You are on page 1of 30

Article

International Journal of Damage


Mechanics
A simple implementation of 2022, Vol. 31(10) 1562–1591
! The Author(s) 2022
localizing gradient damage Article reuse guidelines:
sagepub.com/journals-permissions
model in Abaqus DOI: 10.1177/10567895221109622
journals.sagepub.com/home/ijd

Yi Zhang1 , Yanjie Xu1, Yihe Wang2 and


Leong Hien Poh1

Abstract
With the localizing gradient enhancement, a damage model for quasi-brittle materials is able to achieve
regularized softening responses, with localized damage profiles corresponding to the development of
macroscopic cracks, to resolve the numerical spurious effects induced by the conventional gradient
enhancement. The typical implementation strategy for a gradient enhanced model is to solve the
system of governing equations simultaneously. Focusing on the finite element (FE) package Abaqus, a
user element subroutine is required to define the finite elements with additional degrees of freedom for
the nonlocal field. Moreover, with user elements, additional effort is required to visualize the numerical
results. To an inexperienced engineer/researcher, these requirements can be challenging. In this paper, a
simple implementation of the localizing gradient damage model is elaborated. By utilizing the in-built
coupled thermo-mechanical elements in Abaqus, the user only needs to define the material constitutive
laws, as well as the sensitivity terms with respect to the field variables. Post-processing of results can be
done directly in Abaqus. The applicability and ease of implementation are demonstrated via several
examples, including those that utilize the Abaqus features of element deletion, contact between surfaces,
as well as the incorporation of cohesive elements. Sample files can be downloaded from https://github.
com/leonghien/Localizing-Gradient-Damage-with-UMAT-UMATHT

Keywords
Localizing gradient damage, coupled thermo-mechanical analysis, quasi-brittle fracture, Abaqus

Introduction
The nonlocal gradient enhancement is widely used to regularize the strain softening behavior of
continuum damage models. Typically, a higher order balance equation is incorporated in the

1
Department of Civil and Environmental Engineering, National University of Singapore, Singapore, Singapore
2
Ocean Academy, Zhejiang University, Zhejiang, China

Corresponding author:
Leong Hien Poh, Department of Civil and Environmental Engineering, National University of Singapore, Block E1A, Level 07,
Room 03, 1 Engineering Drive 2, (S)117576, Singapore.
Email: leonghien@nus.edu.sg
Zhang et al. 1563

formulation to capture the interactions between micro-cracks, with a length scale parameter char-
acterizing the size of fracture process zone. The conventional gradient enhancement, however,
suffers from a spurious damage growth phenomenon. This has led to the development of a so-
called localizing gradient enhancement, to give sharp damage profiles corresponding to the devel-
opment of macroscopic cracks. Numerically, the system of governing equations defining equilibri-
um and higher order balance in a gradient enhancement is commonly solved simultaneously, using
finite elements (FE) with additional degrees of freedom (dof). The need for additional dofs com-
plicates its implementation into the standard FE framework of commercial FE software. In this
paper, we present a simple implementation of the localizing gradient enhancement in the commer-
cial FE software Abaqus, by utilizing its in-built elements for coupled thermo-mechanical analyses.
The conventional nonlocal enhancement was first proposed as an integral-type formulation
(Pijaudier-Cabot and Bazant, 1987), where the influence of underlying micro-processes is captured
via the spatial averaging of the driving field within the damage process zone. While the nonlocal
integral approach has been adopted successfully for many problems (Havlásek et al., 2016; Jirásek
and Desmorat, 2019; Pereira et al., 2017), its numerical implementation in a standard FE frame-
work can be difficult. The gradient approach forms another big class of nonlocal enhancement,
which can be understood as a reformulation of the integral enhancement (Peerlings et al., 1996,
1998, 2001). Here, a higher order balance equation governs the evolution of the nonlocal field, to
drive the damage process. This additional differential equation can be assimilated into a standard
FE framework more easily than the integral approach, and is thus widely adopted to regularize the
softening response (Ahmed et al., 2021a, 2021b; Mobasher et al., 2017; Schreter et al., 2018; Zreid
and Kaliske, 2018). While the conventional nonlocal enhancement is able to regularize the structural
response well, both the integral and gradient approaches induce numerical artefacts such as a wrong
damage initiation and propagation behavior (Geers et al., 1998; Giry et al., 2011; Nguyen, 2011;
Simone et al., 2004), or a spurious boundary effect resulting in an inability to reproduce the size
effect phenomenon of quasi-brittle fracture (Gregoire et al., 2013; Havlásek et al., 2016; Marzec and
Bobi nski, 2019; Wosatko et al., 2018). This has motivated the development of localizing gradient
enhancement, where the energetic contribution from active micro-process zone decreases with
damage evolution, to result in a slightly modified higher order balance equation compared to the
conventional expression (Poh and Sun, 2017). The localizing gradient enhancement was shown to
work well for different problems, e.g. quasi-static fracture of quasi-brittle materials (Castillo et al.,
2021; Sarkar et al., 2019; Shedbale et al., 2021; Tong et al., 2021b), size effects in quasi-brittle
fracture (Negi et al., 2021; Zhang et al., 2021), thermo-mechanical fracture (Sarkar et al., 2020),
continuous-discontinuous modelling (Sarkar et al., 2021), dynamic fracture (Wang et al., 2019) and
ductile fracture (Xu et al., 2020; Xu and Poh, 2019).
The gradient enhancement is typically solved as a system of equations in the weak form. While
the numerical framework can be derived easily with only C0 continuity requirement, its implemen-
tation in a commercial FE software can be complicated, especially to an inexperienced researcher/
engineer. Focusing specifically on the FE software Abaqus, the common strategy is to write a user
element (UEL) subroutine to define additional dofs, as well as to determine the elemental contri-
bution to the overall tangent stiffness matrix and residual force vector. The writing of UEL can be
cumbersome for practical applications, especially when a new subroutine has to be written whenever
there is a change in shape function and/or element type. Moreover, the visualization of UEL
analyses in Abaqus has to be done indirectly, e.g. with an overlaying mesh of standard elements
generated during the pre-processing stage (Sarkar et al., 2019; Tong et al., 2021a, 2021b), or having
a script to extract the relevant result data and updating an odb file generated from dummy Abaqus
in-built elements (Roth et al., 2012).
1564 International Journal of Damage Mechanics 31(10)

In the literature, there are attempts to simplify the implementation of the conventional gradient
enhancement in Abaqus. The general approach is to utilize the in-built coupled temperature-
displacement elements in Abaqus. Since the higher-order term in the balance equation of conven-
tional gradient enhancement involves only a Laplacian term, it is analogous to the heat equation
assuming Fourier’s law, where the heat flux is linearly dependent on the temperature gradient. In
Abaqus, such thermo-mechanical analyses can be implemented conveniently using the UMAT and
HETVAL subroutines to define the mechanical constitutive laws and heat flux relation respectively.
A proper definition of parameters in HETVAL subroutine will recover the balance equation in con-
ventional gradient enhancement, with the temperature field corresponding to the nonlocal variable. This
strategy has been utilized successfully for damage modelling (Azinpour et al., 2018), and plasticity-
driven ductile fracture (Papadioti et al., 2019; Seupel et al., 2018). Although not necessary, the more
general UMATHT subroutine was utilized by Ostwald et al. (2019) to define the linear (thermal) flux
relation, in their implementation of the conventional gradient enhancement for a large deformation
isotropic hyperelastic damage model. It is also worth mentioning that the same strategy has been
adopted by researchers to implement the phase field fracture models, another large class of regulariza-
tion method. Since the higher order term in the phase field model also involves only a Laplacian term,
the parameters in HETVAL subroutine can similarly be redefined to recover the phase field balance
equation (e.g. Azinpour et al., 2018; Navidtehrani et al., 2021; Wu and Huang, 2020).
The simplified Abaqus implementation of conventional gradient enhancement/phase field
damage models has motivated us to consider a similar strategy for the localizing gradient enhance-
ment. Since the higher order term in the localizing gradient enhancement involves the divergence of
a nonlinear flux relation, the HETVAL subroutine that is applicable for (linear) Fourier’s law
cannot be utilized here. Departing from the common strategy by other researchers, the more general
UMATHT subroutine is adopted in this paper for defining the nonlinear flux relation, to recover
the localizing gradient enhancement balance equation. The paper is structured as follows. A sum-
mary of the localizing gradient damage model is provided in the next section. In Section
‘Implementation using UMAT-UMATHT subroutines’, the analogy between a coupled thermo-
mechanical analysis and the localizing gradient enhancement is elaborated, and implementation
details via the UMAT-UMATHT subroutines provided. In Section ‘Validation and discussions with
compact tension specimen’, a validation of the proposed simplified implementation approach is
demonstrated, and the ease of analysis in conjunction with other Abaqus in-built features illustrated
in Section ‘Numerical examples considering other Abaqus in-built features’. Finally, the model
extension to 3D is provided by taking advantage of the simplified implementation approach, and
its good performance in capturing non-planar mixed mode fracture behavior demonstrated in
Section ‘3D modelling of prismatic skew edge notched beam in torsion’.

