Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

Journal of Building Engineering 86 (2024) 108975

Contents lists available at ScienceDirect

Journal of Building Engineering


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

Self-heating performance of conductive textile-reinforced


cement-based composites
Zeyue Xie a, Josep Claramunt b, c, Monica Ardanuy a, c, *, Heura Ventura a, c
a
Department of Materials Science and Engineering, Universitat Politècnica de Catalunya, Colom, 1, 08222, Terrassa, Spain
b
Department of Agri-Food Engineering and Biotechnology, Universitat Politècnica de Catalunya, Esteve Terradas, 8, 08860, Castelldefels, Spain
c
Institute of Textile Research and Industrial Cooperation of Terrassa-INTEXTER, Universitat Politècnica de Catalunya, Colom, 15, 08222, Terrassa,
Spain

A R T I C L E I N F O A B S T R A C T

Keywords: The regular distribution of conductive networks in electrically conductive cement composites
Conductive cement (ECCC) continues to be a challenge. This study attempted to develop an ECCC with optimal
Cement composite mechanical, electrical, and thermal properties, combining a non-woven conductive layer (made of
Non-woven fabric stainless steel and polyester fibres –SS/PES-) with five additional non-woven reinforcing layers
Stainless steel fibre made of recycled textile waste fibres. Four ECCC were prepared, having distinct SS fibre contents
Recycled fibre
(20%, 30%, 40%, 50%) in the conductive layer. Electrical resistivity, self-heating properties,
Electrical resistivity
flexural behaviour, and coefficient of linear thermal expansion were assessed. In addition, the
Self-heating
specimens having the best heating performance from each sample underwent heating stability
tests. Observations confirmed that all ECCCs adhered to Ohm’s law. ECCC displayed the highest
resistivity at 20% SS fibre (around 21.3 × 10− 3 Ω m) and the lowest resistivity at 50% SS fibre
(around 4.9 × 10− 3 Ω m). In terms of heating performance, ECCC containing 20% SS fibre dis­
played the best performance despite having the highest resistivity, reaching a temperature in­
crease of approximately 46.2 ◦ C under 21 V within 30 min and maintaining a residual
temperature of approximately 8.7 ◦ C after 30 min of cooling. In addition, for all of the ECCC
samples, the electrical and heating effects remained consistent over 20 cycles. In mechanical
terms, the ECCC had Modulus of Rupture (MOR) values ranging between approximately 9 and 13
MPa. The coefficient of linear thermal expansion (CLTE) was ~13 (10− 6/◦ C). Overall, the study
highlights the potential of non-woven fabric as a conductive material in ECCC, offering stable and
repeatable conductive networks, noting the ECCC’s broader applications.

1. Introduction
Cement plays an important role in construction given that it is an inexpensive and durable traditional building material. In the
winter, however, due to lower temperatures, cement constructions may suffer from scaling [1], spalling [2], and cracking [3], thereby
affecting their structure. Furthermore, snow and ice covering cement-based floors may cause traffic accidents and aircraft delays [4],
seriously endangering the safety of people and property. Traditional methods of melting snow and ice [5] include mechanical and
manual snow removal and salting. However, these methods require considerable labour and resources and may have a negative
environmental impact [6,7], while also failing to prevent structural damage caused by low temperatures.

* Corresponding author. Department of Materials Science and Engineering, Universitat Politècnica de Catalunya, Colom, 1, 08222, Terrassa, Spain.
E-mail address: monica.ardanuy@upc.edu (M. Ardanuy).

https://doi.org/10.1016/j.jobe.2024.108975
Received 23 November 2023; Received in revised form 24 February 2024; Accepted 28 February 2024
Available online 29 February 2024
2352-7102/© 2024 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/4.0/).
Z. Xie et al. Journal of Building Engineering 86 (2024) 108975

Thanks to additional research, new functional cement-based materials are gradually appearing to address these issues. Electrically
conductive cement composite (ECCC) uses cement as the matrix material and conductive material as the reinforcing material. This
permits electrical conductivity and self-heating [8]. The self-heating performance of ECCC is closely related to conductivity. When the
current goes through ECCC, electrical energy may be converted to heat.
The conductive materials used in ECCC can be divided into three main categories, including i) conductive particles, such as carbon
black (CB) [9], graphite powder (GP) [10], carbon nanotubes (CNTs) [11]; ii) conductive fibres, such as stainless steel fibres [12],
carbon fibres (CF) [13], carbon nanofibres (CNF) [14]; and iii) conductive fabrics [15]. In these cases, conductivity depends on the
networks formed by these discontinuous conductive materials. Two or more conductive materials are often applied to form complete
conductive networks, thus improving the conductivity. Previous studies have verified the creation of conductive networks related to
the tunnelling effect [16] and percolation theory [17] in cement-based composites. Understanding these two theories helps to
determine the optimal amount of conductive materials required to form a complete conductive network, thereby providing ECCC with
better mechanical, conductive, and thermal properties.
According to the tunnelling effect, when conductive materials are very close but not in contact with one another, their electrons will
break through the potential barrier and undergo a transition phenomenon once attaining sufficient energy, thus forming electric
currents. According to percolation theory, conductive networks are created through contact with conductive materials. When the
conductive material content reaches a certain value, ECCC resistivity may change significantly. This is the so-called percolation
threshold. As shown in Fig. 1, when the content of the conductive material is low (Fig. 1a), the contact probability is low, thus re­
sistivity is high. As the conductive material content increases, resistivity decreases, leading to the creation of an initial conductive
network when the percolation threshold is reached (Fig. 1b). Continuing to add conductive materials will increase the probability of
overlap between them, thereby improving the integrity of the conductive network, so resistivity decreases significantly (Fig. 1c).
However, after reaching a certain threshold value, continuing to increase the content of the conductive material does cause significant
changes in resistivity (Fig. 1d).
Furthermore, ECCC conductivity is also affected by environmental factors, especially humidity [18–21]. Humidity affects not only
the resistivity of the conductive material but also the cement’s moisture content. This leads to alterations in the connectivity of the
conductive network [20]. The influence of humidity has a greater effect on ECCCs with incomplete conductive networks since it may
connect or create a new conductive path in ECCC. In the case of ECCC with a completed conductive network, variations in humidity
primarily affect the resistivity of the conductive material. Therefore, when constructing a conductive network, it is important to
consider not only the conductive material content but also the effect of humidity on the conductive properties.
The heating property depends greatly on the conductivity. According to Joule’s law, the generated heating is proportional to
current resistance and time (Equation (1)).

Q = I 2 Rt (1)

where Q, I, R, and t are generated heat, current, resistance, and time, respectively.
For pure resistance circuitry (circuits with only resistive elements in addition to the power supply), the equation can be written as
Equation (2), where the generated heating is proportional to the square of the voltage and the time and inversely proportional to the
resistance.

