You are on page 1of 15

Construction and Building Materials 240 (2020) 117989

Contents lists available at ScienceDirect

Construction and Building Materials


journal homepage: www.elsevier.com/locate/conbuildmat

Yield stress criteria to assess the buildability of 3D concrete printing


R. Jayathilakage ⇑, P. Rajeev, J.G. Sanjayan
Swinburne University of Technology, Department of Civil and Construction Engineering, Melbourne, Australia

h i g h l i g h t s

 Modified Mohr-Coulomb failure criterion was implemented to predict the plastic collapse of 3D printed concrete structures.
 Time-dependent material models were experimentally developed for 3D printable concrete.
 Developed Mohr-Coulomb based plastic failure criterion is more accurate for 3D printable concrete material.
 The friction angle of the concrete mix has a considerable effect on the plastic failure height.

a r t i c l e i n f o a b s t r a c t

Article history: 3D concrete printing (3DCP) using the extrusion method is a commonly used additive manufacturing
Received 9 October 2019 technique to construct the structures layer by layer without using formwork. Therefore, the green (wet
Received in revised form 19 December 2019 or fresh) strength of the printable concrete determines the possible printable height of the structure
Accepted 30 December 2019
and the rate of printing to avoid collapse during the printing process. This paper aims to identify the
Available online 7 January 2020
buildability criteria based on the green strength of concrete and the effect of early age material properties
on the stability of printed structures. The commonly available strength-based failure models based on the
Keywords:
material yield stress and the vertical stresses induced were developed considering higher aspect ratios
Concrete 3D printing
Mohr-Coulomb
(height to width ratio). This may be not suitable for lower aspect ratios used in 3DCP layer geometries.
Yield stress Also, the frictional behaviour due to the applied vertical stresses should be considered in the material
Viscosity used for 3DCP applications. In this work, the Mohr-Coulomb based buildability criterion was developed
Finite difference method and validated with laboratory experiments and numerical simulations. The laboratory experiments were
carried out to establish the time-dependent material and rheological properties of 3D printable concrete
mixes. Further, 3D printing of hollow circular sections was carried out to study the failure modes and to
obtain the build height at the time of failure. The nonlinear finite difference model of the 3D printed sec-
tions was developed with the user-defined time-dependent material models to assess the accuracy of the
proposed buildability criterion. Additionally, the proposed buildability criterion was further validated
with the 3DCP experimental and numerical data presented in literature and all confirm that the proposed
criterion shows higher accuracy over existing methods in assessing the buildability.
Ó 2019 Elsevier Ltd. All rights reserved.

1. Introduction of cementitious material should have special rheological character-


istics in order to extrude the material through a given nozzle con-
3D concrete printing (3DCP) is a novel construction automation figuration without any defects and blockages. Also material should
technology to build using digital technologies and additive manu- have adequate early strength and stiffness to support the subse-
facturing concepts. One of the main advantages in 3DCP is the quent layers (i.e., buildability) without excessive layer deformation
reduced formwork cost and post-construction wastages. Also, the and failures [5,6]. Further, the printable concrete should have
3D printing technique gives the capability to construct geometri- enough workability and fluidity at the pumping and extrusion
cally complex shapes instead of rectilinear shapes easily and accu- stage.
rately [1–4]. The layer-wise extrusion method is the commonly The early-age material properties are significantly important in
used concrete printing method, in which concrete layers were the context of the buildability of fresh 3D printable concrete. The
extruded through a nozzle to build large scale objects. 3D printing rheology of fresh material, stress–strain behavior of the material
and the variation of the material properties with time are some
⇑ Corresponding author. important considerations discussed in previous researches when
E-mail address: rjayathilakage@swin.edu.au (R. Jayathilakage). considering the buildability and the strength of printed layers

https://doi.org/10.1016/j.conbuildmat.2019.117989
0950-0618/Ó 2019 Elsevier Ltd. All rights reserved.
2 R. Jayathilakage et al. / Construction and Building Materials 240 (2020) 117989

[7–9]. Hence, a better understanding of the material behavior and sy0 is the yield stress at t = 0 and tc is defined as a curve-fitting
its time-dependent variation in 3D printing is required to predict parameter (characteristic time) with experimental data.
the maximum build height for a particular mix. Also, it is useful  t

to predict the failure mode and enhance the material properties, sy ðtÞ ¼ Athix tc etc  1 þ sy0 ð1Þ
printing parameters such as print speed (nozzle moving speed)
Initially, the yield stress increase linearly with concrete age and
or the geometric properties to prevent failure [10]. The experimen-
smoothly transit to an exponential increase for later concrete ages
tal trial and error printing method can be used to understand the
according to the model given in Eq. (1).
failure of a particular structure. However, it would be time-
consuming to carry out trial and error experiments with material
1.2. Buildability
wastage. On the other hand, numerical simulation techniques can
be the most cost and time-effective method to fully understand
The printed layer should withstand its own weight initially.
the possible structural failure modes in extrusion-based 3D
While printing continues, the bottom layers should sustain the
printing.
gravity-induced stresses resulting from the top layers. However,
Previously, Wolfs et al. [7] numerically studied the buckling
the material yield stress is increasing with the concrete age. There-
failure of a 3D printed structure and considered the time-
fore, the specific characteristic is beneficial when printing the
dependent Mohr-Coulomb material behavior. However, the
structure layer by layer.
numerical simulation over-estimated the failure heights achieved
The stability or buckling failure is another failure mode
in their experiments. It is explained that the compaction of the
reported in previous studies [7,14]. Buckling failure can occur
material before testing in an unconfined compression test and
due to the progressive lateral deformations. The eccentric layer
the direct shear test may result in overestimation of the material
placement also may result in an unstable structure and stability
properties, which lead to the overestimation in failure height and
failure [15]. The failure due to buckling is mainly governed by
the underestimation of lateral deformations.
the material properties such as the elastic modulus at fresh state.
In this study, the Mohr-Coulomb based buildability criterion
Additionally, the geometric shape or slenderness of the structure
was developed and validated with laboratory experiments and
printed may affect the buckling failure. Therefore, characterizing
numerical simulations. The laboratory experiments were carried
the elastic modulus at the fresh state is important to determine
out to establish the time-dependent material and rheological prop-
the stability failure height.
erties of 3D printable concrete mixes. The usage of non-
conventional measurement techniques to measure the early-age
1.3. Review of existing models to assess buildability
material properties was implemented. Specifically, care was taken
to reduce the effects of compaction when preparing the samples
The analytical model developed by Roussel et al. [16] is com-
for testing. The time-dependent elastic modulus of a printed layer
monly used to determine the plastic collapse (failure due to
was trialed using a new method in the current study. Further, 3D
exceeding material static yield stress) height in most of the 3D con-
printing of hollow circular sections was carried out to study the
crete printing literature published [8,13,17,18]. The model is based
failure modes and to obtain the building height at the time of fail-
on ASTM Abrams cone and mini cone and assuming pure elonga-
ure. The nonlinear finite difference model of the 3D printed sec-
tional flow for lower slump values, where the stress variation in
tions was developed with the user-defined time-dependent
the vertical direction is comparatively higher than those in the
material models to assess the accuracy of the proposed buildability
radial direction. Von-Mises yield criterion was used for the analy-
criterion. Additionally, the proposed buildability criterion was fur-
sis. However, one of the major limitations of the model is, that the
ther validated with the 3DCP experimental and numerical data
material is considered as a continuum, and the results are valid
presented in the literature.
only if the sample height to radius ratio (aspect ratio) is sufficiently
large. Also, the derivation of the analytical model was based on a
1.1. 3DCP material rheology
cylindrical sample. However, if larger particles were included in
the mix, the analytical model developed by Roussel et al. may have
Concrete used in 3DCP applications is commonly considered as
some limitations. Also, in 3D printing applications, aspect ratios
Bingham material [11] with visco-plastic behavior. The flow occurs
(layer height to width ratio) used may result in higher confinement
when the material is subject to shear stress larger than the yield
stresses in printed layers [19], which were neglected in the model
stress (sy ), then, the shear stress (s) is linearly proportional to
by Roussel et al. [16].
the shear rate (c_ ). The plastic viscosity (l) of the material is defined
Further, according to the model discussed by Roussel et al.
as the gradient of the shear stress and the shear rate curve. Also,
above, Roussel [8] discussed the buildability requirements exten-
3DCP cementitious material can be generally considered as thixo-
sively for 3D printing concrete based on Bingham yield stress.
tropic material where the time-dependent shear thinning (i.e. vis-
According to [8], when the layer-wise construction process contin-
cosity reduction with shear rate) happens when agitated or highly
ues, the initial shear stresses produced by the deposition process in
sheared. The particular shear- thinning behavior can be modeled
a layer, gradually turn into extensional stresses in the underlying
using the Herschel-Bulkley model.
layers. For a total build height of H of the structure, yield stress
Concrete exhibits a time-dependent material behavior where p
of qgH= 3 (is the gravitational constant and q is the density of
the yield stress of the material increases with the concrete age at
concrete) should be achieved by the bottom-most layer [8].To pre-
the fresh state. This phenomenon is happening due to the floccula-
vent the plastic failure, the Athix (i.e., the yield stress increasing rate
tion process of cement particles and due to a built-up of an internal
with concrete age), neglecting the initial yield stress can be written
_ p 3. Here, the H_ is the vertical build rate of the struc-
structure immediately after mixing and kept at rest. Further, with
increasing concrete age, nucleation and formation of CSH within as Athix  qg H/
the contact points of cement grains may occur. This results in ture. Simply, the following conditions for the yield stress should be
increasing the yield stress further [12]. Perrot et al. [13] developed satisfied after time ‘t’ at the bottom most layer to prevent the plas-
a model to explain the increase in yield stress, sy ðtÞ, with resting tic failure.
time, t, as given in Eq. (1). The parameter, Athix , is the structuration _ p
sy ðtÞ  qg Ht= 3 ð2Þ
rate. This is assumed as a constant for a considered material. The
R. Jayathilakage et al. / Construction and Building Materials 240 (2020) 117989 3