Summary of the localizing gradient damage enhancement


In this section, a brief summary of the localizing gradient damage model is provided. Detailed
discussions can be found in Poh and Sun (2017).

Governing equations and constitutive relations


:
For a given strain rate e , the internal power statement is postulated as
Z
:
Pint ¼ r e~_ þ ~n re~_ ÞdV
ðr : e þ~ (1)
R
Zhang et al. 1565

where r denotes the standard Cauchy stress. Additional contributions are associated with a micro-
morphic strain e~_ and its gradient characterizing the deformation in the (micro) damage process
~ and ~n.
zone, together with the respective conjugate stresses r
The external power is given by
Z
:
Pext ¼ ðt  u þ fe~_ÞdA (2)
@R

where t and f are respectively the traction and higher-order traction acting at the boundary.
Equating Pint ¼ Pext and applying the divergence theorem, we obtain a system of governing
equations given by the standard equilibrium condition, as well as a higher order balance equation

r  r ¼ 0; ~ ¼ r~n
r (3)

The tractions at the external boundary with outward normal n are also determined as t ¼ r  n
and f ¼ ~ n  n.
The free energy density comprises the elastic strain energy of the undamaged material, augment-
ed by a coupling contribution between macro and micro deformations, which are respectively
characterized by an equivalent macro strain measure e and a scalar micro strain e~. Moreover, an
additional contribution arises from the interactions between active micro processes characterized
via r~e , to give

1 1~ 1 ~2
w ¼ ð1  xÞe : C : e þ hðe  e~Þ2 þ ghl r~
e  r~
e (4)
2 2 2

where C is the forth-order elastic stiffness tensor, 0  x  1 an isotropic damage variable capturing
the extent of material degradation, h~ a coupling modulus accounting for the micro-macro inter-
actions, and l a length scale parameter characterizing the initial size of damage process zone.
Note that an interaction function g is appended to the energetic contribution from the interac-
tions between active micro cracks. Since the failure process typically involves an initially diffused
region of micro cracks before transiting into localized macro cracks, the interaction function has the
following characteristics

1; x ¼ 0
g ¼ gðxÞ ¼ (5)
R; x ! 1

where R  0 describes a residual interaction between micro-processes. In this paper, the interaction
function proposed in Poh and Sun (2017) is adopted
ð1  RÞexpðgxÞ þ R  expðgÞ
g¼ (6)
1  expðgÞ

where g is a material parameter describing the rate which the interaction domain decreases with
damage. Discussions on the influence of R and g can be found in Sarkar et al. (2019). In the
following, R ¼ 0:005 and g ¼ 5 are utilized for all examples.
1566 International Journal of Damage Mechanics 31(10)

From the Coleman-Noll procedure, the constitutive relations are determined as

@w ~  e~Þ @e
r¼ ¼ ð1  xÞC : e þ hðe (7a)
@e @e

@w ~

r ¼ hð~
e  eÞ (7b)
@~
e

~n ¼ @w ¼ ghl
~ 2 r~
e (7c)
@r~e

Following Sarkar et al. (2019), the coupling modulus h~ is assumed to be much smaller than the
Young’s modulus, such that the standard constitutive relation for a strain driven damage model is
recovered, i.e.

r ¼ ð1  xÞC : e (8)

The microforce balance governing the micro-macro interactions is also obtained by substituting
equations (7b) and (7c) into equation (3b), to give
 
e~  e ¼ r  gl2 r~
e (9)

This balance equation defines the localizing gradient enhancement. Note that the conventional
gradient enhancement is recovered for the special case of a constant interaction size, i.e. g ¼ 1.

Equivalent strain measure and damage evolution


To complete the model definition, an equivalent strain measure (e) that can adequately characterize
the underlying deformation state is required, together with a suitable damage evolution law.
A detailed discussion on some common equivalent strain measures for the mixed mode fracture
of quasi-brittle materials can be found in Shedbale et al. (2021).
In this paper, the modified von Mises equivalent strain measure (de Vree et al., 1995) is adopted,
defined as
" #0:5
k1 1 ðk  1Þ2 2 12k
e¼ I1 þ I þ J2 (10)
2kð1  2Þ 2k ð1  2Þ2 1 ð1 þ Þ2

where  is Poisson’s ratio, and k is a material parameter characterizing the ratio between
compressive and tensile strengths. The first and second strain invariants are respectively
given as

1
I1 ¼ trðeÞ; J2 ¼ edev : edev (11)
2
Zhang et al. 1567

In addition, we adopt a commonly utilized exponential damage evolution law, given by


(
0 if j < j0
x ¼ xðjÞ ¼ j0    (12)
1 1  a þ aexp bðj  j0 Þ if j  j0
j

where a and b are material parameters controlling the residual stress and the rate of softening
respectively, j0 is a damage initiation threshold, and j is a history variable following the Kuhn-
Tucker relations
j_  0; e~  j  0; j_ ðe~  jÞ ¼ 0 (13)

In this paper, we adopt a ¼ 1 for all examples shown.


The numerical implementation details of this model in Abaqus using UEL subroutines, as well as
the subsequent visualization method, can be found in Sarkar et al. (2019). Below, a simpler
approach is demonstrated.

Implementation using UMAT-UMATHT subroutines


The localizing gradient enhancement can be implemented easily in Abaqus, by taking advantage of
its capability to solve thermo-mechanical problems with user defined relations, via the UMAT and
UMATHT subroutines. This is elaborated below.

Heat equation
The general heat equation is given by

qch_ ¼ r  f þ r (14)

where q is the material density, c is the specific heat capacity, h is the temperature field, f is a heat
flux vector, and r is the spatial heat source.
If the Fourier’s law is assumed, the heat flux vector will be linearly dependent on the temperature
gradient, i.e.

f ¼ krh (15)

where k is a thermal conductivity constant. Substituting this into equation (14) will result in a
Laplacian term in the heat equation, which bears similarity to the conventional gradient enhance-
ment (where g ¼ 1 in equation (9)), as well as the balance equation in phase field damage models.
This resemblance was exploited in Azinpour et al. (2018); Navidtehrani et al. (2021); Wu and Huang
(2020); Seupel et al. (2018) in their implementations of the conventional gradient enhancement/
phase field damage models, using the HETVAL subroutine.
A more general heat flux constitutive relation can be considered in equation (15) by defining the
thermal conductivity as a function of temperature, i.e. k ¼ kðhÞ. Substituting this into equation (14)
results in the following balance equation

qch_ ¼ r  ðkðhÞrhÞ þ r (16)


1568 International Journal of Damage Mechanics 31(10)

Since a nonlinear constitutive relation for heat flux is involved, the HETVAL subroutine cannot
be utilized for the analysis. Instead, the more general UMATHT subroutine has to be adopted,
which is the approach in this paper. In Abaqus, the time integration is done using a backward-
difference scheme, i.e.