U2
Q= t (2)
R

where U is voltage.

Fig. 1. The influence of conductive materials on resistivity.

2
Z. Xie et al. Journal of Building Engineering 86 (2024) 108975

In any case, for the development of heating systems using the Joule effect, extremely high or low resistivity values are undesirable
[22]. When the resistivity is too high, the current is barely able to pass through ECCC, making it impossible to heat by means of the
Joule effect. On the other hand, when resistivity is too low, an extremely high current is needed to heat the system to a certain
temperature. Thus, suitable resistivity is needed for Joule heating. The resistivity between the percolation threshold (Fig. 1b) and the
value close to saturation (Fig. 1c) may achieve this balance.
Therefore, although a high content of conductive materials can ensure high conductivity, it may reduce the Joule effect required for
heating and it may also decrease mechanical properties. Many researchers have focused on realising high conductivity and temper­
ature with lower content conductive materials. Gwon et al. [23] used CB and CF as conductive agents. They found that 0.2 vol% CF
mixed with 0.8 vol% CB can drastically reduce electrical resistivity (97 Ω cm) and achieve the highest surface temperature (77 ◦ C).
Although the introduction of CB reduced the mechanical properties, the minimum strength was 44 MPa. Similarly, Farcas et al. [24]
found that a mixture of 1% CNT and 5% GP displayed a resistivity of 1060 Ω cm, reaching a surface temperature of 50 ◦ C. Fiala et al.
[25] evaluated the incorporation of GP into alkali-activated slag mortar. A GP dosage of 8.75 wt% led to a self-heating capability of 9.9
W. The material displayed an electrical conductivity of 4.62⋅10− 2 S/m and thermal conductivity of 1.36 W/m⋅K, suitable for many
self-heating applications. Ding et al. [12] found that stainless steel wire-reinforced cement composite displayed consistent electrical
conductivity and powerful self-heating capabilities. The electrical resistivity of the composite was 2.58 Ω cm, permitting an increase
from 21.4 ◦ C to 82.4 ◦ C within 30 min under 30 V. Moreover, the heating performance of the material was largely unaffected by
temperature fluctuations or repeated heating and de-icing cycles. Galao et al. [26] examined the potential of CNF integrated into
cement matrices. They found that a 5% CNF content on the cement paste mix could reach temperatures of 138 ◦ C, with the system
initially heating at a rate of 10 ◦ C/min.
In these works, however, the need to include a relatively large amount of conductive elements in the form of randomly dispersed
fibres and particles limits the composite’s mechanical performance. One approach to increase both electrical conductivity and me­
chanical performance could involve the use of conductive fabrics instead. As for mechanical performance, it has been verified that
using woven [27–29] and non-woven fabrics [30–32] as reinforcement materials can improve the mechanical properties of cement
composites. This approach, however, has not been explored to a large extent. Abedi et al. [15] introduced a fabric-reinforced cement
composite material that can self-sense and self-heat. They used coated cotton, polyester, and cotton/polyester fabrics coated with a
dispersion of CNT and graphene nanoplatelets (GNP) in polyurethane-based polymeric resins. The best performance was found for a
polyester fabric coated with four layers of resins containing 4% CNT/GNP. Under 25 V, where the temperature was increased to 44 ◦ C
at a heating rate of 0.44 ◦ C/s and a Joule heating power of 0.7 W/◦ C. Moreover, flexural strength was increased by 50%, and the strain
at failure was increased by 1.18 times.
Previous studies in our research group have shown that six layers of recycled non-woven fabrics made from textile waste fibres may
improve the flexural strength of cement composites [33]. To achieve both conductivity and heating capabilities, and to maintain the
ECCC’s mechanical properties, this study replaced one layer of recycled non-woven fabric with a conductive non-woven fabric made of
stainless steel/polyester fibres (SS/PES). The SS/PES content was varied to examine its impact on electrical, thermal, and mechanical
attributes. Evaluations of these properties were carried out using resistance, self-heating, three-point flexure, and linear thermal
expansion tests, respectively. Furthermore, for each varying content of SS/PES in the ECCC, samples displaying the most optimal
thermal properties were subjected to a heating cycle experiment. This was aimed at assessing the stability and repeatability of the
ECCC.

2. Experimental
2.1. Materials
Bekinox conductive stainless steel (SS) fibres provided from Bekaert (Belgium) in the form of mixtures containing 50 wt% of PES
and 50 wt% of SS fibre (Fig. 2 left) were used to prepare the conductive non-woven fabrics. Virgin PES fibres were used for blending
purposes to change the proportion of conductive fibres in the mixture (Fig. 2 middle).
Recycled fibres from textile waste obtained from mechanical recycling processes were used to produce the reinforcing fabrics.
These fibres were mainly a mixture of cellulosic and PES fibres containing a few non-defibrillated yarns (Fig. 2 right). The main
characteristics of these raw materials are listed in Table 1.
Portland cement Type I 52.5R from Cementos Molins Industrial, S.A. (Spain) was used as the matrix.

Fig. 2. Raw materials used to prepare the non-woven fabrics: Stainless steel/polyester mixed fibre (left); virgin polyester fibre (middle); recycled fibres (right).

3
Z. Xie et al. Journal of Building Engineering 86 (2024) 108975

Table 1
Characteristics of textile raw materials used.

Count (dtex) Length (mm) Diameter

SS/PES mixture from Bekaert


PES Bekaert fibres 3.3 82.3 ± 3.8 18.9 ± 1.0 μm
Bekaert SS fibres 9.1 47.7 ± 3.8 12 μm
PES fibre 6.7 60 13.2 ± 0.8 μm
Recycled fibre mixture (70% cellulose 30% PES)
Recycled fibres (70%) n.a. 0.5-3.5 12–22 μm
Recycled threads (30%) n.a. 10–30 0.37 ± 0.13 mm

2.2. Non-woven preparation


Two types of non-woven fabrics were prepared: one with electrical conductivity, containing variable conductive fibre content (see
nomenclature and proportions in Table 2), and one intended for mechanical reinforcement, made of 100% recycled material (R–NW).
The air-laid processing followed by needle-punching was applied for the web formation and web consolidation in both non-woven
types due to the simple operation, low-energy mechanical processing, and avoidance of chemical product addition. A scheme of the
entire process is shown in Fig. 3.
The air-laid procedure involved a fibre opening followed by an air-laid, executed in a KTRL 600 cotton waste recycling machine line
from QINGDAO KINGTECH MACHINERY CO., LTD (China). The machine parameters used for the web preparation are displayed in
Table 3.
The fibres underwent an initial opening process to ensure an even blending and a reduction in fibre bundles, increasing the final
product’s homogeneity (web). After opening, the well-dispersed fibres were fed to the web-forming unit, where, given the action of
airflow and suctioning of the web-forming cylinder, the fibres were compacted on top of the cylinder’s surface, forming the web.
Afterwards, the formed webs were fed into the needle-punching unit for consolidation. This consolidation was carried out in a DILO
OUG-II6 needle-punching unit from DILO MACHINES (Germany). The fibres were mechanically entangled with one another through
barbed needle action (SNF 15 × 18 × 36 RB30 type) during the needle-punching process, resulting in the consolidation of the webs into
non-woven fabrics of a low thickness (~1.6 mm). The web consolidation processing parameters are based on previous studies [34,35]
and are listed in Table 3.

