Download as pdf or txt
Download as pdf or txt
You are on page 1of 22

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/344441257

Mechanical, thermal and physico-chemical behavior of virgin and


hydrothermally aged polymeric pipes

Article  in  Journal of Thermoplastic Composite Materials · October 2020


DOI: 10.1177/0892705720962167

CITATIONS READS

10 919

5 authors, including:

Houcine Jemii Dalila Hammiche


Ecole Nationale d'Ingénieurs de Sfax Université de Béjaïa
7 PUBLICATIONS   30 CITATIONS    102 PUBLICATIONS   380 CITATIONS   

SEE PROFILE SEE PROFILE

A. Boubakri Noamen Guermazi


Ecole Nationale d'Ingénieurs de Sfax Higher Institute of Applied Science and Technology of Kasserine
15 PUBLICATIONS   426 CITATIONS    73 PUBLICATIONS   1,071 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Characterization of a composite based on poly-lactide acid reinforced with waste from an agrifood product View project

biocomposites View project

All content following this page was uploaded by Houcine Jemii on 01 October 2020.

The user has requested enhancement of the downloaded file.


Original Article
Journal of Thermoplastic Composite
Materials
Mechanical, thermal 1–21
ª The Author(s) 2020
and physico-chemical Article reuse guidelines:
sagepub.com/journals-permissions

behavior of virgin DOI: 10.1177/0892705720962167


journals.sagepub.com/home/jtc

and hydrothermally
aged polymeric pipes
Houcine Jemii1 , Dalila Hammiche2 , Abid Boubakri1,
Nader Haddar1 and Noamen Guermazi1

Abstract
In this paper, an experimental study was conducted to characterize industrial PVC pipes
and to investigate the effect of hydrothermal aging on their physico-chemical, thermal,
and mechanical behavior. Three temperature (25 C, 60 C and 90 C) and full immersion
in distilled water were retained as accelerated hydrothermal conditions. Kinetic of water
absorption was examined and Fickian behavior was observed. The aging temperature
was found to influence the water uptake behavior of PVC samples. Thermogravimetric
analysis (TGA) has proved that the pipe material is not pure, while it consists of PVC
reinforced with calcium carbonate (CaCO3). After exposure to accelerated aging, TGA
and FTIR analysis exhibit preliminary signs of degradation of PVC samples under the
retained conditions. Changes affecting the shape and the color of aged samples were
examined. Mechanical properties have been characterized, after immersion of 30 days,
with an improvement of strength and stiffness of the aged samples, in particular at
elevated aging temperature. However, the aging response is accompanied by a loss of
ductility for the aged material. These results, even for brief exposure, could help to
understand the behavior of PVC composite pipes under hydrothermal conditions.

Keywords
PVC pipes, durability, hydrothermal aging, physico-chemical properties, mechanical
behavior

1
Laboratoire de Génie des Matériaux et Environnement (LGME), Ecole Nationale d’Ingénieurs de Sfax (ENIS),
Université de Sfax, Sfax, Tunisia
2
Laboratoire des Matériaux Polymères Avancés, Faculté de Technologie, Université de Bejaia, Bejaia, Algeria

Corresponding author:
Houcine Jemii, Laboratoire de Génie des Matériaux et Environnement (LGME), Ecole Nationale d’Ingénieurs
de Sfax (ENIS), Université de Sfax, BP 1173-3038, Sfax, Tunisia.
Email: houcine@webmails.com
2 Journal of Thermoplastic Composite Materials XX(X)

Introduction
The use of polymeric pipes is conceivable in various applications due to both environ-
mental, and economic benefits (low cost of installation and maintenance). Over the years,
these polymer pipes were widely used for petroleum, gas and water transportation and
distribution.1–3 From the used polymer materials, the thermoplastics are the best candidate,
such as the polyvinyl chloride (PVC),2 the polyethylene (PE),4–6 the polypropylene (PP).7
However, the service life of these pipes remains dependent on the service conditions.
They can be exposed to several factors or loadings reducing their service life, such as
internal, or external pressure, attack of hot or cold fluid (gas or liquid). In this view,
several studies were conducted about the investigation of the mechanical characteriza-
tion of polymeric pipes and the evaluation of properties deterioration.2–4 Here, the main
common goal is to enlarge their potential application. In 2005, Krishnaswamy8 carried
out creep testing on high-density polyethylene (HDPE) pipes, the results have shown that
they are ductile and brittle materials. It is also to be noted that for most PVC applications,
fillers are added to reduce formulation cost and to enhance the material performance. For
instance, calcium carbonate (CaCO3) particles have been used as fillers in thermoplastics
and especially in PVC. The purpose of adding CaCO3 particles is to lower the material
costs of the finished products.9 The CaCO3 particle size is generally comparable in size
with particles of PVC and hence are capable of generating interparticle friction with PVC
particles, increasing shear, and promoting gelation.10 CaCO3 could also act as a fusion
promoter. One possible explanation for the observed trend could be that the finer par-
ticles having a larger surface area would have a larger contact surface with the PVC
particles, creating a higher shear heating and increasing the fusion speed.11 CaCO3 is
usually surface treated with stearic acid, producing a layer of calcium stearate on the
filler surface. This results in improved processing and dispersion (mechanical properties)
and improved moisture resistance.12
In 2016, Guermazi et al.2 have focused on the effects of the content of CaCO3 fillers
in PVC matrix and environmental conditions on water absorption, thermal stability,
mechanical properties, and wear properties of PVC/CaCO3 composite pipes. Besides,
many researchers have focused on the study of the durability of the polymer pipes which
were initially manufactured based on the service life of 50 years, as reported in Ref.13
Here, the use of accelerated conditions is more suitable method to analyze the durability
of pipes. In particular, various studies have retained hydrothermal aging methodology as
aggressive conditions to provoke the rupture or failure of pipes damage. Different
temperatures or/and different aging environments (humidity, distilled water, seawater)
were widely used at laboratory scale to study materials aging.5,14
From the literature review, polymeric materials become plasticized, softened, and
swelled when exposed or subjected to humidity.15 Studies dealing with the effects of
moisture absorption on the properties have revealed that moisture absorption affects the
material strength, leading to the degradation of the material’s mechanical properties.16
Recently, Mei et al.17 have found that mechanical properties, in terms of average
strength and average stiffness, of carbon fiber reinforced plastics (CFRP) composite
decrease considerably as a function of aging time under 50 C hydrothermal environment.
Jemii et al. 3

