You are on page 1of 37

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

net/publication/365891489

Development of ultra-high-performance concrete with low environmental


impact integrated with metakaolin and industrial wastes

Article in Case Studies in Construction Materials · December 2022


DOI: 10.1016/j.cscm.2022.e01724

CITATIONS READS

7 118

6 authors, including:

Mohamed Abd Ellatief Walid Elemam


Mansoura University Mansoura University
11 PUBLICATIONS 58 CITATIONS 9 PUBLICATIONS 63 CITATIONS

SEE PROFILE SEE PROFILE

Hani Alanazi Gamal Elgendy


University of Nebraska at Lincoln Mansoura University
35 PUBLICATIONS 374 CITATIONS 7 PUBLICATIONS 86 CITATIONS

SEE PROFILE SEE PROFILE

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

Self Healing Concrete View project

EFFECT OF CARBON NANOTUBES AND SILICA FUME ON MECHANICAL PROPERTIES OF CONCRETE View project

All content following this page was uploaded by Mohamed Abd Ellatief on 01 December 2022.

The user has requested enhancement of the downloaded file.


Journal Pre-proof

Development of ultra-high-performance concrete


with low environmental impact integrated with
metakaolin and industrial wastes

Mohamed Abdellatief, Saeeb M. AL-Tam, Walid


E. Elemam, Hani Alanazi, Gamal M Elgendy,
Ahmed M. Tahwia

PII: S2214-5095(22)00856-7
DOI: https://doi.org/10.1016/j.cscm.2022.e01724
Reference: CSCM1724

To appear in: Case Studies in Construction Materials


Received date: 31 August 2022
Revised date: 12 November 2022
Accepted date: 29 November 2022
Please cite this article as: Mohamed Abdellatief, Saeeb M. AL-Tam, Walid E.
Elemam, Hani Alanazi, Gamal M Elgendy and Ahmed M. Tahwia, Development
of ultra-high-performance concrete with low environmental impact integrated
with metakaolin and industrial wastes, Case Studies in Construction Materials,
(2022) doi:https://doi.org/10.1016/j.cscm.2022.e01724
This is a PDF file of an article that has undergone enhancements after acceptance,
such as the addition of a cover page and metadata, and formatting for readability,
but it is not yet the definitive version of record. This version will undergo
additional copyediting, typesetting and review before it is published in its final
form, but we are providing this version to give early visibility of the article.
Please note that, during the production process, errors may be discovered which
could affect the content, and all legal disclaimers that apply to the journal pertain.
© 2022 Published by Elsevier.
Development of ultra-high-performance concrete with low
environmental impact integrated with metakaolin and industrial wastes

Mohamed Abdellatief 1,Saeeb M. AL-Tam 2, Walid E. Elemam1 , Hani Alanazi*3,Gamal M Elgendy 1,


Ahmed M. Tahwia 1
1
Department of Structural Engineering, Faculty of Engineering, Mansoura University, Mansoura, 35516, Egypt
besttime703@std.mans.edu.eg; walidelmetwaly@mans.edu.eg; atahwia@mans.edu.eg;
2
Dept. of Structural Engineering, Faculty of Engineering, Amran University, Yemen; saeebmutlaq1994@gmail.com;
3
Department of Civil and Environmental Engineering, College of Engineering, Majmaah University Al-Majmaah, 11952, Saudi Arabia
*Correspondence: hm.alanazi@mu.edu.sa

Abstract

of
Industrial waste contributes to serious environmental issues. However, these issues can be
solved by utilizing them as building materials. This paper investigated the effects of

ro
substituting the cement with a variety of industrial wastes and metakaolin (MK) on ultra-high-
performance concrete (UHPC) properties. The industrial wastes used in this study include silica
-p
fume (SF), granulated blast-furnace slag (GGBS), and fly ash (FA). The prepared mixtures
were created by varying the amounts of Portland cement, GGBS (30-50%), FA (20-30%), or
re
MK (15-25%), while maintaining a constant amount of SF of 15%. The mechanical, durability,
and micro-structural characteristics of the mixtures were investigated. This study also uses life
lP

cycle assessment to estimate the reductions in embodied energy consumption, life cycle cost,
and carbon footprint associated with integrating metakaolin and industrial waste. The results
a

indicated that the inclusion of 15%MK led to an increase in the compressive, flexural, and
rn

splitting tensile strengths. Also, the chloride permeability resistance was enhanced. The
mechanical properties and chloride-ion permeability resistance of the ternary mixtures with
u

50%GGBS and 25%MK were the lowest compared to other mixtures. These results were
supported by SEM micrographs and EDX analysis for the UHPC structure, which demonstrate
Jo

the high-density microstructure and extremely thin interfacial transition zone thickness. The
inclusion of industrial waste and metakaolin can lower environmental impact of UHPC without
compromising the mechanical performance. Also, the environmental assessment shows that the
UHPC containing industrial waste has a lower impact on the environment, which means they
could make cleaner products in the near future.

Keywords: UHPC - Mechanical properties – Durability -Ternary mixtures - Life cycle cost-
CO2 emission.
1
1. Introduction

The production of ultra-high-performance concrete (UHPC), with its unique advantages


over conventional concrete, was one of the remarkable developments in concrete technology
[1]. The beneficial properties of UHPC can keep pace with the rapid development of
construction work and infrastructure requirements [2]. UHPC has been used in several projects
around the world and has sparked significant interest in more complex applications such as
architectural features, bridge rehabilitation and repair, and the construction of vertical
components such as utility towers for gas and oil industries, windmill towers, and hydraulic
structures [3-5]. Experimental investigations for producing UHPC required constrained
conditions for the mixing process, curing, and high-quality compositions, as well as coarse

of
particles' elimination [6-8]. UHPC has a low permeability coefficient and a dense
microstructure; its porosity ranges from 2% to 6%, and it typically has a permeability

ro
coefficient of about 5×10ˉ⁴ mm²/s [9].

-p
The Portland cement (PC) content of UHPC is usually 2 to 3 times higher than that of
normal concrete [1,11]. The continuing exponential rise of civil infrastructure and
re
industrialization has resulted in a massive demand for the use of PCs [11,12]. The rising
manufacturing of PCs causes financial and environmental risks because the global annual
lP

production of PCs for the construction industry has reached 84.5 million tonnes [13-14]. One
approach is to employ a large amount of extra cementitious materials to partly replace cement.
This need might be fulfilled by adding industrial waste/by-products to UHPC, which will aid in
a

the development of UHPC and a better environment [15,16]. On the other hand, industrial
rn

wastes such as silica fume (SF), GGBS, and fly ash (FA) are undesirable materials for the
environment. However, SF is one of the most common extra cementitious materials used in
u

UHPC production. Due to its tiny particles, SF is essential for filling the pores in the cement
Jo

matrix in addition to taking part in the hydration of the cement [3,17].

Additionally, SF is currently costly in many places, so it may not effectively lower the
initial cost of UHPC. For example, some other industrial byproducts, but not limited to, FA
[18], GGBS [3], metakaolin (MK) [19-20], and rice husk ash [21], etc., are proposed to be
incorporated as partial alternatives to PC in the production of UHPC. These industrial
byproducts were successfully utilized to produce UHPC in several previous studies [15-17].
Along with raw material selection, it is critical to provide an appropriate particle packing

2
method (both for fine and coarse if it exists) to generate a dense microstructure, which is
critical for mechanical and durability properties [22,23].
A ternary blend with cement, SF, and GGBS was reported to have positive effects on fresh
properties and early-age strength of UHPC due to the low water demand of GGBS and
accelerated hydration by SF [9], but it has negative impacts on late-age strength and porosity
due to a dilution effect. It was determined that a ternary binder with cement-SF-limestone has a
high potential to enhance the sustainability and performance of UHPC combinations by
replacing a portion of the cement with silica fume and limestone powders [24]. Satisfactory
results were obtained due to the pozzolanic effect of silica fume and the filler effect and high
sustainability of limestone powder. Advantages of quaternary blends in conventional mortar

of
and concrete were also revealed [25], such as improved strength and chloride resistance using
PC and supplementary cementitious materials (SCMs) by optimal composition combination.

ro
It's worthwhile to note that nano and micro silica powder were only used to replace 3% and 5%
of the cement, respectively. Therefore, this had little effect on cost or the environment.
-p
Nevertheless, to date, there are no comparison studies yet on the ternary systems with low
re
cement dosage and SCMs (like GGBS, FA, and MK) in UHPC blends with environmental
assessments under standard curing conditions [20]. The use of SCMs-containing ternary
lP

systems is important for ecological and low-CO₂ policies, which can be done by replacing as
high a quantity as possible of high-CO₂-emission materials like cement with these SCMs [13].
a

2. Research significance
rn

Based on the above literature, this study primarily focuses on the novel application of using
high quantities of locally accessible SCM as a cement replacement to produce eco-friendly
u

UHPC while maintaining its mechanical characteristics. The effects of SCMs incorporation on
the durability, mechanical, and micro-structural characteristics of UHPC were investigated. The
Jo

SCMs were used in this study as partial replacements for cement that contained FA (20–30%),
MK (15–25%), and GGBS (30–50%) by weight. A comparison between plain UHPC mixtures
and mixtures with SCMs was done during the study. Plain UHPC, including PC and SF, was
employed as the control mix.