1
h_ tþDt ¼ ðhtþDt  ht Þ (17)
Dt

The general strategy is to redefine the terms in the heat equation to (almost) recover the gradient
enhancement balance equation, such that the temperature field ðhÞ can be interpreted as the non-
local field e~. First, the spatial heat supply term (r) in UMATHT is defined via the RPL1 variable in
UMAT as

rpl ¼ e  h (18)

Next, the thermal conductivity parameter is made dependent on the temperature field, defined as

kðhÞ ¼ gðhÞl2 (19)

where g ¼ gðhÞ is the interaction function in equation (6), and l the length scale parameter in
equation (9).
Substituting equations (18) and (19) into equation (16), the modified heat equation is obtained as

qch_ ¼ r  ðgl2 rhÞ þ e  h (20)

Comparing between equations (9) and (20), it is easily observed that the steady state heat equa-
tion recovers the localizing gradient balance equation, with h as an analogous field to e~. For the
transient analysis, a steady state approximation can be achieved numerically by setting the term ‘qc’
to a small value. In the original heat diffusion problem, parameter c denotes the specific heat
capacity with units of J/kg/K. In the modified diffusion equation resembling the localizing gradient
governing equation, the combined term qc has unit of s, which represent a characteristic time. For
all problems, the default loading time for quasi-static analysis (T ¼ 1) is utilized. Discussions of the
loading rate effect on the numerical solution are provided in Appendix 1. To avoid confusion, we
shall assume q ¼ 1 and simply state the small value of c in the following. The influence of parameter
c is discussed later in Section ‘Validation and discussions with compact tension specimen’.

Numerical implementation in Abaqus


The localizing gradient enhanced damage model can now be implemented easily in Abaqus, using
the UMAT and UMATHT subroutines to respectively specify the mechanical and thermal consti-
tutive relations, as well as the tangent contributions, for a Gauss point underlying an element. The
assembly of contributions from each Gauss point at the element level, which is then used to build
the global system of equations, is done internally by Abaqus. Compared to the typical implemen-
tation of gradient enhanced models using the user element subroutine UEL, the adoption of
UMAT-UMATHT subroutines greatly simplifies the implementation requirements. For ease of
reference to the localizing gradient damage model summarized in Section ‘Summary of the
Zhang et al. 1569

localizing gradient damage enhancement’, the temperature field h in the thermo-mechanical element
will be written as its analogous field e~, in the elaborations below.
A flowchart of the UMAT-UMATHT implementation is depicted in Figure 1. Focusing first on
UMAT, the stress constitutive relation has to be specified, which referring to equation (8) gives

def
STRESS ¼ r ¼ ð1  xÞC : e (21)

The corresponding tangent contributions with respective to the field variables are obtained as

def dr
DDSDDE ¼ ¼ ð1  xÞC (22a)
de

def dr @x
DDSDDT ¼ ¼ C:e (22b)
d~
e @j

Additionally, Abaqus expects information on the mechanical heat generation and the associated
sensitivity terms to be provided in UMAT. The mechanical heat source is represented by the var-
iable RPL, defined as

def
RPL ¼ rpl ¼ e  e~ (23)

Figure 1. Flowchart of the UMAT-UMATHT implementation, where K is the tangent stiffness matrix, da is the
vector of incremental nodal degrees of freedom, and R is the residual force vector. Variables to be computed in
UMAT are defined in equations (21) to (24). Solution dependent variables to be computed in UMAT and stored in the
STATEV array, as per equations (6), (25), are passed into UMATHT. Finally, the variables to be defined in UMATHT
are provided in equations (26) to (29).
1570 International Journal of Damage Mechanics 31(10)

The sensitivity of rpl with respect to the field variables are correspondingly

def drpl @e
DRPLDE ¼ ¼ (24a)
de @e

def drpl
DRPLDT ¼ ¼ 1 (24b)
d~e

Finally, the sensitivity of f with respect to the field variables, to be defined later in UMATHT in
equation (29), will involve some terms to be passed from UMAT. At each iterative step, the inter-
action function g defined in equation (6) and the following variables are computed in UMAT

@g gðR  1Þ
¼ expðgxÞ (25a)
@x 1  expðgÞ

@x j0   j0   
¼ ab exp bðj  j0 Þ þ 2 1  a þ aexp bðj  j0 Þ (25b)
@j j j

These solution dependent variables are stored in the array STATEV of UMAT, which will be
passed to UMATHT. The variables to be specified and computed in UMAT are thus completed
at this juncture.
Referring to the flowchart in Figure 1, we now focus on the definitions required in UMATHT
based on the transient heat equation in (20). Within this subroutine, the internal energy per unit
mass U and the heat flux vector f have to be specified, as well as their sensitivities with respect to the
field variables.
The internal energy is defined as a linear function of e~ with a constant specific heat capacity c

def
U ¼ U ¼ c~
e (26)

The corresponding sensitivities with respect to the field variables are thus

def dU
DUDT ¼ ¼c (27a)
d~
e

def dU
DUDG ¼ ¼0 (27b)

dðr~

The heat flux vector is defined as


def
FLUX ¼ f ¼ gl2 r~
e (28)

with the corresponding sensitivities as

def df @g @x
DFDT ¼ ¼ l2 r~
e (29a)
d~
e @x @j
Zhang et al. 1571

def df
DFDG ¼ ¼ gl2 I (29b)
dðr~ eÞ

@g
Recall that the solution dependent variables g, @x and @x
@j are obtained from UMAT in equations (6)
and (25), and I is the identity matrix.
A summary of the variables to be returned in the UMAT and UMATHT subroutines is listed in
Table 1.

Validation and discussions with compact tension specimen


In this section, we consider the mode I fracture of a compact tension specimen (CTS) with a pre-
existing sharp crack, and benchmark the results from the UMAT-UMATHT implementation
against those reported in Poh and Sun (2017) using UEL.
The CTS of dimension d  d is depicted in Figure 2. Here, we adopt the same material parameters
as Poh and Sun (2017), where E ¼ 1 GPa,  ¼ 0:2, k ¼ 10, j0 ¼ 0:002, b ¼ 9 and l ¼ 0:2d. Plane
strain condition is assumed.

Table 1. Summary of the variables to be defined in UMAT-UMATHT subroutines.

Variables Notations Definitions

UMAT
STRESS r ð1  xÞC : e
DDSDDE dr=de ð1  xÞC
DDSDDT dr=d~e ð@x=@jÞC : e
RPL rpl e  ~e
DRPLDE drpl =de @e=@e
DRPLDT drpl =d~e –1
UMATHT
U U c~e
DUDT dU=d~e c
DUDG dU=dðr~e Þ 0
FLUX f gl2 r~e
DFDT df=d~e l2 r~e ð@g=@xÞð@x=@jÞ
DFDG df=dðr~e Þ gl2 I

Figure 2. Geometry and boundary conditions of the CTS. Only half of the specimen geometry is considered due to
symmetry about y ¼ 0.
1572 International Journal of Damage Mechanics 31(10)

Plane strain analysis


We first consider the CPE8T plane strain element (8-node element with biquadratic displacement
and bilinear e~). As highlighted in equation (20), the transient term has to be made negligible
(achieved by having a small value for parameter c) in order to recover the localizing gradient
enhancement balance equation.
The structural responses with different specific heat capacity values of c ¼ 0.005, 0.002 and 0.0005
are depicted in Figure 3(a). It is easily observed that a better match with the reference UEL solution
is achieved, when the value for c decreases. The damage profiles at the final loading stage
(u ¼ 0:006d) are also provided in Figure 4(a). Again, the damage profiles compare well with the
reference UEL solution, to give a localized damage band ahead of the pre-existing crack tip. As
expected, the smallest value for c produces a damage profile that is almost identical to that of UEL
model. In general, as long as parameter c is small, the results will not be significantly affected by the
loading rate. It is shown in Figure 3(a) and Figure 4(a) that c ¼ 0.005 is already a small enough
value, to induce a marginal rate effect. The rate effect becomes negligible for c < 0.002. For generic

(a) (b)

Figure 3. Structural responses using (a) CPE8T elements with different c values and (b) mesh convergence analysis
using CPE3T elements with c ¼ 0.002. The UEL reference solution is obtained from Poh and Sun (2017).

