Bending Tests: 3-Point and 4-Point Bending

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 20

Bending Tests

Related terms:

Young's Modulus, Acoustic Emission, Deflections, Tensile Strength, Tensile Test,


Tensiles, Bending Strength

View all Topics

Volume 1
Yusuf Khan, in Encyclopedia of Biomedical Engineering, 2019

3-Point and 4-point bending


Bending tests are conducted by placing a length of material across a span and
pushing down along the span to bend the material until failure. Bending tests reveal
the elastic modulus of bending, flexural stress, and flexural strain of a material.
3-Point bending involves placing the material across a span supported on either
ends of the material and bringing down a point source to the center of the span
and bending the material until failure while recording applied force and crosshead
displacement (see Fig. 5). The 3-point description comes from the two points of
support at the ends of the material and the one point of deflection brought down
to the middle of the material. 4-point bending tests are conducted similarly to
3-point bending tests except that rather than one point source being brought down
to the center of the span of material two points slightly separated from the center of
the material are brought down in contact with the material. This separation of the
two point sources spreads the region of bending out from the center such that a
larger portion of the material is tested than with only one point of deflection. Hard
materials like bone and implants can be tested using bending tests as a measure of
the tissue or material in both tension (the bottom of the sample as it is tested) and
compression (the top of the material as it is tested).
Fig. 5. 3-Point (left) and 4-point (right) bending test apparatus. 3-point bending
provides three points of contact; two supports and one center point where the
loading is applied. 4-point bending provides four points of contact; two supports
and two points where loading is applied. The two points spread the loading region
along the specimen so a larger portion of the material is tested in bending.

> Read full chapter

Mechanical properties of multifunc-


tional foam core materials
Amir Fathi Volker Altstädt, in Multifunctionality of Polymer Composites, 2015

8.2.2 Sandwich Bending Tests


Bending tests, on the other hand, provide means of loading the sandwich sys-
tem in authentic in-service conditions. Therefore, flexural loadings have been ex-
tensively used in the literature in order to determine the core shear response
[20,30,43,44,50,68,75,77,85–95]. In bending tests, it is possible to obtain a constant
level of shear stress in large sections of the core, as shown in Figure 8.12 [43]. This
means that by choosing correct span lengths (in order to activate core shear failure),
one can accurately measure the shear strength of the core. In Ref. [75], a comparison
has been made between the results obtained from a modified four-point bending
setup with the data from direct foam shear tests. In that work, several advantages
are mentioned for bending experiments. Higher values of shear modulus and yield
stress are reported when measured using bending tests, which were attributed to a
more uniform stress distribution in the four-point bending specimens.
Figure 8.12. Bending moment (M) and transverse force (T) in a sandwich beam
loaded in three-point bending (left) and four-point bending (right).

However, extraction of shear strain and shear modulus is not straightforward using
bending experiments. One way to use flexural tests to determine the shear stiffness
of a sandwich beam (and consequently core shear modulus) is to test the beam using
a series of three-point bendings with different span lengths as thoroughly explained
in Ref. [96]. For a given sandwich beam loaded in three-point bending, the midpoint
deflection is expressed by Eq. (8.2).

(8.2)

wt: total midpoint deflection


P: applied force
L: testing span length
D: bending stiffness of the sandwich beam
U: shear stiffness of the sandwich beam

The first term on the right side of Eq. (8.2) is the deflection due to bending and the
second term corresponds to deflection caused by shear deformations. One should
note that larger span lengths drive higher midpoint deflections by influencing both
bending and shear components. A rearrangement of Eq. (8.2) leads to Eq. (8.3),
where C is the sandwich compliance :

(8.3)

C: sandwich beam compliance


L: testing span length
D: bending stiffness of the sandwich beam
U: shear stiffness of the sandwich beam

Equation (8.3) shows a linear relationship between C/L and L2 in which 1/48D is the
slope and the y-intercept is equal to 1/4U. Therefore, the beam’s bending stiffness,
D, and shear stiffness, U, can be determined by testing the sandwich beam over a
series of different span lengths and plotting the values of C/L against L2, followed
by a data fitting step.