On a macroscopic scale, mechanical characteristics, permeability, and sorptivity were


evaluated. The composition, morphology, and thermal analysis of the UHPC samples
containing SCMs were quantitatively examined at the microscopic level. At the end, using the

3
green concrete lifecycle assessment (LCA) tool to compare embedded CO2 emissions,
embedded energy consumption, unit cost, and the sustainability of prepared UHPC including
environmental and economic values is calculated. The results of this study were analyzed and
compared to the results of other studies that investigated low-cost and environmentally friendly
UHPC.

3. Experimental techniques

3.1 Raw materials

In this study, cementitious materials, fine aggregate, steel fibres, tap water, and
superplasticizer (SP) were used for producing the UHPC. The Portland cement (PC) is CEM I

of
52.5 N, which meets BS EN 197-1/2011 [27]. The industrial wastes (including silica fume (SF),
fly ash (FA), and slag (GGBS)) and metakaolin (MK) were employed, which are named SCMs.

ro
The SF, with an average diameter of 0.3 μm was added at a constant ratio of 15% (by cement
mass).
-p
Table 1 lists the chemical and physical characteristics of the raw materials used in this
re
study. The particle size distributions of the raw materials examined by the laser particle
analyzer are shown in Fig. 1. The transmission electron microscopy (TEM) micrographs of the
lP

SCM particles are shown in Fig. 2. In MK and SF micrographs, a sizable number of ultra-fine
particles with flocculation morphology were seen. The usual morphology of FA particles,
which are round and smooth, was observed. Additionally, it was revealed that GGBS particles
a

were found in the irregular grains.


rn

Silica sand was used as fine aggregate in UHPC, and it was used in three different particle
sizes: fine (0.15-0.30 mm), medium (0.30-0.60 mm), and coarse (>0.60 mm) (0.60–1.20 mm).
u

The silica sand has a specific gravity of 2.65 and a bulk density of 1730 kg/m 3. The used quartz
Jo

powders (QP) have a specific gravity of 2.61, Blaine fineness of 305 m2/kg, and very small
particles ranging from 0.1 to 50 μm. The QP plays a main function in the matrix, which is to
complete the particle-size distribution of the silica sand and to compensate for the lack of silica
sand in the matrix [18, 28].

The amount of high range water reducer (Viscocrete3425) is 2.32% by mass of cementitious
materials. Viscocrete3425 complies with the requirements of ASTM C494/C494M-17 Type F
and was used in this investigation to increase consistency and flowability. Hooked-end steel

4
fibres with a length of 13 mm, a diameter of 0.2 mm, and a tensile strength of 2000 MPa were
used. Their properties are provided in Table 2. Also, the ratio of water to binder was kept at
0.16 to get a dense microstructure and very high strength.

Table 1. Chemical and physical properties of the used materials.


Chemical properties (oxides, PC SF MK FA GGBS QP
% by weight)
CaO 63.02 0.27 0.78 7.5 45.88 0.02
SiO2 20.05 97.80 53.3 46.44 30.38 96.76
Al2O3 4.26 0.34 30.0 38.01 9.05 0.08
Fe2O3 2.82 0.17 4.33 3.12 3.82 0.05
SO3 3.15 - - .69 1.78 3.14
Na2O 0.38 0.23 0.26 0.330 0.52 0

of
K 2O 1.24 0.15 0.62 0.88 0.31 0
MgO 3.80 0.21 0.16 0.23 5.39 0

ro
P2O5 - - - 0.76 - -
MnO - - - 0.11 - -
TiO2 - - - 1.17 - -
L.O.I. (%)
Physical properties
Blaine fineness (m2/kg)
2.3

360
2.0
-p
23,000
0.98

18000 420
- 1.41

400 305
-
re
Specific Gravity 3.15 2.21 2.57 2.29 2.88 2.61

Table 2. Main parameters of steel fiber


lP

Fiber diameter Length Aspect Tensile strength Elastic modulus Density Melting point
3 o
(mm) (mm) ratio (GPa) (GPa) (g/cm ) C
0.2 13 65 2000 210 7.79 1400
a
rn

Fig. 1. Particle size distribution of raw materials.


u
Jo

Fig. 2. TEM images of SCMs.


3.2 Mix proportions of UHPC mixtures
Three series of UHPC incorporating SCMs, Mix-I, Mix-II, and Mix-III, were developed to
explore the impact of SCM inclusion on the characteristics of UHPC. In addition, the control
mix contains only PC and SF. In several experiments for developing UHPC mixtures, the
control mix was designed and prepared with PC, which was substituted with silica fume at a
fixed ratio of 15% (by cement mass). The trial mixes were only performed for the control mix
to determine the optimum percent of SF that produces the desired value of compressive

5
strength. The selection of reactive mineral powders in this study was adopted from studying the
relationship between performance and environmental impact. Also, the ratios of SCMs were in
line with previous studies [7,10,15,20]. After trial tests, 15% SF content was found to produce
the highest value of compressive strength. Based on the mixture proportion of the control mix,
all series were performed to obtain optimum performance with low environmental impact, not
to compare the effect of a certain substitution ratio using a set of variables. For Mix-I, the PC in
the control mix was substituted by GGBS at different levels (30%, 40%, and 50% by mass).
The PC was substituted by FA at different levels (20%, 25%, and 30% by mass) for Mix-II.
Finally, for Mix-III, the PC presence in the control mix was substituted by MK at different
levels (15%, 20%, and 25% by mass). For all series, the water/binder was 0.16 and the fine

of
aggregate-to-binder ratio was 0.71. Besides, steel fibers were included in UHPC by 2% (by
vol.). SP was added in the appropriate proportions to preserve the mixes' workability within

ro
accepted limits. The mix proportions of all UHPC mixtures are presented in Table 3. The
numbers in each series represent the percentage of substitution of PC by SCMs, while Mo
-p
refers to the control mix. For example, mix GGBS30 is the UHPC samples with PC replaced by
30% GGBS.
re
Table 3. Mixtures proportions of UHPC samples.
Components (kg/m3)
lP

Group Mix ID PC SF GGBS FA MK QP Sand SP Water Steel fiber


Control
M0 900 135 - - - 180 1271 24 166 156
mix
Mix-I GGBS30 630 135 270 - - 180 1271 24 166 156
a

GGBS40 540 135 360 - - 180 1271 24 166 156


rn

GGBS50 450 135 450 - - 180 1271 24 166 156


Mix-II FA20 720 135 - 180 - 180 1271 24 166 156
FA25 675 135 - 225 - 180 1271 24 166 156
u

FA30 630 135 - 270 - 180 1271 24 166 156


Mix-III MK15 765 135 - - 135 180 1271 24 166 156
Jo

MK20 720 135 - - 180 180 1271 24 166 156


MK25 675 135 - - 225 180 1271 24 166 156

3.3 Mixture preparation, casting, and curing


Four distinct types of UHPC samples were developed to meet the demands of the
mechanical, permeability, and microstructure properties of UHPC. For compressive and
flexural strength tests, prisms with a size of 100 mm × 100 mm × 400 mm were prepared.
Broken beams (from the flexure test) with dimensions of 100 ×100 ×100 mm were used for the
compressive strength test. For splitting tensile strength, cylindrical samples with a diameter of
6
100 mm and a height of 200 mm were fabricated. Cylindrical slices with a diameter of 100 mm
and a height of 50 mm were prepared for the rapid chloride penetration (RCP) and sorptivity
tests. In each group, three specimens were produced for each type of test. A horbit mixer with
multi-speed control was used to prepare the mixtures. The adopted preparation and mixing
processes are illustrated in Fig. 3 to thoroughly mix the UHPC mixtures with steel fibres
[18,28].

