0% found this document useful (0 votes)
24 views126 pages

JBE D 25 03660 Reviewer

Uploaded by

lileshgautam.set
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
24 views126 pages

JBE D 25 03660 Reviewer

Uploaded by

lileshgautam.set
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 126

Journal of Building Engineering

Graphene Oxide reinforced cement-based composites: A Comprehensive Review on


Mechanical, Thermal, and Durability Properties towards practical applications
--Manuscript Draft--

Manuscript Number: JBE-D-25-03660

Article Type: Review Article

Section/Category: Building materials

Keywords: Graphene oxide; Concrete nanoreinforcement; Cement-Based Composites;


Scalability of GO; Mechanical Strength Enhancement

Abstract: Graphene oxide is a transformative additive for concrete, providing multifaceted


enhancements on mechanical, thermal, electrical, and durability properties. This
comprehensive review consolidates the recent research findings on the integration of
GO into cement-based materials, that contribute to substantial improvements in
mechanical strength, thermal conductivity, and resistance to freeze-thaw cycles.
Additionally, GO's role in reducing porosity and enhancing the interfacial transition
zones (ITZ) within the concrete matrix is explored, showcasing its potential to mitigate
microcracking. Unlike complex composites that are difficult to synthesize and scale,
graphene oxide (GO) presents a simpler and more scalable solution. However, its
adoption in real-world applications remains limited. This study addresses the scalability
challenges associated with GO incorporation, focusing on the production techniques
for large-scale implementation. Additionally, economic and environmental implications
of GO in concrete are analyzed, with an emphasis on sustainable production and cost-
effective adaptation for harnessing GO’s full potential in industrial and infrastructure
development.

Powered by Editorial Manager® and ProduXion Manager® from Aries Systems Corporation
Graphical Abstract
For Publication - Latest Manuscript (Clean Version Only)

Graphene Oxide reinforced cement-based composites: A


Comprehensive Review on Mechanical, Thermal, and
Durability Properties towards practical applications

Stanleydhinakar Mathan a, Vijayan Murugesan a, Balaji Chettiannan a, Manickam


Selvaraj b,c, Mohammed A. Assiri b,c, Gowdhaman Arumugam a, Ramesh Rajendran a*

a Department of Physics, Periyar University, Salem, Tamilnadu 636011, India


b Department of Chemistry, Faculty of Science, King Khalid University, Abha, 61413, Saudi Arabia
c Research Centre for Advanced Materials Science (RCAMS), King Khalid University, PO Box 9004,
Abha 61413, Saudi Arabia

Abstract:

Graphene oxide (GO) has garnered significant attention as a transformative additive for
concrete, providing multifaceted enhancements on mechanical, thermal, electrical, and
durability properties. This comprehensive review consolidates and analyzes recent research
findings on the integration of GO into cement-based materials, highlighting its high surface
area, excellent dispersibility, and surface functional groups that contribute to substantial
improvements in compressive and flexural strength, thermal conductivity, and resistance to
freeze-thaw cycles. Additionally, GO's role in reducing porosity and enhancing the interfacial
transition zones (ITZ) within the concrete matrix is explored, showcasing its potential to
mitigate microcracking and improve structural integrity. Unlike complex composites that are
difficult to synthesize and scale, graphene oxide (GO) presents a simpler and more scalable
solution. However, its adoption in real-world applications remains limited. This study
addresses the scalability challenges associated with GO incorporation, focusing on the
contemporary production techniques for large-scale implementation. Additionally, economic
and environmental implications of GO in concrete are analyzed, with an emphasis on
sustainable production and cost-effective adaptation. This review seeks to bridge the gap by
consolidating recent advancements and addressing key implementation challenges to provide
a foundation for harnessing GO’s full potential in industrial and infrastructure development.

Keywords: Graphene oxide, Concrete nanoreinforcement, Cement-Based Composites,


Scalability of GO, Mechanical Strength Enhancement
Contents
Abstract: .................................................................................................................................. 1
1. Introduction ......................................................................................................................... 3
2. Scope and Objectives ............................................................................................................ 5
3. Interaction mechanism of GO with Cement ........................................................................... 6
3.1 Dispersion Techniques and Challenges ..................................................................................... 9
3.2 Impact of GO on Workability................................................................................................... 15
4. Mechanical Properties of GO-Enhanced Concrete ................................................................ 18
4.1 Compressive strength.............................................................................................................. 20
4.2 Flexural Strength and Toughness ............................................................................................ 28
4.3 Influence of GO on Porosity and Microcracking ..................................................................... 31
4.4 Elastic Modulus and Deformation Behaviour ......................................................................... 36
5. Thermal and Electrical properties ........................................................................................ 46
5.1 Thermo-Electrical Response and Heat Transfer ...................................................................... 47
5.2 Dimensional Stability and Shrinkage ....................................................................................... 54
5.3 Electrical properties ................................................................................................................ 58
6. Durability Performance ....................................................................................................... 61
6.1 Water sorptivity and Permeability .......................................................................................... 61
6.2 Chloride Ion Penetration and Corrosion Resistance ............................................................... 65
6.3 Sulfate Ion Resistance ............................................................................................................. 68
6.4 Freeze-thaw resistance ........................................................................................................... 72
7. Scalability ........................................................................................................................... 76
8. Case Studies ....................................................................................................................... 83
9. Economic Viability and Environmental Impacts ................................................................... 89
10. Summary and Future Prospects ......................................................................................... 96
11. Conclusion ...................................................................................................................... 100
Conflict of interest ................................................................................................................ 101
Acknowledgments ................................................................................................................ 102
References ........................................................................................................................... 102

2
1. Introduction

Cement consumption is expected to rise substantially as the number of construction projects

increases to meet the growing global demand for buildings and infrastructure [1]. Cement is

considered to be the second most consumed product in the world, next to water. However,

traditional concrete suffers from its inherent brittle nature and low tensile strength that cause

crack formation, resulting in poor durability and high maintenance costs [2]. In order to

overcome these issues, reinforcements are generally used. Fibre reinforcements such as glass,

steel, and carbon are widely used to resist crack propagation [3], [4]. However, these

reinforcements are predominantly relied on physical mechanisms that do not alter the cement

hydration products. These fail to prevent crack initiation at the nanoscale. In contrast,

nanomaterials having small sizes and high surface-to-volume ratios can influence the properties

of the matrix at the nano-scale level. Many attempts have been made to incorporate different

nanomaterials, including nano-silica and nano-titanium dioxide, into cement matrix that can

significantly improve its mechanical properties and durability [5]. However, the development

of new nanomaterials with superior properties remains essential. [6], [7]

In cement production, depending on the required properties, construction conditions, and the

desired application, five main types of common cement are used. These include Portland

cement (CEM I), Portland cement with supplementary materials such as blast furnace slag, fly

ash, pozzolana, limestone and silica fume (CEM II), blast furnace cement (CEM III),

pozzolanic cement (CEM IV), and composite cement (CEM V) [8]. While supplementary

cementitious materials can be tailored for specific properties, ordinary Portland cement still

remains the most widely used concrete binder.

