed to be 𝑟𝐸 (𝜀̇𝑐𝑝𝑙 ) = 𝑟𝐸 (𝜀̇𝑐 ) since
𝐸𝑇 (𝜀𝑐 ) = 0 as per Eq. (4).

4.9.5 Fracture energy of concrete


Ideally, the dynamic fracture energy should be directly related to crack opening
velocity instead of strain rate since strain after crack initiation is not objective and
depends on the gauge length of specimens. Experimental results with true crack
opening velocity in the literature are scarce. One contribution factor for the scarcity of
the experimental results has been the technical difficulty in controlling crack opening at
a constant velocity. Lukic (2018) compiled many experimental results of dynamic
fracture energy of concrete in terms of strain rates. It is noted that the strain rate is the
reference rate at the crack initiation in stress rate-controlled experimental programs.
One way to approximately convert the strain rate into crack opening velocity is to
assume that crack opening velocity in these experimental programs is constant and can
be obtained by multiplying the strain rate by the specimen gauge length. Two sets of
experimental results by Cadoni et al (2013) and Brara and Klepaczko (2007), which are
the largest datasets of the experimental results compiled by Lukic (2018), are used for
the rate conversion. Fig. 9 shows the DIF of the fracture energy in terms of the
converted crack opening velocity. For the dataset by Brara and Klepaczko (2007), the
static fracture energy used for the DIF calculation was not reported but calculated using
Eq. (26). The DIF of fracture energy in terms of the crack opening velocity is proposed
to take the similar form as for the tensile strength:

𝑤̇
𝐺̃𝑓 (𝜀̇ 𝑔 )
(𝑤̇ )𝛿𝑔 ≥ 1.0 𝑓𝑜𝑟 𝑤̇ ≤ 𝑤̇𝑖
𝑟𝐺 (𝑤̇ ) = = { 0 𝑤̇ (44)
𝐺𝑓
𝛽𝑔 (𝑤̇ )𝜆𝑔 𝑓𝑜𝑟 𝑤̇ > 𝑤̇𝑖
0

where 𝑤̇ is the crack opening velocity; 𝑤̇0 the reference crack opening velocity; 𝑤̇𝑖 the
𝑤̇
crack opening velocity at the trend transition; and 𝛽𝑔 = (𝑤̇ 𝑖 )𝛿𝑔−𝜆𝑔 . To fit the experimental
0
results as displayed in Fig. 6, the model parameters are taken as 𝑤̇0 = 0.01 𝑚𝑚/𝑠, 𝛿𝑔 =
0.08 , 𝜆𝑔 = 0.62 and 𝑤̇𝑖 = 200 𝑚𝑚/𝑠 . The comparison between the experimental results
and the proposed Eq. (44) is illustrated in Fig. 6.

4.10 Dynamic true stress-strain relationship of concrete


Using Eqs. (1) with the Poisson’s ratio given in Eq. (38), the compressive engineering
stress-strain relationship of concrete in Eq. (20) can be converted into the true stress-
strain relationship. With the effect of strain rate on the compressive strength, tensile
strength, strain at peak engineering stress, and facture energy being accounted for, the
dynamic true stress-strain relationships can be obtained using the same formulation in
Eq. (2). Fig. 7 shows the dynamic true stress-strain relationships of a 40 MPa
heavyweight concrete in compression.

20
Brara et al (2007)
25
DIF of fracture enegy
Cadoni et al (2013)
This paper

1
1E+0 1E+1 1E+2 1E+3 1E+4 1E+5
Crack opening velocity (mm/s)
Fig. 6 DIF of fracture energy in terms of crack opening velocity

100
Static test
(Engineering)
80 Static (True)

60
Stress (MPa)

40

20

0
0 0.001 0.002 0.003 0.004 0.005 0.006
Total Strain

Fig. 7 Uniaxial stress-strain relationship of heavyweight concrete

5. Stress-strain relationship of steel reinforcing bars


The stress-strain relationship of a steel reinforcing bar (rebar) embedded in concrete
is different from that of a bare rebar due to the bond-slip behavior at the interface
between the rebar and concrete. The debonding between the rebar and concrete occurs
after concrete cracking but before steel yielding. This slippage of the rebar adds
additional deformation and behaves similar to “yielding” in the average sense. An
explicit representation of the bond-slip mechanism in an analysis model is
computationally expensive, and may be impractical for large and complex structures.
Alternatively, the bond-slip effect is smeared into the deformation of a reinforcing bar
and the averaged stress-strain relationship is used for the practical analysis.

21
5.1 Average stress-strain relationship of rebars
Hsu and Mo (2010) proposed a bilinear average stress-strain curve for rebars
embedded in concrete by fitting experimental data. The bilinear stress-strain
relationship may be sufficient if the rebars experience small to moderate inelastic
deformation. It may overestimate the strength of the rebars under large inelastic strain.
Therefore, a slight modification to this relationship by limiting the stress to the ultimate
stress, 𝑓𝑠𝑢 , is proposed and given by

𝐸𝑠 𝜀𝑠 𝑓𝑜𝑟 𝜀𝑠 ≤ 𝜀𝑠𝑦
𝜎𝑠 = { ′ (45)
(0.91 − 2𝐵𝑎 )𝑓𝑠𝑦 + (0.02 + 0.25𝐵𝑎 )𝐸𝑠 𝜀 ≤ 𝑓𝑠𝑢 𝑓𝑜𝑟 𝜀𝑠 > 𝜀𝑠𝑦

where 𝑓𝑠𝑦 and 𝑓𝑠𝑢 are the engineering yield stress and tensile strength of bare rebars,

𝑓𝑠𝑦
′ ′
respectively; 𝜀𝑠𝑦 = , the average (smeared) yield strain; 𝑓𝑠𝑦 = (0.93 − 2𝐵𝑎 )𝑓𝑠𝑦 , the
𝐸𝑠
1 𝑓
average yield stress; and 𝐵𝑎 = 𝜌 (𝑓 𝑡 )1.5 , in which 𝑓𝑡 is the tensile strength of concrete
𝑠𝑦
and 𝜌 (≥ 0.15%) the reinforcement ratio.
It is noted that the truss stress-strain relationship of rebars is identical to the
engineering stress-strain relationship if the rebars are modeled as truss or beam
elements without considering cross-sectional Poisson’s effect.

5.2 Effect of strain rate on the response of rebars


As strain rate increases, steel shows an increase in compressive and tensile
strengths. The following two sections discuss the effect of strain rate on the yield stress
and tensile strength of rebars.