(a) (b) (c) (d)

Figure 4. Damage profiles at u ¼ 0:006d from (a) UEL reference solution, UMAT-UMATHT implementation with
(b) c ¼ 0.005, (c) c ¼ 0.002 and (d) c ¼ 0.0005.
Zhang et al. 1573

problems, without performing preliminary sensitivity studies of parameter c, we recommend adopt-


ing c < 0.002.
The UMAT-UMATHT approach, however, requires a smaller time step to ensure the stability of
transient heat analysis. Abaqus determines the time increment according to the element size, heat
velocity and the thermal parameters (q; c; k). Referring to Figure 5(a) on the total number of
iterations required to solve the same CTS problem with different values of c, we can infer that
more computational resources is required with smaller c values. Hence, a general balance between
the accuracy (compared to UEL reference solution) and computational time has to be made by
the user.
It is highlighted that the larger number of total iterations required in the UMAT-UMATHT
implementation, is due to the smaller time steps, instead of the need for more iterations within a
time step. To illustrate this, the problem is solved using 2500 CPE8T elements, c ¼ 0.002, with a
fixed time step. The number of iterations per time increment is depicted in Figure 5(b). It is easily
observed that the solution at each time increment typically converges within one iterative step. This
is in contrast to the numerical solution of a typical phase field damage model, where the non-convex
potential energy function with respect to the field variables typically induces convergence issues with
the standard Newton’s method. For example, similar tensile problems solved using different phase
field damage models in Wu and Huang (2020); Navidtehrani et al. (2021) are shown to require a
large number of iterations per increment, especially at the initial part of strain softening. Therewith,
it was found that the quasi-Newton method is more robust for the implementation of phase field
damage models (Kristensen and Martınez-Pa~ neda, 2020; Wu et al., 2020), but unfortunately, is not
available in Abaqus for use with its in-built coupled temperature-displacement elements. We note
that this numerical convergence issue is not encountered with the localizing gradient enhancement.
It is also highlighted that the visualization and post-processing of simulation results, using the
Abaqus in-built elements for thermo-mechanical analyses, can be done easily with Abaqus/Viewer.
In contrast, these tasks cannot be done directly when the UEL subroutine is used. Instead, addi-
tional efforts are required to establish a workaround, e.g. to include a dummy mesh of standard
elements overlaying the user elements, or to regenerate an odb file by transferring the relevant
output to the standard element mesh.

(a) (b)

Figure 5. (a) Total number of iterations in UMAT-UMATHT with different c values, with 2500 CPE8T elements. For
the same number of 8-noded elements, the UEL subroutine requires 42 iterations to solve the problem and (b)
Number of iterations at each time increment using UMAT-UMATHT (c ¼ 0.002) with 2500 elements, with the force-
displacement graph superimposed. Note that the number of iterations corresponds to the global implicit scheme.
1574 International Journal of Damage Mechanics 31(10)

Finally, a customary check for mesh convergence is done using CPE3T triangular elements
(linear displacement and linear e~). We deliberately adopt a different element type here, compared
to the earlier results, to further demonstrate the ease of implementation with UMAT-UMATHT.2
As the change in element type (and shape functions) are accounted internally for by Abaqus, the
same UMAT-UMATHT subroutines that were utilized earlier can be used here as well. The force-
displacement graphs with different number of elements are shown in Figure 3(b), where mesh
convergence is demonstrated.

Numerical examples considering other Abaqus in-built features


In this section, the UMAT-UMATHT implementation is adopted for more complex problems,
involving the use of other in-built features in Abaqus. For the following examples, a conservative
value of c ¼ 0.0005 is adopted.

Example incorporating element deletion


To capture structural failure process, highly damaged elements can be removed via the deletion
option in Abaqus. This feature is useful in cases with complex crack paths, or in situations with
element distortions at large deformation. In this section, the damage variable x is utilized as the user
specified material state variable which controls the element deletion flag. A material point having a
damage value beyond a threshold limit will be marked as deleted. An element will be removed from
the mesh when all of its underlying material points are marked as deleted.
Here, we will demonstrate the use of this element deletion feature together with the localizing
gradient enhancement, by considering the mixed mode fracture of a Double-Edge-Notched (DEN)
specimen as reported in Nooru-Mohamed (1992). The DEN specimen with dimension of 100 mm 
100 mm  50 mm (denoted as DEN100 hereafter) is illustrated in Figure 6. The loading path 6a
(un ¼ us) is considered, where un donates the average relative vertical displacement between points
M-M0 and N-N0 , and us indicating the relative lateral deformation between points S-S0 .

Figure 6. Geometry and boundary conditions of DEN100, where a proportional loading of un ¼ us is applied to the
loading frame (Nooru-Mohamed 1992).
Zhang et al. 1575

The problem is solved using CPS8T elements (8-node plane stress element with biquadratic
displacement and bilinear e~), with very fine mesh size of 0.3 mm at the critical regions along the
crack propagation path, as illustrated in Figure 7(a). The experimentally determined mechanical
properties for DEM specimens: E ¼ 30 GPa and  ¼ 0:2, are directly adopted to our model. Other
numerical parameters are calibrated as k ¼ 15.05, l ¼ 2 mm; j0 ¼ 9:8  105 and b ¼ 0:01. A mate-
rial point is flagged as being deleted once x > 0:99. With these settings, the structural response is
depicted in Figure 7(b), which matches well with the experimental data. The eroded elements,
corresponding to the crack profiles at the final loading step in Figure 7(b), is depicted in
Figure 8(a), which match well with the experimental observations (see Figure 8(b)). This demon-
strates the viability of the element deletion option to be used in combination with the localizing
gradient enhancement.

(a) (b)

Figure 7. (a) FE discretization of DEN100 and (b) Numerical structural response compared against experimental
data in Nooru-Mohamed (1992).