2.3. Non-woven specimen preparation and characterization


The SS/PES (conductive) non-woven fabrics were cut into square shapes (11 cm × 11 cm) to determine their thickness and
electrical resistivity. The sampling method was based on ASTM D 3776-96. To ensure that the small specimens had the same SS fibre
content as the large fabrics, specimens with an areal weight similar to that of the fabric were selected.
For thickness characterization, a JBA N. 2368 fabric thickness gauge was used under the conditions of the ASTM D 5736 standard.
For each conductive fabric sample, six specimens were measured at five distinct points to obtain the average thickness value. These
values are shown in Table 4.
For the characterization of electrical resistance, a custom-made two-point probe device based on the AATCC 76–2011 standard was
used. The measuring system is shown in Fig. 4. It consisted of a ’U’ shaped plate, obtained by 3D printing technology, with the two feet
covered in copper tape acting as electrodes. Both the resistance of copper tape and the wire connected to the electrodes are less than 1
× 10− 4 Ω, considerably less than the resistance of the non-woven fabric. Thus, the contact resistance can be ignored. A weight of 4.5 kg
(recommended by JIS L1094) was loaded on top of the U-plate to ensure firm contact with the copper contacts and the conductive non-
woven. When additional pressure was applied, resistance was not changed. A Keysight 34465A digital multimeter from Keysight
Technologies (USA) was used to track the resistance data. For each conductive sample, three specimens were tested in both machine and
cross directions (MD and CD, respectively) to obtain the average resistance value for each sample, with the measured area of 10 cm ×
10 cm in all cases. In each test, the currents were allowed to flow through the SS/PES fabric for approximately 1 min until the resistance
value was stable. Then, the material’s resistivity (an intrinsic property of materials, hence not affected by the specimens’ size) was
calculated according to Equation (3),
RA RWt
ρ= = (3)
L L

where ρ is resistivity (in Ω⋅m), R is resistance (in Ω), A is cross-sectional area (in m2), and W, L and t are electrode length, distance
between electrodes, and thickness, respectively (all in m). Conductivity is the inverse of resistivity.

Table 2
Fibre content proportion in the conductive non-wovens.

Type of non-wovens SS PES Ref.

Conductive SS/PES non-wovens 20% 80% NW-20SS


30% 70% NW-30SS
40% 60% NW-40SS
50% 50% NW-50SS

4
Z. Xie et al. Journal of Building Engineering 86 (2024) 108975

Fig. 3. Process of non-woven preparation (top) showing fibre opening, air-laid web formation, and needle-punching web consolidation; and fabrics obtained (bottom):
conductive non-wovens (left) and reinforcement non-wovens (right).

Table 3
Air-laid and Needle-punching parameters.

Air-laid machine
Opener feeder (Hz) 39.88
Opener cylinder (Hz) 18.26
Conveying blower (Hz) 35.90
Hopper pin-lattice (Hz) 27.90
Output conveyor 1 (Hz) 1.17
Airlay web former cylinder (Hz) 31.13
Output conveyor 2 (Hz) 10.11
Needle-punching machine
Number of needles 6840
Stroke frequency (rpm/min) 597
Web feeding speed (m/min) 1.14
Fabric output speed (m/min) 1.15
Distance between needle boards (cm) 1
Upper needle penetration depth (cm) 1
Lower needle penetration depth (cm) 1
Needle punch density (punches/cm2) 591.8

Table 4
Thickness of the SS/PES conductive non-wovens.

Non-wovens Thickness (mm)

NW-20SS 1.81 ± 0.07


NW-30SS 1.78 ± 0.03
NW-40SS 1.57 ± 0.05
NW-50SS 1.65 ± 0.03

2.4. Cement composite preparation


In this research, the matrix consisted of hardened cement paste. The reinforcement layers consisted of five layers of R–NW fabric
(mechanical reinforcement non-woven) and one layer of conductive non-woven fabric, having a different SS content on each sample.
For each ECCC sample, three specimens of 4 cm × 16 cm × 16 cm were prepared. The references and descriptions of the samples are

5
Z. Xie et al. Journal of Building Engineering 86 (2024) 108975

Fig. 4. Resistance measured system.

listed in Table 5.
The preparation procedure of ECCC is shown in Fig. 5 and consists of the following steps. First, SS/PES conductive non-woven and
R–NW were cut into rectangular pieces of 3.5 cm × 15.5 cm. 100% SS yarns were sewn onto the two ends of the rectangular SS/PES
fabric pieces to act as electrodes. Both SS/PES conductive and R–NW fabrics were soaked in water at 60 ◦ C to saturate their water
absorption and remove impurities. To ensure the mechanical properties of cement and allow the cement paste to penetrate into the
non-wovens, two pastes with different water/cement (w/c) ratios were prepared for use as a matrix: one was more fluid (higher w/c
ratio of 1:1) to embed the non-woven fabrics in the core of the plate; and one was denser (lower w/c ratio of 0.35:1) to be placed on the
outer bottom and top layers of the plate. In both cases, the cement paste was mixed for 5 min to ensure a good mixture.
Following these preparations, the mould was impregnated with a demoulding agent, and then the stacking procedure began. First, a
2 mm layer of dense cement paste (0.35:1 w/c) was poured into a micro-grilled mould, and a vibration platform was used to flatten it.
Then, the wetted non-woven fabrics were soaked in the fluid cement paste (1:1 w/c) for 5 min to ease the paste penetrating the fabrics.
A vacuum pump was used to eliminate excess water. Then, after flattening the bottom cement paste layer, the non-woven fabrics were
stacked layer by layer, with the last layer being the SS/PES conductive non-woven one, which remained on top of the other five R–NW
layers. Finally, a top layer of dense cement paste (0.35:1 w/c) was poured over this conductive fabric and was flattened again.
Once the stacking was finished, the mould was assembled and compressed under 5 MPa for 15 min until the pressure remained
constant so that the cement’s water could be further squeezed out by pressure. Then, the plates were left inside the mould, under room
conditions of pressure, temperature and humidity for one day. Afterwards, the ECCC plates were demoulded and cured in water for
seven days. Then, the ECCC plates were dried in an oven for eight days until reaching a stable weight. The structure of the obtained
ECCCs is shown in Fig. 6.