However, its shortcomings, such as brittleness and cracking, along with the growing global

demand for lighter and more slender concrete structures, have accelerated the extensive

3
research efforts into the use of nanomaterials in concrete. The incorporation of nanomaterials

in the cement matrix offers a wide range of potential improvements by significantly altering

the intrinsic properties of cement composites, paving the way for the development of

innovative and multifunctional materials. [9], [10]

Currently, graphene and its derivatives are gaining attention, owing to its outstanding properties

even in their variations, such as Graphene oxide (GO) and reduced graphene oxide (rGO).

Though they are the same material, they exhibit different properties [11]. For instance,

graphene is hydrophobic and electrically conductive, which must be modified with a surfactant

to be dispersed in a medium, but GO is hydrophilic but electrically insulating, offering unique

advantages for specific applications. [12], [13]

GO must be reduced to reduced graphene oxide (rGO) before being extensively used for

electrical conductivity applications. Despite these differences, graphene oxide is widely studied

a lot due to its easy synthesis, cost-effectiveness, and direct addition into the matrix compared

to that of pristine graphene. While concrete with GO has demonstrated promising results, its

broader adaptation remains a challenge. This is partly due to the differences in materials,

preparation protocols, and experimental methods that present challenges in drawing consistent

conclusions. Factors like lateral size, defects, dispersion methods, and workability can also

significantly impact the properties and outcomes of the cement matrix. Moreover, multiple

explanations have been proposed for the mechanism by which GO influences cement

properties, with some being consistent and some contradictory [14][15]. Therefore, a

systematic summary of current research on the effects of GO on cement matrices is necessary.

This review highlights the effects of GO on the cement matrix across multiple aspects,

including its mechanical, thermal, and durability properties, with a focus on large-scale

adaptation. Additionally, economic and environmental analyses are also summarized. To ensure

4
the scope and consistency of the work, this review is purposefully limited to pure graphene

oxide (GO), excluding its derivatives and composites. Overall, it aims to assist researchers and

industries in gaining a deeper understanding of GO's role in concrete and to provide valuable

direction for future research on GO applications in cement composites.

Fig.1 Multifunctional Properties of GO enhanced Concrete

2. Scope and Objectives

 To explore the fundamental physical and chemical interactions between GO and

cementitious materials to understand the GO’s influence on performance.

 To assess the effectiveness of graphene oxide (GO) in concrete properties, including

durability, mechanical strength, and thermal performance.

5
 To identify and address current challenges in GO-incorporated concrete research, with

a focus on scalability and practical adaptation.

 To review the current status of GO adaptation in concrete through relevant case

studies, highlighting real-world applications.

 To Provide targeted recommendations for future research to optimize GO

incorporation for potential benefits.

3. Interaction mechanism of GO with Cement

In recent years, the integration of graphene-based materials into cementitious composites

has gained significant interest due to its potential to enhance mechanical properties and

durability, as shown in Fig.1. Understanding the complete mechanisms behind these

interactions is essential for the development of next-generation concrete materials with superior

performance and functionalities [13], [16]. Broadly, the key interactions between graphene-

oxide based materials and concrete can be categorized into two main types, physical

interactions between graphene sheets themselves and chemical interactions between graphene

and the cement hydration products.

i) Physical interaction

In terms of physical interaction, the two-dimensional nature of Graphene, with its large

surface area and high aspect ratio, allows for its uniform distribution within the

cementitious matrix. Achieving a homogeneous and well-dispersed graphene network

is essential for optimizing the composite's performance. This dispersion is primarily

facilitated by physical interactions such as Van der Waals forces and π-π stacking, which

stabilizes the graphene sheets within the matrix. In a dispersed state, the weak Van der

Waals forces govern the interaction between the individual graphene sheets,

maintaining stability through a delicate balance of attractive and repulsive forces. These

6
forces arise from temporary fluctuations in electron distribution within atoms and

molecules, creating transient dipole moments. Adjacent graphene layers experience an

attractive force due to these induced dipoles that draw them together. However, as the

electron clouds of adjacent layers overlap excessively, a repulsive force emerges due to

electron-electron interactions, preventing uncontrolled collapse of the layers. This

delicate balance between attraction and repulsion ensures the individual graphene

sheets remain separate and well-dispersed within the matrix [17].

Additionally, the honeycomb lattice structure of graphene features aromatic rings

formed by delocalized π electrons in the carbon-carbon double bonds. When adjacent

graphene layers adopt a parallel or near-parallel orientation, the aromatic rings in one

layer can interact with those in the neighbouring layer through π-π stacking. This

stacking interaction stabilizes the layered structure, which contributes to the overall

stability of the graphene sheet within the matrix. The strength of these π-π stacking

interactions also determines the spacing between these layers. Well-defined stacking

through π-π interactions can facilitate efficient charge transport, potentially leading to

enhanced electrical conductivity. However, excessive stacking can lead to undesired

aggregation. Therefore, achieving an optimal balance in π-π stacking interactions is

crucial for maintaining a well-dispersed graphene network within the matrix [18].

ii) Chemical Interaction

Chemical interaction between graphene oxide and cementious materials presents a

fascinating example of synergy at the atomic level. These intricate mechanisms take

place at molecular and atomic scales, where the materials interact chemically. Such

7
interactions can significantly influence the properties of the cement matrix, including

its microstructure, hydration kinetics, and mechanical performance.

Calcium silicate hydrate (C-S-H) gel is a key component of cementitious materials that

plays a crucial role in determining the strength of concrete. Graphene's functional

groups, particularly oxygen, can interact chemically with C-S-H gel, promoting its

growth and thereby enhancing the overall properties of the composite. Furthermore, the

presence of graphene oxide might inhibit the formation of detrimental expansive phases

like ettringite, improving the concrete's durability. Studies suggest that negatively

charged graphene oxide (GO) can react with C-S-H to form new phases resembling

tobermorite and jennite, which are known for their superior strength and stability,

potentially due to its consumption of calcium ions (Ca2+) [19].