According to Larbi et al.18 and Aldajah et al.,19 the hydrothermal aging of polymeric
materials can induce a loss of rigidity and a loss of resistance to breaking, while a gain of
ductility (larger elongation or strain at break) due to swelling phenomenon.
Bao20 has recently shown that with the prolonging of hydrothermal aging period, the
mechanical properties of CMR/PLA bio-composites were decreased. In particular, the
bending and tensile strength characteristics of composites were dramatically affected by
both time and aging temperature: the higher the temperature, the more significant
decrease in the material performance. In general, absorption or diffusion of water into
polymeric material degrades and weakens the strength material. Also, if the polymeric
material plasticizes, mechanical properties will be affected but this depends on some
parameters such as time, temperature, relative humidity, and external stress.21,22
Our goal through this work is (i) to conduct physico-chemical, thermal, and
mechanical investigations of industrial PVC pipes and (ii) to investigate their properties
under the combined effects of two parameters namely temperature and relative humidity.

Materials and techniques


Material and samples
The material used in the current study consists of polymeric pipes based on polyvinyl
chloride (PVC) reinforced with calcium carbonate (CaCO3) fillers, and some other
additives (plasticizer and stabilizer). In particular, the commercial dioctyl phthalate
(DOP) was used in the formulation of pipe material as plasticizer.
The polymeric pipes, which used for water boreholes, are produced by the extrusion
process making use of an extruder with twin-screw at counter-rotating mode. Other details
were given in our previous paper.23 Samples of PVC pipes were supplied for this study
having a 200 mm external diameter and 8 mm of thickness. Samples, dedicated to char-
acterization, were prepared by a DMTC-PX1-CNC milling machine according to ISO
2818:1994 after cutting strips from the pipe section according to ISO 6259-1:2015 standard.
Dimensions of tensile samples were chosen according to the ASTM D638-M-I
specifications. Figure 1 illustrates the dimensions of the pipes, the dimensions of the
samples, and how they were extracted.

Water uptake measurements


Here, a set of pipe specimens were undergone to accelerated hydrothermal aging. Hal-
tered and flat samples were immersed in an aqueous environment (distilled water) and
aged at three different temperatures (25 C, 60 C and 90 C) for around 90 days. Here, it
is true that the maximum of service temperature for PVC pipes does not exceed 90 C, but
such temperature was selected to simulate accelerated weathering conditions since the
study of natural aging is time consuming.14
The moisture diffusion into the sample was then measured. Water uptakes were
measured through a gravimetric method in which samples were taken out from the water,
wiped with filter paper to eliminate surface water, weighted at regular intervals with the
digital balance of 104 g accuracy, and then immersed again in water.
4 Journal of Thermoplastic Composite Materials XX(X)

Figure 1. Extraction method and dimensions (in mm) of PVC pipe and samples. (a) Dimensions of
pipe; (b) extraction of samples from tube; (c) sample and strip; and (d) sample dimensions.

The water uptake rate, (Mt(%)), in the samples was determined using the standard
gravimetric equation (equation (1)),20
 
Wt  W0
Mtð%Þ ¼  100 ð1Þ
W0

where Wt and W0 represent the weight at time t and t ¼ 0 (before immersion), respectively.
Jemii et al. 5

Thermogravimetric analysis (TGA)


Thermogravimetric analysis (TGA) and derivative thermal gravimetry (DTG) were
investigated by using a TGA-DTG (STA PT 1600, LINSEIS, GERMANY). The analysis
was performed under nitrogen flow (20 ml/min) at a temperature interval of 25 at 700 C.
The used weight was about 10–20 mg and the rate of heating is 20 C/min.

Fourier transform infrared spectroscopy (FTIR)


Fourier transform infrared spectroscopy (FTIR) spectra of the unaged and aged PVC
samples were recorded using an FTIR (Cary 630 FTIR, Agilent technologies) in the
range of 4000–400 cm1 with a resolution of 4 cm1. Thin films, which were between
0.5 and 1 mm in thickness, were cut from the tube sample for the analysis.