2.5. Cement composite characterization


2.5.1. Electrical resistivity measurement
The ECCC specimens prepared in the previous step were used for the electrical resistivity measurement. The sewn SS yarn elec­
trodes, which were connected with the SS/PES non-woven fabric inside the cement, were clamped with crocodile clips for approxi­
mately 1 min until the resistance became stable. Then, resistivity could be calculated through Equation (3). A Keysight E36311A DC
power supply from Keysight Technologies (USA) was used to test the Current/Voltage relationship, evaluated by increasing the applied
voltage by 1 V each time, from 0 V to 6 V. The output current under each applied voltage was recorded after 1 min.

2.5.2. Heating performance


For heating performance, three specimens per ECCC sample were evaluated. The aforementioned DC power supply was used to
provide voltage. A FLIR TG267 IR camera from Teledyne FLIR (Canada), equipped with a laser position indicator, was used to measure
the temperature in the middle of the specimen’s surface. The IR camera was placed 15 cm from the testing area. Due to the different
ECCC resistances, the applied voltage had to be adapted. Pre-measurements were carried out before ensuring the suitable voltage. The
voltage applied to all ECCC samples was from 3V to 21V in 3V increments. When a voltage exceeding the ECCC’s tolerance range was

Table 5
Composite components and name codes.

Ref. Conductive layer Reinforcement layer Specimen code

Composition Number of layers Composition Number of layers

CC-20SS 20% SS 1 100% Recycled 5 20A


80% PES 20B
20C
CC-30SS 30% SS 1 100% Recycled 5 30A
70% PES 30B
30C
CC-40SS 40% SS 1 100% Recycled 5 40A
60% PES 40B
40C
CC-50SS 50% SS 1 100% Recycled 5 50A
50% PES 50B
50C

6
Z. Xie et al. Journal of Building Engineering 86 (2024) 108975

Fig. 5. ECCC preparation procedure.

Fig. 6. Structure of ECCC

applied, resulting in an output current exceeding 0.8A, the yarn electrodes were damaged (fused or blackened) due to overcurrent.
Thus, to protect the samples, the voltages listed in Table 6 were used in each case.
To explore the heating ability, each specimen was heated constantly for 30 min, and the voltage was then removed to allow the
ECCC to cool down. The temperature was recorded with the IR camera every 5 min during both the heating and cooling process.
Moreover, to investigate the heating stability, the specimen with the best thermal performance of each sample was selected to repeat
the previous heating cycle twenty times under the highest voltage that could be withstood. Temperatures were recorded on the 5th,
10th, 15th and 20th occasions. A spacing of 2 h was used between cycles to ensure that the samples returned to room temperature prior
to each test. Sample resistivity was also tested after each heating cycle to check the effects caused by the heating.

2.5.3. Mechanical performance measurements


Flexural tests were performed to determine the mechanical properties of the ECCC samples. An Incotechnic press, equipped with a
2 kN cell load and a 3-point flexure test device with 10 cm of bending span (distance between supports), was used to test the specimens
at a 10 mm/min cross speed. The measurement system is shown in Fig. 7.

Table 6
Voltage applied to ECCC.

Ref. Specimen (plate) code Voltages tested for heating

CC-20SS 20A, 20B, 20C 3V, 6V, 9V, 12V, 15V, 18V, 21V
CC-30SS 30A, 30B, 30C 3V, 6V, 9V
CC-40SS 40A, 40B, 40C 3V, 6V
CC-50SS 50A, 50B, 50C 3V, 6V

7
Z. Xie et al. Journal of Building Engineering 86 (2024) 108975

From the stress-displacement curve, in which the limit of proportionality (LOP) is defined as the strength value at the upper point of
the linear portion of the curve and the modulus of rupture (MOR) is defined as the maximum stress reached by the material [36], the
modulus of elasticity (MOE) and toughness were calculated.
The flexural strength values for each force value were calculated according to Equation (4):
3Fl
Flexural strength = (4)
2bh2

where F is the maximum load (in N), l is the length of span (in mm), b is the width of the specimen (in mm), and h is the thickness of the
specimen (in mm), respectively.
The modulus of elasticity (MOE) was calculated by Equation (5),

Fl3
MOE = (5)
4fbh3

where f is the maximum deflection (in mm), and other parameters were listed in equation (4).
In addition, the strain energy stored in sample beam during the test was obtained by calculating the area under the force-deflection
curve between the origin and the vertical limit at 0.4 of the maximum force obtained, according to the RILEM TRF4 [37]. The
toughness value was then calculated by dividing the strain energy by the cross-sectional sample area.

2.5.4. Linear thermal expansion coefficient measurements


The specimens after the 3-point flexure test were used to measure the coefficient of linear thermal expansion (CLTE). The fracture
section of specimens was cut by a fully enclosed saw from Mecánica Científica, S.A. (Spain). An E077 length comparator with a dial
gauge from MATEST S.p.A. (Italy) was employed to detect the changes in length caused by heating. Two cross lines were marked in the
middle of the top surface of the sample to ensure the measurements were always taken at the same point. The device is shown in Fig. 8.

Fig. 7. 3-point flexure test system.

Fig. 8. Length comparator with dial gauge.

8
Z. Xie et al. Journal of Building Engineering 86 (2024) 108975

The measurement method is based on UNE-EN 14581, two complete heating and cooling cycles were performed. The cut specimens
were placed in the oven at 20 ◦ C. When the temperature was stable, measure the length of the CC-20SS-20A using a micrometer as a
reference and check the length difference of other specimens using the length comparator. Increase the oven temperature to 80 ◦ C,
after the specimens reach this temperature, detect the length difference. Then decrease the temperature to 20 ◦ C and repeat this
procedure. The thermal expansion coefficient can be calculated through Equation (6).
Δl
α= (6)
l ⋅ ΔT

Where α is CLTE (in 10− 6/◦ C), l is length of the specimen, Δl is the change in length, and ΔT is change in temperature, respectively.