This interaction highlights the importance of understanding the initial composition of

cement, as its mineral phases play a crucial role in these reactions. In its anhydrous

state, cement comprises various blends of compounds, including C3S - Tricalcium

silicate (Ca3SiO5), C2S - Dicalcium silicate (Ca2SiO4), C3A - Tricalcium aluminate

(Ca3Al2O6) and other components. During hydration, these compounds react to form

various phases, including Ettringite (Ca6Al2(SO4)3(OH)12·26H2O), Mono sulfate

(Ca4Al2(OH)6SO4·12H2O), Calcium hydroxide (Ca(OH)2), and Calcium silicate

hydrate gel (Ca3SiO2·nH2O), which significantly influences concrete strength. [13]. The

hydration process involves a series of chemical reactions that lead to the formation of

calcium silicate hydrates (C-S-H), calcium hydroxide (CH), ettringite, and other phases,

all of which contribute to the overall behaviour of the composite.

In addition to influencing hydration, functionalized graphene, such as Graphene oxide,

offers a broader spectrum of possibilities for various chemical interactions. For

8
example, the hydroxyl groups of GO can form hydrogen bonds with hydroxyl groups

present in the cement hydrates, further influencing the properties of the composite.

Electrostatic interactions and ion exchange processes may also occur between the

charged functional groups on graphene oxide and charged species within cement

hydrates. The surface charge of graphene oxide, which can be manipulated through

functionalization, impacts its interaction with ions.

3.1 Dispersion Techniques and Challenges

The incorporation of any nanomaterials into cement composites heavily relies upon

achieving homogeneous dispersion within the cement matrix. Uneven dispersion of these

materials can negatively impact the microstructure, mechanical properties, and durability of

the material. This challenge becomes particularly significant for 2D materials like graphene,

which faces inherent dispersion problems due to its hydrophobicity. The Strong Van der Waals

forces and the hydrophobic nature of pristine graphene make it difficult to achieve uniform

dispersion within aqueous solutions commonly used in cement [20].

This is where Graphene oxide gains an edge [21]. Unlike pristine graphene, GO readily

disperses in solutions due to the presence of oxygen-containing functional groups. These

functional groups introduce polarity to the GO sheets, which promotes hydrogen bonding and

electrostatic interactions with water molecules and other polar solvents. Fig.2a shows the

uniform dispersion of GO throughout the matrix by 3D CT images of both the control matrix

(C) and the GO-enhanced matrix (CG4). This polarisation characteristic allows for good

dispersion in water, supported by electrostatic repulsion and strong hydrophobicity. However,

this dispersion is not entirely stable over time and is prone to reaggregation. Several factors

contribute to this limitation. A study by Sabziparvar et al. [22] highlighted the mechanism of

9
GO aggregation in a cement matrix through experimental and computational simulations and

concluded that the GO dispersion is strongly affected by cations, particularly in the high-

alkaline environment of cement pore solutions. The alkaline cement pore solution contains

charged ions, such as Ca2+, Mg2+, Na+, and OH−, that can cause agglomeration of GO sheets

by restricting the liquid penetration between the layers [23]. Additionally, GO has a low

tolerance for high Ca2+ concentration, which leads to bridging and agglomeration of GO sheets

when the concentration exceeds a certain threshold [24]. Furthermore, the oxygen-containing

functional groups on GO can form complexes with available calcium ions, further promoting

aggregation. Recent studies suggest that the deprotonation of GO's carboxyl groups in the

highly alkaline environment increases the amount of adsorbed Ca²⁺ ions, exacerbating the issue

even further [25].

Fig.2a 3D X-ray CT images of cement paste showing uniform dispersion of GO in the cement
matrix compared to a control sample; without GO (left), with 0.4wt% GO (right) [18]; Fig.2b
Dispersion mechanism of GO-PCE in cement paste at various concentrations, highlighting
steric hindrance and reduced aggregation [26].

This aggregation issue in GO significantly hinders the benefits of them in cement composites.

When GO aggregates, it loses its high specific surface area, which is a key property for effective

10
interaction with the cement matrix. The reduced surface area weakens the GO’s ability to act

as a nucleation site, hindering cement hydration and the development of a dense microstructure.

As a result of this, the strength of the composite material diminishes. Studies have shown that

GO aggregates larger than 1.56 μm lose their nano or micro-filler properties and behave more

like macro-scale materials, negatively impacting the matrix and reducing mechanical

performance [27], [28]. Therefore, achieving good dispersion with smaller GO aggregates and

a larger overall specific surface area is essential for maximising the effectiveness of GO within

the cement composite.

To address these limitations caused by aggregation, researchers have developed various

fabrication techniques for GO-cement composites in recent years. These techniques primarily

focus on achieving homogeneous dispersion of GO within the cement matrix. In general, two

main approaches are used, good physical dispersion and chemical modification. The physical

dispersion method relies on applying mechanical energy through high-speed shearing or

ultrasonic waves, providing the additional energy that disrupts the chemical interactions

between GO sheets, which weakens Van der Waals forces to promote better dispersion.

Ultrasonication is particularly effective in achieving good dispersion in aqueous solutions.

However, scaling up this process for larger applications becomes impractical due to time

constraints and energy consumption [29]. To overcome the limitations of physical methods,

surfactant-assisted dispersion techniques are widely explored. Surfactants adsorb onto the GO

surface, lowering its overall surface energy. This facilitates better distribution of GO particles

within the cement matrix through electrostatic repulsion between the surfactant-coated GO

sheets [30]. Additionally, Surfactants can be used to graft or copolymerize additional functional

groups onto the GO surface. This further enhances the hydrophilicity of GO, promoting better

compatibility with the cement matrix and reducing the tendency for aggregation [30].

11
While silica fume and fly ash are commonly used mineral admixtures in cementitious materials,

they are not ideal for addressing GO dispersion challenges. Their primary benefits lie in their

filling effect, nucleation effect, and pozzolanic reaction, which enhance the overall properties

of the composite. Fly ash, in particular, offers additional advantages like reducing temperature

shrinkage and improving slurry consistency. Due to its ball effect, it could solve the GO’s

fluidity issue by improving them. However, the entire mechanism of silica fume and fly ash in

GO agglomeration still needs clarity. Studies have shown that incorporating silica fume can

actually worsen GO agglomeration, which complicates the desired dispersion. [31], [32]

Agglomerated GO, rather than well-dispersed sheets, negatively impacts the final properties of

the composite [32], [33]. Therefore, researchers have explored alternative surfactants like

lignosulfonate (LS), polycondensate β-naphthalene sulfonate formaldehyde (PNS), and

polycarboxylate superplasticizer (PC). Among these, polycarboxylate ether (PCE), a type of

PC, has emerged as the most promising option [34]. PCE is considered a promising surfactant

for GO-cement composites due to its multi-functional benefits. It exhibits good dispersion

properties, ensuring its effectiveness within a highly alkaline cement environment. It not only

reduces the aggregation of GO sheets but also ensures the fluidity of the cement composite

during mixture and placement, which can be negatively affected by GO incorporation. This can

be attributed to the unique comb-like structure of PCE molecules that provide steric hindrance.