Fig. 3. Mixing procedure of UHPC mixtures

After the casting process was finished, the moulds were electrically vibrated for compaction

of
and then placed in the normal curing room (25±2oC) with a relative humidity of (60 ± 5%).
Then, all the samples were demolded after 24 hours and placed into a standard curing room

ro
with a temperature of (25 ± 2 oC) and a relative humidity of more than 95% for curing until
testing.
3.4 Test procedure -p
re
3.4.1 Workability of fresh mixtures
To assess the workability of fresh mixtures, the slump flow was performed in line with
ASTM C143/C143M -15a [30]. A mini-slump cone with an upper diameter of 7 mm, a bottom
lP

diameter of 100 mm, and a height of 50 mm was used to assess the workability.
3.4.2 Mechanical tests
The mechanical strengths of prepared samples were tested at 3,7, 28, and 90 days according
a

to ASTM C1609/C1609M-12 [31] and ASTM C39/C39M-17 [32], respectively. The three-
rn

point bending method was used to evaluate the flexural strength. The compressive strength of
six tested pieces from the flexural strength specimens was calculated. The mixtures' splitting
u

tensile strength was also evaluated using the Brazilian test in accordance with ASTM
Jo

C496/C496M-17 [33].
3.4.3 Sorptivity test
Sorptivity is a key indicator of UHPC's durability since it reveals how easily chemicals
containing moisture can penetrate through the concrete layer. It also shows the UHPC's
capacity to absorb water through its capillary pores. The sorptivity test conducted in this study
complied with ASTM C1585-13 [34].
3.4.4 Chloride permeability test
The method used in this study to evaluate UHPC's resistance to chloride ion diffusion was
the rapid chloride permeability test (RCPT), in accordance with ASTM C1202-12 [35].
7
Cylindrical samples were put into the test cell after a full saturation and then sealed firmly. The
samples were supplied at a DC voltage of 60-V for 6 hours using the apparatus. The cathode
terminal was attached to one reservoir that was filled with 3% NaCl solution, and the anode
terminal to another reservoir that was filled with 0.3 M NaOH solution. The total charge passed
is estimated. The test instrument automatically recorded the total charge (in coulombs) that was
passed over a period of 6 hours.
3.4.5 Microstructure
Using a FE-SEM NOVA® instrument, scanning electron microscope (SEM) imaging of the
broken UHPC surfaces was done at various magnifications to investigate the microstructure
formation. Energy Dispersive X-ray Spectroscopy (EDX) was also used to examine the

of
elemental compositions. A diffractometer (Panalytical X' pert Pro) was used to conduct an X
Ray Diffraction (XRD) test to assess the mineralogy of hydration products of the studied mixes.

ro
To comprehend the influence of SCMs replacement on the hydration products of UHPC
samples, Thermogravimetric – Differential Thermal Analysis (TGA-DTA) tests were

-p
performed. The powdered specimens were heated from 27 to 1000 oC at a rate of 10 oC/min.
For XRD and TGA tests, hand grinding was used after a blender mill to produce acceptable
re
powdered samples that were homogeneous and fine enough.

3.4.6 Life cycle assessment (LCA)


lP

This study provided a suggestion for producing environmentally friendly UHPC through the
incorporation of SCMs. Consequently, there is a justification for the ecological assessment of
the developed UHPC [36]. Life cycle assessment (LCA) is a technique for calculating the
a

ecological footprint of a product across its entire life cycle by analyzing the system's inputs and
rn

outputs [36-37]. LCA has evolved into one of the most essential and widely acknowledged
approaches for assessing, reducing, or improving a product's, process's, or activity's
u

environmental effects. The LCA impacts of replacing Portland cement with SCMs to generate
Jo

sustainable UHPC were assessed. Herein, this paper estimates the embedded CO2 emissions (e-
CO2), embedded energy, and life cycle cost of ternary mixtures to assess their ecological
footprint. With this method, both the environmental impact and the strength-normalized
ecological footprint were evaluated.

8
4.0 Results and discussions

4.1 Slump flow and dry density


Fig. 4 shows the effect of SCMs on the workability of UHPC mixtures. The slump flow for
control mix M0 was 195 mm. For series Mix-I, a decrease in the UHPC flow is visible,
particularly for GGBS50, which had a slump drop of 8.9%. This drop in flow can be attributed
to the irregular shape, sharp edges, and angles of the GGBS particles as seen in Fig. 2 [18]. In
contrast, as the amount of FA is increased from 20 to 30% in series Mix-II, the workability of
the UHPC mixtures improves. This improvement is most likely due to the spherical shape of
FA, which might make the mixture easier to spread, as reported by [38,39]. Finally, for Mix-III,
the incorporation of MK lowered the slump flow by 96.9%, 11629%, and 1464.% at

of
replacement ratios of 15%, 20%, and 25%, respectively, compared to the control mix. The
hydrophilic nature of the MK particle plays a main role in decreasing the workability of ternary

ro
mixtures. This was mostly because of the higher specific surface area of MK in comparison
with the cement, fly ash and GGBS used, as well as the MK particle's angular shape. These
-p
results are consistent with the findings of Norhasri et al. [40].
Fig. 4 presents the dry density of UHPC mixes. Generally, the density of the control mix
re
was 2584 kg/m3, while the density of the series Mix-I varied between 2510 kg/m3 and 2566
kg/m3. The density of series Mix II varied between 2500 kg/m3 and 2540 kg/m3. Similar
lP

findings were also reported by Teixeira et al. [41]. Finally, the series Mix III was reduced to a
range between 2408 kg/m3 and 2421 kg/m3. The reason for this reduction in the density of these
a

mixes is the lower density of MK compared to cement.


rn

Fig.4. Slump flow values and density of UHPC.


u

4.2 Mechanical characteristics


Jo

4.2.1 Mechanical characteristics of Mix-I


Fig. 5 shows the compressive, flexural, and splitting tensile strengths of UHPC samples
with various GGBS replacement levels at 3, 7, 28, and 90 days. It is noticeable that when the
replacement ratio of GGBS is increased, the compressive strength, splitting tensile, and flexural
strength decrease at the specific moist-curing age. In particular, the flexural strength at 28 and
90 days was reduced by 6.15% and 14.62% (Fig. 5b), and the compressive strength at 28 and
90 days was reduced by 11.60% and 12.88%, respectively, compared to M0 (Fig. 5a). Both
compressive strengths and splitting tensile strengths show a very similar pattern, with the

9
highest splitting tensile strength being attained by 30% GGBS replacement at a specific curing
age. Also, compared to M0, the splitting tensile strength dropped by 22.1%, 24.16%, 12.41%,
and 3.74% at 3, 7, 28, and 90 days, respectively (Fig. 5c). This finding is consistent with that of
Wang et al. [42] and Qian et al. [5].

Fig. 5. Influence of GGBS replacement on the mechanical properties of UHPC.