> Read full chapter

Advances in Applied Mechanics


Pascal Forquin, François Hild, in Advances in Applied Mechanics, 2010

3.2 Weibull parameters of the six reference materials


Bending tests (at least a dozen) were performed for each reference materials and a
distribution of failure stresses F was deduced. According to Eq. (3.1), the failure
probability of a structure Ω is function of the two Weibull parameters, namely, the
Weibull scale parameter (S0m/ 0) and the Weibull modulus (m). One way to obtain
these parameters is to deduce the Weibull modulus, for example, from Fig. 3.1
knowing the average failure stress and the standard deviation of failure stresses (Eqs.
3.5 and 3.6). It is then possible to compute the stress heterogeneity factor (Eq. 3.4)
knowing the stress field in the structure Ω. The Weibull scale parameter (S0m/ 0) is
deduced afterwards from Eq. (3.5) or (3.6). A classical alternative is to resort to the
so-called Weibull (1939) diagram in which ln[−ln(1−PF)] versus ln( F) is interpolated
by a linear function, the slope of which is the Weibull modulus m.

Three-point flexure tests were carried out on SiC-100, R-SiC ceramics, MB50 and
Ductal® concretes, and crinoidal limestone samples. The sizes of the latter and the
number of tests performed are given in Table 3.1. The Weibull modulus, the average
failure stress, and the effective volume are also reported. For the six studied ma-
terials, the Weibull parameters are significantly different. The Weibull parameters
are representative of the material microstructure and more precisely of the defect
distribution and local toughness properties (Hild et al., 1992; Jayatilaka & Trustrum,
1977). For example, Fig. 3.2 (right) shows the load/displacement curves obtained
from bending tests carried out on Ductal® concrete without fibers. The behavior
is perfectly elastic–brittle (i.e., no loss of linearity is observed until the maximum
load is reached). Consequently, the distribution of maximum loads allows one to
characterize the population of defects (in terms of activation stresses) that induce the
failure of specimens. The pictures of Fig. 3.2 (left) show two failure causes. A large
porosity (0.6 and 2 mm in diameter, respectively, for B1 and B5) is observed close
to the tensile surface of each specimen (upper surfaces of Fig. 3.2 (left)). The largest
porosities in the bulk of the specimens are the likely cause of failure. Concerning the
specimens made of crinoidal limestone, ceramics, and glass, the analysis of failure
patterns did not always allow one to identify the actual defect from which each failure
originated. The failure of R-SiC specimens is probably due to a pore as those visible
in Fig. 2.1. The failure of limestone specimens may originate from inclusions made
of silica, magnesia, sulfur, ferrous sulfide, or potassium oxides that are present in
the material. As demonstrated by Brajer, Forquin, Gy, and Hild (2003), the failure
of specimens made of soda-lime silicate glass is due to surface defects (such as
microscratches). Therefore, an effective surface is computed instead of an effective
volume (Table 3.1).

Table 3.1. Mechanical properties and Weibull parameters of the six reference mate-
rials

Material SiC-100 ce- Porous Ductal® Dry MB50 Crinoidal Soda-lime


ramic R-SiC concrete microcon- limestone silicate glass
ceramic without crete rock
fibers
Elastic properties, density, compression tests
Young’s 410 260 57 31 78 70
modulus, E
(GPa)
Poisson’s ra- 0.15 0.16 0.18 0.2 0.28 0.22
tio,

Density, 3.15 2.64 2.4 2.2 2.7 2.5

Porosity 1.8% 17.1% 1–2% ~12% <1% 0


Compres- –6500 (- Not mea- −220 (SC) −70 (SC) −147 (SC) ~−4000
sive SC) (- sured (Bernier (HEL)-
strengtha Forquin et Cagnoux,
(MPa) and Dalle,
al., 2003a) (1985)
1998)

Bend tests and Weibull parameters


Height ×- 3 × 4 × 30, 3 × 3 × 20, 11 × 10 ×- 15 × 15 × 60, 50 × 50×- Disk bend-
 width ×- 65 19  130, 18 20  150, 40 ing, 400
 span (mm3-
), number of
tests
Mean ten- 360 113 21.9 9.99 19.0 ~94
sile failure
stress, w(Z-
eff) (MPa)
Effective 1.25 0.084 53 39.9 360 Seff = 100-
volume
(mm3)/sur-  cm2
face (mm2)
Mean ten- 368.5 [1- 103 [1 mm3] 31.6 [1 mm3] 13.6 [1 mm3] 24.8 [1 mm3] 350 [1 mm2]
sile failure  mm3]
stress (MPa)-
[Veff ], [Seff ]
Weibull 9.6 26.5 11 12 22 ~7
modulus, m

References Hild et al., Forquin et Grange et Oakley


(2003) al., (2003) al., (2008) (1996)
Forquin Forquin
and Hild., and Erzar,
(2008) (2009)

a SC: minimum nominal axial stress under simple compression; HEL: Hugoniot
elastic limit (corresponding to the axial stress level during plates impact when
the elastic limit is reached).