PCE’s carboxyl groups can bind Ca²⁺ ions, with one Ca²⁺ binding to multiple carboxyl groups.

This interaction allows PCE adsorbed on GO sheets to form complexes with Ca²⁺, reducing

free Ca²⁺ concentration around GO and preventing agglomeration. After adsorption, PCE’s

side-chain steric hindrance disperses the GO sheets and blocks Ca²⁺ from reaching their

surfaces. Longer side chains or higher PCE charge density enhance GO dispersibility, as shown

in Fig.2b [26]. This means the bulky structure of the PCE molecules physically separates GO

sheets, further preventing aggregation. In addition to that, PCE molecules containing functional

12
groups like –COO– and –OH interact with the GO surface through electrostatic repulsion,

preventing the sheets from clumping together and aggregating [35].

In a work by Jing et al. [36] X-ray computed tomography of GO distribution is developed to

study the effect of PCE in concrete. Fig.3 illustrates 3D CT scans comparing the distribution

and aggregation of graphene oxide (GO) in a cement matrix with and without polycarboxylate

ether. In the controlled sample without PCE (Fig.3 a,b,c), the distribution of GO is limited, with

visible agglomerations. Fig.3a presents a grayscale CT scan where darker areas indicate GO

particles, dispersed yet prone to clustering within the matrix. The constructed 3D view in

(Fig.3b) uses colour coding to illustrate the size and distribution of these GO clusters,

highlighting the tendency of GO to aggregate. The (Fig.3c) further emphasizes this clustering,

revealing the spatial extent of GO within the matrix.

In contrast, (Fig.3 d,e,f) depicts the GO distribution in a matrix with PCE added. Here, the PCE

effectively reduces GO agglomeration, leading to a more uniform distribution. The grayscale

CT scan in (Fig.3d) shows fewer and smaller darker clusters compared to the controlled sample,

while the 3D view in (Fig.3e) displays a more even spacing of GO particles, indicating

improved spatial distribution. The spatial view in (Fig.3f) confirms this trend, with GO

particles more evenly dispersed throughout the matrix. These images demonstrate that the

addition of PCE helps to prevent GO particles from aggregating, resulting in a more consistent

and homogeneous integration of GO within the cement matrix.

13
Fig.3 3D CT images of GO distribution and aggregation in cement matrix without PCE (a–c)
vs with PCE (d–f), showing improved uniformity [36]

Numerous studies have shown that polycarboxylate ethers (PCE) are highly effective

dispersants for GO in cement composites [37], [38], [39]. However, some factors need to be

considered. Ghazizadeh et al. [40] suggests that the effectiveness of PCE depends heavily on

the dosage, as insufficient addition may have minimal impact on dispersion. Conversely,

excessive use of it can negatively impact the fluidity of the concrete [41]. Li et al. [10] also

emphasized the importance of good dispersion for functionalities like self-sensing in cement

composites, as GO aggregates lack electrical conductivity. These findings highlight the need

for further optimization of PCE usage for consistent and effective GO dispersion.

Despite the advantages of PCE, several crucial factors beyond its selection can also influence

the effectiveness of GO dispersion. Studies suggest that the sonication frequency (energy

14
input), the timing of PCE addition, and the pH of the medium play a major role in achieving

optimal dispersion [42],[43]. In addition to this, the practical implementation of PCE also faces

challenges. Although ultrasonic dispersion and chemical surface modification are effective,

these methods are complex and time-consuming, making them less feasible for large-scale

applications. Therefore, future research should focus on developing advanced dispersants that

effectively disperse GO sheets without compromising other concrete properties. Additionally,

there is a lack of standardized methods for quantifying the degree of GO dispersion within the

cement matrix. Establishing such quantitative methods for evaluating the degree of GO

dispersion in cementitious composites is essential for optimizing dispersion techniques and

ensuring consistent performance at scale.

3.2 Impact of GO on Workability

Workability refers to the ease with which fresh concrete can be placed and compacted.

Essentially, it measures the concrete's flowability and consistency, which directly influences

the overall quality and strength of the final product. It is a crucial property for ensuring the

performance of cementitious materials. Good workability translates to better homogeneity,

fewer air voids, and improved strength and durability of the final concrete products.

Workability is a complex property that includes factors like rheology (flow behaviour),

contractility (shrinkage), and thermal diffusion. For successful casting and achieving the

desired structural properties, fresh concrete must exhibit good plasticity or moldability. The

flowability of fresh concrete is typically assessed using the mini-slump cone test, which is a

standardized test involving the measurement of the vertical distance a concrete sample slumps

after the mould is removed. A higher slump value indicates better flowability. Standards for

this test include IS 1199:2018 (India) and ASTM C1437-07 (US).

15
In terms of GO, several studies have shown that GO has a negative influence on the workability

of concrete as measured by the slump test. For instance, Gong et al. [44] observed a significant

reduction of 34% in slump value when incorporating 0.3 wt% of GO compared to the control

sample, as shown in Fig.4a. This reduction was attributed to the formation of calcium cation

agglomerates through chemical cross-linking with GO sheets. These agglomerates trap free

water within the cement matrix, hindering its flow and reducing the overall fluidity of the

matrix. Similar findings were observed by Akarsh et al. [45], they observed a 20% reduction

in slump value with a 0.2 wt% GO addition. This decrease can also be attributed to the high

specific surface area of GO. The large surface area of GO flakes acts like a sponge, absorbing

and entrapping water molecules, which reduces the amount of free water available for

lubrication during mixing and placement, consequently affecting the workability.

Wang et al. [46] employed a more scientific approach using the Bingham model to study the

rheological behaviour (flow properties) of cement paste with GO, and the results further

confirmed a decrease in fluidity with increasing GO content. Rheometer measurements

indicated a rise in both plastic viscosity and yield stress, as shown in Fig.4b, signifying a more

rigid and less flowing material.

Fig.4a Slump test results showing reduced workability with GO-added cement [44]; Fig.4b
Increase in yield stress and plastic viscosity of cement with increasing GO dosage [46]

16
While incorporating GO can significantly reduce the workability of cement paste, the addition

of naphthalene-based superplasticizers and polycarboxylic ethers (PCEs) offers a promising

solution to overcome this limitation. Studies have shown a remarkable improvement in

workability when PCE is used with GO compared to GO alone [47], [48]. For instance, Roy et

al. [49] observed a minimal reduction of only 2.8% inflow value with PCE-added GO samples,

which is significantly lower compared to the reductions reported by Gong et al. al. [44] and

Akarsh et al. [45] without PCE.