4.2.2 Mechanical characteristics of Mix-II


Fig. 6 illustrates the compressive, flexural, and splitting tensile strengths of UHPC with
various FA replacement levels at 3, 7, 28, and 90 days. FA has a negative impact on the

of
mechanical characteristics of UHPC because the compressive strength of the UHPC matrix
without FA is higher than that of other matrices with FA at each age. However, at a particular

ro
age, Mix-II had mechanical properties noticeably higher than Mix-I. For instance, the FA20
mixture achieved 116 MPa and143 MPa at 28 and 90 days, respectively, which is higher than
-p
GGBS30 by 8.9% and 9.61%, respectively. Both the splitting and flexure strength of UHPC at
curing age showed a similar trend. It is worth noting that the flexure strength of FA20 Mixture
re
was approximately comparable to the control sample (M0). Additionally, the splitting tensile
strength at 90 days of curing decreased in comparison to M0 by 0%, 3.69%, and 8.33% for
lP

20%, 25%, and 30% FA content, respectively. These results are consistent with those of Du et
al. [39] and Qian et al. [5].
a
rn

Fig. 6. Influence of FA replacement on the mechanical properties of UHPC.


u

4.2.3 Mechanical characteristics of Mix-III


Jo

The compressive, flexural, and splitting tensile strengths of UHPC with different MK
substitution levels at 3, 7, 28, and 90 days are presented in Fig. 7. It can be shown that when the
substitution ratio of MK was 15%, the compressive strength and the flexural strength both
reached their highest levels. Specifically, flexural strength at 28 and 90 days increased by
4.61% and 5.62%, respectively, compared to M0, while compressive strength increased by
10.2%, 6.43%, 4.95%, and 3.7% at 3,7, 28, and 90 days, respectively. Similar findings were
reported by Zhang et al. [43], Tahwia et al. [10], and Zhan et al. [2]. It was determined that the
mechanical performance of UHPC containing 15% was superior to Mix-II and Mix-I at 90 days

10
of curing. For instance, the compressive strength of the MK15 mixture exceeds the strength of
FA20 and GGBS30 mixtures by 9.09% and 19.23%, respectively, at 90 days of curing.

Fig. 7. Influence of MK replacement on the mechanical properties of UHPC.


Based on the main objective of the research, which was to produce UHPC with a low
environmental impact utilizing local SCMs, this goal was achieved by obtaining concrete that is
approximately equal in its properties to or exceeds the control mixture. The use of GGBS, FA,
and MK as partial substitutes for PC has a significant effect on the mechanical properties of
UHPC concrete. The use of GGBS as a partial replacement for PC illustrated a significant
decrease in the compressive strength, flexural strength, and splitting tensile strength of concrete

of
at all ages, as shown in Fig. 5. On the contrary, the inclusion of MK led to achieving desirable
mechanical properties, especially when replacing cement with 15% MK (the mechanical

ro
characteristics exceed the control mixture), as shown in Fig. 7. According to Fig. 6, the use of
FA as a partial replacement of cement resulted in a slight decrease in compressive and tensile

-p
strengths compared to the control mixture, while the flexural strengths increased at ages 28 and
90 days. In addition, the influence of adding local SCMs into UHPC in terms of life cycle cost
re
(LCC), carbon dioxide emissions (E-CO2), and embodied energy consumption (EE) is reported
in Section 4.5.
lP

4.3 Durability characteristic


4.3.1 Chloride permeability
Fig. 8 shows a comparison of the RCPT values. It is worth noting that all UHPC samples
a

have a very high degree of compactness and are dense in microstructure. Commonly, all
rn

UHPCs, including SCMs, have total electric charges of less than 100 coulombs, which is a
good indicator of their durability [2,44]. Additionally, it was observed that the RCPT results of
u

the UHPC specimens were influenced by SCM incorporations [45]. For instance, the average
Jo

values over three samples of Mix-I were 91, 93, and 96 Coulombs for 30%, 40%, and 50%
GGBS replacement, respectively, which is higher than the control mix (M0). Furthermore, it
was observed that FA inclusion has the opposite trend to Mix-I performance, which decreased
the RCPT values by roughly 6.15% to 15.38% for Mix-II samples compared to the control mix.
MK15 mixture had the lowest electric charge, 48 Coulombs, whereas GGBS50 had the highest,
about 99 Coulombs, as confirmed by [2,34].

Fig. 8. Effect of replacement of SCMs on the RCPT value of UHPC at 28 days.

11
4.3.2 Sorptivity
Fig. 9 presents the results of the initial sorptivity values of UHPC samples at 28 days. The lower
average values as evaluated on 3-specimens of each UHPC were FA30 mixture (Mix-II), which was
obtained at 8.2 × 10-4 mm/√sec. The highest values were GGB50 mixture (Mix-II), which was obtained
at 16.60 × 10-4 mm/√sec. According to Tahwia et al. [18], active pozzolanic materials play an
important part in diminishing UHPC sorptivity; consequently, the degradation in sorptivity may be
related to the reduction of the pores in the ternary mixtures or the disconnect of the linked pore network.
Furthermore, the results obtained are consistent with the findings of Prem et al. [46], which examined
effects in ternary cementitious systems.
Fig. 9. Initial sorptivity values for UHPC samples at 28 days.

of
4.4 Microstructure
Most of the internal microstructure of UHPC is made up of unhydrated cement particles,

ro
quartz sand, and the main hydration product (C-S-H) [47]. Pore structure, hydration products
morphology, and interfacial transition zone (ITZ) microstructure can all influence the

-p
microstructure of UHPC [48]. In the current work, SEM, XRD, and TGA tests were used to
evaluate the microstructure, morphology, and thermal analysis of the control mix as well as the
re
mixes that had the highest mechanical properties in each series. Therefore, the mixtures
GGBS30 from Mix-I, FA20 from Mix-II, and MK15 from Mix-III were chosen.
lP

4.4.1 Morphology
The morphology of the UHPC as determined by SEM and EDX is shown in Fig. 10. It
a

was noticed that the UHPC microstructure changed with the type of integrated SCMs.
rn

GGBS30, FA20, and MK15 micrographs appear to be less compact and denser than control
mixture (M0) micrographs (Fig. 10a), which are characterized by unreacted particles and
u

interfacial fissures. These weaknesses point to the microstructure brittleness and an incomplete
hydration process as compared to the M0. As shown in Fig. 10b, demonstrating that the
Jo

GGBS30 samples had the largest pore size when compared to the other ternary mixes and
control mix, this result is compatible with those of the mechanical and durability properties. It
was indicated that when GGBS was added, the ITZ between the binder and aggregate
deteriorated and became less dense than that of the control mix. It was also observed that the
presence of unreacted particles at the microstructure level. A similar trend in the FA20 mixture
was observed, which negatively affected the microstructure of the UHPC (Fig. 10c).
In comparison to the GGBS30 mixture, the microstructure of the FA20 mixture, as shown
in Fig. 10c, was less porous. This density occurred due to the increased proportion of SiO2
12
supplied by FA in addition to SF compared to cement. Moreover, there were fewer interfacial
voids. Cement hydration output from the reaction between the SF and FA components and
portlandite made up for the lack of C-S-H [48]. The C-SH gel was spread out evenly in the
UHPC paste structure, which is clear from the SEM images. This finding was corroborated by
XRD patterns as discussed later (Fig. 11). On the other hand, SEM imaging revealed that 15%
of MK replaced with PC in the control mix plays a major part in enhancing the sample
microstructure, which was reflected in their mechanical and durability performance as
presented in Fig. 10d. Studies [10,49] in previous literature supported the development of
additional phases, such as C-S-H gel, which could improve the microstructure of the matrix.
Fig. 10(e-h) shows the EDX result on spot A, demonstrating the effect of SCM

of
incorporation on the UHPC composition. The Ca, Si, and O elements, which are the primary
ingredients of cementitious products, were the prominent peaks in the M0 and MK15 mixtures.

ro
Consequently, the resulting mixture possesses a hydraulic activity that can react with Ca (OH) 2
to create more C-S-H gel [18,47]. The EDX observations also demonstrate the change in
-p
element content because of SCMs' incorporation. The addition of MK increased the amount of
Si, Fe, and Ca, while FA and GGBS decreased the concentration of silicon and iron
re
components in the structure of produced mixes, which also constituent with the chemical
composition presented in Table 1. This result indicates that the microstructure and engineering
lP

properties of UHPC are considerably improved by high fineness and a chemically balanced
composition. These findings agree with the previous studies [18,50].
a

Fig.10. SEM and EDX of UHPC mixtures.


rn

4.4.2 Mineralogy.
u

Fig. 11 shows the XRD patterns of the samples incorporating SCMs at 28 days. The
Jo

crystalline phases of UHPC mixtures were studied using XRD, and the X-ray scan rate had an
influence on the intensity of the diffraction peaks [51-52]. And the intensity of the distinctive
peaks in their XRD spectra is directly proportional to the quantity of hydration products or
certain minerals. The distribution of peaks in the M0, GGBS30, FA20, and MK15 mixtures
after XRD in the 2𝜃-range from 20◦ to 44◦. The peak SiO2 in the UHPC with both FA or MK
inclusion dropped, while the peak calcium carbonate and other hydration products produced
rose to varying degrees with GGBS inclusion, confirming the trend of the mechanical
properties of UHPC in the previous sections [3,52]. According to the hydration product of the

13
control mix, the primary phases following a hydration period are Ca (OH) 2 and ettringite.
Furthermore, there were still C3S, C2S, and C4AF peaks attributable to non-hydrated cement
particles in the cement rich UHPC mixes, which were influenced by SCM substitution. This
conclusion was consistent with prior studies on the effects of SCMs on the characteristics of
UHPC [53].