3. Results and discussion


3.1. Electrical properties
Fig. 9 presents the evolution of the resistivity with respect to the SS fibre content for the conductive non-woven fabrics and ECCC
samples. As seen in Fig. 9 (a) and as expected, the resistivity of the non-wovens decreased with the increase in SS fibre content. The
range of values varied from the peak resistivity of the NW-20SS sample (around 8.7 × 103 Ω m) to the lowest resistivity of the NW-50SS
sample (around 4 Ω m). Compared with NW-20SS, the resistivity of NW-30SS plunged sharply to 61 Ω m, and for NW-40SS, it further
decreased to approximately 23 Ω m. Clearly, there were major resistivity differences between NW-20SS and NW-30SS. Resistivity
decreased by approximately 99%, meaning that the full conductive pathway was formed when the SS fibre content was 30%. As a
three-dimensional (3D) structure material, SS fibres were randomly distributed in non-woven with some distance existing between
fibres. Increased SS fibre content enhanced the probability of fibre contact, thus fostering the creation of a conductive pathway. The
integrity of the conductive network facilitates current transport, leading to improved conductivity.
The resistivity of ECCC samples is displayed in Fig. 9 (b). The ECCC samples’ resistivity also displayed a decreasing trend with
increased SS fibre content. The highest resistivity was found for the CC-20SS sample (approximately 21.3 × 10− 3 Ω m) and the lowest
resistivity for the CC-50SS sample (approximately 4.9 × 10− 3 Ω m). As occurred with the non-woven fabrics, the resistivity of CC-30SS
also displayed a relatively wide change from CC-20SS, decreasing to approximately 8.5 × 10− 3 Ω m by 60%. The resistivity of CC-40SS
was approximately 5.2 × 10− 3 Ω m. According to the percolation theory, the percolation threshold was between CC-20SS and CC-30SS.
CC-40SS and CC-50SS were beyond the percolation threshold since further addition of SS fibres produced only weak changes in re­
sistivity values. All ECCC samples displayed substantially lower resistivity as compared to SS/PES non-woven fabrics due to the
compaction of the non-wovens within the composite. This bonding compaction adds constant pressure for the non-woven fabric,
reducing the distance between fibres and increasing the chance of fibre contact, allowing it to reach lower resistivity values.
Moreover, the resistivity difference between varied SS content ECCC narrowed significantly, with the 20%–50% range presenting a
differential of only about 16.4 × 10− 3 Ω m. These observations may be attributed to the dispersed state of fibres in the non-wovens.
Under pressure from the cement matrix, a greater number of SS fibres establish robust contacts, thereby increasing the conductivity
network and decreasing resistivity. Therefore, although NW-20SS displayed low conductivity, CC-20SS samples could achieve suffi­
cient low resistivity for heating applications using a very low content of SS fibres.
Fig. 10 highlights the linear relationship between current and voltage for the ECCC within the three specimens prepared from each
sample, showing that they all displayed a linear relationship (R2 ≥ 0.99) between current and voltage, in line with Ohm’s law, and
similar electrical behaviour, suggesting the homogeneity of the materials. Furthermore, according to Ohm’s law, at a consistent
voltage level, a higher SS fibre content provides a lower resistance, facilitating an easier current flow.

Fig. 9. Resistivity values with respect to SS content for the (a) non-woven fabrics and (b) ECCC.

9
Z. Xie et al. Journal of Building Engineering 86 (2024) 108975

Fig. 10. Current/Voltage relationship of (a) CC-20SS (b) CC-30SS (c) CC-40SS (d) CC-50SS

Fig. 11. Heating performance of the cement composites with different conductive fibre content (a) CC-20SS (b) CC-30SS (c) CC-40SS (d) CC-50SS

10
Z. Xie et al. Journal of Building Engineering 86 (2024) 108975

3.2. Thermal performance of the ECCC


3.2.1. Heating behaviour
Fig. 11 illustrates the heating performance of the ECCC, revealing an increase in temperature with rising voltage and time. As seen
in Fig. 10, the lower the resistivity, the easier it is to pass through currents. To avoid potential damage from excessive currents,
different voltages were applied based on sample resistivity. Consequently, the voltage range for CC-30SS, CC-40SS, and CC-50SS,
which displayed optimal resistivity, was capped at 6V–9V. In contrast, CC-20SS with relatively higher resistivity tolerated voltages
up to 21V. Given these varying voltage tolerances, the resultant maximum temperatures for each sample differed significantly. The
maximum temperature increases reached without damaging the specimens due to excessive currents were approximately 46.2 ◦ C for
CC-20SS at 21V. For CC-30SS, CC-40SS, and CC-50SS, maximums were 10.5 ◦ C at 9V, 13.3 ◦ C at 9V, and 5.6 ◦ C at 6V, respectively.
According to Joule’s law, in a purely resistive system such as the ECCC, the rise in temperature is directly proportional to the square
of voltage and time applied and it is inversely proportional to resistance. Resistivity is proportional to resistance. Therefore, longer
durations, higher voltages, and lower resistances lead to greater temperature increases. For each sample, the temperature was
increased with the raising of voltage. When equal low voltage (3V–6V) was applied across the samples, the temperature differences
were minimal due to closely matched resistivities. When the voltage was 3V, no significant temperature change took place in any of the
samples due to the low voltage. At 6V over 30 min, CC-50SS, having the lowest resistivity, recorded a temperature increase of
approximately 5.6 ◦ C, while CC-20SS, having the maximum resistivity, registered a rise of approximately 3 ◦ C. CC-30SS displayed a
median temperature increase of approximately 4.5 ◦ C. CC-40SS, mirroring the resistivity of the 50% variant, displayed a similar
thermal profile. However, as the applied voltage increased, the current disparity grew, leading to distinct temperature profiles for each
sample. The increased temperature of CC-20SS under 9V at 30 min was approximately 6.4 ◦ C, CC-30SS and that of CC-40SS were
previously mentioned.
Additionally, the rate of temperature increase is influenced by both the ECCC’s resistivity and the voltage applied. Lower resistivity
and higher voltage ensure a quicker temperature rise. Therefore, at a consistent voltage, the heating rate of CC-50SS surpassed the
others. In the case of CC-20SS, the steepening slope of its curve indicated a substantial acceleration in its heating rate with increasing
voltage.
Of these ECCC samples, although CC-20SS had the highest resistivity, it was similar to the percolation threshold and could
withstand higher voltages than the others. Therefore, it was concluded that CC-20SS display the best heating performance.

3.2.2. Cooling performance


Fig. 12 illustrates the cooling performance of ECCCs with different SS fibre contents, revealing their thermal storage capacities.
While all of the samples displayed similar cooling patterns, those reaching higher temperatures in heating were able to cool down
faster. Furthermore, samples that had been heated to higher temperatures maintained a higher temperature during the cooling process.
When heated at a low voltage of 3V, all ECCC samples displayed minimal temperature changes, and this minimal change persisted
even during the cooling process. At a slightly higher voltage of 6V, the temperature differences began to appear, but each sample
returned to room temperature after 30 min of cooling.
Examining the highest achievable temperatures of the samples at their respective voltages (21V for CC-20SS, 9V for both CC-30SS
and CC-40SS, and 6V for CC-50SS), temperature reductions after 30 min of cooling were approximately 32.1 ◦ C, 7.7 ◦ C, 8.5 ◦ C, and
2.4 ◦ C, respectively. Consequently, the residual temperatures in gradient for these samples were approximately 8.7 ◦ C, 3.8 ◦ C, 4.8 ◦ C,
and 1.8 ◦ C, respectively. When CC-20SS was heated at 15V and 18V, it achieved a similar residual temperature (around 6.8 ◦ C and
8.5 ◦ C) to that at 21V after 30 min, with a temperature drop of approximately 17 ◦ C and 23.5 ◦ C during cooling.
In short, although under high voltage ECCC samples could be heated to a higher temperature, they could also cool down faster.
Furthermore, the temperature difference became progressively smaller as cooling time increased. And within a specific time, the
residual temperatures became similar.