5.2.1 Yield stress of rebars


The dynamic increase factor (DIF) of the yield stress of steel rebars in terms of
uniaxial strain rate, 𝜀̇𝑠 , presented by Malvar (1998) and adopted in CSA S850 (2012), is
given by

𝑓̃𝑠𝑦 (𝜀̇ 𝑠 ) 𝜀̇
𝑟𝑠𝑦 (𝜀̇𝑠 ) = 𝑠
= (10−4 )𝛼𝑠𝑦 ≥ 1.0 (46)
𝑓𝑠𝑦

𝑓𝑠𝑦
where 𝛼𝑠𝑦 = 0.074 − 0.04 414 . The DIF of the yield stress can be expressed in terms of
𝑝𝑙
the plastic strain rate, 𝜀̇𝑎𝑥𝑖𝑎𝑙 , which is assumed to be 𝑟𝑠𝑦 (𝜀̇𝑠𝑝𝑙 ) = 𝑟𝑠𝑦 (𝜀̇𝑠 ) since 𝐸𝑇 (𝜀𝑠𝑦 ) =
0 for rebars at the yield stress plateau regime.

5.2.2 Tensile strength of rebars


The dynamic increase factor (DIF) of the tensile strength of steel rebars in terms of
uniaxial strain rate, 𝜀̇𝑠 , presented by Malvar (1998) and adopted in CSA S850 (2012), is
given by

𝑓̃𝑠𝑢 (𝜀̇ 𝑠 ) 𝜀̇
𝑟𝑠𝑢 (𝜀̇𝑠 ) = 𝑠
= (10−4 )𝛼𝑠𝑢 ≥ 1.0 (47)
𝑓𝑠𝑢

22
𝑓𝑠𝑦
where 𝛼𝑠𝑢 = 0.019 − 0.009 414 . The DIF of the tensile strength can be expressed in
terms of the plastic strain rate, 𝜀̇𝑠𝑝𝑙 , which is assumed to be 𝑟𝑠𝑢 (𝜀̇𝑠𝑝𝑙 ) = 𝑟𝑠𝑢 (𝜀̇𝑠 ) since
𝐸𝑇 (𝜀𝑠𝑢 ) = 0 for rebars at the peak engineering stress (also true stress as discussed in
Section 5.1).

6. Constitutive model for ductile metals


6.1 Yield surface surface function and plastic flow potential
For ductile metals, the classical von Mises yield surface as defined in the meridional
plane is commonly used:
𝑝𝑙
𝐹(𝜎𝑖𝑗 ) = 𝑞(𝜎𝑖𝑗 ) − 𝜎𝑦 (𝜀 ) = 0 (48)

𝑝𝑙
where 𝜎𝑦 (𝜀 ) is the evolved yield stress with the hardening variable being taken as the
𝑝𝑙
equivalent plastic strain, 𝜀 . As the associated plastic flow shows no plastic volume
change, the equivalent plastic strain simply coincides with the uniaxial plastic strain, i.e.,
𝑝𝑙 𝑝𝑙
𝜀 = |𝜀𝑎𝑥𝑖𝑎𝑙 | , which is given by a uniaxial test.
For cases involving gross plastic straining such as plastic deformation under impact
loading, an isotropic hardening rule can be assumed. For ductile metals, associated
flow rule is commonly applied, and the plastic flow potential function is identical to the
yield surface function.

6.2 Poisson’s ratio of ductile metals


For ductile metals, the elastic Poisson’s ratio can be taken as 0.3. After initiation of
yielding, the Poisson’s ratio reaches 0.5 since the volume of the metals remains
approximately unchanged.

6.3 Fracture plastic strain of ductile metals


Under triaxial stress states, the criterion for the onset of fracture of ductile metals can
𝑝
be specified using the initial equivalent plastic strain at fracture, 𝜀̅𝑐𝑟 . In accordance with
ASME BPVC Section VIII.2 (ASME, 2017b), the initial plastic fracture strain is given in
the following form:

𝑝 𝑝 𝛼𝑠1 1
𝜀̅𝑐𝑟 = 𝜀𝑐𝑟 𝑒𝑥𝑝 [− (1+𝑚 ) (𝜂 − 3)] (49)
2

where 𝜂 = −𝑝(𝜎𝑖𝑗 )/𝑞(𝜎𝑖𝑗 ) is the stress triaxiality factor defined as the ratio of the
hydrostatic pressure stress to the equivalent stress (𝜂 = 1/3 corresponds to uniaxial
tension); 𝛼𝑠1 = 2.2 for ferritic steel, or 𝛼𝑠1 = 0.6 for stainless steel; 𝑚2 = 0.6(1 − 𝑓𝑠𝑦 /𝑓𝑠𝑢 )
for ferritic steel, or 𝑚2 = 0.75(1 − 𝑓𝑠𝑦 /𝑓𝑠𝑢 ) for stainless steel; and 𝑓𝑠𝑦 and 𝑓𝑠𝑢 the
𝑝
engineering yield stress and engineering ultimate strength, respectively. 𝜀𝑐𝑟 is the
critical plastic strain at the onset of fracture to be determined based on uniaxial test
result of a standard specimen, or determined using Eq. (50) as per ASME BPVC
(ASME, 2017b):

23
𝑝 𝐸𝑓 𝑓𝑠𝑦
𝜀𝑐𝑟 = 𝐴𝜀 𝑙𝑛 (1 + 100) ≥ 𝐵𝜀 (1 − 𝑓 ) (50)
𝑠𝑢

where 𝐸𝑓 is the % elongation of a standard specimen; and 𝐴𝜀 = 2 for ferritic steel, or


𝐴𝜀 = 3 for stainless steel. The fracture criterion in Eq. (49) can be incorporated into a
progressive damage and failure rule in conjunction with the plasticity models for ductile
metals.
It is noted that the fracture initiation strain may be element-type-dependent and mesh-
size-dependent (Korgesaar et al, 2014; and Pack and Mohr, 2017). The fracture
initiation strain for the corresponding element type and mesh size must be calibrated
based on experimental results such that the force-displacement curve is correlated with
the experimental results by performing a nonlinear finite element analysis of the test
specimen.