Tensile tests
Tensile tests were performed at room temperature using a universal testing machine of
the type WDW-10 E with a crosshead rate of 5 mm/min. Five measurements were
conducted and an average for the final result was considered.
Mechanical properties obtained from tensile tests of PVC samples were determined for
different immersion temperatures (25 C, 60 C and 90 C) after regular intervals of time.

Results and discussion


Kinetics of water absorption
Figure 2 shows the behavior of PVC samples immersed in the water according to the
aforementioned conditions. Curves show that the specimens immersed followed the
same tendency of the weight change. The three curves exhibit the same behaviors with
two stages of absorption: in the first one, the sorption curve is a linear function of the
square root of time (Fick’s law of diffusion). Here, it is to note that such curve slope was
used to estimate the diffusion coefficient (D), as follows14,24:

 
M ðt Þ 4 D:t
¼  ð2Þ
M ð1Þ h p
where M1 represents the maximum moisture content, h is the thickness, and t is the
aging time.
Afterward, in the second stage, the equilibrium water uptake seems to be reached and
consequently, the sample is saturated. It is seen that the higher the immersion tem-
perature, the greater the mass gain. According to Ref.,25 the temperature favors the
movement of the polymer chains which facilitates the penetration of the water. Thereby,
the temperature of the absorption process has sway on the water absorption curves.
Table 1 regroups the obtained data of diffusivity (D), saturation time (ts) and maxi-
mum moisture content, (M1) of water for the three temperatures.
6 Journal of Thermoplastic Composite Materials XX(X)

Figure 2. Moisture content (%) as function of immersion root time (h0.5) of PVC samples aged at
different temperatures (25 C, 60 C and 90 C) in distilled water for around 90 days.

Table 1. Characteristics extracted from absorption curves for PVC pipe (8 mm).

Temperature, Water diffusion Saturation time, Maximum moisture


T ( C) coefficient, D (m2/s) ts (days) content (M1)

25 1.49  107 15 1.52


60 7.73  107 10 1.75
90 2.49  106 6 2.19

As has been pointed from Table 1, the maximum moisture content increased con-
siderably when hydrothermal aging temperatures increased. This tendency agrees with
the result obtained by Bao when studying the hydrothermal aging behavior of CMR/PLA
biocomposites at different temperatures.20 In addition, the diffusion coefficient increases
significantly with the immersion temperature and reaches a remarkable value at 90 C.
This confirms that the diffusion is a thermally activated process and the diffusivity is
very sensitive to temperature.26 The increase of aging temperature improves segmental
mobility which induces a greater zone of activation, consequently, the rate of water
absorption raises.
In other words, the diffusion mechanism in polymer materials is shown to be highly
dependent on the conditioning temperature.27–29 Indeed, the temperature plays a role of
Jemii et al. 7

-12
y = -4.689x + 0.008
-12.5 R² = 1

-13 Ea = 39.01 kj mol-1


-13.5

Ln(D)
-14
-14.5
-15
-15.5
-16
2.6 2.8 3 3.2 3.4
1/T ( x 0,001/°K)

Figure 3. Arrhenius plot for PVC samples aged in distilled water at three temperatures (25, 60
and 90 C).

activator of the absorption mechanism obeying the law of Arrhenius whose activation
energy “Ea” can be determined by drawing the curve Ln(D) ¼ f(1/T), using the Arrhenius
relation:
 
Ea
D ¼ D0 exp ð3Þ
RT
where D0 is a constant, Ea, R, and T are activation energy, the gas constant, and absolute
temperature, respectively.
Figure 3 shows a plot of lnD versus reciprocal of absolute temperature “1/T.” The
diffusivities show a linear Arrhenius plot over the three temperatures. The obtained
activation energy is about 39.01 kJmol1 and this value seems to be reasonable compared
with literature values.30,31 Similarly, in Ref.,26 it was reported that typical activation
energies of moisture diffusion in polymer composites vary from 20 to 80 kJmol1.

Thermogravimetric analysis of PVC pipes


Thermogravimetric analysis (TGA) was conducted to assess the thermal stability of the
samples at different temperatures. Dynamic thermogravimetric and the derivative
thermogravimetric (DTG) curves are shown in Figures 4 and 5.
Over the analyzed samples, until 250 C, all curves remain constant and there is no
variation in weight, this means the samples are thermally stable in this range of temperature.
However, for the unaged PVC sample, which is considered as reference, it can be
observed that there are two weight loss stages. In the first stage, the decomposition starts
at about 253 C and ends around 330 C, as summed up in Table 2. From 400 to 500 C, a
second decomposition stage is observed, much shorter than the first one, and corre-
sponding to polyacetylene cracking.32 Finally, a stable residue can be formed when the
temperature is beyond 550 C.
It is to note that the DTG curves show a peak in the range of 253–330 C, which
corresponds to the “temperature of maximum rate degradation.” Such peak arises at a
8 Journal of Thermoplastic Composite Materials XX(X)

Figure 4. TGA curves of unaged and aged PVC samples after 30 days of immersion in distilled water.

Figure 5. DTG curves of unaged and aged PVC samples after 30 days of immersion in distilled water.

bass temperature for aged samples. This can be explained by the effect of water on the
polymer structure. The later seems to be softer with the absorption of water, which
directly affects the thermal stability.
Jemii et al. 9

Table 2. Thermal properties of unaged and aged samples (30 days) of PVC pipe obtained from
TGA and DTG analysis.