Laser Confocal Scanning Microscopy (LCSM) by Wang et al. [46] offers valuable insights into

how PCE enhances the performance of cement paste containing graphene oxide (GO). They

revealed that the PCE promotes the formation of flocculation structures within the cement

paste, which are significantly smaller and looser compared to those formed with GO alone.

These structures don’t hinder water movement, which results in improved workability. Also,

its positive impact arises from the PCE’s structure and its unique interaction with the cement

particles and GO. The hydrophobic groups of PCE actively attach themselves to the cement

particle surface, while the comb-branched chains with hydrophilic groups extend outwards,

causing steric hindrance and electrostatic repulsion. This prevents the GO from adhering to

cement particles and potentially causing water entrapment. Additionally, it disrupts existing

large structures and liberates trapped water, which significantly improves the workability of

GO cement paste. This highlights the importance of PCE as a potential solution to counteract

the negative effects of GO on workability.

However, it's crucial to remember that achieving optimal concrete performance requires careful

consideration of several factors, including the water-to-cement ratio, setting time, aggregate

properties, testing environment, plastic shrinkage, and segregation. Also, excessive use of PCE

for improved workability may cause a delay in cement hydration properties. So, finding the

right balance between these aspects is key to achieving optimal performance. Future research

17
should focus on optimizing PCE dosage and exploring alternative dispersants that can

effectively enhance workability without compromising the hydration properties of concrete.

4. Mechanical Properties of GO-Enhanced Concrete

Concrete remains a dominant construction material due to its affordability, durability,

and versatility. However, it suffers from limitations like low tensile strength, brittleness, and

cracking. These limitations arise from its inherent microstructural defects and heterogeneities

within the cementitious matrix that compromise its structural integrity and service life. These

flaws make them vulnerable to degradation processes such as carbonation, alkali-silica reaction

(ASR), and freeze-thaw cycles [50], [51].

Traditional methods to enhance concrete performance include steel reinforcement, fiber

reinforcement, and polymer modification [52], [53]. However, these approaches often have

drawbacks, including increased weight, susceptibility to corrosion, and reduced workability.

The pursuit of enhancing these mechanical characteristics of concrete led to innovative

explorations in materials [54], [55], [56]. Wherein the integration of graphene-based materials,

particularly Graphene oxide, emerges as a promising solution. Studies have demonstrated that

incorporating GO into concrete matrices significantly enhances tensile strength, flexural

strength, and toughness, effectively addressing the traditional concrete’s deficiencies. For

instance, A study by Bheel et al. [57] found that even a small amount of GO, such as 0.05 wt%,

can lead to notable improvements, such as its compressive strength increased by 63.78 MPa,

its tensile strength by 4.98 MPa, its flexural strength by 6.92 MPa, and Modulus of elasticity

by 41.50 GPa, and its drying shrinkage reduced by approximately 45.71% compared to control

samples. Similarly, in a study conducted by Bagheri et al. [58] the incorporation of graphene

oxide (GO) demonstrated significant enhancements in both the compressive and flexural

18
strength of concrete. Specifically, the addition of GO increased compressive strength by up to

46% after 7 days and by 28% at 28 days. Long-term improvements were also observed in

flexural strength, with an increase of up to 50% at 28 days, emphasising GO’s substantial

impact on the performance of concrete. The effects of GO on compressive and flexural strength

across various concentrations are shown in Fig.5a.

Fig.5a Improvements in compressive and flexural strength at various GO concentrations [58];


Fig.5b (i) Compressive strength vs. GO content, (ii) percentage increase in strength with GO
incorporation [59].

The incorporation of Graphene Oxide in concrete offers a unique opportunity to modify its

nanostructure, resulting in superior mechanical performance compared to traditional concrete.

This emerges as a critical factor for why most of the existing studies prefer to use GO to

improve the mechanical properties and durability performance of cement-based composites

[60], [61]. As previously discussed, the exceptional hydrophilicity and water dispersibility

nature of GO is favourable for its incorporation into cement to form a uniform composite.

While the potential of GO is evident, a deeper understanding of its mechanical effects on

concrete is crucial for its mass adoption in real-world applications. The following sections will

delve into the mechanisms by which GO contributes to improved mechanical strength.

19
4.1 Compressive strength

The compressibility of the concrete stands as a critical parameter in the ability of concrete to

withstand compressive loads without catastrophic failure. It plays a crucial role in the design

of infrastructure and serves as an important indicator of concrete quality. A huge consideration

is made directly into its ability to withstand load and its performance under various situations

and tensions.

Concrete's resistance to compressive forces is largely due to its dense, interlocking

microstructure formed during hydration, which exhibits remarkable resilience under

compressive stresses [62]. However, it suffers from inherent brittleness, which limits its ability

to withstand dynamic or varying loads without cracking. To enhance its compressive strength,

the concrete is often reinforced with materials and additives, and the integration of graphene

oxide in concrete matrices is emerging as a promising avenue.

Studies have consistently shown that incorporating graphene oxide (GO) into concrete

enhances its compressive strength [61], [63], [64]. A study by Fonseka et al. [59] demonstrated

that GO enhances the performance of concrete, as shown in Fig.5b. In Fig.5b(i), the relationship

between compressive strength and GO content is presented, while Fig.5b(ii) illustrates the

percentage increase in compressive strength with GO addition. The results indicate that the

concrete mix containing 0.08% GO achieved the highest compressive strength at both 28 and

56 days, reaching 47.26 MPa and 54.05 MPa, respectively. This corresponds to a 21%

improvement at 28 days and a 19% improvement at 56 days compared to the control mix.

Jing et al. [65] reported a 15.05% increase in compressive strength with 0.6 wt% GO, while

Gong et al. [44] reported a remarkable improvement of over 40% in the mechanical

performance of cement composites with just 0.03wt% GO, attributing this to the GO’s

functional groups and excellent dispersibility that promote strong bonding within the matrix.

20
Similarly, Devi et al. [66] achieved a 49% increase with 0.08wt% of GO. These findings

collectively demonstrate the significant potential of GO as a strength-enhancing agent for

concrete.

While the positive impact of GO is clear, understanding the underlying mechanism is crucial

for optimizing its use. The observed improvement in compressive strength in GO-enhanced

concrete can be attributed to two key factors, enhanced interfacial bonding and improved

hydration kinetics.

Strong interfacial bonding between the cementitious matrix and GO is crucial for maximising

the concrete strength. When the cementitious matrix forms a strong connection with the added

GO, the applied load is effectively transferred throughout the material. This even distribution

strengthens the overall performance of the concrete. Conversely, weak interaction between the

additive and matrix poses a significant risk. If GO sheets are poorly bonded, under load, they

can detach from the matrix, causing stress to concentrate at specific points and potentially

leading to cracks or failure [67], [68].