6.4 Static engineering stress-plastic strain relationship of ductile metals


Soroushian and Choi (1987) proposed an engineering stress-strain relationship up to
peak engineering stress taking the form

𝑓𝑠𝑦
𝐸𝑠 𝑒 𝑓𝑜𝑟 𝑒 < 𝐸𝑠
𝑓𝑦𝑠
𝑠 = 𝑓𝑠𝑦 𝑓𝑜𝑟 𝐸𝑠
≤ 𝑒 ≤ 𝑒𝑠ℎ (51)
112(𝑒−𝑒 )+2 𝑒−𝑒𝑠ℎ 𝑓𝑠𝑢
𝑓𝑠𝑦 [ 60(𝑒−𝑒 𝑠ℎ)+2 + 𝑒 ( − 1.7)] 𝑓𝑜𝑟 𝑒𝑠ℎ < 𝑒 ≤ 𝑒𝑢
{ 𝑠ℎ 𝑢 −𝑒𝑠ℎ 𝑓𝑠𝑦

where 𝑒 and 𝑠 are uniaxial engineering strain and stress, respectively; 𝑓𝑠𝑦 and 𝑓𝑠𝑢 the
uniaxial engineering yield and peak stresses, respectively; 𝐸𝑠 the modulus of elasticity;
and 𝑒𝑠ℎ and 𝑒𝑢 the engineering strains at onset of strain hardening and that at the peak
engineering stress, respectively.

6.5 Static true stress-plastic strain relationship of ductile metals


If steel components are modeled using solid elements, the true stress-strain
relationship is required. It was found that the Ramberg-Osgood stress-strain relationship
has difficulty to closely fit the experimental result of the true stress-strain curve for the
mild steel CSA G40.20/21 300W provided by Chen (2010). To fit the experimental result
for general mild steels, the John-Cook constitutive model is expanded by adding a linear
term and a general true stress-plastic strain relationship is proposed for ductile metals
with an engineering strain hardening as follow:
𝜀𝑝
𝑓𝑜𝑟 𝜀𝑠𝑝 < 𝜀𝑠ℎ
𝑝
𝑝 1 + 𝜀𝑠ℎ
𝑝 𝜀𝑠
𝜎𝑠 (𝜀𝑠 ) 𝑠ℎ
={ (52)
𝑓𝑠𝑦 𝛿
𝐸1 𝜀𝑠𝑝 𝐸2 (𝜀𝑠𝑝 ) 𝑝
𝜀𝑠𝑝 𝑝
𝑟𝑖 + + 𝑓𝑜𝑟 𝜀𝑠ℎ ≤ ≤ 𝜀𝑐𝑟

where 𝐸1 , 𝐸2 , and 𝛿 are material constants, which are used for a combination of a
Ramberg-Osgood exponential strain hardening at small to moderate plastic strains and
𝑝
a linear strain hardening correction at large plastic strains; 𝑟𝑖 = 1 + 𝜀𝑠ℎ − 𝐸1 𝜀𝑠ℎ −

24
𝑝 𝛿 𝑝
𝐸2 (𝜀𝑠ℎ ) ; 𝜀𝑠ℎ = 𝜀𝑠ℎ − 𝑓𝑠𝑦 /𝐸𝑠 ; 𝜀𝑠ℎ the strain at onset of strain hardening in the
engineering stress-strain curve (e.g., the strain at the end of the yield stress plateau in
the engineering stress-strain curve); 𝐸𝑠 the modulus of elasticity; 𝑓𝑠𝑦 the yield stress
defined as the elastic limit stress below which stress-strain behavior is linear (it is noted
𝑝
that yield stress is not the proof or offset yield stress); and 𝜀𝑐𝑟 is determined by test or
given by Eq. (50).
For mild steels, the material constants 𝐸1 , 𝐸2 and 𝛿 in Eq. (52) are proposed to be
determined for the true stress-plastic strain curve to cross through the following three
true plastic strain-stress points: 〈𝜀𝐿𝑝 , 𝜎𝐿 〉; 〈𝜀𝑢𝑝 , 𝜎𝑢 〉; and 〈𝜀ℎ𝑝 , 𝜎ℎ 〉:
112𝑒 +2 𝑒𝑠ℎ
𝜎𝐿 = (1 + 2𝑒𝑠ℎ ) [ 60𝑒 𝑠ℎ+2 𝑓𝑠𝑦 + 𝑒 (𝑓𝑠𝑢 − 1.7𝑓𝑠𝑦 )]
𝑠ℎ 𝑢 −𝑒𝑠ℎ
𝜎𝐿
𝜀𝐿𝑝 = ln(1 + 2𝑒𝑠ℎ ) −
𝐸𝑠

𝜎𝑢 = 𝑓𝑠𝑢 (1 + 𝑒𝑢 ) (53)
𝜎
𝜀𝑢𝑝 = 𝑙𝑛(1 + 𝑒𝑢 ) − 𝐸𝑢
𝑠

1+𝑒
𝜎ℎ = 0.5𝑓𝑠𝑢 exp(2𝜀𝑢𝑝 )(1 + √exp(2𝜀𝑢𝑝 ))
𝑢

𝜀ℎ𝑝 = 2𝜀𝑢𝑝
in which 𝜀𝐿𝑝 and 𝜀𝑢𝑝 are the true plastic strains at the true stress, 𝜎𝐿 , corresponding the
engineering stress determined using Eq. (51) with respect to the engineering strain of
2𝑒𝑠ℎ and the true stress, 𝜎𝑢 , corresponding to the peak engineering stress, 𝑓𝑠𝑢 ,
𝑝
respectively. The true stress, 𝜎ℎ , is derived using Eq. (3) along with the fact that 𝜀ℎ ≅ 𝜀ℎ
for the large strain and that the engineering stress is very close to 𝑓𝑠𝑢 at the true plastic
strain of 𝜀ℎ𝑝 or approximately the corresponding engineering strain of 1 − 1/(1 + 𝑒𝑢 )2 .