(a) (b)

Figure 8. (a) Crack (eroded elements) patterns of DEN100 using localizing gradient enhancement (a magnification
factor of 20 is utilized), with element deletion at x > 0:99 and (b) Crack paths of DEN100 observed in the
experiment (Nooru-Mohamed 1992).
1576 International Journal of Damage Mechanics 31(10)

Example incorporating contact between surfaces


In this section, we refer to the compressive tests of 3D printed rock-like specimens with pre-existing
flaws, as reported in Sharafisafa et al. (2018). Each specimen is placed between a pair of loading
platens, and subjected to a displacement controlled applied force at 4 mm/s. Here, we consider
single-flawed specimens with two different inclination angles (a ¼ 45 and 90 ). Each disc has an
in plane diameter of 40 mm and a thickness of 15 mm, as well as a pre-existing flaw of
8 mm  0:6 mm. The geometrical details are depicted in Figure 9. More information on the exper-
imental set up can be found in Sharafisafa et al. (2018, 2019); Aliabadian et al. (2021).
The following mechanical properties: E ¼ 1:35 GPa and  ¼ 0:285, are experimentally deter-
mined for the 3D printed material, and are adopted in the numerical analyses. The length scale
parameter is assumed as l ¼ 0:5 mm. Damage related parameters of k ¼ 10, j0 ¼ 1:75  103 and
b ¼ 7, are calibrated.
The specimen is meshed using CPS8T plane stress elements (8-node element with biquadratic
displacement and bilinear e~). Since the steel platens are much stiffer than the specimen, each of them
is modelled as a 2-node rigid link (R2D2). At the interface between specimen and platen, the
Abaqus feature of hard contact in normal behavior is selected. To prevent disc rotation, the top-
most and bottom-most nodes of the specimen are constrained horizontally.
A mesh convergence check is first carried out for the specimen with a ¼ 45 considering element
sizes of 0.06 mm and 0.08 mm in the critical regions, as depicted in Figure 10(a). The structural
responses in Figure 10(b) show that mesh convergence is achieved with element size of 0.08 mm. For
the following analyses in this section, an element size of 0.08 mm will be used in the critical regions.
 
The structural responses of the single-flawed discs with a ¼ 45 and 90 are provided in Figure 11
(a). Compared to the experimental data, the numerical results have higher elastic stiffness, and are
more brittle. We note that Chen and Shen (2021), the same group that had conducted the exper-
iment, has the same observations from their numerical simulations using smoothed particle hydro-
dynamics, as compared in Figure 11(a-i). It was postulated in Chen and Shen (2021) that the
inconsistency may be due to the inadequacy of a 2D model in capturing the damage develop-
ment/stiffness degradation along the thickness direction. Notwithstanding these differences, the
peak forces obtained numerically match well with the experimental data for the two specimens.
For each case, the fracture process initiates from the tips at both ends of the flaw, followed by an

(a) (b)

Figure 9. Geometry details for the compressive test of a rock-like specimen, with the pre-existing flaw orientating
at (a) a ¼ 45 and (b) a ¼ 90 .
Zhang et al. 1577

(a) (b)

Figure 10. (a) Meshing details and (b) mesh convergence for the specimen with a ¼ 45 , with element sizes of
0.06 mm and 0.08 mm in the critical regions.

(a) (b) (c)

Figure 11. (a) Comparison between experimental data (Sharafisafa et al., 2018) and numerical structural responses
obtained using localizing gradient enhancement, and smoothed particle hydrodynamics (Chen and Shen, 2021); failure
patterns from (b) localizing gradient damage model and (c) experimental DIC images (Aliabadian et al., 2021), for the
specimens with (i) a ¼ 45 and (ii) a ¼ 90 .

upward/downward propagation to the specimen surface, before the disc splits open in tension. This
failure process is well captured by the localizing gradient damage model, as seen from the final
damage profiles for both specimens in Figure 11(b), which compare well with the experimentally
obtained digital image correlation (DIC) images in Figure 11(c).
1578 International Journal of Damage Mechanics 31(10)

Example incorporating cohesive elements


In this section, the meso-scale failure process of fiber-reinforced composites is considered. Referring
specifically to carbon-epoxy systems, it was observed experimentally in Parıs et al. (2007) that the
failure process starts with the debonding of fiber-matrix interfaces at semi-angles of 60 –70 , leading
to the initiation and propagation of macro cracks into the matrix. A schematic representation of the
failure process is shown in Figure 12(a). In the literature, this failure process has been modelled with
different numerical methods, e.g. discontinuous Galerkin/cohesive zone model (Nguyen, 2014),
augmented Lagrangian functional cohesive zone model (Labanda et al., 2018) and coupled cohesive
phase field model (Li et al., 2020).
In the following, we consider a unit cell comprising a single fiber embedded within the matrix
material, with the unit cell subjected to a tensile load. This problem was solved in Nguyen et al.
(2017) assuming plane stress condition, with the geometry and boundary conditions provided in
Figure 12(b). Therewith, the fiber is elastic, and the matrix modelled using a kinematically enriched
constitutive formulation that smeared a cohesive fracture over an element. At the fiber-matrix
interface, cohesive elements following a bilinear traction-separation law are adopted.
In our model, the Abaqus in-built cohesive elements are adopted to describe the fiber-matrix
interfacial response. The fiber is elastic, and the matrix described using the localizing gradient
damage model. Elastic properties of both materials are adopted from Nguyen et al. (2017),
where Ef ¼ 40 GPa, f ¼ 0:33; Em ¼ 4 GPa and m ¼ 0:4. Additional parameters in the localizing
gradient damage model are calibrated as j0 ¼ 7:8  103 , l ¼ 0.02 mm and b ¼ 0:01. Modelling
parameters of interface are summarized in Table 2. Details on the traction-separation law adopted
are provided in Appendix 2. The unit cell discretization is shown briefly in Figure 13, with 14693
CPS4T elements (4-node plane stress element with bilinear displacement and bilinear e~) for the
matrix material, 1296 CPS4 elements (4-node plane stress element with bilinear displacement) for
the fiber, and 264 COH2D4 interface elements (4-node two-dimensional cohesive element) for the
fiber-matrix interface.
The reaction force versus displacement curve in Figure 14(a) shows a good agreement with the
reference results from Nguyen et al. (2017). The failure pattern of the composite at u ¼ 10 mm is

(a) (b)

Figure 12. (a) Experimentally observed failure process of fiber-reinforced composites (Parıs et al., 2007): (i) initial
debonding of the fiber-matrix interface, (ii) debonding at semi-angles of 60 –70 and (iii) kinking of interface cracks
into the matrix and (b) Geometry and boundary conditions of a single fiber embedded within matrix material.
Zhang et al. 1579

Table 2. Material properties of interface.

Property Value

Knn 1014 (N/m)


Kss 1014 (N/m)
ft;n 10 (MPa)
ft;s 10 (MPa)
Gc 50 (N/m)

Figure 13. Mesh detail of the fiber-matrix composite: the matrix is discretized using CPS4 elements, the fiber
discretized using CPS4T elements, and the interface discretized using COH2D4 interface elements.

(a) (b)

Figure 14. (a) Numerical structural response compared to reference results and (b) Interfacial crack and damage
(x > 0:95) propagation in the matrix material, at u ¼ 10 mm (a magnification factor of 2 is utilized).

provided in Figure 14(b), with the (damaged) cohesive elements removed from the deformed shape
for better visualizing of the debonding phenomenon. The general failure process is captured cor-
rectly, with a kinked crack initiating from the interfacial crack and propagating into the matrix. The

semi-debonding angle of 66 obtained numerically is also within the range of 60 –70 suggested in
Parıs et al. (2007).
1580 International Journal of Damage Mechanics 31(10)

To better illustrate the failure process, the evolution of stress component along x axis is plotted
in Figure 15, corresponding to the loading steps indicated in Figure 14. The fracture development at
each loading step is briefly described as follows:

• Step a: The start of a nonlinear response. The failure process is expected to first initiate and
propagate at the fiber-matrix interface, according to the stress distribution in Figure 15(a).
• Step b: Debonded surfaces of matrix and fiber are created, with kinks developing at the tips of
cohesive cracks.
• Step c: Interfacial debonding continues. One of the kinks starts to dominate the failure process
(here, the one on the right), with the crack starting to extend into the matrix material.
• Step d: The kinked crack at the cohesive crack tip propagates upwards/downwards through the
matrix material.