Hydration kinetics, the chemical reaction between cement and water, plays a vital role in

concrete's structural performance. To understand the hydration in detail, Ordinary Portland

Cement (OPC), the most commonly used cement, consists of four main minerals: Calcium

trisilicate (C3S), Calcium disilicate (C2S), Tricalcium aluminate (C3A) and Tetracalcium

alumino ferrite (C4AF). Of these, C3S is the most reactive phase and is particularly important

in the early stages of hydration. It readily reacts with water to form calcium silicate hydrate (C-

S-H), a very fine and highly cohesive material responsible for the binding properties of cement.

The C3S hydration leads to increased thixotropy (resistance to flow) due to the formation of C-

S-H bridges between cement grains. As hydration progresses, a slower dormant period allows

for further C-S-H precipitation, gradually filling the spaces between cement particles. This

21
progressive filling contributes to the overall setting and hardening of the concrete, transforming

it from a fluid state to a solid structure [69], [70]. Thus, the hydration process plays a crucial

role in determining the properties of concrete. Enhancing this process can directly improve the

overall performance of the material. In this context, the role of graphene oxide (GO) in

enhancing concrete properties becomes significant.

Hou et al. [71] employed simulations and experiments to investigate the chemical interaction

between GO and C-S-H gel. Their findings revealed that the interfacial strength between these

elements plays a significant role in boosting the mechanical properties of the concrete. Beyond

the interfacial effect, GO's influence extends to the formation of hydration crystals themselves.

The oxygen-containing functional groups on GO's surface actively interact with cement

components (C3S, C2S, and C3A), creating growth points that serves as templates for the

continuous and organized formation of hydration crystals. This process is well illustrated in the

work of Lv et al. [72] as shown in Fig.6a. These growth points promote the continuous

formation of column-shaped crystals (Aft, AFm, CH, and C-S-H gel) that grow outwards.

These crystals then transform into interlocking floral patterns, effectively filling pores, cracks,

and loose areas within the concrete. By filling these voids, the floral-patterned crystals prevent

cracks from propagating, leading to a denser and more robust microstructure throughout the

concrete composite. This densification is further confirmed by Quereshi et al. [73] who

observed that the ~10mm pores in GO cement samples were filled with hydration products.

Essentially, GO acting as a conductor, directing the growth of hydration crystals to strengthen

the overall structure and promoting hydration [61].

22
Fig.6a Template effect in hydration mechanism of GO added concrete, leading to flower-like
crystal formations [72]. Fig.6b SEM images showing hydration progression in GO added
cement at 1,3,7,28,60,90 days of curing (A,B,C,D,E,F) [61].

In addition, Lv et al. [61] observed the formation of unique snowflake-like crystals upon

incorporating 0.03wt% of GO, leading to a remarkable 60.7% improvement in the flexural

strength of the cement paste. Fig.6b shows the SEM images of GO-incorporated concrete

illustrating the progression of cement hydration at different curing periods (Fig.6b A–F

representing 1, 3, 7, 28, 60, and 90 days, respectively). Image A-C demonstrates the initial

formation and growth of hydration products, with GO providing nucleation sites that promote

the development of hydration structures. While D–E shows a denser microstructure as

23
hydration products continue to form and fill micropores. By 90 days (Fig.6b F), it reveals a

more compact and interconnected structure, indicating substantial maturity in hydration. This

microstructural refinement, aided by GO, significantly contributes to the long-term strength

and durability of the cement composite.

These studies reveal a fascinating link between GO dosage and the shape of hydration crystals

in cement paste. A lower dosage of 0.02wt% yielded flower-like crystals with large petals,

enhancing the cement paste's toughness, but a higher dosage of 0.05wt% resulted in irregular

polyhedral crystals, boosting the compressive strength. The correlation between GO dosage

and crystal density is clearly depicted in Fig.7, further supporting the role of GO as a template,

influencing the hydration process and promoting the formation of flower-like crystals within

voids. This observation aligns with the work of Lv et al. [72], who reported similar crystal

shapes forming on GO sheets.

Beyond crystal morphology, the addition of GO positively impacts hydration kinetics. Studies

from Zhao et al. [74] observed a significant increase of 24.89% in the mean chain length (MCL)

of C-S-H gel with the addition of 0.05 wt% GO, indicating a higher degree of polymerization.

This aligns with the findings of Vallurupalli et al. [47]. This implies that the GO promotes

hydration. Furthermore, Gong et al. [44] also reported an increase in Non-Evaporable Water

(NEW) and CH content in samples containing 0.03 wt% GO, indicating enhanced hydration.

This observation is similar to the results of Long et al. and Li et al. [75]. This accelerated

hydration process is further supported by XRD patterns [73], [75] and SEM images [76], [77]

that visually confirm a more uniform and compact formation of hydration products, likely due

to a seeding effect of GO. These findings collectively concludes that the GO acts as a catalyst,

accelerating hydration and promoting the formation of stronger and more uniform hydration

products, ultimately leading to improved performance.While the potential of GO to enhance

concrete strength is evident, achieving optimal performance requires careful consideration.

24
Inconsistencies in research findings across various studies highlights the need for a deeper

understanding of GO's behaviour within the hydration process.

Fig.7 SEM images of hydration crystals at different dosages and periods, illustrating density
and morphology changes [78], [79]

25
The exact role of GO in cement hydration remains uncertain, with studies reporting conflicting

observations. In contrast, many findings suggest improved hydration kinetics, and some

research works, like Kaur et al. [80] and Horszczaruk et al. [81] observed minimal to no

changes in the structure of the hydrated products. For instance, Long et al. [28] study on

hydration reported that GO influences the formation of C-S-H gel, resulting in denser

microstructure not by changing the nanostructure but by increasing the volume fraction. The

volume fractions of high-density C–S–H rose from 16% to 26%, while loosed-packing C–S–

H decreased from 40% to 24%. These results suggest that GO influences hydration by

densification of C-S-H and not by altering the structure of C-S-H, which requires further

investigation.

Contrastingly, some studies propose that the high content of GO can impede cement hydration

by absorbing more water [82]. However, some take an alternative perspective of water

molecules absorbed by functional groups of GO to facilitate water migration, promoting the

hydration process [71]. Wang et al. [83] emphasized the increase in calcium hydroxide crystal

size with the inclusion of GO. By applying the Debye–Scherrer equation, with 0.5wt% of GO,

they observed the increase in the size of calcium hydroxide crystal size from 95 to 175nm,

potentially linking this to improved mechanical strength. However, the underlying mechanism

remains unclear.