3.2.3. Heating cycle


Fig. 13 presents the results of repeated heating tests for the ECCC samples at their maximum voltages (21V, 9V, 9V, and 6V,
respectively). Of the samples, the specimen displaying the most effective heating was selected and chosen for the heating repeatability
examination. For all of the SS fibre content ECCC, the temperature remained consistent despite repeated heating cycles. This stability is
attributed to the SS non-woven material, the sole conductive material in the ECCC, which creates direct and straight conductive
pathways.
Fig. 14 illustrates a comparison of the ECCC’s resistivity measured before and after 5, 10, 15 and 20 heating cycles. The resistivity of
all of the ECCC samples was stable. After the 20th heating, all ECCC samples maintained a resistivity consistent with the initial
resistivity.
It is clear that using SS/PES non-woven fabric as the reinforced material in ECCC may ensure stable electrical and heating prop­
erties. High temperatures barely affected the stability of the ECCC, and it was able to maintain stable performance after considerable
repeated heating.

3.3. Mechanical properties


Fig. 15 presents a representative flexural stress-deflection curve for each of the different ECCC samples, and Fig. 16 shows the main
mechanical parameters, including LOP, MOR, MOE, and toughness obtained from the curves. As seen in Fig. 15, all composites pre­
sented similar behaviour in which the ECCC deflection process is divided into three stages. The first stage was the elastic deflection. In
this case, the mechanical properties were mainly governed by the cement matrix. At a certain point, the LOP was reached, and the

11
Z. Xie et al. Journal of Building Engineering 86 (2024) 108975

Fig. 12. Cooling performance of the cement composites with different conductive fibre content (a) CC-20SS (b) CC-30SS (c) CC-40SS (d) CC-50SS

Fig. 13. Results of the repeated heating tests on cement composites with different conductive fibre content (a) CC-20SS (b) CC-30SS (c) CC-40SS (d) CC-50SS

12
Z. Xie et al. Journal of Building Engineering 86 (2024) 108975

Fig. 14. Resistivity of ECCC before and after 5, 10, 15 and 20 heating cycles (a) CC-20SS (b) CC-30SS (c) CC-40SS (d) CC-50SS

initial cracks appeared. In the second stage, strain-hardening behaviour is displayed. The ECCC continued to rupture with the increase
of stress, at which point both the non-woven reinforcement and cement matrix were the materials primarily affecting the mechanical
properties. Finally, after reaching the MOR (the bending stress peak point), the stress decreased at a slower negative trend, leading to
the opening of the principal crack with massive pull-out fibres, causing specimen failure. The bending behaviour is consistent with the
use of 100% waste fibres in the non-woven [38].
Regarding the parameters of LOP, MOR, MOE, and toughness calculated from the flexural strength curves, these data exhibited a

Fig. 15. Flexural strength representative curves of ECCC with different content of SS fibre.

13
Z. Xie et al. Journal of Building Engineering 86 (2024) 108975

Fig. 16. Variation of main parameters obtained from the flexural tests of the ECCC with different SS fibre contents (a) LOP (b) MOR (c) MOE (d) Toughness.

high degree of uniformity across the specimens. Although, according to the ANOVA calculation, the values of LOP, MOR, and MOE
showed significant differences depending on the composition these differences are mainly related to the production process that can
lead to variations in the thickness of the cement layers rather than the effect of the composition of the conductive since the five layers
R–NW were taking the leading role in the reinforcement capacity.

3.4. Coefficient of linear thermal expansion


Table 7 indicates the results of CLTE of all ECCC. As the temperature increases, all ECCC specimens show an increasing trend in

Table 7
Results of linear thermal expansion coefficient.

Ref. Cycle 1 Cycle 2

Length Length Unitary α Average unitary α Length Length Unitary α Average unitary α
20 ◦ C (mm) 80 ◦ C (mm) (10− 6/◦ C) (10− 6/◦ C) 20 ◦ C (mm) 80 ◦ C (mm) (10− 6/◦ C) (10− 6/◦ C)

20A 67.84 67.946 26.042 16.811 ± 8.982 67.840 67.947 26.287 18.359 ± 7.866
20B 68.554 68.621 16.289 68.553 68.628 18.234
20C 67.890 67.923 8.101 67.885 67.928 10.557
30A 69.404 69.462 13.928 10.546 ± 2.929 69.397 69.457 14.410 13.740 ± 0.630
30B 69.686 69.723 8.849 69.665 69.720 13.158
30C 69.593 69.63 0 8.861 69.591 69.650 13.651
40A 63.091 63.138 12.416 9.978 ± 4.765 63.097 63.130 8.717 8.920 ± 3.746
40B 62.675 62.724 13.030 62.680 62.728 12.763
40C 63.148 63.165 4.487 63.140 63.160 5.279
50A 66.216 66.262 11.578 10.830 ± 0.835 66.219 66.256 9.313 13.054 ± 3.911
50B 65.467 65.506 9.929 65.446 65.496 12.733
50C 65.260 65.303 10.982 65.242 65.309 17.116
Average α 12.78 ± 5.309
(10¡6/◦ C)

14
Z. Xie et al. Journal of Building Engineering 86 (2024) 108975

thermal strain. CC-20SS displayed a higher CLTE than other samples, but it is still within the CLTE of cement [39]. The CLTE of ECCC is
affected by both the reinforcing material and the cement matrix. The reinforcement material is tightly bonded to the matrix and such
behaviour can contribute to dimensional stability, which results in a lower CLTE. Besides, as SS fibre has lower CLTE, the higher
contents of SS fibre can reduce the CLTE. Moreover, since CC-20SS can reach higher temperatures, the high temperature may cause
microcracks. Especially for CC-20SS-20A, which has been subjected to the 20th heating stability test, so it shows the highest CLTE in
the study.