3D modelling of prismatic skew edge notched beam in torsion


The prevention of failure behavior in quasi-brittle materials is a major concern in the engineering
field. Numerically, whenever appropriate, quasi-brittle fracture problems are modelled and solved
either in 1D or 2D, to reduce computational cost. In many cases, however, the loading conditions
may induce complex damage evolution processes to give non-planar crack surfaces and twisted
crack trajectories. These problems require the use of a 3D model. This motivates the current section,
where the localizing gradient damage model is employed for the 3D mixed mode fracture, by taking

(a) (b)

(c) (d)

Figure 15. Profiles of the stress component along x-axis for the fiber-matrix composite, corresponding to the four
loading stages (a)–(d) of Figure 14. A magnification factor of 5 is utilized.
Zhang et al. 1581

advantage of the simplified implementation approach. From the UMAT-UMATHT (2D) subrou-
tines, the extension to 3D requires only the revision of mechanical constitutive laws, as well as the
sensitivity terms with respect to the field variables.

3D implementation of localizing gradient damage model


In this section, an extension of the gradient enhancement from 2D to 3D is considered. The same
notations and definitions of the UMAT-UMATHT variables follow those of Section ‘Numerical
implementation in Abaqus’. Modifications required in the subroutines for 3D implementation are
highlighted below.
Focusing first on the mechanical constitutive behavior defined in UMAT, the stress and strain
tensors have six components for 3D problems in Voigt notation. According to the element types
adopted, the change in dimensions of stress and strain arrays is also accounted for internally by
Abaqus. Thus, the stress constitutive law and the corresponding sensitivities with respect to the field
variables, which were defined earlier in 2D, can be utilized for 3D implementation by redefining the
elastic stiffness matrix, such that

def
STRESS ¼ r ¼ ð1  xÞC : e (30a)

def dr
DDSDDE ¼ ¼ ð1  xÞC (30b)
de

def dr @x
DDSDDT ¼ ¼ C:e (30c)
d~
e @j

where
2 3
1   0 0 0
6  1  7
6 0 0 0 7
6   1 7
6 0 0 0 7
6 1  2 7
E 6 0 0 0 0 0 7
C¼ 6 2 7
ð1 þ Þð1  2Þ 6 1  2 7
6 0 0 7
6 0 0 0 7
6 2 7
4 1  2 5
0 0 0 0 0
2

Further, the mechanical heat generation RPL and the corresponding tangent contribution with
respect to the strain field should be revised accordingly, i.e.

def
RPL ¼ rpl ¼ e  e~ (31a)

def drpl @e
DRPLDE ¼ ¼ (31b)
de @e

where the equivalent strain measure e in equation (10) should be computed based on 3D condition.
1582 International Journal of Damage Mechanics 31(10)

Note that the UMATHT subroutine, which defines the constitutive thermal behavior, can be
adopted directly for 3D analyses without any modifications.

Experimental setup and observations


We consider the prismatic skew edge notched beam with square cross section, as reported in
Jefferson et al. (2004). Geometry details are depicted in Figure 16. In the experiment, the horizon-
tally positioned plane concrete beam is clamped with collars at both ends. Three of the four steel
arms are constrained in the vertical direction, while the last one is subjected to a downward con-
centrated load. The clamping frames are designed to make perfect contact with the concrete speci-
men to ensure transferring of the eccentric load to the specimen, which leads to a torsional moment
aligned with the axis of the beam. Figure 17(a) illustrates the final damage of the prismatic beam
under the torsional loading, where a complex fracture behavior is observed. The fracture surfaces
are spatially twisted and warped, which initiate from the notch front and intersect the bottom
surfaces with a kind of S-shaped curve.

Numerical results and discussions


Due to the complex fracture surface that wraps around the beam’s longitudinal axis, the
problem has to be solved in 3D. To apply the torsional loading, one of the right steel arms is
vertically supported in the middle, the left two arms are fixed at a single point. A downward
displacement u is imposed on the last steel arm, as depicted in Figure 18. The CMOD is defined
as norm of the relative distances between points A and B along x and z directions. To reduce
computational cost, only the critical region highlighted in Figure 18 is modelled using the localizing
gradient enhancement. The standard elastic model is adopted for the remaining parts of the speci-
men, as well as the steel collars. The experimentally obtained mechanical properties for concrete,
E ¼ 35 GPa and  ¼ 0:2 are adopted. The damage related parameters, i.e. k ¼ 10, j0 ¼ 7:2  103 ,
l ¼ 1 mm and b ¼ 120, are calibrated accordingly. A detailed discussion on the influence of the
parameter c in equation (20) can be found in Section ‘Plane strain analysis’. Here, c ¼ 0.002 is
utilized.

Figure 16. Geometry and boundary conditions for the notched prismatic torsion test (Jefferson et al., 2004).
Zhang et al. 1583

(a) (b)

Figure 17. Comparison of the fracture surfaces for the notched prismatic concrete beam under torsional loading
between (a) experimental observations (Jefferson et al., 2004) and (b) numerical results using localizing gradient
damage enhancement (considered as x > 0:95), from different views of (i) and (ii).

Figure 18. Boundary conditions and critical damage region in the numerical simulation for the prismatic skew edge
notched beam under torsional loading.

Considering two different discretization methods, the specimen is meshed using:

1. 462580 C3D8T elements (8-node brick element with trilinear displacement and e~) and
268200 C3D8 elements (8-node brick element with trilinear displacement), as depicted in
Figure 19(a);
1584 International Journal of Damage Mechanics 31(10)

(a) (b)

Figure 19. Mesh details of the prismatic skew edge notched beam: (a) the critical damage region is discretized using
C3D8T elements, the remaining parts discretized using C3D8 elements and (b) the critical damage region is dis-
cretized using C3D4T elements, the remaining parts discretized using C3D4 elements.

(a) (b)

Figure 20. (a) Comparison between experimental data (Jefferson et al., 2004) and numerical results with two
different mesh types using the localizing gradient damage model for the prismatic skew edge notched beam and
(b) Profiles of the displacement magnitude for the prismatic skew edge notched beam under torsional loading,
corresponding to Step c. A magnification factor of 50 is adopted.

2. 2646554 C3D4T elements (4-node tetrahedron element with linear displacement and e~) and
709144 C3D4 elements (4-node tetrahedron element with linear displacement), as depicted in
Figure 19(b).

A converged structural response is obtained (see Figure 20(a)). It is shown that the reaction force
versus CMOD curve matches well with the two sets of experimental data. The displacement mag-
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
nitude ( u2x þ u2y þ u2z ) of the specimen, corresponding to the final loading step indicated in
Figure 20(a), is visualized in Figure 20(b), where the twisting of the concrete beam around z-axis
is clearly observed, together with the relative opening and sliding of both notch edges.
Zhang et al. 1585

Figure 21. The growth of crack (considered as x > 0:95) in the prismatic skew edge notched beam, corresponding
to the three loading stages (a)-(c) indicated in Figure 20.