Fig. 3.2. Results of three-point bend tests performed with Ductal® concrete without
fibers and two postmortem observations.

The Weibull modulus is mainly related to the decay for large defect sizes of the
corresponding probability density function (Hild et al., 1992; Jayatilaka & Trustrum,
1977), which is usually described by a power law whose exponent is a linear function
of the Weibull modulus. Consequently, different Weibull moduli, as observed for
the studied materials, are an indication of different defect populations, in particular,
for the largest defect sizes, i.e., the ones likely to initiate failure under quasi-static
loading conditions. For a given reference density 0, different stresses S0 are mainly
induced by different toughnesses and average defect sizes. Different defect pop-
ulations will therefore lead to different Weibull parameters. Conversely, different
Weibull parameters are indicators of different defect distributions. When dealing
with dynamic fragmentation, the Weibull model itself can no longer be used since
a weakest link hypothesis does not apply; however, the following microstructure
model using the Weibull parameters is still considered.

A microstructure model is now introduced to account for random distributions


of defects. It has to be valid for describing single and multiple fragmentation
regimes. Its advantage then lies in the fact that quasi-static experiments may be
used to identify the material parameters. Once they are determined, they may be
used in a situation where single fragmentation no longer occurs (e.g., dynamic
fragmentation). The material is assumed to contain point defects of density t. Such
hypotheses are those of a Poisson point process of intensity t(Denoual & Hild, 2000;
Gulino & Phoenix, 1991; Jeulin, 1991). Consequently, the probability of finding
defects within a uniformly loaded domain Ω of size Z reads

(3.8)

so that t corresponds to the average number of defects within Ω. By definition,


t is related to the density t by t = Z t for a uniformly loaded domain. Let us now

assume that t is a function of the applied stress 1. The larger the applied stress,
the greater the number of defects that will initiate cracks. One possible choice to
account for this trend is given by a power-law function of the maximum principal
stress:

(3.9)

where m and S0m/ 0 are interpreted as the Weibull parameters when single fragmen-
tation occurs. By using the weakest link hypothesis, the failure probability PF is the
probability of finding at least one defect within Ω when F =  1 > 0:

(3.10)

when a uniform stress is applied. If the stress field is heterogeneous, t is related to


t by

(3.11)

and the failure probability reduces to Eq. (3.1) and describes the fact that the larger
the volume, the smaller the mean failure stress. The Poisson–Weibull model allows
one to relate the Weibull parameters to microstructural properties describing the
population of initiation sites. The latter is the key for understanding probabilistic
features related to the fragmentation of brittle materials.

> Read full chapter

Monitoring Load Events


Carosena Meola, ... Giovanni maria Carlomagno, in Infrared Thermography in the
Evaluation of Aerospace Composite Materials, 2017

5.5 Quasistatic Bending Tests


Quasistatic bending tests are performed owing to a three-points configuration with
the load applied at the centre of the specimen, which is subjected to bending under
the pushing force until collapse. In this case, an infrared imaging device may be used
to visualize temperature variations over the specimen surface or over a lateral side
(through its thickness). Any temperature rise (hot spot) is a symptom of formation
of cracks. More specifically, it is possible through acquisition of images in time
sequence to follow the evolution of cracks from their origin and successive grow up.
This may be useful to get information about the material performance.

As an example, some thermal images, taken through the thickness, during bending
at a speed of about 1 m/s are shown in Fig. 5.26 together with the distribution of
ΔT values against the displacement DF. For DF = 8 mm, on the left of the punch, it is
possible to recognize the presence of light tracts that account for local temperature
rise due to formation of cracks. These tracts become lighter as the temperature
increases and the cracks enlarge until collapse of the specimen for DF = 20 mm. It is
worth noting that these are only preliminary tests; some results have been presented
for the sake of completeness.