Fig. 11. X-ray diffraction pattern for UHPC tested at 28 days.

4.4.3 Thermal analysis

Fig. 12 shows the DTA-TG curves of UHPC samples. These curves indicate the stages of

of
weight loss. It was seen that each group of samples had the same tendency to lose weight.
However, they exhibit varied weight loss ratios at each changing temperature range, indicating

ro
that the quantity of material change varies at each stage of the reaction. Generally, the curves
show three distinctive temperature ranges. The first stage is between 27 and 200 oC, which
-p
indicates that the C-S-H, ettringite, and alumina, ferric oxide, and monosulfate (AFm) phases
have lost some of their moisture [54]. Hence, hydrated cementitious material pastes go through
re
a process wherein they lose free water, dehydrate, and then transfer the hydration products
during high temperature processing [55]. Therefore, the primary peak at 140 °C was weakened
lP

and moved to a little lower temperature because of GGBS replacement (GGBS30). While the
peak for FA20 and MK15 mixtures was found at or near 127 °C and 119 °C, respectively. This
a

is mostly explained by a decrease in the amount of unreacted gypsum and a decrease in the
amount of ettringite seen in XRD patterns. The second stage, a rather moderate gradient, which
rn

ranges in temperature from 400 to 550 oC, represents the dehydroxylation of Ca (OH)2. The
final stage of transformation is the degradation of calcium carbonate at temperatures ranging
u

from 600 °C to 1000 °C. This is in line with earlier findings, which demonstrated a similar
Jo

trend of these stages [17,54].

It's interesting to note that as the temperature rises above 600 °C, the rate of weight loss
from dehydration increases. Table 4 shows the effect of the SCMs on the mass loss of UHPC
after being exposed to 300, 600, 800, and 1000°C. The results of the thermal analysis also show
that the use of GGBS can limit cement hydration and reduce the amount of hydration that
happens, while the addition of MK and FA can speed up cement hydration, which is in line
with the results of Zhang et al. [43].

14
Fig. 12. TGA-DTA curves of UHPC.

Table 4. Effect of the SCMs on the mass loss of UHPC (%)

Mix ID 300 oC 600 oC 800 oC 1000 oC

M0 6.35 7.14 7.42 8.83

GGBS30 4.73 5.73 5.72 6.35

FA20 5.35 7.95 8.48 8.92

MK15 4.78 6.39 7.07 7.64

of
4.5 Life cycle assessment (LCA)

ro
Recent studies have confirmed that the use of cement in large quantities poses serious
environmental risks and increases global warming. Therefore, the trend was to reduce the

-p
amount of cement consumed by using green materials such as SCMs [5] and geopolymer
systems [11]. Incorporating SCMs into UHPC opens new possibilities for the economization
re
and recycling of industrial solid waste [46]. This use of SCM is based on its potential to reduce
the cement content used in UHPC production and is compared with non-SCM control mixtures
lP

in terms of life cycle cost (LCC), carbon dioxide emissions (E-CO2), and embodied energy
consumption (EE). The LCC, E-CO2, and EE of the used material according to previous studies
are summarized in Table 5 [1,56,57].
a

Table 5 Inventory of carbon footprint, embodied energy, and unit cost of raw materials.
rn

No. Raw material E-CO2 (kg/kg) EE (MJ/kg) Cost ($/kg)


u

1 Cement 0.83 4.8 0.11


2 SF 0.0140 0.1000 0.04
Jo

3 GGBS 0.019 1.588 0.08


4 FA 0.0090 0.1 0.04
5 MK 0.40 3.48 0.06
6 Sand 0.001 0.022 0.02
7 HRWR 0.25 18.1 3.20
8 Water 0.0002 0.01 0.00
9 Steel fiber 1.49 20.59 4.78

15
4.5.1 Carbon footprint (CF)
Using the inventory data in Table 5, Eq. (1) may be used to compute the CF of each
ternary and binary mix:

∑ Eq. (1) [13]

Where CF is a combination's carbon footprint; fi is the E-CO2 unit of materials (i), n= (1


to 9), as mentioned in Table 5; and r was the weight of the used material (i), as described in
Table 3. The unit CF of each combination was computed using Eq. (1).

The strength-normalized carbon footprint (kg/m3/MPa) of each matrix was estimated

of
using its compressive strength (MPa) at 28 days and can be expressed as (ST-CF). The
outcomes of the binary (M0), ternary mixes' unit CF and strength-normalized CF are displayed

ro
in Fig. 13. As the addition of FA influenced the compressive strength of UHPC samples
compared to M0, the unit CF declined from 988 to 767 kg/m3, and the ST-CF decreased from
-p
8.98 kg/m3/MPa to 7.04 kg/m3/MPa as the FA content rose to 30%. Also, because the addition
of GGBS (50%) reduced the compressive strength compared to M0, the unit CF decreased from
re
988 to 623 kg/m3, whereas the ST-CF decreased from 8.98 kg/m3/MPa to 6.93 kg/m3/MPa.
Finally, with MK inclusion decreasing the compressive strength of ternary mixes compared to
lP

M0, the unit CF decreased from 988 to 892 kg/m3, whereas the ST-CF rose from 8.98
kg/m3/MPa to 10.25 kg/m3/MPa. Generally, when compared to the control mixture, the usage
a

of all SCM materials decreased the unit CO2 emission.


rn

Fig. 13. Unit CF and ST-CF of the UHPC samples.


u

4.5.2 Embodied energy (EE)


Jo

The EE of each binary and ternary combination may be computed using Eq. (2) using the
inventory data in Table 5:

∑ Eq. (2) [13]

where EE is the energy of a combination; ei denotes the EE unit of the materials (i) and n
= (1 to 9), as mentioned in Table 5; and r denotes the weight of the material (i), as mentioned in
Table 3. The unit EE of each M0 and ternary mixture was estimated using Eq. (2). The

16
strength-normalized EE (MJ/m3/MPa) of each matrix was estimated using the compressive
strength (MPa) at 28 days and can be expressed as (ST-EE). The EE and ST-EE of the
examined mixes are shown in Fig. 14. Because the introduction of FA influenced the
compressive strength, the CF declined from 8013.52 MJ/m3 to 6744 MJ/m3, and the ST-EE
decreased from 66.23 MJ/m3/MPa to 61.88 MJ/m3/MPa as the FA content rose to 30%. Also,
because the inclusion of MK lowered the compressive strength, the EE decreased from 8013.52
MJ/m3 to 7716.52 MJ/m3, whereas the ST-EE rose from 66.23 MJ/m3/MPa to 88.70
MJ/m3/MPa as the MK content rose to 25%. Also, because the introduction of GGBS lowered
the compressive strength, the EE decreased from 8013.52 MJ/m3 to 6568 MJ/m3, whereas the
ST-EE rose from 66.23 MJ/m3/MPa to 72.98 MJ/m3/MPa as the GGBS content rose by up to

of
50%. The ternary mixture containing 30% FA achieved the lowest ST-EE. In this investigation,
the best ternary matrix was FA30.

ro
Fig. 14. EE and ST-EE of the UHPC samples.