Similarly, conflicting findings exist regarding the GO’s influence on the Mean Chain Length

(MCL) of C-S-H, a crucial parameter influencing mechanical properties. Long et al. [28]

reported a decrease in MCL with GO addition, Whereas Li et al. [74] observed an increase in

MCL. In contrast, Yang et al. [84] found no significant influence of GO on MCL. This debate

extends to the very nature of flower-shaped crystals observed in some studies. Cui et al. [85]

challenged this notion by analysing the crystals' chemical composition. Their findings suggest

these formations are likely due to the carbonation of existing hydration products during sample

26
preparation for SEM analysis, not a direct consequence of GO. And proposed that the flower

shapes might be artifacts arising from the methods used to prepare samples for scanning

electron microscopy (SEM) analysis.

In terms of compressive strength, several studies, including those by Gholampour et al. [86],

observed a decrease in strength with higher GO dosages exceeding 0.1 wt%. This reduction is

attributed to the aggregation of GO particles within the cement matrix, which may limit the

availability of nucleation sites for hydration crystals. Similar trends were reported in other

research [60], [66], [87], [88] suggesting a potential sweet spot around 0.1 wt% GO for optimal

strength benefits. However, studies by Antonio et al. [89] and Shareef et al. [90] contradict

these findings. They observed significant strength improvements up to 57.4% even with higher

GO dosages of 2 – 4wt%. These contrasting results highlight the complexity of GO's interaction

with concrete and the potential influence in determining its effects on compressive strength.

Likewise, the impact of GO on long-term compressive strength remains a subject of debate.

For instance, some studies, like Jing et al. [65] observed a significant initial strength increase

of 15.07% at 3 days with 0.6 wt% of GO, but this gain diminished over time, dropping to just

1.30% at 28 days. Similar trends were reported in other studies [91], [92], suggesting that GO

has a strong effect on early hydration, followed by a strength regression. However,

contradictory findings exist. Devi et al. [66] observed a consistent rise in strength over curing

time. Similarly, Fonseka et al. [59] reported sustained strength improvements with a 0.08% GO

addition showing a 19% increase at 28 days and 21% at 56 days, with similar trends observed

in other studies [77], [78],[93], [94], [95], [96], [97], highlighting the inconsistencies in GO's

effects on long-term performance.

In summary, While GO’s potential to enhance concrete performance is promising, its precise

influence on hydration and mechanical properties remains obscure. Researchers have observed

27
various effects, such as changes in crystal morphology, accelerated hydration, and inconsistent

impacts on key parameters like the Mean Chain Length (MCL) of C-S-H. These

inconsistencies can be attributed to a multitude of factors, including the type of cement, GO

properties, interfacial bonding, and experimental methodologies. A comprehensive

understanding of these factors and how they interact with GO is crucial for optimizing its use

in concrete. Further research is needed to reconcile conflicting observations and establish clear

guidelines for GO dosage and interaction with concrete components.

4.2 Flexural Strength and Toughness

While compressive strength is a crucial aspect of concrete performance, an ideal concrete

should also excel in its tensile and flexural strength. Traditional concrete often falls short in

these areas, exhibiting low flexural strength and a brittleness [98], [99]. Brittleness refers to

the tendency of the concrete to fail suddenly without any significant deformation when

subjected to stress. This involves an energy conversion process where the material releases the

maximum elastic energy before reaching the fracture critical point as surface energy along the

main crack. The fracture surface energy is influenced by factors like composition and

microstructure, indicating the material's brittleness.

While compressive strength has long been the primary focus in concrete design, modern

infrastructure demands have highlighted the importance of enhancing flexural strength.

Compared to flexural and tensile strength, the compressive strength of concrete is typically 8–

10 times higher. However, structures such as bridges, pavements, and tall buildings are

subjected to dynamic loads, vibrations, and bending forces that require improved flexural

performance beyond what traditional concrete offers. Therefore, the development of advanced

concrete materials with high tensile and flexural strength has become necessary. This is where

28
GO emerges as a promising solution to improve the flexural strength of the concrete [46]. For

instance, Chen et al. [100] investigated the tensile strength of GO-enhanced concrete at various

concentrations and found that at an optimal concentration of 0.03 wt%, the tensile strength

increased by 18.05% at 28 days, as shown in Fig.8. This improvement is attributed to the

enhanced formation of hydration crystals. Moreover, the tensile strength improved further with

longer curing periods.

Fig.8 Flexural strength improvements of GO-enhanced concrete with different concentrations


[100]

Similarly, Lv et al. [78], observed that with a 0.04 % addition of GO increased the flexural

strength by 67.1 % and attributed this improvement to the GO influence on modifying the

morphology of cement hydration products, likely promoting the formation of denser flower-

like crystals. Hou et al. [71] also demonstrated significant improvements in flexural strength

through both simulation and experimental work, reporting an 11.62% increase with 0.16 wt%

GO addition. Their stress-strain analysis revealed that GO enhances the interfacial strength

29
between the CSH gel and the cement matrix, contributing to better overall mechanical

behaviour.

Which is similar to the results by Wang et al. [101], their microstructural and crack propagation

analysis using scanning electron microscopy (SEM) and energy-dispersive spectroscopy (EDS)

demonstrated GO’s ability to positively influence ductility and energy absorption. Collectively,

these studies underscore the influence of GO in enhancing the flexural strength of concrete.

Also, GO plays a significant role in enhancing load transfer within the concrete matrix. It acts

as a nano-scale reinforcement that interlocks the cement matrix by connecting different

hydration products such as Calcium Silicate Hydrate (C-S-H) and portlandite. This interlocking

effect enhances the load transfer efficiency between the hydration phases and reduces the stress

concentration in the matrix. Improper distribution of load on the matrix can cause specific load

points failure that compromises the entire structural integrity of the structure. Unlike other

fillers, GO can efficiently redirect or alter the trajectory and orientation of cracks around them,

thus reducing the formation of fine cracks and improving the material's load-bearing capacity

[102]. Additionally, the load transfer efficiency is also influenced by the chemical bonding at

the GO-matrix interface, which is mediated by the functional groups on the GO surface and the

C-S-H or Calcium Hydroxide (COH) in the matrix. This chemical reaction results in strong

covalent bonds that increase the flexural and tensile strength of the matrix [71], [103].

Gholampur et al. [104] demonstrated that the incorporation of GO sheets in the Cement matrix

increased the elastic modulus and axial strain of the composite. This enhancement was

attributed to the increased bond strength and load transfer efficiency between the GO sheets

and the matrix. Additionally, the wrinkled morphology of the GO sheets also contributed to the

mechanical interlocking between the GO reinforcement and the matrix, which further enhanced

the load transfer and bond strength.