The crack profiles obtained using the localizing gradient damage model are compared against the
experimental results of Jefferson et al. (2004), as depicted in Figure 17. For ease of visualization,
elements with damage x  0:95 are removed, to reveal the fracture surface. As discussed, the
numerical spatially distorted fracture surfaces and curved crack paths are very similar to the exper-
imental results.
Finally, fracture evolution of the specimen corresponding to different loading stages of Figure 20
is examined in Figure 21. The results match well with the experimental descriptions, in which a
localized damage band initiates from the bottom edges of the notch, before propagating with a
rotational effect to intersect with the base to form a S-curve.
1586 International Journal of Damage Mechanics 31(10)

Conclusion
In this contribution, a simple implementation approach of the localizing gradient damage model in
the commercial FE software Abaqus is presented, by utilizing its in-built coupled temperature-
displacement elements. The balance equation governing the localizing gradient enhancement
involves the divergence of a nonlinear flux term. Accordingly, the UMAT and UMATHT subrou-
tines are adopted for defining the mechanical and thermal constitutive relations, as well as the
associated sensitivities with respect to the field variables. With the proper definitions of material
parameters, the temperature field can be made analogous to the nonlocal variable, such that the
heat equation almost recovers the balance equation of the gradient enhanced model. Compared to
the standard implementation method by writing a user element subroutine, this simplified approach
is more manageable for an average user.
To recover the balance equation of the localizing gradient enhancement, the transient term in the
heat equation described by UMATHT has to be set as small as possible. However, the time step
required for numerical stability is dependent on the magnitude of this transient term. The conflict-
ing requirements to closely approximate the gradient damage model and having lower computa-
tional time can be thus considered as a drawback. It is shown that convergence occurs rapidly
within a time-step, which helps to mitigate the effect of small time steps. Moreover, the simplified
implementation approach enables the direct visualization and post-processing of results in Abaqus/
Viewer – a feature that is not possible with the standard implementation method via user element
subroutine.
The versatility of the simplified implementation approach is demonstrated with the tensile load-
ing of a compact tension specimen, solved using different element types. Additionally, it is shown
that the localizing gradient damage model can be used in conjunction with other Abaqus features,
e.g. element deletion, contact between surfaces and the incorporation of cohesive elements, with
good numerical predictions obtained for the problems considered. Finally, it is demonstrated that
the simplified implementation approach can be extended to 3D analyses easily. Considering the
torsional loading of a prismatic concrete beam, the localizing gradient enhancement is shown to
capture well the complex 3D twisting fracture behavior.

Declaration of conflicting interests


The author(s) declared no potential conflicts of interest with respect to the research, authorship, and/or pub-
lication of this article.

Funding
The author(s) received no financial support for the research, authorship, and/or publication of this article.

ORCID iDs
Yi Zhang https://orcid.org/0000-0002-7429-6102
Leong Hien Poh https://orcid.org/0000-0002-7670-937X

Notes
1. In a standard thermo-mechanical model, this quantity captures the volumetric heat generation caused by
mechanical working.
2. Modified (M) or reduced integration (R) elements, such as CPE6MT and CPS8RT, are not recommended
for use with UMATHT when the thermal and mechanical fields are coupled through RPL – see Abaqus
requirements in Systemes (2014).
Zhang et al. 1587

References
Ahmed B, Voyiadjis GZ and Park T (2021a) Local and non-local damage model with extended stress decom-
position for concrete. International Journal of Damage Mechanics 30(8): 1149–1191.
Ahmed B, Voyiadjis GZ and Park T (2021b) A nonlocal damage model for concrete with three length scales.
Computational Mechanics 68(3): 461–486.
Aliabadian Z, Sharafisafa M, Tahmasebinia F, et al. (2021) Experimental and numerical investigations on
crack development in 3D printed rock-like specimens with pre-existing flaws. Engineering Fracture
Mechanics 241: 107396.
Azinpour E, Ferreira JPS, Parente MPL, et al. (2018) A simple and unified implementation of phase field and
gradient damage models. Advanced Modeling and Simulation in Engineering Sciences 5(1): 1–24.
Castillo D, Nguyen TH and Niiranen J (2021) Spatially random modulus and tensile strength: Contribution to
variability of strain, damage, and fracture in concrete. International Journal of Damage Mechanics 30(10):
1497–1523.
Chen Z and Shen L (2021) A modified smoothed particle hydrodynamics for modelling fluid-fracture interac-
tion at mesoscale. Computational Particle Mechanics 9(2): 277–297.
de Vree J, Brekelmans W and Van Gils M (1995) Comparison of nonlocal approaches in continuum damage
mechanics. Computers & Structures 55(4): 581–588.
Geers MGD, de Borst R, Brekelmans WAM, et al. (1998) Strain-based transient-gradient damage model for
failure analyses. Computer Methods in Applied Mechanics and Engineering 160(1–2): 133–153.
Giry C, Dufour F and Mazars J (2011) Stress-based nonlocal damage model. International Journal of Solids and
Structures 48(25–26): 3431–3443.
Gregoire D, Rojas-Solano LB and Pijaudier-Cabot G (2013) Failure and size effect for notched and unnotched
concrete beams. International Journal for Numerical and Analytical Methods in Geomechanics 37(10):
1434–1452.
Havlásek P, Grassl P and Jirásek M (2016) Analysis of size effect on strength of quasi-brittle materials using
integral-type nonlocal models. Engineering Fracture Mechanics 157: 72–85.
Jefferson AD, Barr BIG, Bennett T, et al. (2004) Three dimensional finite element simulations of fracture tests
using the Craft concrete model. Computers and Concrete 1(3): 261–284.
Jirásek M and Desmorat R (2019) Localization analysis of nonlocal models with damage-dependent nonlocal
interaction. International Journal of Solids and Structures 174-175: 1–17.
Kristensen PK and Martınez-Pa~ neda E (2020) Phase field fracture modelling using quasi-Newton methods and
a new adaptive step scheme. Theoretical and Applied Fracture Mechanics 107: 102446.
Labanda NA, Giusti SM and Luccioni BM (2018) Meso-Scale fracture simulation using an augmented
Lagrangian approach. International Journal of Damage Mechanics 27(1): 138–175.
Li G, Yin BB, Zhang LW, et al. (2020) Modeling microfracture evolution in heterogeneous composites: a
coupled cohesive phase-field model. Journal of the Mechanics and Physics of Solids 142: 103968.
Marzec I and Bobi nski J (2019) On some problems in determining tensile parameters of concrete model from
size effect tests. Polish Maritime Research 26(2): 115–125.
Mobasher ME, Berger-Vergiat L and Waisman H (2017) Non-local formulation for transport and damage in
porous media. Computer Methods in Applied Mechanics and Engineering 324: 654–688.
Navidtehrani Y, Beteg on C and Martınez-Pa~neda E (2021) A simple and robust Abaqus implementation of the
phase field fracture method. Applications in Engineering Science 6: 100050.
Negi A, Singh U and Kumar S (2021) Structural size effect in concrete using a micromorphic stress-based
localizing gradient damage model. Engineering Fracture Mechanics 243: 107511.
Nguyen GD (2011) A damage model with evolving nonlocal interactions. International Journal of Solids and
Structures 48(10): 1544–1559.
Nguyen VP (2014) Discontinuous Galerkin/extrinsic cohesive zone modeling: Implementation caveats and
applications in computational fracture mechanics. Engineering Fracture Mechanics 128: 37–68.
Nguyen VP, Nguyen GD, Nguyen CT, et al. (2017) Modelling complex cracks with finite elements: A kine-
matically enriched constitutive model. International Journal of Fracture 203(1-2): 21–39.
1588 International Journal of Damage Mechanics 31(10)