The Eq. (2) is mainly based on the yield stress of the material based on a rectangular cross-section of a vertical slender column
and use to predict plastic failure. Fore mentioned model developed with a unit length. Therefore, can be used for rough estimations
by Roussel et al. [16] was the basis of developing Eq. (2) consider- of critical parameters required.
ing the pure elongational flow. In a recent study done by Wolfs et al. [15], a set of parametric
First attempts in studying the buildability of 3D printing con- equations with numerical and experimental results were pre-
crete were done by Di Carlo [20] through an unconfined compres- sented to study the failure heights of straight, vertical walls. The
sion test. An analytical and numerical simulation was done to two major criteria for wall failure were identified as plastic failure
model the buildability of a 3D printed layer considering the early and elastic buckling failure. The time-dependent material proper-
age material behavior. A numerical Finite Element model based ties (specified by linear and exponential curing functions) as well
on the linear Drucker-Prager yield criterion was used to assess as the wall widths, wall end restrains, and print speed effects were
the failure of printed layers. The analytical model developed was considered in the study. Additionally, the effect of initial imperfec-
a simple stress–strain model based on vertical stresses applied to tions in printed structure for stability also was assessed using the
the bottom-most layer. The unconfined compressive strength parametric model as well as the numerical simulation. According
was considered as the strength of the material to overcome the to [15], for a straight wall structure, the competition between elas-
vertical stresses. tic buckling failure and plastic collapse can be identified using the
Wangler et al. [17] developed an analytical model to predict the following parameters;
buildability requirement of 3D printed concrete layers based on
the rheological yield stress. The failure stresses are determined  13
lcr h sy0
by the yield stress behavior with concrete age and the structura- < 2
¼ x Elastic buckling
lp D0 ðqg Þ3
tion rate (Athix ).
An analytical model was developed by Panda et al. [9] to find
the vertical deformation of a printed layer due to the applied pres-  13
lcr h sy0
sure from the above layers. The strength-based failure of a layer > 2
¼ x Plastic collapse
was considered in the study. However, the model only captured lp D0 ðqg Þ3
the elastic deformation effectively rather than the higher deforma-
tions occurring in the plastic state. The experimental vertical defor- In Eq. 5, lcr and lp are dimensionless buckling height and plastic
mation values were higher than the analytically predicted values in collapse height. The values for lcr and lp can be found using a set of
[9]. According to the authors, the main reason for the discrepancy Equations given in [15] in detail. h and D0 are wall thickness and
between analytical and numerical results is the assumption they the initial wall bending stiffness respectively. sy0 is the initial yield
made, that the deformation is elastic. However, the plastic flow strength of the mix considered.
before failing may give a higher deformation value and the authors Finally, a lower bound analytical model was developed by Kru-
believed that the tangential modulus of plastic state should be con- ger et al. [19] recently to assess the plastic failure of 3D printed
sidered to calculate the plastic deformations. concrete layers. A Tresca yield criterion was used considering a
Another critical failure mode occurring in 3D printing is stabil- strength correction factor to address the confinement of the
ity failure. Wolfs et al. [7] have done extensive and initial studies bottom-most layer. The authors mentioned that the low aspect
on stability failure using experimental studies as well as numerical ratios used in 3D printed layers may result in a tri-axial stress state
methods. The key failure mode identified in the particular study is rather than a uniaxial stress state. Also, they have considered the
the buckling failure with locally occurred strength-based failure. structuration as well as the re-flocculation rates of the material
Progressive deformation of printed cylindrical structures was to define the thixotropic behavior of the bottom-most layer. How-
recorded experimentally to track buckling failure. The failure mode ever, the model only predicts a lower bound limit of plastic failure.
was effectively predicted by the numerical simulations as well in
the study done by Wolfs et al. [7]. The time-dependent material
behavior was considered in that study. 2. Experimental work
Roussel [8] discussed the buckling failure limits and the plastic
failure limits by analyzing the results published by Wolfs et al. [7]. Material flow properties and failure modes for 3D printed struc-
Depending on the modulus of elasticity (E) of the material, a criti- tures were determined using laboratory experiments. A part of the
cal buckling failure height (Hc ) can be defined as in [8] due to the experimental results were published by the authors elsewhere.
self-weight of a column. Furthermore, Roussel [8] defined a transi-
tion height (HT ) between two failure criteria (strength-based fail-
ure or buckling failure), where strength based failure is dominant 2.1. Material and mix
for print heights lower than the transition height and buckling fail-
ure is dominant if the print heights are higher than the transition General-purpose (GP) cement (ASTM C595/C595M) was used in
height. Eq. (4) gives the transition height between two failure the mix considered. Graded coarse and fine sand (coarse sand with
criteria. a maximum particle size of 1.18 mm and fine sand with a maxi-
mum particle size of 600l m) were used. Coarse aggregate to
Hc ¼ ð8EI=qgAÞ1=3 ð3Þ cement and fine aggregate to cement ratios used are 0.56 and
sffiffiffiffiffiffiffiffiffiffiffiffiffiffi 1.11 respectively. Silica fume was used in a ratio of 0.11 with
1þt cement. The water to cement ratio used is 0.25. Note that all the
HT ¼ 2d pffiffiffi ð4Þ ratios mentioned are as weight ratios. Furthermore, 0.9 ml of
3 3cc
superplasticizer was used per 100 g of binder and 0.6 ml of retar-
In Eq. (3), I and A denotes the 2nd moment of area of the printed ders were used per 100 g of the binder. Here, cement and silica
cross-section and the horizontal cross-section area respectively. In fume were considered as the binder.
Eq. (4), t and d are the Poisson’s ratio and width of a printed layer Initially, dry mixing was done for about 5 min. Slow mixing was
respectively. cc is the critical shear strain at flow onset. The critical done up to 5 min after adding water slowly. Afterward, high-speed
shear strain can be found using the expression, cc = 2sy ð1 þ tÞ=E. It mixing was done about another 5 min after adding superplasticiz-
should be noted that the equations derived in [8] and Eq. (4) is ers and retarders.
4 R. Jayathilakage et al. / Construction and Building Materials 240 (2020) 117989