> Read full chapter

Atomistic-Continuum Theory
K.M. Liew, ... Lu-Wen Zhang, in Mechanical Behaviors of Carbon Nanotubes, 2017

4.5.12 Bending Test


A bending test is carried out by incrementally rotating the two end planes of the
SWCNTs in opposite directions (Fig. 4.25). During the rotation of the two end planes,
the axial movement of one end is prohibited, but that of the other is unconstrained.
A (15, 0) SWCNT is considered. It has 30 hexagonal cells along the axial direction,
and the original length is 12.87 nm. The rotation angle is 1 degree per loading step,
and a total of 21 steps are applied. The purpose of the present computations is to
test the convergence of the solutions and the effect of the scaling factor .

Figure 4.25. A SWCNT is bent by rotating the two end planes in opposite directions.
First, is fixed as 2.0. Fig. 4.26 shows the variation of the system energy versus the
bending angle for the cases of , , , and nodes. Good convergence can be seen.
To check the effect of the DOI, detailed computations are carried out for Dmax=2.0,
2.5, 3.0, and 3.5. Table 4.7 shows the dependence of the solution on Dmax and the
number of nodes. When fewer nodes are used, Dmax=2.0 has a bad result. However,
when more nodes are used, Dmax in the range of (2.0, 3.0) does not show very large
differences.

Figure 4.26. The strain energy versus the bending angle.

Table 4.7. The Dependence of the Strain Energy on the Scaling Factor and the
Number of Nodes

Node Numbers Dmax=2.0 (eV) Dmax=2.5 (eV) Dmax=3.0 (eV) Dmax=3.5 (eV)
26.4863 19.2940 19.1104 19.0163
19.9848 18.1101 18.1191 17.9542
18.3417 17.9120 17.9631 17.8662
18.3100 17.8917 17.9441 17.8388
18.0450 17.8907 17.9241 17.8660
18.0212 17.8756 17.8947 17.8037

The results are obtained at a bending angle of 20 degrees.

> Read full chapter

Experimental study of dynamic and


static damage failure of concrete dam
based on acoustic emission technology
Chen Houqun, ... Dang Faning, in Seismic Safety of High Arch Dams, 2016
13.2.4.1 Analysis of acoustic emission activity
Dynamic bending test was carried out on wet-sieved concrete in Xiaowan arch dam
with different loading rates, with the load time 201.10, 21.20, 2.33, and 0.29 s,
respectively. The mechanical properties and acoustic emission activity characteristics
in the flexural tensile failure process under different loading rates are shown in
Table 13.12. The acoustic emission hits, strength-loading rate curve, and cumulative
acoustic emission hits–load curve throughout the loading process are shown in Figs
13.23 and 13.24. It can be seen in Fig. 13.23 that the resonant sensor adopts more
acoustic emission hits than the broadband sensor does, primarily due to the higher
sensitivity of the resonant sensor.

Table 13.12. The Mechanical Properties and Acoustic Emission Activity Characteris-
tics in the Flexural Tensile Failure Process Under Different Loading Rates

Specimen Loading Ultimate The Cumu- The Cumu- Hit rate Level Stress
Number Time (s) Load (kN) lative Num- lative Num- Peak Occurs
ber of WD ber of RD (times/s) When AE (%)
Hits (times) Hits (times)
DC-01 201.1 49.3 1577 3340 215 17.6 (8.7 kN)
DC-02 21.2 52.7 481 1245 460 48.6 (25.6
kN)
DC-03 2.33 57.3 174 433 859 73.5 (42.1
kN)
DC-04 0.29 65.6 51 69 1400 73.5 (48.2
kN)

Note: The table “WD” represents broadband sensor results; “RD” represents average
value of multiple resonant sensor results.
Figure 13.23. The Correlation Curve of AE Hits, Strength, and Loading Rate in the
Flexural Tensile Failure Process of Wet-Sieved Concrete Under Different Loading
Rates
Figure 13.24. The Correlation Curve of the Cumulative Number of AE Hits and the
Load in the Flexural Tensile Failure Process of Wet-Sieved Concrete Under Different
Loading Rates

It can be seen from these figures and tables that the higher the loading rate is,
the less the acoustic emission hits are, the higher the peak hit rate is, the later
the concentration of emergence of acoustic emission hits are, and the higher the
corresponding stress level is. Accordingly, with an increase of the loading rate, the
whole failure time of concrete becomes shorter, and its flexural strength becomes
higher. The result shows a significant strain rate effect.