6.6 True stress-strain relationship of mild steel CSA G40.20/21 300W


Chen (2010) experimentally studied the uniaxial true stress-strain relationship of the
mild steel. After correction for non-uniaxial stress condition in the necking region by Eq.
(3), the experimental result of the true stress-plastic strain relationship is displayed in
Fig. 11. To calibrate the material constants in Eq. (52) for this tested steel, let the true
stress-plastic strain curve cross through the proposed three points as presented in
Section 6.5 with Eqs. (53) using the experimental results of the material properties of
the group G specimens: 𝑓𝑠𝑦 = 352 𝑀𝑃𝑎, 𝑓𝑠𝑢 = 536 𝑀𝑃𝑎 , 𝑒𝑠ℎ = 0.017 , and 𝑒𝑢 = 0.16
(Chen, 2010). The resulted material constants are 𝐸1 = −0.282 , 𝐸2 = 125.9, and 𝛿 =
2.774 × 10−3. The derived true stress-plastic strain relationship is displayed in Fig. 8,
which agrees well with the experimental result.
Consider the same steel grade CSA G40.20/21 300W but with the code-specified
material properties 𝑓𝑠𝑦 = 300 𝑀𝑃𝑎 and 𝑓𝑠𝑢 = 450 𝑀𝑃𝑎 and assuming the same 𝑒𝑠ℎ =
0.017 and 𝑒𝑢 = 0.16. Letting the true stress-plastic strain curve cross through the three
points results in the corresponding material parameters for Eq. (52): 𝐸1 = −0.216, 𝐸2 =
120.9 and 𝛿 = 2.758 × 10−3. The resulted true stress-plastic strain relationship with
these specified material properties is shown in Fig. 8.

25
800

Stress (MPa) 600

400
Engineering (Chen, 2010)
True (Chen, 2010)
200 This paper
True with specified properties
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
Plastic Strain
Fig. 8 Stress-plastic strain relationship of CSA G40.20/21 300W

6.7 Uniaxial plastic stress-strain relationship of stainless steels


Stainless steels do not exhibit a noticeable yield stress plateau in the engineering
stress-strain curve before strain hardening. The general form of the true stress-plastic
𝑝
strain relationship in Eq. (52) can still be applicable to stainless steels. With 𝜀𝑠ℎ = 0 , Eq.
(52) reduces to the following equation for stainless steels:
𝑝
𝜎𝑠 (𝜀𝑠 ) 𝛿
= 1 + 𝐸1 (𝜀𝑠𝑝 ) + 𝐸2 (𝜀𝑠𝑝 ) (54)
𝑓𝑠𝑝

where 𝑓𝑠𝑝 is the proportionality limit taken as the proof stress at 0.01% strain offset.
Based on the engineering stress-strain relationship included in Annex C of Eurocode 3
(2006) for the constitutive modeling of stainless steel material behavior, 𝑓𝑠𝑝 falls in the
range of 0.6 to 0.7 times the yield or proof stress at 0.2% strain offset, 𝑓𝑠𝑦 .
Approximately, it can be taken as 𝑓𝑠𝑝 = 0.65𝑓𝑠𝑦 . 𝐸1 , 𝐸2 and 𝛿 are material constants
which are proposed to be determined for the true stress-plastic strain curve to cross
through the following three true plastic strain-stress points: 〈𝜀𝑎𝑝 , 𝜎𝑎 〉; 〈𝜀𝑢𝑝 , 𝜎𝑢 〉; and
〈𝜀ℎ𝑝 , 𝜎ℎ 〉:
𝜎𝑎 = 0.5(𝑓𝑠𝑦 + 𝑓𝑠𝑢 )(1 + 𝑒𝑎 )
𝜎𝑎
𝜀𝑎𝑝 = 𝑙𝑛(1 + 𝑒𝑎 ) −
𝐸𝑠
𝜎𝑢 = 𝑓𝑠𝑢 (1 + 𝑒𝑢 ) (55)
𝜎
𝜀𝑢𝑝 = 𝑙𝑛(1 + 𝑒𝑢 ) − 𝐸𝑢
𝑠

1+𝑒
𝜎ℎ = 0.57𝑓𝑠𝑢 exp(2𝜀𝑢𝑝 )(1 + √exp(2𝜀𝑢𝑝 ))
𝑢

𝜀ℎ𝑝 = 2𝜀𝑢𝑝

26
𝑓 𝑓𝑠𝑦 𝑓 1−𝑓𝑠𝑦 /𝑓𝑠𝑢
𝑒𝑎 = 0.007 𝑓𝑠𝑢 − 0.005 + 𝑠𝑢
+ 2𝐸 + 1+3.5𝑓𝑠𝑦 /𝑓𝑠𝑢
𝑠𝑦 𝐸𝑠 𝑠 2

where 𝜀𝑦𝑝 and 𝜀𝑢𝑝 are the true plastic strains at the true yield stress, 𝜎𝑦 , and the true
stress, 𝜎𝑢 , corresponding to the peak engineering stress, respectively. 𝑒𝑎 is the
engineering strain corresponding to the engineering stress of 0.5(𝑓𝑠𝑦 + 𝑓𝑠𝑢 ) which is
determined based on the engineering stress-stain curve adopted in Annex C of
Eurocode 3 (2006). The proposed strain, 𝜀𝑢𝑝 , is derived based on the engineering strain,
𝑒𝑢 = 1 − 𝑓𝑠𝑦 /𝑓𝑠𝑢 , corresponding to the peak engineering stress, , adopted in Annex C
of Eurocode 3 (2006). The true stress, 𝜎ℎ , is derived using Eq. (3) along with the fact
that 𝜀ℎ ≅ 𝜀ℎ𝑝 for the large strain and that the engineering stress approximately equals to
𝑓𝑠𝑢 at the true plastic strain of 𝜀ℎ𝑝 or approximately the corresponding engineering strain
2
of 1 − [𝑓𝑠𝑢 /(2𝑓𝑠𝑢 − 𝑓𝑠𝑦 )] ≈ 𝑒𝑢 for 𝑓𝑠𝑢 = (1 𝑡𝑜 3)𝑓𝑠𝑦 .
For AISI 304L, Blandford et al (2007) performed experimental investigation of the true
stress-strain relationship. The tensile properties of the 304L (average of three tests) are
𝑓𝑠𝑦 = 277 𝑀𝑃𝑎 and 𝑓𝑠𝑢 = 660 𝑀𝑃𝑎. Using these tensile properties, and letting the true
stress-strain curve given in Eq. (54) cross through the proposed three stress-strain
points with Eqs. (55), the material parameters in Eq. (54) are determined to be 𝐸1 =
2.532, 𝐸2 = 5.727 and 𝛿 = 0.569. The comparison between the resulted true stress-
plastic strain curve and the experimental result by Blandford et al (2007) is shown in
Fig. 9.
2400

2000

1600
Stress (MPa)

1200
This paper
800 Blandford et al…
400

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 1.1 1.2 1.3 1.4 1.5 1.6 1.7
Plastic Strain
Fig. 9 True stress-strain relationship of stainless steel AISI 304L

As per ASME/BPVC Section II, Part D (ASME, 2017a), the specified tensile properties
for AISI 304L are 𝑓𝑠𝑦 = 170 𝑀𝑃𝑎 and 𝑓𝑠𝑢 = 485 𝑀𝑃𝑎. Accordingly, the corresponding
material parameters for the true stress-plastic strain relationship in Eq. (54) are 𝐸1 =
−45.549, 𝐸2 = 55.778 and 𝛿 = 0.941.