Nooru-Mohamed MB (1992) Mixed-mode fracture of concrete: An experimental approach. PhD thesis, Delft
University of Technology. Delft: The Netherlands.
Ostwald R, Kuhl E and Menzel A (2019) On the implementation of finite deformation gradient-enhanced
damage models. Computational Mechanics 64(3): 847–877.
Papadioti I, Aravas N, Lian J, et al. (2019) A strain-gradient isotropic elastoplastic damage model with J3
dependence. International Journal of Solids and Structures 174: 98–127.
Parıs F, Correa E and Mantic V (2007) Kinking of transversal interface cracks between fiber and matrix.
Journal of Applied Mechanics 74(4): 703–716.
Peerlings RHJ, de Borst R, Brekelmans WAM, et al. (1996) Gradient enhanced damage for quasi-brittle
materials. International Journal for Numerical Methods in Engineering 39(19): 3391–3403.
Peerlings RHJ, de Borst R, Brekelmans WAM, et al. (1998) Gradient-enhanced damage modelling of concrete
fracture. Mechanics of Cohesive-Frictional Materials 3(4): 323–342.,
Peerlings RHJ, Geers MGD, de Borst R, et al. (2001) A critical comparison of nonlocal and gradient-enhanced
softening continua. International Journal of Solids and Structures 38(44-45): 7723–7746.,
Pereira LF, Weerheijm J and Sluys LJ (2017) A numerical study on crack branching in quasi-brittle materials
with a new effective rate-dependent nonlocal damage model. Engineering Fracture Mechanics 182: 689–707.
Pijaudier-Cabot G and Bazant ZP (1987) Nonlocal damage theory. Journal of Engineering Mechanics 113(10):
1512–1533.
Poh LH and Sun G (2017) Localizing gradient damage model with decreasing interactions. International
Journal for Numerical Methods in Engineering 110(6): 503–522.
Roth S, Hütter G, Mühlich U, et al. (2012) Visualisation of user defined finite elements with ABAQUS/viewer.
GACM Rep 5: 7–14.
Sarkar S, Singh IV and Mishra BK (2020) A thermo-mechanical gradient enhanced damage method for frac-
ture. Computational Mechanics 66(6): 1399–1426.
Sarkar S, Singh IV and Mishra BK (2021) A simplified continuous–discontinuous approach to fracture based
on decoupled localizing gradient damage method. Computer Methods in Applied Mechanics and Engineering
383: 113893.
Sarkar S, Singh IV, Mishra BK, et al. (2019) A comparative study and ABAQUS implementation of conven-
tional and localizing gradient enhanced damage models. Finite Elements in Analysis and Design 160: 1–31.
Schreter M, Neuner M and Hofstetter G (2018) Evaluation of the implicit Gradient-Enhanced regularization of
a damage-plasticity rock model. Applied Sciences 8(6): 1004.
Seupel A, Hütter G and Kuna M (2018) An efficient FE-implementation of implicit gradient-enhanced damage
models to simulate ductile failure. Engineering Fracture Mechanics 199: 41–60.
Sharafisafa M, Shen L and Xu Q (2018) Characterisation of mechanical behaviour of 3D printed rock-like
material with digital image correlation. International Journal of Rock Mechanics and Mining Sciences 112:
122–138.
Sharafisafa M, Shen L, Zheng Y, et al. (2019) The effect of flaw filling material on the compressive behaviour
of 3D printed rock-like discs. International Journal of Rock Mechanics and Mining Sciences 117: 105–117.
Shedbale AS, Sun G and Poh LH (2021) A localizing gradient enhanced isotropic damage model with Ottosen
equivalent strain for the mixed-mode fracture of concrete. International Journal of Mechanical Sciences 199:
106410.
Simone A, Askes H and Sluys LJ (2004) Incorrect initiation and propagation of failure in non-local and
gradient-enhanced media. International Journal of Solids and Structures 41(2): 351–363.
Systemes D (2014) Abaqus User Subroutines Reference Guide, Version 6.14. Providence, RI, USA: Dassault
Systemes Simulia Corp..
Tong T, Hua G, Liu Z, et al. (2021a) Localizing gradient damage model coupled to extended microprestress-
solidification theory for long-term nonlinear time-dependent behaviors of concrete structures. Mechanics of
Materials 154: 103713.
Tong T, Yuan S, Wang J, et al. (2021b) The role of bond strength in structural behaviors of UHPC-NC
composite beams: Experimental investigation and finite element modeling. Composite Structures 255:
112914.
Zhang et al. 1589

Wang Z, Shedbale AS, Kumar S, et al. (2019) Localizing gradient damage model with micro inertia effect for
dynamic fracture. Computer Methods in Applied Mechanics and Engineering 355: 492–512.
Wosatko A, Pamin J and Winnicki A (2018) Numerical prediction of deterministic size effect in concrete bars
and beams. In: Computational modelling of concrete structures: Proceedings of the conference on compu-
tational modelling of concrete and concrete structures (EURO-C 2018), Bad Hofgastein, Austria. New
York: CRC Press, pp.447–456.
Wu JY and Huang Y (2020) Comprehensive implementations of phase-field damage models in Abaqus.
Theoretical and Applied Fracture Mechanics 106: 102440.
Wu JY, Huang Y and Nguyen VP (2020) On the BFGS monolithic algorithm for the unified phase field
damage theory. Computer Methods in Applied Mechanics and Engineering 360: 112704.
Xu Y, Biswas R and Poh LH (2020) Modelling of localized ductile fracture with volumetric locking-free
tetrahedral elements. International Journal for Numerical Methods in Engineering 121(12): 2626–2654.
Xu Y and Poh LH (2019) Localizing gradient-enhanced Rousselier model for ductile fracture. International
Journal for Numerical Methods in Engineering 119(9): 826–851.
Zhang Y, Shedbale AS, Gan Y, et al. (2021) Size effect analysis of quasi-brittle fracture with localizing gradient
damage model. International Journal of Damage Mechanics 30(7): 1012–1035.
Zreid I and Kaliske M (2018) A gradient enhanced plasticity–damage microplane model for concrete.
Computational Mechanics 62(5): 1239–1257.

Appendix 1. Loading rate effect


The transient heat equation (20) introduces a characteristic time qc. In general, the numerical
solution may thus be influenced by the loading rate. A quick illustration of this loading rate
effect is provided in Figure 22 below. For a given displacement applied at the loaded surface of
an element, the strain rate is characterized by the loading time. Figure 22 shows that the numerical
solutions can be affected by the loading time, as seen from the graphs of T ¼ 1 and T ¼ 2, both
having qc ¼ 0:002. For loading conditions with the same T=qc ratios, the numerical results are the
same.

(a) (b)

Figure 22. (a) A single CPE8T plane strain element (8-node element with biquadratic displacement and bilinear
~e ) in tension and (b) Numerical tensile response obtained with different combinations of total loading time (T) and
characteristic time qc.
1590 International Journal of Damage Mechanics 31(10)

Extending the above consideration to generic problems, our strategy is to adopt a sufficiently
small characteristic time (qc) against the default setting of T ¼ 1 in our implementation. The rec-
ommended value was shown to work well for the range of problems considered.

Appendix 2. Traction-separation law


As shown in Figure 23, a cohesive element consists of a pair of top and bottom faces. The consti-
tutive response of the relative motion of the two faces along the stack/thickness direction is defined
in terms of the local coordinates of the element. Since “zero-thickness” cohesive seams are consid-
ered in Section ‘Example incorporating cohesive elements’, the two faces are overlapping each other
at the initial state, as depicted in Figure 13.
The traction linearly increases with the relative separation until the damage initiation is reached.
The traction tensor consists of sn (along the local-2 direction) and ss (along the local-1 direction),
with the corresponding separations denoted by dn and ds respectively. The nominal strain compo-
nents are thus defined as

dn ds
en ¼ ; es ¼ (32)
T0 T0

Figure 23. Stack direction and the coordinate systems of the 2D cohesive elements.

Figure 24. A linear damage evolution law.


Zhang et al. 1591

where T0 ¼ 1 is the analytical thickness of the cohesive element. Considering an uncoupled elastic
behavior between the normal and shear components, we have

sn Knn 0 en
¼ (33)
ss 0 Kss es

It is assumed that the damage initiation of the interface material follows a quadratic nominal
strain criterion, given by
( )2  2
hen i es
þ ¼1 (34)
e0n e0s

where hi is the Macaulay bracket, e0n and e0s represent the corresponding nominal strain measure-
ment under the tensile strength ft;n and ft;s .
An effective displacement which combines both the normal and shear separations is defined as
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
dm ¼ hdn i2 þ d2s (35)

Damage of the composite material follows a linear evolution law as depicted in Figure 24, with d0m
denoting the threshold for damage initiation. In Abaqus, the damage evolution is specified by the
separation dfm at complete failure, or the dissipated energy Gc during the failure process.

You might also like