Acoustic emission rate variation with time is called acoustic emission mode (Hong-
guang et al., 2000). The curve of acoustic emission hit rate varying with stress level
under the different loading rates is shown in Fig. 13.25. It can be seen from the
figure that when the loading rate increases by 10-fold of the magnitude, the peak
acoustic emission hit rate increases by twofold of the magnitude. This indicates that
there is a good correlation between acoustic emission hit rate and loading rate.
Figure 13.25. The Curve of Acoustic Emission Hit Rate Varying with Stress Level
Under the Different Loading Rates(a) Loading speed: 0.25 kN/s (b) Loading speed:
2.50 kN/s (c) Loading speed: 25.00 kN/s (d) Loading speed: 250.00 kN/s.

> Read full chapter

POSITION BUTT WELDING OF


TECHNOLOGICAL PIPELINES OF
POWER ENGINEERING STRUC-
TURES
I.K. Pokhodnya, ... L.N. Orlov, in Welding in Energy-Related Projects, 1984

The impact bending tests in the temperature interval from 20 down to −60°C
were performed to determine the critical transition temperature. The notch was
oriented to the weld center and along the HAZ. Fig. 3 and 4 give the curves of the
impact toughness dependence on the testing temperature. The critical transition
temperature was determined from the impact toughness level of 50 J/cm2. The
welded joint made on 20 steel has the lowest transition temperature −43°C (notched
along the weld); −40°C (notched along the HAZ).

Fig. 3. Dependence of impact toughness of the welded joint on test temperature


(notch along the weld): 1-15 0 steel; 2-16 C steel; 3-20 steel.

Fig. 4. Dependence of impact toughness of a welded joint on test temperature (notch


in the HAZ). 1-15 C steel; 2- 16 C steel; 3-20 steel.

> Read full chapter

Interfacial and Nanoscale Fracture


W. Gerberich, W. Yang, in Comprehensive Structural Integrity, 2003
8.01.5.2.3 Bending mode and rippling
The continued bending test of Wong et al. (1997) of the nanotubes led to an interest-
ing elastic buckling pattern, featuring a rippling mode for multiwall nanotubes. MD
found some difficulties to tackle on this problem, since the interatomic potential
among different layers of a multiwall nanotube remains elusive. Liu et al. (2001)
investigated the burst of an unusual bending mode during the multiwall nanotube
measurement that corresponded to rippling on the inner area of the bent nanotubes.
They carried out nonlinear analysis featuring highly anisotropic behavior: stiff against
tension parallel to the wall and compliant against shearing and tension normal to
the wall. They were able to capture the rippling mode (shown in Figure 23(a)) that
suggested that the effective Young's modulus could indeed decrease substantially
with increasing diameter, and that the results from the classical linear theory may
be invalid in such measurements (shown in Figure 23(b)).

Figure 23. Rippling of a multiwall nanotube under bending and the corresponding
nonlinear response (source Liu et al., 2001; courtesy of Q. S. Zheng).

> Read full chapter

Acoustic emission and impact–echo


techniques for evaluation of reinforced
concrete structures: a case study
M. Ohtsu, in Non-Destructive Evaluation of Reinforced Concrete Structures:
Non-Destructive Testing Methods, 2010

Outline of experiments
During the bending test of the beam shown in Fig. 24.15a, several surface cracks
were observed in the bending span. Cracks were extended in a zigzag manner with a
crack depth of approximately 200 mm. Here, two surface cracks are selected as illus-
trated in Fig. 24.25. One is a multiple crack, and the other is a zigzag crack. Before
the test, the velocity of the P wave was measured as 4065 m s−1. After shooting the
aluminum bullet against the top surface of the specimen with a compressed pressure
of 0.05 MPa, surface displacements resulting from the impact were recorded by an
accelerometer on the top.