4. Conclusions
In this study, an electrically conductive cement composite (ECCC) was developed by combining one layer of conductive stainless
steel/polyester (SS/PES) non-woven fabric and five layers of non-woven reinforcement fabric made from (non-conductive) recycled
fibres. The main objective was to verify the effects of the SS fibre content in the conductive layer on the electrical and thermal
properties, maintaining the material’s mechanical performance. According to the results, the following conclusions have been reached.
1. Both non-woven fabrics and ECCC showed the highest resistivity at 20% SS fibre (8.7 × 103 Ω m and 21.3 × 10− 3 Ω m) and the
lowest resistivity at 50% SS fibre (4 Ω m and 4.9 × 10− 3 Ω m), whereas the constraining effect of the cement matrix reduced the
resistivity of the conductive non-woven layer by approximately 100%. In addition, all of the ECCC samples complied with Ohm’s
law, regardless of the SS fibre content.
2. When the SS fibre content was close to the percolation threshold and lower than the limit value, the ECCC achieved the best thermal
properties, demonstrating that for heating properties, neither the highest nor the lowest resistivity was desirable. CC-20SS met this
condition, with its surface temperature increasing to approximately 46.2 ◦ C when applying 21 V for 30 min.
3. The higher the temperature reached by the ECCC, the faster its cooling rate, and the higher the maintained temperature. At the
same voltage, the higher the SS fibre content, the higher the temperature reached by the ECCC.
4. The heating properties of all ECCC samples were stable. Under the highest applicable voltage, the temperature was reached, and the
resistivity values were the same even after repeating the heating cycle 20 times.
5. The developed material combining the conductive non-woven layer with waste fibre non-woven reinforcement layers displays
strain-hardening behaviour (MOR between 10 and 13 MPa and toughness between 1.4 and 2 kJ/m2).
6. ECCC has dimensional stability and a low coefficient of linear thermal expansion (around 13 (10− 6/◦ C)) due to the tight bonding of
the reinforcement and the matrix.
7. CC-20SS provided a good balance of thermal properties, with the minimum amount of conductive fibres being the most suitable
proportion of reinforced conductive material for electrical and self-heating applications.
8. The ECCCs developed and characterized in this study displayed stable and reproducible properties, suggesting their great possi­
bilities in self-heating applications.

CRediT authorship contribution statement


Zeyue Xie: Writing – review & editing, Writing – original draft, Methodology, Investigation, Formal analysis, Data curation,
Conceptualization. Josep Claramunt: Writing – review & editing, Resources, Methodology, Investigation, Funding acquisition.
Monica Ardanuy: Writing – review & editing, Supervision, Resources, Project administration, Methodology, Investigation, Funding
acquisition, Conceptualization. Heura Ventura: Writing – review & editing, Writing – original draft, Supervision, Methodology,
Investigation, Conceptualization.

Declaration of competing interest


The authors declare that they have no known competing financial interests or personal relationships that could have appeared to
influence the work reported in this paper.

Data availability

Data will be made available on request.

Acknowledgements
The authors express their gratitude to the Government of Spain, Ministerio de Ciencia e Innovación (MCIN), and Agencia Estatal de
Investigación (AEI) for the financial support received under the scope of the RECYBUILDMAT project (PID2019-108067RB-I00MICIN/
AEI/10.13039/501100011033). Also, they acknowledge the funding of the research group TECTEX (2021 SGR 01056) from the
Department de Recerca i Universitats de la Generalitat de Catalunya. Zeyue Xie gratefully acknowledges the financial support from the
China Scholarship Council (CSC). The author Heura Ventura is a Serra-Húnter fellow.

References
[1] K. Amini, S. Sadati, H. Ceylan, P.C. Taylor, Effects of mixture proportioning, curing, and finishing on concrete surface hardness, ACI Mater. J. 116 (2) (2019)
119–126, https://doi.org/10.14359/51714457.