Samples T5% ( C) T20% ( C) T50% ( C) Tmax ( C) Char residue (%)

Unaged PVC samples 278 298 540 288 42.5


Aged at 25 C 274 296 454 286 34.3
Aged at 60 C 266 275 430 276 31.9
Aged at 90 C 215 239 393 270 28.6

T5%: temperature at 5% weight loss; T20%: temperature at 20% weight loss; T50%: temperature at 50% weight
loss; Tmax: temperature of maximum reaction rate.

The decrease in the thermal stability of the PVC after long-term water immersion was
attributed to the network structure change, driven by plasticization which allows
migration of the stabilizers. This finding is conforming well to those found in Ref.33
Also, it can be observed from Figure 4, even at high temperature, 700 C, a percentage
of material samples remains intact, around 40%. It is probably an inorganic material in
the PVC pipes.
This suggests an inorganic filler used in the PVC matrix. According to the literature,34
such residual mass can correspond to the carbonate of calcium (CaCO3). Therefore, it
will be important to confirm this point in the following section.
The results referred to in Table 2 show a lowering in the thermal parameters (T20, T50,
and Tmax) at a higher aging temperature (90 C) compared with that of the 25 C. The
temperature has a significant effect on thermal stability as previously described in Ref.35
Thermal stability and HCl release reduction in PVC formulation is improved by using
fillers (CaCO3).36 From Table 2, we note that the char rate decreases as exposure
temperature increased indicating the decrease of the amount of filler in PVC after
migration caused by the mobility of macromolecules. This explains the lowering of the
stability of the polymeric samples.

Infrared spectroscopy characterization of PVC pipes


Figure 6 represents the FTIR spectrum for the unaged PVC sample. The absorption
bands observed at 610, 635, and 690 cm1 are attributed to the stretching vibration
modes of the (C–Cl) bond. The peaks at 720 and 729 cm1 reveal that all the previous
structures belong to long-chain hydrocarbon of four or more methylene units. These
vibrations have been identified by others to belong to the long aliphatic chains existing in
the lead-based stabilizer as reported in Refs.37–39
The vibrations at (847, 874 cm1) are those of the bonds of the (C–O–S) struc-
ture.40,41 In the same region, there is the CO3 (875 cm1) belonging to calcium carbonate
(CaCO3).42,43 This funding confirms that the material of the PVC tube is not pure. It
consists of a PVC matrix containing CaCO3 used as fillers.
An intense peak at 1425 cm1 assigned to (CH2, COO) of the stabilizer system, and
(C–O) of calcium carbonate (CaCO3). Similarly, this result indicates that the material of
10
Figure 6. FTIR spectra of unaged PVC samples.
Jemii et al. 11

Table 3. Characteristic bands for the used plasticizer.

Frequency (cm1) Assignment

1725 C¼O (ester)


1580 C–C stretch aromatic
1390 CH3 bend

the PVC pipe also contains a lead-based stabilizer system. Finally, the absorption bands
at (2850–2918 and 2874–2956 cm1) belong to symmetric stretching vibration and anti-
symmetric stretching vibration of the (C–H) bond. Such observed peak was already
described in a similar study.44
The spectrum also shows the typical doublet of the phthalates at 1580 cm1. The
bands corresponding to the carbonyl and ester group, as well as those corresponding to
the aromatic and aliphatic C–H bonds can also be observed and Table 3 shows these
bands, according to Refs.45,46
The spectrum of PVC before and after aging was compared. This allowed the iden-
tification of some characteristic bands which are related to the migration of plasticizer
present in the formulation. The intensity of characteristic bands will change. An increase
in the intensity corresponds to the absorption of liquid through the PVC samples and
inversely, a decrease in intensity corresponds to the migration of one or more compo-
nents in the water.47
There is a decrease in absorbance after aging. This indicates that the migration of
plasticizer occurred at both temperatures aging. In the case of high temperature, it seems
that the penetration of the water which was observed in the rate of mass variation favored
the migration of additives. Indeed, the lowest absorbance were obtained for the for-
mulation aged at 90 C; therefore, the highest rates of migration.
The superposition of the spectra of unaged and aged samples in Figure 7 shows that all
the bands present in the unaged one are also recorded in the aged specimen but with some
difference in intensity of absorbance.
It was noticed, from Figures 6 and 7, that the intensity of the (C–Cl) peaks at (635 and
690 cm1) and those of (CH, CH2) at (955, 1197, 1253, 1329, 1425 cm1) decrease with
increasing the immersion temperature (from 25 to 60 C). Further, Figure 7 indicates that
the peak intensity between 1600–1680 cm1 increases when the aging temperature
grows. This region corresponds to the stretching vibration of the C¼C conjugated group
and the C¼C in conjugated group stretching vibration.48 This can be explained by a
dehydrochlorination process.44
The increase in peak intensity (875, 965, 1101, and 2853 cm1) for the sample aged at

90 C can be explained by the fact that the additives (plasticizer and stabilizer) have
migrated from the core of the samples to the surface. The same result was found by
Jakubowicz et al.,49 when studying the effects of accelerated and natural aging on
plasticized polyvinyl chloride (PVC). They have confirmed the mass loss of PVC
samples by a dominant loss of plasticizer.
12
Figure 7. FTIR spectra of unaged and aged PVC samples.
Jemii et al. 13

Figure 8. Discoloration of aged PVC after immersion in distilled water for 30 days at different
temperatures.