2.2. Measurement of time-dependent yield stress rent test also, the condition is satisfied as the maximum particle
size (around 1 mm) is nearly one-tenth of the smallest dimension
The yield stress variation with concrete age was measured (i.e. 10 mm height) of the sample.
using a vane shear apparatus as in [22]. The standard ASTM The main drawback of the current method is that the sample
D4648/D4648M test procedure was followed. Vane shear appara- height is too low to achieve failure in the weakest failure plane.
tus is commonly used for the measurement of the shear strength The end effects from the two compression plates can increase com-
of clayey soils. However, it was implemented in the current study pression stress significantly. Therefore, the method is not appropri-
as a cost-effective method to measure the yield stress variation ate to determine the unconfined compressive stress or failure
with time. The vane used has a height of 70 mm and a width of stress. However, in the current study, only the elastic modulus of
90 mm. the material is of interest and it can be achieved by considering a
Angle measurements in degrees were taken in 15 min time low strain limit of the specimen.
intervals from 0 to 60 min, using a dial gauge fixed on top of the A 500 N load cell attached to an MTS testing system was used to
vane. It should be noted, that different samples were used in each measure very low loads applied in the test. The vertical displace-
time interval. Also, all the samples were prepared initially, before ments were measured using the LVDT sensor in the MTS loader.
conducting the vane shear test at 0 time period. Three trials were The test setup is shown in Fig. 2.
done at a particular time period (in 3 locations of the sample Test trials were conducted for three samples at the same time
according to the standards) to get the average angle value and later period to get the average stress–strain values. Material mixing,
converted into shear strength values using the equations given in printing, and preparation of the anticipated sample were done
[23], for the measurements taken near to the free surface of the quickly enough to measure the elastic modulus.
material. A very low displacement rate (0.5 mm.min1) was applied for
the samples. The test was done up to 2 mm total vertical deforma-
2.3. Time-dependent modulus of elasticity tion (20% strain limit) results in 4 min to complete a single test.
Therefore, the selected displacement rate to conduct the test
The conventional and the standard method to measure the assumed to be fast enough to neglect the thixotropic build-up
modulus of elasticity is to perform an unconfined compression (The initial re-flocculation state defined by Kruger et al. [19] is
test for cylindrical specimen according to ASTM D2166 standards neglected in the current study. Only the thixotropic build rate
as done in [7] and [24] for fresh concrete. This can be performed (Athix ) effect was considered in the particular study).
in several time intervals to acquire the variation of elastic modu- The elastic modulus at 5 min and at a more than 5 min time per-
lus with time. However, it was highlighted in the previous studies iod can only be considered (due to the time taken for the specimen
[7,9] that the material elastic modulus and compressive strength preparation and to conduct the test). It is difficult to determine the
are over-estimated due to the compaction of the samples tested elastic modulus experimentally at 0 min with the above condi-
in the unconfined compression test. Therefore, in the current tions. 5 min to 60 min resting times were considered in the current
study, a new non-standardized approach was trialed to attain study.
the modulus of elasticity variation with time as described below
(Fig. 1). 2.4. Mohr-Coulomb parameters for 3D printable concrete
Instead of testing cylindrical samples, printed samples were
tested in compression. 100 mm long layers extruded from the noz- The Direct shear test (according to the ASTM D3080/D3080M-
zle (which were used in the printing experiment) were used for the 11 standards) can be used to find the Mohr-Coulomb parameters
test as shown in Fig. 2. The printed samples had 10 mm thickness and the dilation angle value of the mix considered. In the direct
and a 30 mm width. The length selected for the specimen is com- shear test method, different normal stresses can be applied to
paratively larger than the width and the thickness of the layer. This achieve the failure shear stress in a pre-determined horizontal
is to assume, that the strain in length direction is negligible com- plane. In the current study, peak shear stress was considered as
pared to the other two directions of the sample (plane strain the failure shear stress. A complete description of the direct shear
conditions). test method and how to implement it to measure the friction angle
According to the ASTM D2166/D2166M standards, the sample values, cohesion stresses and dilation angle values in 3D printing
used for the unconfined compression test should have a maximum concrete were explained and published by the authors elsewhere
particle size of one-tenth of the diameter of the sample. In the cur- [25].

Fig. 1. a) Vane shear apparatus b) Schematic diagram of the test setup.


R. Jayathilakage et al. / Construction and Building Materials 240 (2020) 117989 5

LVDT attached

MTS actuator (compression)


10 mm
(30 mm width) Specimen
Platform
Load cell
100 mm
(a) (b)
Fig. 2. a) Elastic modulus measuring test setup b) Schematic diagram.

Circular direct shear test box with 72.5 mm diameter and a 300 mm inner diameter were printed using two different nozzle
43 mm depth was used for the experiment. The direct shear box sizes. A 30 mm and 20 mm circular nozzles were used for printing
was filled with 3D printing samples in 3 layers and each layer and the layer height was kept as 10 mm in each case. Three trials
was slightly tamped to achieve a continuous sample with mini- were conducted for each nozzle size to achieve an average failure
mum air voids. height. A printing speed of 2244 mm.min1 was used, which gives
In the current study, a Shear Trac II direct shear test setup was a 0.42 min time gap between two consecutive layers.
used. Normal stresses varying from 1 to 15 kPa were applied and
the shear stress variation, horizontal displacement, and the vertical
displacement were recorded. A displacement rate of 8 mm.min1 3. Numerical simulation for layer by layer construction and
was used for shearing the samples. It was found from a previous simplified failure criterion
study that the displacement rates used in the direct shear test
(i.e. around 0.5 to 15 mm min1) have no significant effect on mea- A non-linear Finite Difference Method Software (FLAC 3D) [26]
suring the Mohr-Coulomb parameters [25]. was used as the numerical simulation tool. The elastic modulus,
However, measuring the cohesion values at 0 to 15 min of con- EðtÞ of the fresh mix should be determined as an input parameter
crete age is difficult in the direct shear test method. The reason is for the numerical simulation to determine the elastic stress–strain
the time taken to complete the shearing process after mixing the behavior. The elastic modulus was found to be time-dependent by
material is typically higher than 10 min. In the study [25], the Wolfs et al. [7]. Also, the time-dependent Mohr-Coulomb failure
cohesion stress was found to be equivalent to the yield stress of criterion can be written as in Eq. (6) where cðtÞ, rn and £ are cohe-
the material. Therefore, the time-dependent yield stress values sion, normal stress and the friction angle of the material
obtained from the vane shear test can be used as the cohesion respectively.
stress values of the mix. The similarity between vane shear yield
stress results and direct shear cohesion stress results for cementi- sy ðtÞ ¼ cðtÞ þ rn tan£ ð6Þ
tious material were further discussed by Assaad et al. [23].
The friction angle can be assumed as a time-independent value
at the early ages of concrete [7]. The main input parameters con-
2.5. 3D concrete printing of circular section sidered in the simulations are the time-dependent cohesion values
and time-independent friction angle value.
One of the 3D concrete printing setups at Swinburne University The Mohr-Coulomb failure criterion used in FDM software is a
of Technology is shown in Fig. 3. Hollow cylindrical structures with composite Mohr-Coulomb criterion with tension cutoff. Eq. (7)
and Eq. (8) gives the shear and tensile yield functions. The major
steps of calculating the elastic stresses, strains and plastic stresses
and strain values can be found in [21].
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1 þ sin£ 1 þ sin£
f s ¼ r1 þ r3  2c ð7Þ
1  sin£ 1  sin£