24.25. Bending cracks observed in the reinforced concrete beam: (a) multiple crack
and (b) zigzag crack.

Firstly, the impact test of the conventional SIBIE method was conducted by
one-point detection. The distance between input and output was set at 100 mm
across the crack. Secondly, the impact test for the scanning SIBIE method was
conducted with three pairs of impact and detection points with 5 cm interval. Fourier
spectra of accelerations were analyzed by FFT. Sampling time was 4 μs and the
number of digitized data for each waveform was 2048.

> Read full chapter

FAILURE ANALYSIS AND EXPERI-


MENTAL STRESS ANALYSIS OF A
THREADED ROTATING SHAFT
R.B. TAIT, in Failure Analysis Case Studies II, 2001

3.1 Bending test series


For the bending test series five tests were initially undertaken with the trip limit
switch overridden, so that the shaft rotation was nominally arrested by the action of
disc brake against the brake pads. The output from the strain bridge was recorded
on the digital storage scope which also had the capability of producing a hard copy
of the stored image. The amplitude settings were 0.5 s/cm horizontally and 0.2 V/cm
vertically. It was expected that these tests would yield an oscillating signal about the x
axis which would gradually increase in stress amplitude, as the brake engaged, until
the wormshaft was arrested.

This type of strain trace was indeed obtained, of which test number 3 (Fig. 4(a)) is
typical, but with one major variation—the trace exhibited a substantial offset strain
and did not oscillate about the mean position. These traces may be regarded as
composed of 3 regimes which may be characterised as follows (refer to Fig 4(a)).
Region I corresponds to the free rotation of the shaft under full speed conditions.
Small (noise) oscillations about the mean indicate the gauges were responding but
there was no bending in the shaft. As the torque limiter brake was applied (region II)
individual oscillations in the trace at the frequency of rotation (period 0.042–0.05 s)
were observed. These increased in amplitude as the brake engaged, until the brake
effectively arrested the shaft and load was maintained. The curious observation,
however, was the large offset strain of approximately 178 μ (microstrain) in this case
(Fig. 4(a)). This will be discussed in greater detail later. Region III is the appropriate
constant load condition, once the shaft had been arrested. When the power was
dropped completely the strain trace returned to zero and the x axis.
Fig. 4. Strain time traces for the strain gauged shaft on the Koeberg Rotork test bench
for (a) bending and arrest against the disc brake, (b) tension and arrest against the
disc brake. Note the significant offset.

When the limit switch system was engaged (i.e. not overridden) then the disc brake
torque limiter was not put under load. The recording of the bending stress for this
case indicated minimal stresses (less than 5 MPa) (Fig. 5(a)) as expected.
Fig. 5. Strain time traces for limit switch operation and arrest for (a) bending and (b)
tension. In both cases the resultant stresses are small.

Although the number of bending tests was limited due to time constraints, it is
nonetheless useful to interpret these bending strains in terms of (i) the measured
offset; (ii) cyclic amplitude; as well as (iii) peak strain, together with the consequent
derived stresses. These stresses, with the torque switch limiter overridden, are ef-
fectively equivalent whether derived from a modulus viewpoint or from a calibration
curve approach and the data is summarised in Table 1. The data in the table is
considered to be accurate to within approximately 8% and from this data it can be
inferred that, under nominally “normal” bench test loading conditions it is possible
for significant peak bending stresses of approximately 170–200 MPa to occur.
Such cyclic stresses are not trivial, especially when one takes into account the stress
concentrating effects of the threads, from which it would appear fatigue at the thread
roots is highly likely. The inherent cyclic stress is, however, typically less than 50 MPa,
which seems reasonable from a design viewpoint.

Table 1. Bending strains and stresses from the bending bridge configuration, in-
cluding the disc brake torque limiter

Test number Measured Measured Measured Stress of- Cyclic Peak stress
strain cyclic strain peak strain fset (MPa) stress (MPa) (MPa)
offset after (μ ) (μ )
stall (μ )

1 1567 502 2310 130 42 191


2 1732 630 2420 143 52 200
3 1780 583 2169 147 48 180
4 2024 575 2288 168 48 189
5 1970 620 2222 163 51 184
6 1582 625 2156 131 52 179
Mean 1776 589 2261 147 48.8 187

> Read full chapter

ScienceDirect is Elsevier’s leading information solution for researchers.


Copyright © 2018 Elsevier B.V. or its licensors or contributors. ScienceDirect ® is a registered trademark of Elsevier B.V. Terms and conditions apply.

You might also like