6.8 Strain rate effect on the dynamic stress of ductile metals


Alves and Jones (1999) performed monotonic tests on notched specimens of mild
steel at strain rates from 4E-3 S-1 to nearly 2000 s-1. Results from these tests show that

27
the strain rate does not have a significant effect on the engineering strain at the peak
engineering stress and fracture strain.

6.8.1 Strain rate effect on mild steel CSA G40.20/21 300W


Based on the investigation of relevant experimental results, the effect of the strain rate
on the dynamic stress is most prominent at the initial yield stress. This effect decreases
with the increase of the plastic strain. In this paper, the general form of the DIF of the
dynamic stress in terms of plastic strain rate and plastic strain is proposed to take the
form

𝑝 𝑝 𝑝 𝛿1 𝑝 𝛿2
̃𝑠 (𝜀̇ 𝑠 ,𝜀𝑠 )
𝜎 𝜀̇ 𝜀
𝑟𝑠 (𝜀̇𝑠𝑝 , 𝜀𝑠𝑝 ) = 𝑝 = 1 + (𝜀̇ 𝑝𝑠 ) [1 − (𝜀𝑝𝑠 ) ] (56)
𝜎𝑠 (𝜀𝑠 ) 𝑠0 𝑠0

𝑝 𝑝
where 𝜀̇𝑠0 , 𝛿1 , 𝜀𝑠0 , and 𝛿2 are material constants. For the mild steel CSA G40.20/21
𝑝 𝑝
300W, the material constants are determined to be 𝜀̇𝑠0 = 44.409 , 𝛿1 = 0.267 , 𝜀𝑠0 =
0.55 , and 𝛿2 = 0.4 to fit the experimental results by Chen (2010).

6.8.2 Strain rate effect on stainless steel AISI 304L


For the AISI 304L stainless steel, Stout and Follansbee (1986) and Lichtenfeld et al
(2006) experimentally studied the effect of strain rate on the uniaxial stress-strain
relationship of the AISI 304L stainless steel. Eq. (56) can be used to define the DIF of
the dynamic stress of stainless steels. For the stainless steel AISI 304L, the material
𝑝
constants are determined to be 𝜀̇𝑠0 = 24,000 , 𝛿1 = 0.20 , 𝜀𝑠0 = 0.44 , and 𝛿2 = 0.82 to
fit the experimental results by Lichtenfeld et al (2006).

6.9 Uniaxial plastic stress-strain relationship of welds


6.9.1 Static true stress-strain relationship
Weld metals do not exhibit a noticeable yield stress plateau in the engineering stress-
strain curve before strain hardening. Hertele et al (2014) proposed a uniaxial true
stress-strain relationship for welds based on the Ramberg-Osgood expression:

𝜎 𝜎 𝑛
𝜀 = 𝐸 + 0.002 (𝑓 ) (57)
𝑠 0.2

where 𝜀 and 𝜎 are uniaxial true strain and stress, respectively; 𝑛 = 2.4 +
2.9(𝑓0.2 /𝑓𝑢 )/(1 − 0.95𝑓0.2 /𝑓𝑢 ) , in which 𝑓0.2 is the 0.2% proof stress and 𝑓𝑢 the peak
engineering stress. In the absence of the 0.2% proof stress, this ratio can be taken as
𝑓0.2 /𝑓𝑢 = 1/[1.07 + (350/𝑓𝑢 )4.8 ].

6.9.2 Effect of strain rates


Barsom and Rolfe (1999) proposed the effect of plastic strain rates on the dynamic
increase factor (DIF) of the true stress of the weld metal using the electrode E70T-4:

𝑝 0.04
̃ (𝜀 )
𝜎 𝜀̇
𝑟𝑠 (𝜀̇𝑠𝑝 , 𝜀𝑠𝑝 ) = 𝜎𝑠 (𝜀𝑠 ) = (1 + 0.2
𝑠
) (58)
𝑠 𝑠

28
where 𝜎𝑠 (𝜀𝑠 ) is the static true stress and 𝜎̃𝑠 (𝜀𝑠 ) the dynamic true stress. This DIF is
assumed to be applicable to general weld metals due to scarcity of experimental results.

7. Numerical example of dynamic analysis of a transport package


7.1 Description of the FE model
An example of dynamic analysis of a transport package is presented in this paper,
which consists of an outer packaging using two impact limiters and an inner packaging
using a dry storage container (DSC) containing irradiated fuels. Materials used in the
transport package include stainless steel shell, crushable foam and connection
components of the impact limiters as well as mild steel liners, weld metal, concrete,
reinforcing steel, and fuel modules of the dry storage container. In order to satisfy the
IAEA Regulations SSR-6 under accident conditions, the dynamic analysis for a 9-meter
accident corner drop condition is performed for the transport package.
The transport package is modeled using the software ABAQUS (2017) with explicit
dynamic analysis. Concrete, crushable foams, closure weld, steel liners and flanges,
and fuel module are meshed using solid elements with reduced integration (C3D8R).
Reinforcing bars (rebars) are meshed using truss elements (T3D2). Impact limiter shells
are meshed using shell elements with reduced integration (S4R). Impact limiter
connectors are meshed using truss elements (T3D2). Total weight of the fuel module,
dry storage container, and impact limiters are approximately 100 tonnes. The transport
package impacts onto a rigid surface with an initial velocity of 13.3 m/s (9-meter free
drop). Fig. 10 shows the FE model under the 9-meter accident drop condition.

Fig. 10 Transport package subjected to 9-meter accident corner drop

29
Mechanical contacts among all package components are established. The contact
setup is assigned to the interfaces between the concrete-steel liners, cask-lid flanges,
steel liners-impact limiter shells, impact limiter shell-shell, crushable foam-impact limiter
shell, rigid plane-impact limiter shell, and self-contact of materials. Rebars are
embedded in the host concrete by using the embedded constraint with the rebar steel
being assigned with a smeared stress-strain relationship.
Strain-rate-dependent material constitutive models in terms of true stress and train are
specifically established based on the equations presented in this paper. Fracture criteria
are incorporated in the progressive failure rule in conjunction with the material
constitutive models. Attention is brought to the tensile stress-crack opening relationship
of the concrete constitutive model. This relationship is converted into the tensile stress-
crack strain relationship based on the mesh size.