Figure 9. Shape change of aged PVC after immersion in distilled water for 30 days at different
temperatures.

The spectrum of the aged PVC specimen at 25 C and 90 C present specific char-
acteristics of absorbed water put in evidence by the picks at 3400 and 1600 cm1. That is
obvious by the presence of the stretching vibrations and the bending modes of the water
molecules as reported by Hammiche and their collaborators.25 These results correlate
well with the water uptake ones as described in the absorption section.

Mechanism of yellowing and shape change


In this context, the identification of the mechanism of PVC samples yellowing and shape
change was studied after 60 days of immersion at different temperatures. This investi-
gation displays that there are obvious signs of degradation of PVC samples on view to
temperature and the time of immersion, including the color and shape changes as illu-
strated in both Figures 8 and 9, respectively.
As it is clear from Figure 8, at room temperature (25 C) and after medium
immersion periods, PVC samples do not change its color. However, when exposed
to high temperatures (90 C) in full immersion PVC samples change its color. So,
extended exposure can cause yellowing too. Based on the literature,50 such dis-
coloration can result from the interaction of additives or compounds in the polymer
14 Journal of Thermoplastic Composite Materials XX(X)

formulation and their accumulation on the material surface. It equally reveals a


degradation process (breaking and rearranging of chemical bonds). Beside, such
discoloration endures after the air drying. This result was in good agreement with
the finding of Krauklis and Echtermeyer51 in the case of the hydrothermal aging of
epoxy. These authors have observed a change in the color of aged versus unaged
epoxy and have confirmed that the turn in color was irreversible. Similarly, Simar
et al.22 have highlighted polymer color changes during hydrothermal aging (yellow,
brown, and dark) and they have explained this by the network chemical modifica-
tions provoked by oxidation or hydrolysis phenomena.
Additionally, there is another sign of degradation: it consists of a change of the shape
of the aged sample, especially at the higher temperature (see in Figure 9). This underlies
the mechanism of shape change in PVC pipes; thus, our studied material can be con-
sidered a shape-changing polymer under hydrothermal conditions. Here, one can con-
clude that, depending on the conditions and extent of hydrothermal exposure, the
observed aging effects affecting the color and the shape of aged samples are considered
as irreversible changes that provoke permanent property alterations within the thermo-
plastic. These findings are in good accordance with the dramatic symptoms of degra-
dation (deformation and discoloration) previously reported by Guermazi et al.,2 in the
case of PVC/CaCO3 composites exposed to hydrothermal aging. Also, the severity of the
exposing conditions is found to increase with the energy when increasing the aging
temperature, as reported by Grammatikos et al.52

Mechanical properties of PVC pipes


To confirm the results of the previous sections, the effect of the aging temperature on
mechanical behavior was investigated by tensile tests.
Superposed stress-strain curves of unaged samples and aged ones at 25, 60, and 90 C
are subjected in Figure 10. All the curves present three steps, the first step of the
deformation corresponds to the elastic linear deformation. After the limit point of
elasticity, the stress goes through a maximum corresponding to ultimate stress “su”
followed by a plastic deformation until rupture.53,54
The aging temperature has an impact on mechanical properties. The slope of the
elastic part of the curve increases remarkably from 25 C to 90 C, showing a stiffening
arise of the material. The ultimate strength also increased with the immersion tem-
perature and reaches the maximum value at 90 C.
The experimental tensile properties, extracted from Figure 10, for PVC samples are
given in Table 4 with standard deviations. Here, it is to note that the average values of
each characteristic are based on the results of three tensile samples at least.
To better show the impact of the aging temperature on mechanical behavior, the
variation of E, su, eb, and sb are presented in Figure 11(a) and (b).
It can be observed that E increased with aging temperature. The increase was
clear from 25 C to 90 C with a rate of 34.61%.The su evolved similarly to E with
the increasing aging temperature. From 25 C to 60 C the increasing rate of su was
2.62%, while from 60 C to 90 C it was 18.21%. The increasing of E and su reveals
Jemii et al. 15

Figure 10. Stress-strain curves of unaged and aged samples of PVC.

Table 4. Tensile properties of unaged and aged samples (30 days) of PVC pipe.

E (GPa) su (MPa) eb (%) sb (MPa)

Unaged PVC 1.8 + 0.5 24.3 + 0.6 12.0 + 0.5 23.8 + 0.6
Aged at 25 C 1.7 + 0.4 22.3 +1 11.4 + 0.6 21.7 + 0.7
Aged at 60 C 2.4 + 0.4 22.9 + 0.4 8.1 + 0.6 22.4 + 0.6
Aged at 90 C 2.6 + 0.3 28.0 + 0.4 7.2 + 0.8 27.9 + 0.9

E: Young’s modulus; su: ultimate strength; eb: strain at break; sb: stress at break.