f t ¼ r3  rt ð8Þ
In Eqs. (7) and (8), r1 , r2 , r3 are the principal stresses (r1  r2
 r3 ) and rt is the tensile stress. f s and f t are the shear and tensile
failure functions where f s = 0 is Mohr-Coulomb failure envelope
and f t = 0 is the tensile failure envelop. rt in Eq. (8) is the tensile
strength which has a maximum value (rt;max ) of c/tan£. However,
the default value for rt is taken as 0 in the FDM software. The ten-
Fig. 3. Printing setup (Swinburne University of Technology, Australia). sile stresses were considered as positive in the FLAC analysis.
6 R. Jayathilakage et al. / Construction and Building Materials 240 (2020) 117989

Authors recommend the readers to follow the detailed descrip- 25


Concrete age = 5 min
tion of the failure criteria in FLAC 3D documentation [26] and other Concrete age = 10 min
references [27]. Concrete age = 15 min
According to the theory of plasticity and Mohr-Coulomb failure Concrete age = 25 min
Concrete age = 35 min
criterion, the failure happens when f s = 0 in Equation (7). There- 20
Concrete age = 60 min
fore, from Eq. (7), we can get;
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1 þ sin£ 1 þ sin£
r1 þ r3 ¼ 2c ð9Þ
1  sin£ 1  sin£ 15

Stress (kPa)
If we consider the stresses acting on the bottom-most layer, the
vertical stresses due to the top layers can be simply written as qgH,
neglecting the self-weight of the bottom layer. If one considers the
10
compression as negative, this will be the minimum principal stress
(i.e. r1 ). The maximum principal stresses (r3 ) acting in the bottom-
most layer (which are in the radial direction) can be assumed as
zero if we assume an ideal smooth surface on the print bed. Also,
5
if the stability failure is not initiating, it is reasonable to assume,
that the maximum principal stresses are zero. Therefore, Equation
(9) can be simplified to Equation (10) as follows (the unknown
maximum principal stresses were omitted for simplification);
0
2ctanð£ þ p4Þ 0 0.01 0.02 0.03 0.04 0.05
Hf ¼ 2
ð10Þ Strain (mm/mm)
qg
Fig. 4. Average stress–strain diagrams for different concrete ages.
If the Mohr-Coulomb parameters are known, Equation (10) can
be simply used to determine the plastic failure height (Hf ) of the
bottom-most layer due to gravity-induced stresses. A straight wall (20% strain), a linear stress–strain diagram can be achieved. How-
with different end restrictions was studied by Suiker. In the study ever, a 5% strain limit (0.5 mm deformation) was considered to cal-
[14], it was mentioned that the lateral stresses are zero in an culate the elastic modulus and presented in Fig. 4.
unconstrained wall and is equal to Poisson’s ratio (t) times the ver- The achieved elastic modulus for 3 trials, the calculated average
tical stress in a fully constrained wall. Therefore, the magnitude of and the relative standard deviation (RSD) value are shown in
r3 value in the current study also should be in the range of 0 r3  Table 1.
t.r1 . Using the average elastic modulus values, the variation of elas-
In addition to Mohr-Coulomb parameters, Poisson’s ratio (t) tic modulus with concrete age can be plotted as in Fig. 5.
and the density of the material (q) are also should be considered From Fig. 5, an initial elastic modulus value of 51.6 kPa can be
in the numerical simulation. From previous studies by Wolfs achieved using the linear fit. The elastic modulus value achieved
et al.[7] it was found that these parameters are time-independent. is typically closer to the elastic modulus values for 3D printing con-
Primitive-based grid points, cylindrical shell mesh, and zones in crete found in the literature [7,9,24].
FLAC3D were used to model the printed layers. A unit length of the In the current study, the layers with required length were
circular column was considered for the modeling. A single layer printed and kept covered using a polythene cover for measuring
was divided into more than 30 elements (as per the convergence the elastic modulus. It was observed that the quick hardening of
of the results). The axis-symmetry behavior was considered when layers occurred after around 60 min. Therefore, the elastic modulus
applying the boundary conditions. The radial direction movement was considered only up to 60 min. Also, the printing experiment
and the vertical movement were not restrained. Fixed boundary was done in less than 15 min, and considering the linear response
conditions were used at the bottom layer assuming a rough print of elastic modulus with concrete age will be appropriate for the
bed. particular time period.
The layer height and widths used in the numerical simulation In the current study, the specimen considered is very small
are equal to the ones used in the experimental printing process. compared to a typical specimen used in an unconfined compres-
The time-dependent properties for each layer were input by a sion test. Therefore, the surface areas considered are comparatively
user-defined FISH algorithm specifically used in the software. small and the hardening of the specimen can happen easily. This
phenomenon can be attributed to an actual 3D printing process,
4. Results and discussion which will take more than around 60 min of printing time. The bot-
tom layers can harden quite easily due to low surface areas
4.1. Elastic modulus exposed to the surrounding environment. Due to this reason,
time-dependent elastic modulus values measured using larger
The stress–strain behavior was developed using the force–dis- cylindrical samples may not accurately represent the actual elastic
placement data achieved from the compression tests for printed modulus variation in a printed specimen. However, further inves-
layers. No area correction was done because the variation of the tigations are needed before making any conclusions.
area in the samples are considered negligible due to the small
strain limit considered and the small sample height. Therefore,
the initial surface area (i.e. 30  100 mm2) was used for stress cal- 4.2. Vane shear and direct shear test results
culations. Also, due to the small specimen height (i.e. 10 mm), the
vertical stresses due to the self-weight of the layer (which is Vane shear test results, which are also presented in [21] for dif-
around 0.2 kPa) were neglected. ferent concrete ages are shown in Fig. 6.
Fig. 4 shows the average stress–strain diagrams for different The initial yield stress (0.3 kPa) of the material is a very low
concrete ages and the average stress–strain plots. Up to 2 mm value compared to the mix used in [7]. The yield stress values
R. Jayathilakage et al. / Construction and Building Materials 240 (2020) 117989 7

Table 1
Elastic modulus values.