7.2 Numerical stability of explicit dynamic analysis


In brittle materials such as concrete that undergo large deformations, an element may
become largely distorted which may cause premature termination of the analysis or
unrealistic analysis results. The drastic stress drop following the material descending
branch contributes to the numerical instabilities. There is an inherent limit to how much
deformation a Lagrangian mesh can accommodate without some sort of mesh
smoothing or remeshing taking place. Based on the experience of performing the
dynamic analysis, the following strategies can be used to overcome the numerical
instabilities in an explicit dynamic analysis:
 Avoid to use the full “stable” time increment that the software calculates, which is
based on element size, Young’s modulus, and density. A sufficiently smaller time
increment may be needed to suit the descending branch of the true stress-strain
relationship of brittle materials. It is recommended to use a time increment size of
a small fraction of the software calculated time increment. A time increment
scaling factor of a value between 0.01 and 0.1 may be used.
 Avoid fully-integrated solid elements which tend to be less stable in situations
involving large deformation or distortion. Mesh the materials to generate
elements with the shape as regular as possible.
 Use viscous hourglass control for brittle materials with a small displacement
hourglass control factor. This factor may be taken a value between 0.01 and 0.1.
 Use appropriate strain rate factor to filter or dampen nonphysical high-frequency
oscillations in the calculated plastic strain rate increments associated with strain-
rate-dependent material behavior.
 Use a “softened” contact relationship for the interfaces between brittle materials
and steel liners or shells. The initial contact stiffness can be set as small as
possible but not to be too small to cause significant material penetration. For

30
example, the initial stiffness multiplication factor to scale the default contact
stiffness may be given a value of 0.0001.

7.3 Results and discussion


7.3.1 Maximum strain rates of the materials
The maximum strain rates of the materials used in the transport package subjected to
the 9-meter accident corner drop condition are listed below:

Table 1. Maximum strain rates


Material Maximum strain rate (s-1) Yield stress increase
Stainless steel shells 520 46%
927
Foam FR-3718 71%
(compression)
Steel liners/flanges 126 132%
Closure weld 19 20%
Concrete in compression 332 214%
Concrete in tension 369 990%
Reinforcing steel 85 62%

It can be seen from Table 1 that the effect of strain rate is significant. Constitutive
models for the materials used in the transport packages must include the effect of strain
rate.

7.3.2 Damages to the materials


The maximum equivalent plastic strains experienced in the materials of the transport
package subjected to a 9-meter corner drop condition are listed below:

Table 2. Maximum (equivalent) plastic strain or damage index


Maximum plastic Max. plastic strain /
Material
strain fracture strain
Stainless steel shells 0.72 1.0
0.96
Foam FR-3718 Not applicable
(compression)
Steel liners/flanges 0.012 0.045
Closure weld 0.0034 0.0030
Reinforcing steel 0.038 0.083
Damage index /
Material Damage index
limit value
Concrete in compression 0.78 0.98
Concrete in tension 0.80 1.0

31
Based on the fracture criteria that are stress-triaxiality-dependent as given in Eq. (49),
no fracture occurs in the DSC containment boundary components including steel liners,
flanges, and welds, although local and minor yielding occurs in the DSC steel liner.
Fracture does occur in the stainless steel shells of the impact limiter directly in contact
with the rigid surface. Cracks occur in the lid concrete, but the shielding capacity is
judged not to be significantly affected because the crack widths are small and cracks
will be closed after the dynamic impact since the rebars in the lid concrete are still in the
elastic regime, except one reinforcing tie is slightly yielded.

8. Conclusions
This study proposed a framework for dynamic analysis of radioactive material
transport packages under accident drop scenarios to meet the nuclear licensing
requirements. The true stress and strain, effect of strain rate, triaxial stress states,
realistic constitutive models, and damage and fracture criteria of materials are
incorporated into the dynamic analysis framework to address large deformation of
materials occurring in a very short impact duration. Critical material parameters
associated with material constitutive models, true stress-strain relationships, and effect
of strain rate are studied and determinations of these parameters are proposed. Based
on the study, the following conclusions can be drawn:
 Materials in the transport packages are subjected to large deformation under the
accident 9-meter drop scenario. Therefore, true stress and strain should be used
to realistically reveal the actual strain energy distribution among the package
components.
 Effect of strain rates is significant even with impact limiters being attached to the
radioactive material containers. The strain rates of the materials subjected to the
accident condition can cause the increase of the yield stresses by 50% to 160 %.
 Stress conditions in the transport packages are very complex. Therefore,
material constitutive models should be established for highly triaxial stress states.
Parameters associated with the material constitutive models, true stress-strain
relationships, and effect of strain rate are appropriately calibrated and
determined based on experimental results or theoretically derived based on
observed material behavior in elementary stress states.
 Stress triaxiality-dependent damage and fracture criteria of materials are
incorporated into the dynamic analysis framework to limit the plastic deformation
to material fracture and to safeguard the containment boundary of the DSC that
contains radioactive materials.
 Ignoring effect of strain rate and using engineering stress-strain overestimate the
strain energy absorption in the impact limiters and underestimate the strain
energy distribution to the critical radioactive material containers.
The proposed framework is sufficient to capture the realistic mechanical behavior of
materials and the structural behavior of the transport packages and radioactive material
containers. It can be conveniently implemented in analysis software for the dynamic
analysis of the transport packages under various accident impact scenarios.

32
Acknowledgements

The authors acknowledge Ontario Power Generation Inc. for providing resources to
write this paper. The opinions and conclusions expressed in this paper are of the
authors and do not necessarily reflect the official position of Ontario Power Generation
Inc.

References

ABAQUS version 2017. ABAQUS User’s Manual, Dassault Systems.

M. Alves and N. Jones, 1999. Influence of Hydrostatic Stress on Failure of Axisymmetric


Notched Specimens. Journal o f the Mechanics and Physics o f Solids, Vo. 47, pp.643-
667, 1999.

N. Arioglu, Z. C. Girgin, and E. Ariglu, 2006. Evaluation of ratio between splitting tensile
strength and compressive strength for concrete up to 120 MPa and its application in
strength criterion. ACI Materials Journal, Vol. 103, No. 1, pp.18-24.