the change of the elastic character of the material, which tends to be more stiffen. In
the same way, we have studied the evolution of eb and sb as a function of the aging
temperature for 30 days. From Figure 11(b), it was seen that the stress at break sb
tends to increase with the aging temperature. These results correlate well those
found by Gumargalieva et al.,55 in their investigation about plasticized poly(vinyl
chloride) samples aged in natural and artificial conditions. The latter have found that
the stress at break sb and the strain at break eb, increases from 25 to 30 MPa for 24
months and from 60 to 25%, respectively.
However, Figure 11(b) displays a remarkable decrease of eb with the temperature of
aging. In fact, from 25 C to 90 C, the reduction rate is about 36.84%. This can be
explained by the fact that the material became more brittle. This result is consistent with
that reported by Simar et al.,22 the latter has noted that the resin behavior becomes brittle
due to hydrothermal aging (based on the significant reduction of the failure strain).The
same behavior was observed by Merah et al.,53 when studying the physical and chemical
16 Journal of Thermoplastic Composite Materials XX(X)

Figure 11. Tensile properties: (a) E, su and (b) eb, sb of PVC versus aging temperature.

properties of the powders and films of PVC. Indeed, the later have found that there is a
significant effect of the temperature on the fracture behavior.
To more explain the variation of eb, some photographs of the samples before, during,
and after tensile tests have been performed, as seen in Figure 12. The failure of the tested
samples takes place with brittle rupture. Indeed, there is no significant reduction in the
loaded section (thickness  width). This can be explained by the loss of additives as
previously noted in the FTIR section. Perhaps, these compounds or additives migrate,
during the immersion period, from the core of material to its surface, as described in the
previous section (discoloration and shape change). These results correlate well with sev-
eral studies that have investigated the effect of the rate of plasticizers on the mechanical
properties of PVC.56–58 In particular, Ito and Nagai,57 have shown that the stepwise flow
out of inorganic components and plasticizer can be considered the main mechanism of the
degradation of plasticized PVC under weathering conditions. Accordingly, we can con-
sider that the loss of inorganic components and plasticizer leads to reduce the elongation at
break, whereas it enhances the rigidity and the resistance of aged PVC samples. Finally,
although the aging temperature, in this study, significantly affects the failure strain of PVC
material and tends to improve its strength and stiffness, longer exposure durations can also
significantly affect material behavior than brief exposure. The question of whether
obtained tendencies of mechanical properties changes will remain the same even in the
long-term is then crucial. This point will be developed in future work.
Jemii et al. 17

Figure 12. Illustration of the decrease of the strain at break eb during 30 days.

Conclusions
The work presented in this manuscript is based on the assessment of the physico-
chemical, thermal, and mechanical characteristics of industrial PVC pipes before
and after hydrothermal aging. This characterization was evaluated by several tech-
niques involving water absorption, TGA, FTIR, and mechanical measurements. Based
on the obtained results, during the retained conditions, the following conclusions can
be drawn:

(1) The pipe material has been identified, it is based essentially on thermoplastic
matrix (PVC) reinforced with CaCO3 fillers, and mixed with a commercial
plasticizer (DOP).
(2) Fickian behavior was observed for PVC samples immersed in water at 25 C,
60 C and 90 C. Both of maximum water absorption and the rate of water
absorption increased dramatically with increasing temperature.
(3) Both of TG-DTG and FTIR analysis show preliminary signs of degradation of
PVC samples under the retained conditions.
18 Journal of Thermoplastic Composite Materials XX(X)

(4) The adopted hydrothermal aging induces discoloration and shape changes of the
aged samples.
(5) Higher strength and stiffness were obtained for the aged samples, particularly at
elevated aging temperatures. However, the aged material losses its ductility,
compared to the reference one.

In future work, it will be interesting to conduct a more detailed mechanical and


tribological characterization of PVC pipes before and after long exposure to hydrother-
mal aging.

Acknowledgment
Many thanks are owed to the Engineer Mr Jamel Lachkhem, the Head of Tuyauplast
company (Gafsa, Tunisia) for his technical support in this work.

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

ORCID iD
Houcine Jemii https://orcid.org/0000-0002-6111-8980
Dalila Hammiche https://orcid.org/0000-0001-8663-946X

References
1. Laiarinandrasana L, Gaudichet E, Oberti S, et al. Effects of aging on the creep behaviour and
residual lifetime assessment of polyvinyl chloride (PVC) pipes. Int J Press Vessel Pip 2011;
88(2–3): 99–108.
2. Guermazi N, Haddar N, Elleuch K, et al. Effect of filler addition and weathering conditions on
the performance of PVC/CaCO3 composites. Polym Compos 2016; 37(7): 2171–2183.
3. Benyahia H, Tarfaoui M, El Moumen A, et al. Mechanical properties of offshoring polymer
composite pipes at various temperatures. Compos B Eng 2018; 152: 231–240.
4. Zhang J, Xiao Y and Liang Z. Mechanical behaviors and failure mechanisms of buried
polyethylene pipes crossing active strike-slip faults. Compos B Eng 2018; 154: 449–466.
5. Bredács M, Frank A, Bastero A, et al. Accelerated aging of polyethylene pipe grades in
aqueous chlorine dioxide at constant concentration. Polym Degrad Stabil 2018; 157: 80–89.
6. Wang Y, Lan HQ and Meng T. Lifetime prediction of natural gas polyethylene pipes with
internal pressures. Eng Fail Anal 2019; 95: 154–163.
7. Domaneschi M. Experimental and numerical study of standard impact tests on polypropylene
pipes with brittle behaviour. Proc Inst Mech Eng B J Eng Manuf 2012; 226(12): 2035–2046.
8. Krishnaswamy RK. Analysis of ductile and brittle failures from creep rupture testing of high-
density polyethylene (HDPE) pipes. Polymer 2005; 46(25): 11664–11672.
9. Lin Y and Chan CM. Calcium carbonate nanocomposites. In: G Fengge (ed.) Advances in
Polymer Nanocomposites. Sawston: Woodhead Publishing, 2012, pp. 55–90.
10. Fernando NAS and Thomas NL. Effect of precipitated calcium carbonate on the mechanical
properties of poly(vinyl chloride). J Vinyl Addit Technol 2007; 13(2): 98–102.
11. Azimipour B and Marchand F. Effect of calcium carbonate particle size on PVC foam. J Vinyl
Addit Technol 2006; 12(2): 55–57.
Jemii et al. 19