Concrete age (min) Elastic Modulus (kPa)


Trial 1 Trial 2 Trial 3 Average RSD (%)
5 70.1 76.3 82.5 76.3 6.6
10 113 117.3 98.1 109.5 2.0
15 166.2 152.6 174.6 164.5 5.5
25 246.8 285.6 273.7 268.7 6.0
35 324.2 322.9 336.7 327.9 0.2
60 478.1 482.4 – 480.2 0.4

500 achieved from the vane shear test were considered equivalent to
Experimental
2
Linear fit ( E = 7.4t + 51.6 , R = 0.98 )
the cohesion value as mentioned earlier.
450 The linear Mohr-Coulomb failure envelop was achieved using
the normal stress value and the corrected (according to ASTM
400 D3080/D3080M-11 standards) peak shear stress value from the
direct shear test data. The Mohr-Coulomb envelop for the particu-
350 lar mix is shown in Fig. 7(b). However, it should be noted that the
Elastic modulus - E (kPa)

cohesion value is not equivalent to the initial shear yield stress


300 value considering the time gap between material preparation and
the end of shearing the sample (The time gap between material
250 preparation and end of shearing the sample varies between 15
and 20 min).
200 In Fig. 7(b), it can be seen that the achieved cohesion value is
1 kPa. The higher cohesion values when compared with the initial
150 vane shear yield stress value (0.3 kPa) is due to the time-dependent
behavior. If one considers the experimental vane shear yield value
100
at 15 min (Fig. 6), a yield stress value around 0.8 kPa can be seen,
which is closer to the cohesion value achieved from the direct
shear test.
50
From earlier studies, it was found that the friction angle (£) and
the dilation angle (w) are time-independent for the early ages of
0
0 10 20 30 40 50 60 concrete [7]. However, the friction angle can increase with increas-
Concrete age (min) ing concrete ages (more than 100 min) [28]. Furthermore, it was
found in previous studies that the friction angle value mainly
Fig. 5. Elastic modulus variation with concrete age.
depends on the aggregate proportion in the concrete mix [25,29].
Therefore, no time-dependent behavior was considered for friction
angle and the dilation angle. The friction angle (£) was found as
2 42° from Fig. 7(b). The dilation angle can be found from the vertical
Experimental
Fitted ( y0 = 0.3 kPa , Athix= 0.02 kPa.min-1 )
displacement vs. horizontal displacement diagrams achieved from
1.8 the direct shear test [25]. The average dilation angle was found as
0.7° for the material used.
1.6

4.3. Printing experiment


1.4

The printing experiment and the collapsing of the structure for


Yield stress (kPa)

1.2
the two different layer thicknesses are shown in Fig. 8. Printing
test results for two nozzle sizes and for each trial are shown in
1
Table 2.
Similar results for the failure height can be achieved for a speci-
0.8 fic nozzle size (i.e. layer width) from three trials. Therefore, an
average failure height can be determined for each layer width as
0.6 in Table 2. If compare the two average values, a failure layer num-
ber difference of 2 layers can be seen. The difference between the
0.4 two cases is around 15%. According to the analytical model by Kru-
ger et al. [19], the lower bound failure layer number for 30 mm and
0.2 20 mm layer widths is 2 layers for each case, irrespective of the
layer aspect ratio used. Therefore, it may be reasonable to assume
0 that the aspect ratios used in the current study have no significant
0 10 20 30 40 50 60
Concrete age (min) effect on the failure layer number when compared to the other
material properties (i.e.yield stress, thixotropic build rate, etc.) in
Fig. 6. Vane shear test results for different concrete ages. the particular mix. However, the aspect ratio will have a significant
8 R. Jayathilakage et al. / Construction and Building Materials 240 (2020) 117989

16 16
Normal stress = 1.3 kPa Experimental
Normal stress = 5.3 kPa 0 2
Mohr-Coulomb failure envelop ( =1+ tan(42 ) , R = 0.9)
Normal stress = 9.6 kPa y n
14 Normal stress = 14.5 kPa 14

12 12

Shear stress - y (kPa)


Shear stress (kPa)

10 10

8 8

6 6

4 4

2 2

0 0
0 2 4 6 8 10 12 14 16 0 5 10 15
Horizontal displacement (mm) Normal stress - n (kPa)
(a) (b)

Fig. 7. a) Shear stress vs. horizontal displacement b) Mohr-Coulomb failure envelop.

Fig. 8. a) Printing experiment (30 mm nozzle) b) Collapsing of the structure (30 mm nozzle) c) Printing experiment (20 mm nozzle) d) collapsing of the structure (20 mm
nozzle).

effect on higher build heights and will significantly vary according als were conducted using different batches of concrete (with the
to the mix used. Also, it should be mentioned that a total of six tri- same mixing protocol and the same environmental conditions).
This can be another possible reason for achieving different average
Table 2
failure heights for two different nozzle sizes. However, direct con-
Printing experiment result comparison.
clusions couldn’t be made as the material used has low buildability
Nozzle diameter (layer width) (mm) Trial No: Failure layer No: when compared to the typical 3D concrete printing mixes used in
30 1 14 practice. Therefore, it can be predicted that the failure method of
2 13 the circular column is a strength-based failure irrespective of the
3 13
layer geometry used.
Average ~13
20 1 10
Also, while printing the circular structures, no specific radial
2 11 deformations can be seen and sudden collapse occurred at the final
3 12 layer. In Fig. 8, two immediate instances of before failure and after
Average ~11 failure were shown. No specific buckling can be seen before failure
R. Jayathilakage et al. / Construction and Building Materials 240 (2020) 117989 9

and the bottom-most layer is largely deformed and failed to make The reason for using low buildable material in the current study
the whole structure collapse. However, further validation of the was to achieve strength-based failure without any stability failure
failure method was done using the numerical simulations and ana- or combination of stability and strength-based failure (which may
lytical equations available in the literature. occur in larger heights).

Fig. 9. Numerical simulation results for a) Zone state after 5 layers (20 mm width) b) Zone state at failure (20 mm width) c) Zone state after 5 layers (30 mm width) d) Zone
state at failure (30 mm width).
10 R. Jayathilakage et al. / Construction and Building Materials 240 (2020) 117989

4.4. Numerical modelling results It should be noted that each horizontal line parallel to the time
step axis in Fig. 10 represents a layer construction above the bot-
In the numerical simulation, Mohr-Coulomb plasticity state plot tom layer (1st horizontal line represents the 3rd layer construc-
and the deformations of the layers can be used to identify the fail- tion). Therefore, the failure layer number also can be achieved
ure of the structure. Shear failure plasticity state plot and the ten- using the above diagram considering that the large strain incre-
sile failure plasticity state plot are the two types of plasticity state ments are occurring at the plastic state.
plots described in FDM simulation software. After the convergence Using the maximum and minimum principal stresses and the
criteria and the end of the time-step, the state of each zone can be failure criteria given in Eq. (7) and Eq. (8), the shear failure of the
checked. The state plot indicates that the stresses within a zone are bottom-most layer can be approximated as in Fig. 11.
currently on the yield line as mentioned in Eq. (7) (i.e., the zone is From Fig. 11 also, it is clearly visible that the shear failure and
at active failure now, -n), or the zone has failed earlier in the model plastic flow occurs in similar principal stresses for both layer
run, but now the stresses fall below the yield line (the zone has widths. It should be mention that the tensile failure line (as in
failed in the past, -p). A failure mechanism is indicated if there is Eq. (8)) is equal to r3 = 0 because the default value of rt is zero
a contiguous line of active plastic zones (indicated by either in FDM software [26]. The maximum and minimum principal
shear-n or tension-n) that join two surfaces. The diagnosis is con-
firmed if the velocity plot also indicates motion corresponding to
the same mechanism. 0
ft = 0
The plasticity state plots at 5 layers and at failure are given in
Fig. 9 for both nozzle diameters (layers widths).
In Fig. 9, it can be seen that for 20 mm layer width, Mohr-
Coulomb shear failure happens at 10th layer construction. Like- -0.05 domain 2

wise, for the 30 mm layer width, 10 layers can be considered as

Max. principal stress - 3 (kPa)


fs = 0
the failure layer number. In the numerical simulation, failure hap-
pens irrespective of the layer widths.
As shown in Fig. 10, the minimum and the maximum stress- -0.1
time step and strain–time step plots of the bottom-most layer domain 1

can be plotted considering a zone in the numerical model. The


elastic and plastic response of the element can be clearly seen
for the two layer widths considered (after the plastic state (fail- -0.15
ure), the strain increments are comparatively higher as shown
in Fig. 10(a)).
If one considers the maximum and minimum principal stresses
and strain increments, the 30 mm layer width gives similar results -0.2
to the 20 mm layer width. If we consider the principal strain incre-
ments (Fig. 10(a)), strain increments are occurring when the plastic
20 mm layer width
flow initiates around 8  104 time step. It should be noted that the 30 mm layer width
negative sign of the strain increments indicates the reduction of -0.25
-2.5 -2 -1.5 -1 -0.5 0
the layer height in the vertical direction and positive strain incre- Min. principal stress - 1 (kPa)
ments indicates the increase of layer width in the radial direction.
It should also note, that the failure occurs approximately in a sim- Fig. 11. Principal stress variation for 20 mm and 30 mm layer widths and failure
ilar strain for both layer widths as shown in Fig. 10(a). criterion (1st layer).