F. Aslani and R. Jowkarmeimandi, 2012. Stress-strain model for concrete under cyclic
loading. Magazine of Concrete Research, Vol. 64, No. 8, pp. 673-685.

ASME, 2017a. Boiler and pressure vessel code 2017, Section II: Materials, Part D:
Properties. American Society of Mechanical Engineers.

ASME, 2017b. Boiler and pressure vessel code 2017, Section VIII: Rules for
Construction of Pressure Vessels, Division 2: Alternative Rules. American Society of
Mechanical Engineers.

J. M. Barsom and S. T. Rolfe, 1999. Fracture and fatigue control in structures.


Application for fracture mechanics. Philadelphia. American Society for Testing and
Materials.

Z. P. Bazant and E. Becq-Giraudon, 2002. Statistical prediction of fracture parameters


of concrete and implications for choice of testing standard. Cement and Concrete
Research, Vol. 32, pp. 529-556.

B. Binici, 2005. An Analytical Model for Stress-Strain Behavior of Confined Concrete.


Engineering Structures, Vol. 27, No. 7, pp. 1040-1051.

R. K. Blandford, D. K. Morton, S. D. Snow, and T. E. Rahl, 2007. Tensile stress-strain


results for 304L and 316L stainless steel plate at temperature. Proceedings of 2007
ASME Pressure Vessels and Piping Division Conference, San Antonio, Texas, U.S.A.

A. Brara and J. R. Klepaczko, 2007. Fracture energy of concrete at high loading rates in
tension. International Journal of Impact Engineering, Vol. 34, pp. 424-435.

33
E. Cadoni, G. Solomos and C. Albertini, 2013. Concrete behavior in direct tension tests
at high strain rates. Magazine of Concrete Research, Vol. 65, No. 11, pp. 660-672.

D. C. Candappa, J. G. Sanjayan, and S. Setunge, 2001. Complete Triaxial Stress-Strain


Curves of High-Strength Concrete. ASCE Journal of Materials in Civil Engineering, Vol.
13, No. 3, pp. 209-215.

J. Chen, 2010. An experimental study of strain rate effects on mild steel. Master Thesis,
Carleton University, Ottawa, Ontario, Canada.

J. B. Choi and R. S. Lakes, 1992. Non-linear properties of polymer cellular material with
a negative Poisson’s ratio. Journal of Materials Science. Vol. 27, pp. 4678-4684.

CSA, 2012. Design and assessment of buildings subjected to blast loads. Standard
CSA S850-12. Canadian Standard Association (CSA).

EPRI, 2005a. Spent fuel transportation applications: fuel rod failure evaluation under
simulated cask side drop conditions. Report No. 1009929, Electric Power Research
Institute, Palo Alto, California, USA.

EPRI, 2005b. Spent fuel transportation applications: global forces acting on spent fuel
rods and deformation patterns resulting from transportation accidents. Report No.
1011817, Electric Power Research Institute, Palo Alto, California, USA.

Eurocode 3, 2006. Design of steel structures – Part 1-4: General rules – Supplementary
rules for stainless steels. EN 1993-1-4, European Standad.

E. Ferretti, 2004. On Poisson’s ratio and volumetric strain in concrete. International


Journal of Fracture, Vol. 126, pp. 49-55.

Federation Internationale du Beton (FIB), 2012. CEB-FIP Model Code 2010 – final draft.
FIB Bulletin 65, Vol. 1, Lausanne, Switzerland.

M. F. Ferrotto, L. C., and F. D. Trapani, 2018. FE modeling of partially steel-jacketed


(PSJ) RC columns using CDP model. Computers and Concrete, Vol. 22, No. 2, pp. 143-
152.

P. Grassl, K. Lundgren, and K. Gylltoft, 2002. Concrete in compression: a plasticity


theory with a novel hardening law. International Journal of Solids and Structures. Vol.
39, No. 20, pp. 5205-5223.

S. Hertele, N. O’Dowd, K. V. Minnerbruggen, M. Verstraete, and W. de Waele, 2014.


Fracture mechanics analysis of heterogeneous welds: validation of a weld
homogenisation approach. Procedia Materials Science, Vol. 3, pp. 1322-1329.

34
T. T. C. Hsu and Y. L. Mo, 2010. Unified Theory of Concrete Structures. New York:
John Wiley & Sons.

IAEA Safety Standards, 2012. Regulations for the Safe Transport of Radioactive
Material, Specific Safety Requirements No. SSR-6, International Atomic Energy
Agency.

I. Imran and S. J. Pantazopoulou, 1996. Experimental study of plain concrete under


triaxial stress. ACI Materials Journal, Vol. 93, No. 6, pp. 589-601.

J. F. Jiang and Y. F. Wu, 2012. Identification of material parameters for Drucker-Prager


plasticity model for FRP confined circular concrete column. International Journal of
Solids and Structures, Vol. 49, No. 3-4, pp. 445-456.

T. Jankowiak and T. Lodygowski, 2005. Identification of parameters of concrete damage


plasticity constitutive model. Foundation of civil and environmental engineering, Vol. 6,
pp. 53–69.

E. Kasparek, U. Zencker, R. Scheidemann, H. Volzke, and K. Muller, 2011. Numerical


and experimental studies of polyurethane foam under impact loading. Computational
Materials Science, Vol. 50, No. 4, pp. 1353-1358.

E. M. Kasparek, H. Volzke, R. Scheidemann, and U. Zencker, 2012. Numerical and


experimental investigations of polyurethane foam for use as cask impact limiter in
accidental drop scenarios. Proceedings of Annual Waste Management Symposium
2012, Phoenix AZ, pp. 688-696.

M. Korgesaar, H. Remes, and J. Romanoff, 2014. Size dependent response of large


shell elements under in-plane tensile loading. International Journal of Solids and
Structures, Vol. 51, pp. 3752-3761.

B. Kraus, R. Das, and B. Banerjee, 2013. Anisotropy and variability in polyurethane


foams: experiments and modeling. Technical report, Callaghan Innovation Research
Limited, Auckland, New Zealand, http://imechanica.org/node/14071.

J. Lee and G. L. Fenves, 1998. Plastic-damage model for cyclic loading concrete
structures. Journal of Engineering Mechanics, Vol. 124, No. 8, pp. 892-900.

J. A. Lichtenfeld, M. Mataya, and C. J. Van Tyne, 2006. Effect of strain rate on stress-
strain behavior of alloy 309 and 304L austenitic stainless steel. Metallurgical and
Material Transactions A, Vol. 37A, No. 1, pp. 147-161.