12. Gilbert M and Patrick S. Poly(vinyl chloride). In: G Marianne (ed.) Brydson’s Plastics Mate-
rials. Oxford: Butterworth-Heinemann, 2017, pp. 329–388.
13. ISO 4437-1:2014. Plastics piping systems for the supply of gaseous fuels - Polyethylene (PE).
London: British Standards Institutions, 2014. https://www.iso.org/standard/59128.html.
14. Guermazi N, Elleuch K and Ayedi HF. The effect of time and aging temperature on structural
and mechanical properties of pipeline coating. Mater Des 2009; 30(6): 2006–2010.
15. Eslami S, Honarbakhsh-Raouf A and Eslami S. Effects of moisture absorption on degradation
of E-glass fiber reinforced Vinyl Ester composite pipes and modelling of transient moisture
diffusion using finite element analysis. Corrosion Sci 2015; 90: 168–175.
16. Ellyin F and Maser R. Environmental effects on the mechanical properties of glass-fiber
epoxy composite tubular samples. Compos Sci Technol 2004; 64(12): 1863–1874.
17. Mei J, Tan PJ, Liu J, et al. Moisture absorption characteristics and mechanical degradation of
composite lattice truss core sandwich panel in a hygrothermal environment. Compos A Appl
Sci Manuf 2019; 127: 105647.
18. Larbi S, Bensaada R, Djebali S, et al. Experimental and theoretical study on hygrothermal
aging effect on mechanical behavior of fiber reinforced plastic laminates. Int J Mech Mecha-
tron Eng 2016; 10(7): 1239–1242.
19. Aldajah S, Alawsi G and Rahmaan SA. Impact of sea and tap water exposure on the durability
of GFRP laminates. Mater Des 2009; 30(5): 1835–1840.
20. Bao Y. Hydrothermal aging behaviors of CMR/PLA biocomposites. J Thermoplast Compos
Mater 2018; 31(10): 1341–1351.
21. Böer P, Holliday L and Kang THK. Independent environmental effects on durability of fiber-
reinforced polymer wraps in civil applications: a review. Construct Build Mater 2013; 48:
360–370.
22. Simar A, Gigliotti M, Grandidier JC, et al. Evidence of thermo-oxidation phenomena occur-
ring during hygrothermal aging of thermosetting resins for RTM composite applications.
Compos A Appl Sci Manuf 2014; 66: 175–182.
23. Jemii H, Bahri A, Boubakri A, et al. On the mechanical behaviour of industrial PVC pipes
under pressure loading: experimental and numerical studies. J Polym Res 2020; 27(8): 1–13.
24. Wan YZ, Wang YL, Huang Y, et al. Hygrothermal aging behaviour of VARTMed three-
dimensional braided carbon-epoxy composites under external stresses. Compos A Appl Sci
Manuf 2005; 36(8): 1102–1109.
25. Hammiche D, Boukerrou A, Djidjelli H, et al. Hydrothermal aging of alfa fiber reinforced
polyvinylchloride composites. Construct Build Mater 2013; 47: 293–300.
26. Batista NL, de Faria MCM, Iha K, et al. Influence of water immersion and ultraviolet weath-
ering on mechanical and viscoelastic properties of polyphenylene sulfide–carbon fiber com-
posites. J Thermoplast Compos Mater 2015; 28(3): 340–356.
27. Rajeesh KR, Gnanamoorthy R and Velmurugan R. Effect of humidity on the indentation
hardness and flexural fatigue behavior of polyamide 6 nanocomposite. Mater Sci Eng A
2010; 527: 2826–2830.
28. Taktak R, Guermazi N, Derbeli J, et al. Effect of hygrothermal aging on the mechanical
properties and ductile fracture of polyamide 6: experimental and numerical approaches. Eng
Fract Mech 2015; 148: 122–133.
29. Ksouri I and Haddar N. Long term ageing of polyamide 6 and polyamide 6 reinforced with
30% of glass fibers: temperature effect. J Polym Res 2018; 25: 153.
30. Jakubowicz I. Effects of artificial and natural ageing on impact-modified poly(vinyl chloride)
(PVC). Polym Test 2001; 20(5): 545–551.
20 Journal of Thermoplastic Composite Materials XX(X)