4.5.3 Cost analysis


-p
re
Using the inventory data in Table 4, Eq. (3) may be used to determine the unit cost of
each mixture:
lP

∑ Eq. (3) [13]


a

where C is the unit cost of a matrix/m3 (unit: $/m3); Xi is the unit cost (unit: $/kg) of the
rn

materials (i), where n = (1 to 9), as shown in Table 5; and Yi is the weight of the material (i), as
mentioned in Table 3. The unit cost of each combination was determined using Eq. (3). The
u

strength-normalized cost ($/m3/MPa) was estimated using the compressive strength of each
combination at 28 days and can be expressed as (ST-C). The findings of the unit cost and ST-C
Jo

of the ternary mixes are shown in Fig. 15. Because the employment of FA influenced the
strength, as the FA 30% content, the unit cost decreased from 956 $/m3 to 937 $/m3, and the
ST-C increased from 7.90 $/m3/MPa to 8.60 $/m3/MPa. Also, because usage of GGBS lowered
the compressive strength, the unit cost was reduced from 956 $/m3 to 942.40 $/m3 as the GGBS
content reached 50%, and the ST-C was increased from 7.90 $/m3/MPa to 10.47 $/m3/MPa. The
unit cost decreased from 956 $/m3 to 944 $/m3 as the MK content reached 25%, and the ST-C
increased from 7.90 $/m3/MPa to 10.86 $/m3/MPa, since the MK25 was the lowest compressive

17
strength. The standard cost of strength for GGBS30, FA20, and MK15 is approximately 8.86,
8.13, and 7.91 $/m3/MPa, respectively.

Fig. 15. Unit cost and ST-C of the ternary mixes.

4.5.4 Comparison with other ternary blends


In recent studies, the production of UHPC matrices has mostly focused on lowering
material costs, CF, and EE without compromising mechanical characteristics. In general, higher
mechanical performance UHPC composites have higher material costs, CF, and EE. In contrast,
the GGBS50 mixture had the minimum value of unit CF (623 Kg/m3), as shown in Fig. 13. The

of
estimated CF and strength-normalized carbon emission (Kg/m3/MPa) of the control mixture,
the optimized ternary mixes (GGBS50), and various kinds of ternary systems published in

ro
earlier references [7, 58-65] are shown in Fig. 16.

-p
Fig. 16. Comparison of carbon footprint and ST-CF of the developed ternary systems in
re
this study and previous studies [60,5,7, 58,59,15,5,62,63].
5.0 Conclusions
lP

This study concerns the production of eco-friendly UHPC with supplementary


cementitious materials. In the study, PC was replaced with GGBS, fly ash, or metakaolin. A
a

constant amount (15% of cement mass) of silica fume was also added to all UHPC mixtures.
Three different levels of SCMs were used, and their effects on the mechanical, durability, and
rn

microstructural properties of UHPC were also comprehensively investigated. The key


conclusions in this paper can be summarized as follows:
u

The workability of UHPC was decreased by MK and GGBS additions, while it


Jo

1.
increased with FA inclusion. The compressive, flexural, and splitting strengths of a
mixture containing 15% MK are always the highest at the curing age of 90 days, which
increased by 3.16%, 4.57%, and 5.37%, respectively, compared to the control mix.
2. Besides, the effective role of SCMs in promoting the durability performance of
concrete, high substitution of FA up to 30% results in a lower chloride ion permeation
and sorptivity of 55 C and 8.9 x10-4 mm/s0.5, respectively, while inclusion of 50%
GGBS had the highest chloride ion permeation of 96 C.

18
3. The industrial wastes (fly ash, silica fume, and slag) and metakaolin are required to
manufacture low carbon emission UHPC using only 35 to 65% cement in the ternary
binder system without compromising the mechanical performance. In this approach,
the cement is efficiently utilized, and the manufacturing of UHPC achieves a balance
between performance and environmental impact.
4. Regardless of the type of SCMs employed, as the compressive strength of
UHPCs containing SCMs increases, the e-CO2 index of the UHPC gradually
declines, and the value of embodied energy and cost index gradually increase.
5. In comparison to previous ternary mixtures presented in the literature review,
the produced ternary mixture in this study presents major promise for

of
generating green and cost-effective UHPC under normal curing.

ro
6. The results of the environmental assessment further demonstrate how cement
may be efficiently replaced by SCMs to guarantee UHPC's enhanced

-p
performance while lowering CO2 emissions per unit volume. This suggests that
it is reasonable to properly substitute SCMs for cement in the creation of an
re
eco-friendly UHPC, providing a fresh idea to produce new clean products in the
future.
lP

Acknowledgments
a
rn

The authors extend their appreciation to the deputyship for Research & Innovation,

Ministry of Education in Saudi Arabia for funding this research work through the
u

project number (IFP-2022-X).


Jo

References

[1] Y. Dong, ‘Performance assessment and design of ultra-high-performance concrete (UHPC)


structures incorporating life-cycle cost and environmental impacts’, Constr. Build. Mater.,
vol.167, 2018, doi: 10.1016/j.conbuildmat.2018.02.037.

[2] P. Zhan, J. Xu, J. Wang, and C. Jiang, ‘Multi-scale study on synergistic effect of cement
replacement by metakaolin and typical supplementary cementitious materials on properties

19
of ultra-high-performance concrete’, Constr. Build. Mater., vol.307,2021, doi:
https://doi.org/10.1016/j.conbuildmat.2021.125082.

[3] Ç. Yalçınkaya and O. Çopuroğlu, ‘Hydration heat, strength and microstructure


characteristics of UHPC containing blast furnace slag’, J. Build. Eng., vol.34, 2021, doi:
10.1016/j.jobe.2020.101915.

[4] A. S. Faried, S. A. Mostafa, B. A. Tayeh, and T. A. Tawfik, ‘Mechanical and durability


properties of ultra-high-performance concrete incorporated with various nano waste
materials under different curing conditions’, J. Build. Eng., vol. 43,2021, doi:
10.1016/j.jobe.2021.102569.

[5] D. Qian et al., ‘A novel development of green ultra-high-performance concrete (UHPC)


based on appropriate application of recycled cementitious material’, J. Clean. Prod., vol.

of
261, 2020, doi: 10.1016/j.jclepro.2020.121231.

[6] H. Zhang, T. Ji, B. He, and L. He, ‘Performance of ultra-high-performance concrete

ro
(UHPC) with cement partially replaced by ground granite powder (GGP) under different
curing conditions’, Constr. Build. Mater., vol. 213,2019, doi:
10.1016/j.conbuildmat.2019.04.058.
-p
[7] Y. Shi, G. Long, C. Ma, Y. Xie, and J. He, ‘Design and preparation of ultra-high-
re
performance concrete with low environmental impact’, J. Clean. Prod., vol. 214,2019, doi:
10.1016/j.jclepro.2018.12.318.
lP

[8] A. A. Abadel, M. I. Khan, and R. Masmoudi, ‘Experimental and numerical study of


compressive behavior of axially loaded circular ultra-high-performance concrete-filled tube
columns’, Case Stud. Constr. Mater., vol. 17, 2022, doi:
https://doi.org/10.1016/j.cscm.2022.e01376
a

[9] C. Shi, Z. Wu, J. Xiao, D. Wang, Z. Huang, and Z. Fang, ‘A review on ultra-high-
rn

performance concrete: Part I. Raw materials and mixture design’, Constr. Build. Mater.,
vol. 101, 2015, doi: 10.1016/j.conbuildmat.2015.10.088.
u

[10] A. M. Tahwia, G. M. Elgendy, and M. Amin, ‘Mechanical properties of affordable and


Jo

sustainable ultra-high-performance concrete’, vol. 16, 2022.

[11] A. M. Tahwia, M. Abd Ellatief, A. M. Heneigel, and M. Abd Elrahman,


‘Characteristics of eco-friendly ultra-high performance geopolymer concrete incorporating
waste materials’, Ceram. Int., vol. 48(14),2022, doi:
https://doi.org/10.1016/j.ceramint.2022.03.103.

[12] I. Y. Hakeem, M. Amin, A. M. Zeyad, B. A. Tayeh, A. M. Maglad, and I. S. Agwa,


‘Effects of nano sized sesame stalk and rice straw ashes on high-strength concrete
properties’, J.Clean.Prod., vol.370,2022, doi:
https://doi.org/10.1016/j.jclepro.2022.133542.
20
[13] Hermann, C., & Kara, S. (Eds.), ‘Sustainable Production, Life Cycle Engineering and
Management’. Springer, 2012.

[14] A. M. Tahwia, Ashraf Heniegal, Mohamed S. Elgamal, Bassam A. Tayeh. 'The


prediction of compressive strength and non-destructive tests of sustainable concrete by
using artificial neural networks'. Computers and Concrete, An International Journal,
27(1),2021, DOI: https://doi.org/10.12989/cac.2021.27.1.000.