0.15 0

0.1
-0.5
Principal strain (mm/mm)

0.05
Principal stress (kPa)

-1
0

-0.05
-1.5

-0.1

Max. principal strain (20 mm)


-2
Max. principal stress (20 mm)
-0.15 Max. principal strain (30 mm) Max. principal stress (30 mm)
Min. principal strain (20 mm) Min. principal stress (20 mm)
Min. principal strain (30 mm) Min. principal stress (30 mm)
-0.2 -2.5
2 4 6 8 10 2 4 6 8 10
Time step 104 Time step 104
(a) (b)
Fig. 10. Principal strain and stress variation with time steps for 20 mm and 30 mm layer widths (1st layer).
R. Jayathilakage et al. / Construction and Building Materials 240 (2020) 117989 11

stresses at plastic flow onset is approximately closer to the values


shown in Fig. 10(b).
When comparing the experimental results with the numerical
results, it can be seen that the results are very similar. The average
failure height from the experimental results considering two noz-
zle sizes is 12 layers. It was assumed that the aspect ratio effect is
negligible for the considered mix and for the total constructed
layer height. For numerical simulations, a value of 10 layers was
achieved. A difference of 2 layers (i.e. 20 mm difference in failure
height) can be seen. The deviation is around 17% of the experimen-
tal values (deviation of 9% and 23% respectively considering the
20 mm layer width and 30 mm layer widths separately). The esti-
mated yield stress values and the friction angle values are the main
parameters that decide the failure height of numerical simulation,
according to the Mohr-Coulomb failure criterion. The experimental
cohesion (shear yield stress) values achieved from the vane shear
test and friction angle values achieved from the direct shear test
can deviate from the actual values in the mix used in printing tri-
als. Therefore, the simulation results can be considered accurate
enough to predict the strength-based failure of 3D printing
concrete.

4.5. Comparison of the results with previous studies

Using the current numerical model, further investigations for


failure prediction was done for a previous study done by Wolfs
Strength-based failure
et al. [7]. The same average parameters used in the study were at bottom layers
used in the current simulation to check the failure mode and vali-
date the numerical simulation. Initially, the strength-based failure
height was checked without considering the buckling failure as
shown in Fig. 12 below.
From the analysis, it was found that the strength-based failure Fig. 12. Strength-based failure prediction for the study done by Wolfs et al. [7].
happens in 63 layers, which is higher than the numerical simula-
tion results (i.e. 46 layers) by Wolfs et al. [7]. However, the cap-
tured failure height in [7] is dominated by a buckling failure
rather than the strength-based failure. Therefore, it can be con- checked. The local buckling failure may highly depend on the layer
cluded that buckling failure is dominant in the study [7] and if thickness and the diameter of the circular section [30].
somehow the buckling failure was prevented, a strength-based However, if one considers the buckling shape and the radial
failure can happen at around 63 layers. deformations in the numerical simulations (Fig. 13), the simplified
Furthermore, radial deformations of each layer were tracked for formulas considering a horizontal straight cross-section with 1
15, 30 and 45 layer construction (similar to [7]) to compare the unit length can be implemented, which used by Roussel [8].
results. According to that, a critical failure height around 20 layers
Interestingly, shear failure initiated in the bottom layer locally, (0.2 m) can be achieved for the study done by Wolfs et al. [7].
when constructing the 47th layer as shown in Fig. 13. However, as In the study by Wolfs et al. [7], the buckling failure happens
discussed earlier, the structure collapses (plastic flow occurs) at around 45 layers with 12 mm of maximum radial deformation in
the 63rd layer due to more layers exceeding the shear failure limit. the numerical simulation. In the current study, a radial deforma-
Only due to the consideration of strength-based failure in cur- tion around 10 mm can be achieved for 45th layer construction.
rent numerical simulation, buckling failure should be checked For further validation of the failure mode, the transition height
using Equation (3). A critical height where the buckling failure (HT ) can be calculated for two layer widths used in the current
occurs due to self-weight can be calculated. The cylindrical shape study, using Equation (4). The calculated values for critical strain
was considered for the calculation rather than the simplified equa- (cc ) using the yield stress values and the elastic modulus values
tions used by Roussel [8]. For the current study, critical failure at the end of printing is 0.011 (1.1%) for both cases. Therefore,
heights of 0.62 m (62 layers) and 0.61 m (61 layers) were achieved the HT value can be calculated as 0.286 m and 0.191 m respectively
for 30 mm and 20 mm layer thicknesses respectively. For further for 30 mm and 20 mm layer widths. This is roughly 29 layers and
clarifications for the usage of the equations, the buckling failure 19 layers respectively for 30 mm and 20 mm layer widths. There-
height was calculated for the study done by Wolf et al. [7]. The fore, it can be concluded that the failure happened was purely a
buckling failure height calculated using Eq. (3) is around 1 m, strength-based failure, because the failure occurred at a lower
which is around twice higher than the numerical values (around layer number compared with the transition height.
0.5 m) achieved, and the experimental values achieved (around From the recent study done by Wolfs et al. [15], the transition
0.3 m) in that study. It should be noted, that the elastic modulus between elastic buckling height and plastic collapse height can
(E) values used for the calculations are the initial elastic modulus also be checked for the failure in the current study. From Eq. 5,
lcr
values, and the value can increase further with the concrete age. the calculated value is 2.1 and the value for x in the expression
lp
This may result in higher buckling heights calculated from Eq. lcr
(3). Also, Eq. (3) gives the overall global buckling failure height is 0.23. Therefore, > x, which expresses that a plastic collapse
lp
rather than a local buckling failure height. Therefore, a local buck- may be dominant rather than elastic buckling failure. However, it
ling failure is possible in a lower number of layers and should be should be noted that Eq. 5 is also formulated based on a straight
12 R. Jayathilakage et al. / Construction and Building Materials 240 (2020) 117989

Fig. 13. Radial deformation for layer construction using the data from Wolfs et al. [7].