J. C. Lim, T. Ozbakkaloglu, A. Gholampour, T. Bennett, and R. Sadeghi, 2016. Finite-


element modeling of actively confined normal-strength and high-strength concrete under
uniaxial, biaxial, and triaxial compression. ASCE Journal of structural engineering, Vol.
142, No. 11, 04016113.

35
Q. Liu, and G. Subhash, 2004. A phenomenological constitutive model for foams under
large deformations. Polymer Engineering and Science, Vol. 44, No. 3, pp. 463-473.

R. Lo Frano, G. Pugliese, and M. Nasta, 2014. Structural performance of an IP2


package in free drop test conditions: Numerical and experimental evaluations. Nuclear
Engineering and Design, http://dx.doi.org/10.1016/j.nucengdes.2014.09.034.

W. Y. Lu, 2014. Mechanical characterization of rigid polyurethane foams. Technical


Report, Sandia National Laboratories (SNL-CA), Livermore, California, USA.

J. Lubliner, J. Oliver, S. Oller, and E. Onate, 1989. A plastic-damage model for


concrete. International Journal of Solids and Structures, Vol. 25, No. 3, pp. 299-326.

B. Lukic, 2018. Advanced measuring techniques for characterisation of the concrete


dynamic tensile response. Mechanics of materials [physics.class-ph]. University
Grenoble Alpes.

A. K. Maji, H. L. Schreyer, S. Donald, Q. Zuo, and D. Satpathi, 1995. Mechanical


Properties of polyurethane-foam impact limiters. ASCE Journal of Engineering
Mechanics, Vol. 121, No. 4, pp. 528-540.

R. Malm, 2006. Shear cracks in concrete structures subjected to in-plane stresses.


TRITA-BKN, Bulletin 88, Royal Institute of Technology, Division of Structural Design
and Bridges, Stockholm, Sweden.

R. Malm, 2009. Predicting shear type crack initiation and growth in concrete with non-
linear finite element method. TRITA-BKN, Bulletin 97, Royal Institute of Technology,
Division of Structural Design and Bridges, Stockholm, Sweden.

L. J. Malvar, 1998. Review of static and dynamic properties of steel reinforcing bars.
ACI Materials Journal, Vol. 95, No. 5, pp. 609-616.

L. J. Malvar and C. A. Ross, 1998. Review of strain rate effect for concrete in tension.
ACI Materials Journal, Vol. 95, No. 6, pp. 735-739.

W. S. A Nana, T. T. Bui, A. Limam, and S. Abouri, 2017. Experimental and numerical


modeling of shear behavior of full-scale RC slabs under concentrated loads. Structures,
Vol. 10, pp.96-116.

K. Pack and D. Mohr, 2017. Combined necking & fracture model to predict ductile
failure with shell finite elements. Engineering Fracture Mechanics, Vol. 182, pp. 32-51.

C. Poinard, Y. Malecot, and L. Daudeville, 2010. Damage of concrete in a very high


stress state: experimental investigation. Materials and Structures, vol. 43, pp. 15-29.

36
L. Qiao, U. Zencker, H. Völzke, F. Wille, and A. Musolff, 2012. Validation of numerical
simulation models for transport and storage casks using drop test results. Proceedings
of the Eleventh International Conference on Computational Structures Technology,
B.H.V. Topping, (Editor), Civil-Comp Press, Stirlingshire, Scotland.

K. J. R. Rasmussen, 2003. Full-range stress–strain curves for stainless steel alloys.


Journal of Constructional Steel Research, Vol. 59, pp. 47–61.

H. W. Reindhard, 1984. Fracture mechanics of an elastic softening material like


concrete. Heron, Vol. 29, No.2, pp. 1-42.

H. W. Reindhard, H. A. W. Cornelissen, and D. A. Hordijk, 1986. Tensile tests and


failure analysis of concrete. ASCE Journal of Structural Engineering, Vol. 112, No. 11,
pp. 2462-2477.

A. K. Samani, and M. M. Attard, 2012. A stress-strain model for uniaxial and confined
concrete under compression. Engineering Structures, Vol. 41, pp. 335-349.

A. K. Samani, and M. M. Attard, 2014. Lateral strain model for concrete under
compression. ACI Structural Journal, Vol. 111, No. 2, pp. 441-452.

B. Song, W. Y. Lu, and C. J. Syn, 2009. The effects of strain rate, density and
temperature on the mechanical properties of polyurethane diisocyanate (PMDI)-based
rigid polyurethane foams during compression. Journal of Materials Science, Vol. 44, No.
2, pp. 351-357.

P. Soroushian and K.–B. Choi, 1987. Steel mechanical properties at different strain
rates. ASCE Journal of Structural Engineering, Vol. 113, No. 4, pp. 863-872.

J. L. A. de Oliveira e Sousa and R. Gettu, 2006. Determining the Tensile Stress-Crack


Opening Curve of Concrete by Inverse Analysis. ASCE Journal of Engineering
Mechanics, Vol. 132, No. 2, pp. 141-148.

J. L. Sprung, D. J. Ammerman, N. L. Breivik, R. J. Dukart, F. L. Kanipe, J. A. Koski, G.


S. Mills, K. S. Neuhauser, H. D. Radloff, R. F. Weiner, and H. R. Yoshimura, 2000.
Reexamination of spent fuel shipment risk estimates. NUREG/CR-6672 (prepared for
U.S. NRC), Vol. 1, SAND2000-0234, Sandia National Laboratories, Albuquerque, New
Mexico, U.S.A.

M. G. Stout and P. S. Follansbee, 1986. Strain rate sensitivity, strain hardening, and
yield behavior of 304L stainless steel. ASME Journal of Engineering Materials and
Technology, Vol. 108, pp. 344-353.

J. G. M. Van Mier, 1986. Multiaxial strain softening of concrete. Part I: fracture. Material
s and structures, Col. 19, No. 111, pp. 179-190.

37
J. Weerheijm and I. Vegt, 2010. The dynamic fracture energy of concrete: Review of
test methods and data comparison. Proceedings of the 7th International Conference on
Fracture Mechanics of Concrete and Concrete Structures, Jeju Island, Korea.

S. Yan, X. Zhao, and A. Wu, 2018. Ductile Fracture Simulation of Constructional Steels
Based on Yield-to-Fracture Stress–Strain Relationship and Micromechanism-Based
Fracture Criterion. ASCE Journal of Structural Engineering, vol. 144, No. 3, 04018004.

38

You might also like