31. Jian L, Dafei Z and Deren Z. The photo-degradation of PVC: part I—photo-degradation in air
and nitrogen. Polym Degrad Stabil 1990; 30(3): 335–343.
32. Lu Y, Khanal S, Ahmed S, et al. Mechanical and thermal properties of poly(vinyl chloride)
composites filled with carbon microspheres chemically modified by a biopolymer coupling
agent. Compos Sci Technol 2019; 172: 29–35.
33. Panaitescu I, Koch T and Archodoulaki V-M. Accelerated aging of a glass fiber/polyurethane
composite for automotive applications. Polym Test 2019; 74: 245–256.
34. Guermazi N, Haddar N, Elleuch K, et al. Effect of filler addition and weathering conditions on
the performance of PVC/CaCO3 composites. Polym Compos 2015; 37(7): 2171–2183.
35. Gil-Castell O, Badia JD, Kittikorn T, et al. Impact of hydrothermal ageing on the thermal
stability, morphology and viscoelastic performance of PLA/sisal biocomposites. Polym
Degrad Stabil 2016; 132: 87–96.
36. Luca P, Giuseppe R, Marco M, et al. The use of POSS-based nanoadditives for cable-grade
PVC: effects on its thermal stability. Polymers 2019; 11: 1105.
37. Robinet L and Corbeil MC. The characterization of metal soaps. Stud Conserv 2003; 48:
23–40.
38. Silva BD, Gelbrich T, Hursthouse MB, et al. The characterization of lead fatty acid soaps in
protrusions in aged traditional oil paint. Polyhedron 2003; 22: 3171–3179.
39. Paterakis AB. The influence of conservation treatments and environmental storage factors on
corrosion of copper alloys in the Ancient Athenian Agora. J Am Inst Conserv 2003; 42(2):
313–339.
40. Silverstein RM, Webster FX and Kiemle DJ. Spectrometric identification of organic com-
pounds. 7th ed. Hoboken, NJ: John Wiley & Sons Inc., 2005.
41. Matsuhiro B, Osorio-Román IO and Torres R. Vibrational spectroscopy characterization and
anticoagulant activity of a sulfated polysaccharide from sea cucumber Athyonidium chilensis.
Carbohydr Polym 2012; 88: 959–965.
42. Anddersen FA and Brecevic L. Infrared spectra of amorphous and crystalline calcium carbo-
nate. Acta Chem Scand 1991; 45: 1018–1024.
43. Ennaciri Y, Bettach M, Cherrat A, et al. Conversion of phosphogypsum to sodium sulfate and
calcium carbonate in aqueous solution. J Mater Environ Sci 2016; 7(6): 1925–1933.
44. Amar ZH, Chabira SF, Sebaa M, et al. Structural changes undergone during thermal aging
and/or processing of Unstabilized, dry-blend and rigid PVC, investigated by FTIR-ATR and
curve fitting. Annales de Chimie: Science des Materiaux 2019; 43: 59–68.
45. Dutta PK and Graf KR. Migration of plasticizer in vinyl resins: an infrared spectroscopic
study. J Appl Polym Sci 1984; 29(6): 2247–2250.
46. Beltrán M, Garcia JC and Marcilla A. Infrared spectral changes in PVC and plasticized PVC
during gelation and fusion. Eur Polym J 1997; 33(4): 453–462.
47. Boussoum MO and Belhaneche-Bensemra N. Reduction of the additives migration from poly
vinyl chloride films by the use of permanent plasticizers. J Geosci Environ Protect 2014; 2(4):
49–56.
48. Taghizadeh MT and Dasdarligvan H. The study of the thermal stability of poly(vinyl-chloride)
with different molecular weight. JSUT 2007; 33(2): 25–32.
49. Jakubowicz I, Yarahmadi N and Gevert T. Effects of accelerated and natural ageing on
plasticized polyvinyl chloride (PVC). Polym Degrad Stabil 1999; 66(3): 415–421.
50. Kim MH, Byun DJ, You JE, et al. Discoloration mechanism of polymer surface in contact with
air–water interface. J Ind Eng Chem 2013; 19(3): 920–925.
51. Krauklis A and Echtermeyer A. Mechanism of yellowing: carbonyl formation during hygro-
thermal aging in a common amine epoxy. Polymers 2018; 10(9): 1017.
Jemii et al. 21

52. Grammatikos SA, Evernden M, Mitchels J, et al. On the response to hygrothermal aging of
pultruded FRPs used in the civil engineering sector. Mater Des 2016; 96: 283–295.
53. Merah N, Irfan-ul-Haq M and Khan Z. Temperature and weld-line effects on mechanical
properties of CPVC. J Mater Process Technol 2003; 142(1): 247–255.
54. Moura RT, Clausen AH, Fagerholt E, et al. Impact on HDPE and PVC plates—experimental
tests and numerical simulations. Int J Impact Eng 2010; 37(6): 580–598.
55. Gumargalieva KZ, Ivanov VB, Zaikov GE, et al. Problems of ageing and stabilization of
poly(vinyl chloride). Polym Degrad Stabil 1996; 52(1): 73–79.
56. Gil N, Saska M and Negulescu I. Evaluation of the effects of biobased plasticizers on the
thermal and mechanical properties of poly(vinyl chloride). J Appl Polym Sci 2006; 102(2):
1366–1373.
57. Ito M and Nagai K. Analysis of degradation mechanism of plasticized PVC under artificial
aging conditions. Polym Degrad Stabil 2007; 92(2): 260–270.
58. Benaniba MT and Massardier-Nageotte V. Evaluation effects of biobased plasticizer on the
thermal, mechanical, dynamical mechanical properties, and permanence of plasticized PVC. J
Appl Polym Sci 2010; 118(6): 3499–3508.

View publication stats

You might also like