[15] R. Yu, Q. Song, X. Wang, Z. Zhang, Z. Shui, and H. J. H. Brouwers, ‘Sustainable


development of Ultra-High Performance Fibre Reinforced Concrete (UHPFRC): Towards
to an optimized concrete matrix and efficient fibre application’, J. Clean. Prod., vol. 162,
2017, doi: 10.1016/j.jclepro.2017.06.017.

[16] M. Amin, A. M. Zeyad, B. A. Tayeh, I. Saad Agwa, and I. Saad, ‘Effect of ferrosilicon

of
and silica fume on mechanical, durability, and microstructure characteristics of ultra-high-
performance concrete’, Constr. Build. Mater., vol. 320, 2022, doi:

ro
10.1016/j.conbuildmat.2021.126233.

[17] R. Yu, P. Spiesz, and H. J. H. Brouwers, ‘Effect of nano-silica on the hydration and

amount’, Constr.
10.1016/j.conbuildmat.2014.04.063.
Build. -p
microstructure development of Ultra-High-Performance Concrete (UHPC) with a low
binder Mater., vol. 65, 2014, doi:
re
[18] A. M. Tahwia, G. M. Elgendy, and M. Amin, ‘Durability and microstructure of eco-
efficient ultra-high-performance concrete’, Constr. Build. Mater., vol.303, 2021, doi:
lP

10.1016/j.conbuildmat.2021.124491.

[19] A. Tafraoui, G. Escadeillas, S. Lebaili, and T. Vidal, ‘Metakaolin in the formulation of


UHPC’, Constr. Build. Mater., vol. 23, 2009, doi: 10.1016/j.conbuildmat.2008.02.018.
a

Y. R. Alharbi, A. A. Abadel, O. A. Mayhoub, and M. Kohail, ‘Effect of using available


rn

[20]
metakaolin and nano materials on the behavior of reactive powder concrete’, Constr. Build.
Mater., vol. 269, 2021, doi: https://doi.org/10.1016/j.conbuildmat.2020.121344
u

[21] N. S. Ha, S. S. Marundrury, T. M. Pham, E. Pournasiri, F. Shi, and H. Hao, ‘Effect of


Jo

grounded blast furnace slag and rice husk ash on performance of ultra-high-performance
concrete (UHPC) subjected to impact loading’, Constr. Build. Mater., vol.329, 2022, doi:
10.1016/j.conbuildmat.2022.127213.

[22] K. Wille, A. E. Naaman, S. El-Tawil, and G. J. Parra-Montesinos, ‘Ultra-high-


performance concrete and fiber reinforced concrete: Achieving strength and ductility
without heat curing’, Mater. Struct. Constr., vol. 45(3), 2012, doi: 10.1617/s11527-011-
9767-0.

[23] P. Shen, L. Lu, Y. He, F. Wang, and S. Hu, ‘The effect of curing regimes on the
mechanical properties, nano-mechanical properties and microstructure of ultra-high-
21
performance concrete’, Cem. Concr. Res., vol.118, 2019, doi:
10.1016/j.cemconres.2019.01.004.

[24] J. F. Burroughs, J. Weiss, and J. E. Haddock, ‘Influence of high volumes of silica fume
on the rheological behavior of oil well cement pastes’, Constr. Build. Mater., vol.203,
2019, doi: 10.1016/j.conbuildmat.2019.01.027.

[25] P. P. Li, Y. Y. Y. Cao, H. J. H. Brouwers, W. Chen, and Q. L. Yu, ‘Development and


properties evaluation of sustainable ultra-high performance pastes with
quaternaryblends’,J.Clean.Prod.,vol.240, 2019, doi:10.1016/j.jclepro.2019.118124.

[26] M. A. Mosaberpanah and O. Eren, ‘Effect of quartz powder, quartz sand and water
curing regimes on mechanical properties of UHPC using response surface modelling’, Adv.
Concr. Constr., vol.5(5),2017, doi:https://doi.org/10.12989/acc.2017.5.5.481.

of
[27] EN, BS. Cement-Part 1: Composition, specifications, and conformity criteria for
common cements (2011). British European Standards Specifications. EN BS.

ro
[28] D. Fan et al., ‘A new development of eco-friendly Ultra-High-performance concrete
(UHPC): Towards efficient steel slag application and multi-objective optimization’,
Constr.Build. Mater., -p
https://doi.org/10.1016/j.conbuildmat.2021.124913.
vol.306,2021, doi:
re
[29] ASTM C494/C494M-17, ‘Standard specification for silica fume used in cementitious
mixtures. ASTM International, 2017.
lP

[30] ASTM C143/C143M-15a, ‘Standard Test Method for Slump of Hydraulic-Cement


Concrete’. ASTM International, 2015.
a

[31] ASTM C1609/C1609M-12, ‘Standard Test Method for Flexural Performance of Fiber-
Reinforced Concrete (Using Beam with Third-Point Loading)’. ASTM International, 2012.
rn

[32] ASTM C39/C39M-17. ‘Standard Test Method for Compressive Strength of Cylindrical
Concrete Specimens’. ASTM International, 2017.
u

[33] ASTM C496/C496M-17, ‘Standard Test Method for Splitting Tensile Strength of
Jo

Cylindrical Concrete Specimens’. ASTM International, 2017.

[34] ASTM C1585-13, ‘Standard Test Method for Measurement of Rate of Absorption of
Water by Hydraulic-Cement Concretes’. ASTM International, 2017.

[35] ASTM C494/C494M-17, ‘Standard Test Method for Electrical Indication of Concrete's
Ability to Resist Chloride Ion Penetration’. ASTM International, 2017.

[36] S. Editors, C. Herrmann, and S. Kara, Sustainable Production, Life Cycle Engineering
and Management Progress in Life Cycle Assessment. 2018.

22
[37] Klöpffer, Walter, and Birgit Grahl. ‘Life cycle assessment (LCA): a guide to best
practice’. John Wiley & Sons, 2014.

[38] A. M. Yousef, A. M. Tahwia, and N. A. Marami, ‘Minimum shear reinforcement for


ultra-high-performance fiber reinforced concrete deep beams’, Constr. Build. Mater.,
vol.184, 2018, doi: https://doi.org/10.1016/j.conbuildmat.2018.06.022.

[39] J. Du, Z. Liu, C. Christodoulatos, M. Conway, Y. Bao, and W. Meng, ‘Utilization of


off-specification fly ash in preparing ultra-high-performance concrete (UHPC): Mixture
design, characterization, and life-cycle assessment’, Resour. Conserv. Recycl., vol.180,
2022, doi: https://doi.org/10.1016/j.resconrec.2021.106136.

[40] M. S. M. Norhasri, M. S. Hamidah, and A. M. Fadzil, ‘Inclusion of nano metaclayed as


additive in ultra-high-performance concrete (UHPC)’, Constr. Build. Mater., vol.201, 2019,

of
doi: 10.1016/j.conbuildmat.2019.01.006.

[41] E. R. Teixeira, A. Camões, F. G. Branco, J. B. Aguiar, and R. Fangueiro, ‘Recycling of

ro
biomass and coal fly ash as cement replacement material and its effect on hydration and
carbonation of concrete’, Waste Manag., vol. 94,2019, doi: 10.1016/j.wasman.2019.05.044.

[42] -p
X. Wang, D. Wu, J. Zhang, R. Yu, D. Hou, and Z. Shui, ‘Design of sustainable ultra-
high-performance concrete: A review’, Constr. Build. Mater., vol.307,2021, doi:
re
https://doi.org/10.1016/j.conbuildmat.2021.124643.

[43] B. Zhang, T. Ji, Y. Ma, and Q. Zhang, ‘Effect of metakaolin and magnesium oxide on
lP

flexural strength of ultra-high-performance concrete’, Cem. Concr. Compos., vol. 131, p.


104582, 2022, doi: https://doi.org/10.1016/j.cemconcomp.2022.104582.

[44] M. G. Sohail, R. Kahraman, N. Al Nuaimi, B. Gencturk, and W. Alnahhal, ‘Durability


a

characteristics of high and ultra-high-performance concretes’, J. Build. Eng., vol. 33, p.


rn

101669, 2021, doi: https://doi.org/10.1016/j.jobe.2020.101669.

[45] M. J. Mohd Faizal, M. S. Hamidah, M. S. Muhd Norhasri, I. Noorli, and M. P.