and vertical wall. For a cylindrical shape or a rather complex shape, For the study done by Wolfs et al. [7], the failure should happen
this may not be accurate compared to a straight wall. around 18 min according to the Mohr-Coulomb failure criterion.
This gives a failure layer number of 58 according to the print speed
used in the particular study. If one considers the Roussel [8] model,
4.5.1. Comparison of plastic failure heights using the Mohr-Coulomb failure happens earlier comparatively to the Mohr-Coulomb failure
approach and Roussel [8] model prediction. The calculated failure layer number is 32 layers. How-
The simplified Mohr-Coulomb based equation (Equation (10)) ever, failure happens less than 58 layers due to the buckling of
in the current study and the model based on Roussel [8]’s analysis the structure according to the numerical analysis done by Wolfs
(Equation (2)) were used to find the plastic failure height of the et al. [7]. Furthermore, the experimental results from the Wolfs
printed structures. The study done by Wolf et al. [7], Kruger et al. et al. [7] study show an average failure layer number of 29, which
[19] and the current study were used for the plastic collapse height is due to the buckling of the structure rather than the plastic
evaluation. The time-dependent material behavior also was con- collapse.
sidered. For the study done by Kruger et al., an average friction For the study done by Kruger et al., an average failure number of
angle of 30 degrees was assumed considering the range of friction 64 layers can be achieved from the current, simplified Mohr-
angles used in the current study as well as the study [7]. The ver- Coulomb model. However, Roussel [8] model only gives a failure
tical stresses applied in the bottom layer (according to the print layer number of 29, which is approximately half of the actual
velocities used in each case) and the material strength develop- experimental value achieved in the study (i.e. 60 layers).
ment (considering the two models) were plotted against the print- Therefore, it is reasonable to state, that the Roussel [8] model
ing time to find the failure time (the intersecting location of the underestimates the plastic collapse failure height. The simplified
graphs) as shown in Fig. 14. Mohr-Coulomb failure criterion gives comparatively accurate
For the current study, the Mohr-Coulomb model gives a printing results for the plastic collapse height.
time of around 2.8 min before the failure happens (Fig. 14). This When we compare the above results with the current numerical
gives a number of 7 layers as the failure layer number. However, simulation results, the current study gives a failure layer number of
according to the Roussel [8] model, the failure happens earlier 10 from the numerical results and 63 and 69 layers for the studies
and the failure layer number is approximately 2 layers. done by Wolfs et al. [8] and Kruger et al. [20] respectively. The
R. Jayathilakage et al. / Construction and Building Materials 240 (2020) 117989 13

Yield strength of the bottom layer (Current study) Yield strength of the bottom layer (Current study) Yield strength of the bottom layer (Current study)
Yield strength of the bottom layer (Roussel) Yield strength of the bottom layer (Roussel) Yield strength of the bottom layer (Roussel)
3 Vertical stress on bottom layer 35 Vertical stress on bottom layer 20 Vertical stress on bottom layer

2.5 30

15
25
2
Failure
Stress (kPa)

Stress (kPa)

Stress (kPa)
20
1.5 Failure 10
15
1
10 Failure
5 Failure
0.5
Failure 5 Failure

0 0
0 1 2 3 4 5 0 10 20 30 40 50 0
0 5 10 15 20
Printing time (min) Printing time (min) Printing time (min)
Current study Wolfs et al. Kruger et al.

Fig. 14. Yield strength and vertical stress evolution with printing time.

major reason for higher layer numbers in numerical simulation is 25 Failure layer no. variation with cohesion (n = 24c/cmax )
the effect of fixed boundary conditions applied in the bottom layer. Failure layer no. variation with friction angle ( n = 4e1.5 /
max)

Because of these restraints in radial directions, there will be prin- Failure layer no. variation with yield stress increase rate (n = 8e0.9Athix/Athix,max)

cipal stresses acting in a particular direction (Fig. 11). However,


for the simplified equation, the fore mentioned stresses were 20
Number of printed layers at failure - n

neglected. Only 30%, 8% and 7% differences between numerical


simulation results and simplified approximation can be seen
respectively in the current study, the study done by Wolfs et al.
[7] and Kruger et al. [19]. 15
Therefore, the failure height given by Equation (10) is a better
approximation than by Eq. (2). Hence, the cohesion, as well as
the friction angle values, could be important parameters to con-
sider for 3D printing cementitious materials to determine the plas- 10
tic collapse height.
The results from each analysis can be tabulated and summa-
rized as in Table 3.
5

4.6. Effect of cohesion, friction angle and yield stress increase rate
values on failure height

0
From previous sections, it is conclusive that the cohesion and 0 0.2 0.4 0.6 0.8 1
friction angle both affect the failure height, if one considers the c/cmax , / max, Athix /Athix,max
plastic collapse. When we consider the cohesion values, the
time-dependent increase in the cohesion values (Athix ) also is of Fig. 15. Failure layer number variation with normalized cohesion and friction angle
importance. Therefore, a parametric study was conducted using values.
the validated numerical model to see the variation in failure layer
number (i.e., failure height). The initial cohesion value was chan-
ged from 0 to 0.8 kPa while keeping the other parameters a con- was only considered for the check. Fig. 15 shows the change in
stant (values used in the numerical simulation in Section 3). A failure height with each parameter. Normalized values were used
similar procedure was followed for friction angles and Athix values (ranging from 0 to 1) to compare the parameters with the consid-
to determine the change in failure height (a friction angle range ered range.
of 0 to 60 degrees, and Athix range of 0 to 0.08 kPa.min1 were In Fig. 15, cmax , £max and Athix;max are maximum cohesion value
considered). It should be noted, that the 30 mm layer width (i.e. 0.8 kPa), the maximum friction angle value (i.e. 60 degrees)

Table 3
Comparison of the failure layer number.

Case Failure layer number (plastic collapse)


Experimental Mohr-Coulomb criterion (FDM simulation) Simplified Mohr-Coulomb criterion Model by Roussel [8]
Current study 12 10 7 2
Wolfs et al. [7] – 63 58 32
Kruger et al. [19] 60 69 64 29
14 R. Jayathilakage et al. / Construction and Building Materials 240 (2020) 117989