15
Z. Xie et al. Journal of Building Engineering 86 (2024) 108975

[2] D.F. Cao, et al., Evaluation of prestress losses in prestressed concrete specimens subjected to freeze–thaw cycles, Structure and Infrastructure Engineering 12 (2)
(Feb. 2016) 159–170, https://doi.org/10.1080/15732479.2014.998241.
[3] J. Cui, et al., Evaluation of applicability of minimum required compressive strength for cold weather concreting based on winter meteorological factors,
Materials 15 (23) (2022) 8490, https://doi.org/10.3390/ma15238490.
[4] A. Malakooti, H. Abdualla, S. Sadati, H. Ceylan, S. Kim, K. Cetin, Experimental and theoretical characterization of electrodes on electrical and thermal
performance of electrically conductive concrete, Compos. B Eng. 222 (2021) 109003, https://doi.org/10.1016/j.compositesb.2021.109003.
[5] X. Zha, H. Hu, C. Zhang, T. Luo, Development and laboratory simulation tests of anti-skidding ice melting cushion with carbon fibers for pavement, International
journal of pavement research & technology 15 (5) (2022) 1251–1261, https://doi.org/10.1007/s42947-021-00086-2.
[6] K. Reinosdotter, M. Viklander, K. Reinosdotter, M. Viklander, Road salt influence on pollutant releases from melting urban snow, Water Qual. Res. J. Can. 42 (3)
(2007) 153–161, https://doi.org/10.2166/wqrj.2007.019.
[7] K. Turk, N. Cicek, M. Katlav, I. Donmez, P. Turgut, Electrical conductivity and heating performance of hybrid steel fiber-reinforced SCC: the role of high-volume
fiber and micro fiber length, J. Build. Eng. 76 (2023) 107392, https://doi.org/10.1016/j.jobe.2023.107392.
[8] L. Wang, F. Aslani, A review on material design, performance, and practical application of electrically conductive cementitious composites, in: Construction and
Building Materials, vol. 229, Elsevier Ltd, Dec. 30, 2019, https://doi.org/10.1016/j.conbuildmat.2019.116892.
[9] V. Koci, et al., Preparation of self-heating alkali-activated materials using industrial waste products, J. Clean. Prod. 260 (2020) 121116, https://doi.org/
10.1016/j.jclepro.2020.121116.
[10] C. Farcas, et al., Heating and de-icing function in conductive concrete and cement paste with the hybrid addition of carbon nanotubes and graphite products,
Smart Mater. Struct. 30 (4) (2021) 045010, https://doi.org/10.1088/1361-665X/abe032.
[11] D. Jang, et al., Cyclic heat-generation and storage capabilities of self-heating cementitious composite with an addition of phase change material, Construct.
Build. Mater. 369 (2023) 130512, https://doi.org/10.1016/j.conbuildmat.2023.130512.
[12] S. Ding, et al., Self-heating ultra-high performance concrete with stainless steel wires for active deicing and snow-melting of transportation infrastructures, Cem.
Concr. Compos. 138 (2023) 105005, https://doi.org/10.1016/j.cemconcomp.2023.105005.
[13] S.R. Armoosh, et al., The combined effect of carbon fiber and carbon nanotubes on the electrical and self-heating properties of cement composites, J. Intell.
Mater. Syst. Struct. 33 (18) (2022) 2271–2284, https://doi.org/10.1177/1045389X221077447.
[14] A.L. Pisello, et al., Multipurpose experimental characterization of smart nanocomposite cement-based materials for thermal-energy efficiency and strain-sensing
capability, Sol. Energy Mater. Sol. Cell. 161 (2017) 77–88, https://doi.org/10.1016/j.solmat.2016.11.030.
[15] M. Abedi, et al., A self-sensing and self-heating planar braided composite for smart civil infrastructures reinforcement, Construct. Build. Mater. 387 (Jul. 2023),
https://doi.org/10.1016/j.conbuildmat.2023.131617.
[16] M. Sun, et al., Study on the hole conduction phenomenon in carbon fiber-reinforced concrete, Cement Concr. Res. 28 (4) (1998) 549–554, https://doi.org/
10.1016/S0008-8846(98)00011-8.
[17] P. Xie, P. Gu, J. Beaudoin, P. Xie, P. Gu, J.J. Beaudoin, Electrical percolation phenomena in cement composites containing conductive fibres, J. Mater. Sci. 31
(15) (1996) 4093–4097, https://doi.org/10.1007/BF00352673.
[18] V. Uher, V. Černý, R. Drochytka, Š. Baránek, The effect of exposure conditions on the properties of cementitious composites with reduced electrical resistivity,
Buildings 12 (12) (Dec. 2022), https://doi.org/10.3390/buildings12122124.
[19] G. Hong, S. Choi, D.Y. Yoo, T. Oh, Y. Song, J.H. Yeon, Moisture dependence of electrical resistivity in under-percolated cement-based composites with multi-
walled carbon nanotubes, J. Mater. Res. Technol. 16 (Jan. 2022) 47–58, https://doi.org/10.1016/j.jmrt.2021.11.151.
[20] J. Gomis, O. Galao, V. Gomis, E. Zornoza, P. Garcés, Self-heating and deicing conductive cement. Experimental study and modeling, Construct. Build. Mater. 75
(May 2015) 442–449, https://doi.org/10.1016/j.conbuildmat.2014.11.042.
[21] X. Fan, Effects of environmental temperature and humidity on the electrical properties of carbon fiber graphite cement mortar, in: Advanced Materials Research,
2011, pp. 1022–1026. https://doi.org/10.4028/www.scientific.net/AMR.143-144.1022.
[22] B. Han, L. Zhang, J. Ou, Electrothermal concrete, in: B. Han, L. Zhang, J. Ou (Eds.), Smart and Multifunctional Concrete toward Sustainable Infrastructures,
Springer Singapore, Singapore, 2017, pp. 261–271, https://doi.org/10.1007/978-981-10-4349-9_14.
[23] S. Gwon, H. Kim, M. Shin, S. Gwon, H. Kim, M. Shin, Self-heating characteristics of electrically conductive cement composites with carbon black and carbon
fiber, Cem. Concr. Compos. 137 (2023) 104942, https://doi.org/10.1016/j.cemconcomp.2023.104942.
[24] C. Farcas, et al., Heating and de-icing function in conductive concrete and cement paste with the hybrid addition of carbon nanotubes and graphite products,
Smart Mater. Struct. 30 (4) (2021) 045010, https://doi.org/10.1088/1361-665X/abe032.
[25] L. Fiala, et al., Self-heating alkali activated materials: microstructure and its effect on electrical, thermal and mechanical properties, Construct. Build. Mater. 335
(2022) 127527, https://doi.org/10.1016/j.conbuildmat.2022.127527.
[26] O. Galao, et al., Self-heating function of carbon nanofiber cement pastes, Mater. Construcción 64 (314) (2014) e015, https://doi.org/10.3989/mc.2014.01713.
[27] F. Majstorovic, et al., Impact of metakaolin on mechanical performance of flax textile-reinforced cement-based composites, Cem. Concr. Compos. 126 (2022)
104367, https://doi.org/10.1016/j.cemconcomp.2021.104367.
[28] A. Dogan, N. Karaca, A. Dogan, N. Karaca, Experimental and theoretical behavior of cementitious plates containing ethylene vinyl acetate reinforced with glass
woven fabric under impact load, Materials Science-Poland 39 (4) (2021) 491–506, https://doi.org/10.2478/msp-2021-0041.
[29] S.S. Ajirloo, et al., Investigation on the effect of interlaced yarn structures on bending properties of textile reinforced cement composite with cold plasma treated
polypropylene fabric, J. Textil. Inst. 114 (4) (2023) 601–612, https://doi.org/10.1080/00405000.2022.2054581.
[30] P. Sadrolodabaee, J. Claramunt, M. Ardanuy, A. de la Fuente, Characterization of a textile waste nonwoven fabric reinforced cement composite for non-
structural building components, Construct. Build. Mater. 276 (Mar. 2021), https://doi.org/10.1016/j.conbuildmat.2020.122179.
[31] P. Sadrolodabaee, J. Claramunt, M. Ardanuy, A. de la Fuente, Mechanical and durability characterization of a new textile waste micro-fiber reinforced cement
composite for building applications, Case Stud. Constr. Mater. 14 (Jun. 2021), https://doi.org/10.1016/j.cscm.2021.e00492.
[32] M. Ardanuy, J. Claramunt, R.D. Toledo Filho, Cellulosic fiber reinforced cement-based composites: a review of recent research, in: Construction and Building
Materials, vol. 79, Elsevier Ltd, Mar. 15, 2015, pp. 115–128, https://doi.org/10.1016/j.conbuildmat.2015.01.035.
[33] P. Sadrolodabaee, J. Claramunt, M. Ardanuy, A. de la Fuente, A textile waste fiber-reinforced cement composite: comparison between short random fiber and
textile reinforcement, Materials 14 (13) (2021) 3742, https://doi.org/10.3390/ma14133742.
[34] H. Ventura Casellas, M. Ardanuy Raso, F.J. Capdevila Juan, F. Cano Casas, J.A. Tornero García, Effects of Needling Parameters on Some Structural and Physico-
Mechanical Properties of Needle-Punched Nonwovens, 2014.
[35] J. Claramunt Blanes, H. Ventura Casellas, L. Fernández Carrasco, M. Ardanuy Raso, Tensile and Flexural Properties of Cement Composites Reinforced with Flax
Nonwoven Fabrics, 2017.
[36] Rilem, TFR1 test for the determination of modulus of rupture and limit of proportionality of thin fibre reinforced cement sections, in: RILEM Technical
Recommendations for the Testing and Use of Construction Materials, 1994, pp. 161–163, https://doi.org/10.1201/9781482271362-117.
[37] Rilem, TFR4 the determination of energy absorption in flexure of thin fibre reinforced cement sections, in: RILEM Technical Recommendations for the Testing
and Use of Construction Materials, 1994, pp. 171–173, https://doi.org/10.1201/9781482271362-124.
[38] L. Zhang, et al., Flexural properties of renewable coir fiber reinforced magnesium phosphate cement, considering fiber length, Materials 13 (17) (2020) 3692,
https://doi.org/10.3390/ma13173692.
[39] T. Uygunoglu, I.B. Topcu, Thermal expansion of self-consolidating normal and lightweight aggregate concrete at elevated temperature, Construct. Build. Mater.
23 (9) (2009) 3063–3069, https://doi.org/10.1016/j.conbuildmat.2009.04.004.

16

You might also like