Mohamad Ezad Hafez, ‘Chloride Permeability of Nanoclayed Ultra-High Performance
u

Concrete BT - InCIEC 2014’, 2015, pp. 613–623.


Jo

[46] P. R. Prem, M. Verma, and P. S. Ambily, ‘Sustainable cleaner production of concrete


with high volume copper slag’, J. Clean. Prod., vol. 193, pp. 43–58, 2018, doi:
10.1016/j.jclepro.2018.04.245.

[47] S. Erdem, A. R. Dawson, and N. H. Thom, ‘Impact load-induced micro-structural


damage and micro-structure associated mechanical response of concrete made with
different surface roughness and porosity aggregates’, Cem. Concr. Res., vol. 42, no. 2, pp.
291–305, 2012, doi: https://doi.org/10.1016/j.cemconres.2011.09.015.

23
[48] M. Amin, B. A. Tayeh, and I. S. Agwa, ‘Effect of using mineral admixtures and
ceramic wastes as coarse aggregates on properties of ultrahigh-performance concrete’, J.
Clean. Prod., vol. 273, p. 123073, 2020, doi: https://doi.org/10.1016/j.jclepro.2020.123073.

[49] A. M. Tahwia, O. El-Far, and M. Amin, ‘Characteristics of sustainable high strength


concrete incorporating eco-friendly materials’, Innov. Infrastruct. Solut., vol. 7, no. 1, pp.
1–13, 2022, doi: 10.1007/s41062-021-00609-7.

[50] J. Du et al., ‘New development of ultra-high-performance concrete (UHPC)’, Compos.


Part B Eng., vol. 224, p. 109220, 2021, doi:
https://doi.org/10.1016/j.compositesb.2021.109220.

[51] R. Jing, Y. Liu, and P. Yan, ‘Uncovering the effect of fly ash cenospheres on the
macroscopic properties and microstructure of ultra-high-performance concrete (UHPC)’,

of
Constr. Build. Mater., vol. 286, p. 122977, 2021, doi: 10.1016/j.conbuildmat.2021.122977.

[52] J. Liu, C. Shi, and Z. Wu, ‘Hardening, microstructure, and shrinkage development of

ro
UHPC: A review’, J. Asian Concr. Fed., vol. 5, no. 2, pp. 1–19, 2019, doi:
10.18702/acf.2019.12.5.2.1.

[53] -p
O. M. Abdulkareem, A. Ben Fraj, M. Bouasker, L. Khouchaf, and A. Khelidj,
‘Microstructural investigation of slag-blended UHPC: The effects of slag content and
re
chemical/thermal activation’, Constr. Build. Mater., vol. 292, p. 123455, 2021, doi:
10.1016/j.conbuildmat.2021.123455.
lP

[54] W. Lun Lam, P. Shen, Y. Cai, Y. Sun, Y. Zhang, and C. Sun Poon, ‘Effects of seawater
on UHPC: Macro and microstructure properties’, Constr. Build. Mater., vol. 340, no.
December 2021, p. 127767, 2022, doi: 10.1016/j.conbuildmat.2022.127767.
a

[55] Z. Mo, X. Gao, and A. Su, ‘Mechanical performances and microstructures of


metakaolin contained UHPC matrix under steam curing conditions’, Constr. Build. Mater.,
rn

vol. 268, no. xxxx, p. 121112, 2021, doi: 10.1016/j.conbuildmat.2020.121112.


u

[56] H. Sameer, V. Weber, C. Mostert, S. Bringezu, E. Fehling, and A. Wetzel,


‘Environmental assessment of ultra-high-performance concrete using carbon, material, and
Jo

water footprint’, Materials (Basel)., vol. 16, no. 6, 2019, doi: 10.3390/ma12060851.

[57] K. Scrivener, R. Snellings, B. Lothenbach, and F. Group, A Practical Guide to


Microstructural Analysis of Cementitious Materials. 2018.

[58] N. A. Soliman and A. Tagnit-Hamou, ‘Development of ultra-high-performance


concrete using glass powder – Towards ecofriendly concrete’, Constr. Build. Mater., vol.
125, pp. 600–612, 2016, doi: 10.1016/j.conbuildmat.2016.08.073.

[59] W. Huang, H. Kazemi-Kamyab, W. Sun, and K. Scrivener, ‘Effect of cement


substitution by limestone on the hydration and microstructural development of ultra-high-

24
performance concrete (UHPC)’, Cem. Concr. Compos., vol. 77, pp. 86–101, 2017, doi:
10.1016/j.cemconcomp.2016.12.009.

[60] J. Du, Z. Liu, C. Christodoulatos, M. Conway, Y. Bao, and W. Meng, ‘Utilization of


off-specification fly ash in preparing ultra-high-performance concrete (UHPC): Mixture
design, characterization, and life-cycle assessment’, Resour. Conserv. Recycl., vol. 180, no.
December 2021, 2022, doi: 10.1016/j.resconrec.2021.106136.

[61] M. Radlinski and J. Olek, ‘Investigation into the synergistic effects in ternary
cementitious systems containing portland cement, fly ash and silica fume’, Cem. Concr.
Compos., vol. 34, no. 4, pp. 451–459, 2012, doi: 10.1016/j.cemconcomp.2011.11.014.

[62] A. M. Tahwia, A. M. Heniegal, M. Abdellatief, B. A. Tayeh, and M. Abd Elrahman.


"Properties of ultra-high performance geopolymer concrete incorporating recycled waste

of
glass." Case Studies in Construction Materials 17, e01393, 2022,
https://doi.org/10.1016/j.cscm.2022.e01393.

ro
[63] Alharbi, Yousef R., Aref A. Abadel, Abdulrahman A. Salah, Ola A. Mayhoub, and
Mohamed Kohail. "Engineering properties of alkali activated materials reactive powder
concrete." Construction and
-p
Building
https://doi.org/10.1016/j.conbuildmat.2020.121550.
Materials, 271, 2021,
re
[64] N. S. Ha, S. S. Marundrury, T. M. Pham, E. Pournasiri, F. Shi, and H. Hao, ‘Effect of
grounded blast furnace slag and rice husk ash on performance of ultra-high-performance
concrete (UHPC) subjected to impact loading’, Constr. Build. Mater., vol. 329, no.
lP

December 2021, p. 127213, 2022, doi: 10.1016/j.conbuildmat.2022.127213.

[65] A. Dixit, H. Du, and S. D. Pang, ‘Carbon capture in ultra-high-performance concrete


using pressurized CO2 curing’, Constr. Build. Mater., vol. 288, p. 123076, 2021, doi:
a

10.1016/j.conbuildmat.2021.123076.
u rn
Jo

CRediT authorship contribution statement

Conceptualization, M.A., W.E., and A.T.; methodology, M.A. and S.A.; validation,
W.E., H.A., and A.T.; formal analysis, M.A., S.A., and G.E.; investigation, M.A. and G.E.;
resources, A.T. and H.A.; data curation, G.E. and S.A.; writing—original draft preparation,
M.A., S.A., and G.E.; writing—review and editing, W.E., H.A., and A.T.; visualization, A.T.;
supervision, H.A. and A.T.; project administration, A.T. and H.A.; funding acquisition, H.A.
All authors have read and agreed to the published version of the manuscript.
25
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.

☐ The authors declare the following financial interests/personal relationships which may be
considered as potential competing interests:

of
ro
-p
re
a lP
u rn
Jo

26
Fig. 2
Fig. 1
Jo
u rn
alP

27
re
-p
ro
of
Fig. 4
Fig. 3
Jo
urn
a lP

28
re
-p
ro
of
Fig. 5
Jo
urn
alP

29
re
-p
ro
of
Fig. 7
Fig. 6
Jo
urn
alP

30
re
-p
ro
of
Fig. 9
Fig. 8
Jo
urn
alP

31
re
-p
ro
of
Fig. 10
Jo
urn
alP

32
re
-p
ro
of
of
ro
-p
re
Fig. 11
lP
a
u rn
Jo

Fig. 12

33
of
ro
Fig. 13
-p
re
lP
a
u rn
Jo

Fig. 14

34
of
Fig. 15

ro
-p
re
lP
a
u rn
Jo

Fig. 16

35

View publication stats

You might also like