and the maximum Athix value (i.e. 0.08 kPa.min1) considered in the  The strength-based failure criterion [8], under-estimates the
parametric study. failure height of the structure. The friction angle or the frictional
As can be seen in Fig. 15, it is interesting to note that the failure behavior of the cementitious material affect the failure height
height increase exponentially with friction angle values and Athix considerably. Therefore, a simplified expression based on the
value. The failure layer number is increasing linearly with the Mohr-Coulomb failure criterion can be used to determine the
cohesion values. However, it should be noted, that using a zero plastic collapse height, incorporating the cohesion and friction
friction angle value and increasing the Athix value up to 0.08 kPa, angle values of the material considered.
min1 did not change the failure layer number (i.e. 4 layers). Nev-
ertheless, keeping Athix value zero and increasing friction angle Declaration of competing interests
value increased the failure layer number in an exponential manner.
Therefore, frictional behavior and friction angle values of 3D print- The authors declare that they have no known competing finan-
ing cementitious material also can be considered important in the cial interests or personal relationships that could have appeared to
context of the buildability of the structure and is a major focus of influence the work reported in this paper.
future research. Furthermore, the effect of print bed conditions
(support conditions) is also of interest in future research to quan- CRediT authorship contribution statement
tify the plastic collapse, considering the maximum principal stress
variation. R. Jayathilakage: Conceptualization, Methodology, Software,
Formal analysis, Investigation, Visualization. P. Rajeev: Methodol-
5. Conclusions ogy, Software, Supervision. J.G. Sanjayan: Methodology, Resources,
Supervision, Conceptualization, Formal analysis.
Failure of 3D printed concrete structures at an early age was
experimented and validated with numerical results. Time-
Acknowledgment
dependent material behavior was considered and characterized
experimentally. The values were then implemented in a FDM sim-
Authors gratefully acknowledge the financial support from
ulation. During the printing of the structure, a strength-based fail-
Australian Research Council grants, Infrastructure Grant
ure was observed. The buckling mode of the structure cannot be
LE170100168 and Discovery Grant DP170103521.
seen during printing. Previous studies also were considered for
model verification. Further clarifications were done using the avail-
References
able analytical models in the literature to identify the failure mode.
The commonly used model [8] to predict plastic failure was com- [1] Nematollahi B, Xia M, Sanjayan J Current progress of 3D concrete printing
pared with the Mohr-Coulomb model used in the current study. technologies. In: ISARC. Proceedings of the International Symposium on
It was found, that the Mohr-Coulomb model can predict the plastic Automation and Robotics in Construction, 2017. Vilnius Gediminas Technical
University, Department of Construction Economics.
failure height accurately, compared to the model in [8]. Finally, a [2] Y.W.D. Tay, B. Panda, S.C. Paul, N.A. Noor Mohamed, M.J. Tan, K.F. Leong, 3D
parametric study was done to identify the effect of initial cohesion, printing trends in building and construction industry: a review, Virt. Phys.
friction angle and Athix values on plastic failure heights. Prototyp. 12 (3) (2017) 261–276.
[3] F. Bos, R. Wolfs, Z. Ahmed, T. Salet, Additive manufacturing of concrete in
Following conclusions can be made from the current study; construction: potentials and challenges of 3D concrete printing, Virtual and
Phys. Prototyp. 11 (3) (2016) 209–225.
 The time-dependent Mohr-Coulomb failure criterion and FDM [4] D.D. Camacho, P. Clayton, W. O’Brien, R. Ferron, M. Juenger, S. Salamone,
Seepersad C Applications of Additive Manufacturing in the Construction
numerical simulations can be used effectively to model failure Industry–A Prospective Review, Vilnius Gediminas Technical University,
limits of 3D printed concrete in the green-state. The number Department of Construction Economics, 2017.
of layers at the time of failure was shown to be in reasonable [5] S. Lim, R.A. Buswell, T.T. Le, S.A. Austin, A.G. Gibb, T. Thorpe, Developments in
construction-scale additive manufacturing processes, Autom. Constr. 21
agreement with the numerical simulations.
(2012) 262–268.
 The criterion for purely strength-based failure can be identified [6] R.A. Buswell, W.L. de Silva, S. Jones, J. Dirrenberger, 3D printing using concrete
using different print layer widths with low buildable material. extrusion: a roadmap for research, Cem. Concr. Res. (2018).
[7] R. Wolfs, F. Bos, T. Salet, Early age mechanical behaviour of 3D printed
The increase of the layer width has no significant effect on the
concrete: Numerical modelling and experimental testing, Cem. Concr. Res. 106
failure height of the structure for the considered material. The (2018) 103–116.
slight variance in the experimental was probably due to the [8] N. Roussel, Rheological requirements for printable concretes, Cem. Concr. Res.
aspect ratio difference as reported in previously done research. (2018).
[9] B. Panda, L.J. Hui, M.J. Tan, Mechanical properties and deformation behaviour
 A new method for assessing the elastic modulus of printed of early age concrete in the context of digital construction, Compos. Part B:
material was trialed considering the drawbacks of sample Eng. (2019).
preparation in unconfined compression tests. Printed speci- [10] J. Kruger, S. Cho, S. Zeranka, C. Viljoen, G. van Zijl, 3D concrete printer
parameter optimisation for high rate digital construction avoiding plastic
mens (using the same printer and nozzle) with similar dimen- collapse, Compos. Part B: Eng. (2019) 107660.
sions as in 3D printing experiments were used for the [11] E.C. Bingham, Fluidity and Plasticity, vol 2, McGraw-Hill, 1922.
compression test. The method used in the current study gives [12] N. Roussel, G. Ovarlez, S. Garrault, C. Brumaud, The origins of thixotropy of
fresh cement pastes, Cem. Concr. Res. 42 (1) (2012) 148–157.
considerably good results. It was found that the printed layers [13] A. Perrot, D. Rangeard, A. Pierre, Structural built-up of cement-based materials
dry out quickly after some time period (60 min) due to the used for 3D-printing extrusion techniques, Mater. Struct. 49 (4) (2016) 1213–
low surface area exposed to the environment. The elastic mod- 1220.
[14] A. Suiker, Mechanical performance of wall structures in 3D printing processes:
ulus values jump to higher values after this time. This phe-
Theory, design tools and experiments, Int. J. Mech. Sci. 137 (2018) 145–170.
nomenon may occur in the bottom layers when the printing [15] R. Wolfs, A. Suiker, Structural failure during extrusion-based 3D printing
time is higher (higher than around 60 min). processes, Int. J. Adv. Manuf. Technol. (2019) 1–20.
[16] N. Roussel, P. Coussot, ‘‘Fifty-cent rheometer” for yield stress measurements:
 Buckling failure of structures also can be predicted using the
from slump to spreading flow, J. Rheol. 49 (3) (2005) 705–718.
model by tracking the radial deformations. [17] T. Wangler, E. Lloret, L. Reiter, N. Hack, F. Gramazio, M. Kohler, M. Bernhard, B.
 For 3D printing cylindrical circular structures, global buckling Dillenburger, J. Buchli, N. Roussel, Digital concrete: opportunities and
failure, as well as local buckling failure, should be considered. challenges, RILEM Techn. Lett. 1 (2016) 67–75.
[18] Y. Weng, M. Li, M.J. Tan, S. Qian, Design 3D printing cementitious materials via
However, buckling shape modeling is important to determine Fuller Thompson theory and Marson-Percy model, Constr. Build. Mater. 163
the mentioned failure mode. (2018) 600–610.
R. Jayathilakage et al. / Construction and Building Materials 240 (2020) 117989 15

[19] J. Kruger, S. Zeranka, G. van Zijl, 3D concrete printing: A lower bound analytical American Society of Mechanical Engineers, pp V009T010A062-
model for buildability performance quantification, Autom. Constr. 106 (2019) V009T010A062.
102904. [25] R. Jayathilakage, J. Sanjayan, P. Rajeev, Direct shear test for the assessment of
[20] T. Di Carlo, Experimental and numerical techniques to characterize rheological parameters of concrete for 3D printing applications, Mater. Struct.
structuralproperties of fresh concrete relevant to contour crafting, University 52 (1) (2019) 12.
of Southern California, 2012. [26] F. Itasca, Fast Lagrangian analysis of continua, Itasca Consulting Group Inc,
[21] R.I. Jayathilakage, P. Rajeev, J. Sanjayan, Predication of strength-based failure Minneapolis, Minn, 2000.
in extrusion-based 3D concrete printing, in: Rheology and Processing of [27] M.P. Nielsen, L.C. Hoang, Limit analysis and concrete plasticity, CRC Press,
Construction Materials, Springer, 2019, pp. 391–399. 2016.
[22] T.T. Le, S.A. Austin, S. Lim, R.A. Buswell, A.G. Gibb, T. Thorpe, Mix design and [28] L.K. Mettler, F.K. Wittel, R.J. Flatt, H.J. Herrmann, Evolution of strength and
fresh properties for high-performance printing concrete, Mater. Struct. 45 (8) failure of SCC during early hydration, Cem. Concr. Res. 89 (2016) 288–296.
(2012) 1221–1232. [29] G. Lu, K. Wang, Theoretical and experimental study on shear behavior of fresh
[23] J.J. Assaad, J. Harb, Y. Maalouf, Measurement of yield stress of cement pastes mortar, Cem. Concr. Compos. 33 (2) (2011) 319–327.
using the direct shear test, J. Nonnewton. Fluid Mech. 214 (2014) 18–27. [30] K. Ahn, J.S. Kim, H. Huh, The effects of local buckling on the crash energy
[24] Di Carlo T, Khoshnevis B, Carlson A Experimental and numerical techniques to absorption of thin-walled expansion tubes, in: Numisheet, Numisheet, 2008,
characterize structural properties of fresh concrete. In: ASME 2013 pp. 799–804.
International Mechanical Engineering Congress and Exposition, 2013.

You might also like