30
Despite these benefits, achieving optimal performance in flexural strength with GO also

presents a significant challenge, with studies reporting conflicting findings on the ideal dosage.

For instance, Indukuri et al. [105] and Liu et al. [106] observed increased flexural strength up

to 0.03wt% and 0.05wt% GO additions, respectively, but found a decrease with higher dosages.

In contrast, Chen et al. [107] reported a linear increase in flexural strength with increasing GO

content (0.02wt% - 0.08wt%), while Reddy et al. [96] observed a similar trend with a peak

increase at 0.1wt% GO. These findings further diverge from Antonio et al. [89] who reported

improvements up to 4wt% GO, with a 65.2% increase in flexural strength. These discrepancies

highlight the complexity of GO's interaction with concrete. Beyond dosage, other factors, such

as the type and properties of GO used and the overall concrete mix design, likely play a

significant role in its effectiveness. Further research is necessary to determine the optimal GO

content for achieving consistent and reliable improvements in flexural strength across different

concrete formulations.

4.3 Influence of GO on Porosity and Microcracking

Concrete is a complex composite material where the interplay between its components dictates

its performance. One crucial yet often unseen aspect is the intricate network of pores that

permeate the concrete matrix.

In Concrete’s complex matrix, one crucial yet often overseen aspect is the intricate network of

pores. This pore structure significantly influences its mechanical properties and overall

behaviour. These pores, ranging from large capillary pores to microscopic gel pores, play a

critical role in the concrete's durability and strength. The volume and distribution of these pores

directly impact their mechanical performance. Large capillary pores, with diameters exceeding

50 nanometers, act as pathways for water and other substances to travel through the concrete.

31
A high volume of large pores can create weaknesses within the concrete, making it more

susceptible to cracking and deterioration from processes like freeze-thaw damage caused by

water infiltration.

In terms of micropores, concrete inherently contains numerous smaller pores, typically less

than 50 nanometers in diameter, which reside within the hydration products. These tiny pores,

known as gel pores, exert a significant influence on key properties like shrinkage and creep

behaviour. An excessive volume of these gel pores can lead to the formation of microcracks,

affecting the long-term performance of the concrete. Graphene Oxide (GO) emerges as a

promising solution to address these microcracking concerns. By acting as a nano-filler, GO has

the potential to reduce the volume of these tiny voids, thereby hindering the formation of

microcracks and enhancing the structural stability of concrete structures. Sun et al. [108] in

their study reported that the addition of 0.08wt% of GO significantly reduced the total porosity

of the concrete by 67.16%, 53.26%, and 31.45% for the water-to-binder (w/b) ratio of 0.35,

0.45, and 0.55 respectively, as shown in Fig.9a. The reduction in overall porosity is attributed

to the filler effect [77] and nucleation effect of GO [61]. Liu et al. [109] also found that

incorporating GO decreased the average pore size and total porosity of cement by 16.9% and

25.5%, respectively. Which is in consistent with various studies [82], [110].

32
Fig.9a Effect of GO on the total Porosity reduction of Cement Mortar at Different Water-to-
Binder Ratios [108]; Fig.9b (i) Cumulative pore volume curves for control and GO-modified
samples; Fig.9b (ii) Corresponding differential pore volume curves for control and GO-
modified samples [111]; Fig.9c SEM image showing GO bridging a crack in cement, improving
interlocking within hydration products [71].

However, total porosity alone cannot fully characterize the pore structure of cement

composites, a more detailed understanding emerges from analysing the pore size distribution

within the matrix. Differential and cumulative pore volume curves offer valuable insights into

the size and interconnectedness of pores within the material.

Several studies have explored the effects of GO on the pore structure of cement-based materials

using Mercury Intrusion Porosimetry (MIP) and gas adsorption techniques, [14], [112]. These

33
investigations reveal that GO reduces the total porosity of the cement matrix by decreasing the

number of pores, resulting in a denser microstructure and improved mechanical strength. Two

mechanisms have been proposed for this behaviour. One is the pore-filling effect of the nano-

sized GO sheets, and the other is the seeding effect that accelerates the formation of hydration

[113], [114].

Zeng et al. [111] examined the influence of graphene oxide (GO) on concrete porosity using

Mercury Intrusion Porosimetry (MIP). Their findings revealed that the addition of GO

significantly refined the pore structure of concrete, as shown in Fig.9b, which compares the

cumulative and differential pore volume curves for a control specimen (M00) and GO-modified

specimens (M12, M14, and M16), with GO concentrations of 0.02, 0.04, and 0.06 wt%,

respectively. Specifically, the cumulative pore volume analysis Fig.9b(i) showed a 37.3%

reduction in critical pore diameter with the addition of 0.04 wt% GO, indicating a finer pore

size distribution. The differential intrusion curves Fig.9b(ii) further highlighted this effect,

displaying a shift in the peak value towards smaller pore diameters with higher GO content.

These suggests that GO effectively refines the pore structure, reducing the presence of

macropores and forming a denser matrix that is more resistant to water absorption and chloride

ion migration, thereby improving durability.

Similar effects were observed by Zeng et al. [115] , they observed a decrease in the macro pores

fraction of mortar samples from 95.66% to 93.15% after adding 0.04 wt.% GO. The cumulative

intrusion curves showed that the volume fraction of macropores decreases with an increase in

GO content, suggesting that GO-modified samples contain smaller pores compared to the

controlled samples. These studies suggests that GO can improve pore structure and create more

small pores, which can prevent water from penetrating the matrix through interconnected pores.

34
In terms of microcracking, cement-based materials are inherently heterogeneous and porous,

which leads to the formation and propagation of microcracks under loading, resulting in low

tensile strength and deformation. Even after high compaction, pores or voids ranging from

nanoscale to micro scale are especially prevalent in the Interfacial Transition Zone between the

aggregates and the cement matrix. These nanoscale cracks can develop into microcracks [116],

[117].

Graphene oxide can effectively mitigate this development by acting as a nanoscale

reinforcement and as a crack inhibitor. Wang et al. [118] stated that GO nanosheets can form a

three-dimensional (3D) network structure in the cement matrix by interconnecting GO

monomers. This 3D network structure is strongly bonded to the hydration products of the

cement matrix through adsorption processes and impedes crack propagation by blocking and

altering the crack path. This effect is more pronounced when GO is added in UHPC (Ultra-

High-Performance Concrete), where it acts as a bridge and enhances the fracture toughness of

the material, particularly in the ITZ. The crack-bridging effect of GO is captured through SEM

in a study by Hou et al. [71] as shown in Fig.9c, where GO is actively bridging a crack width

of 500nm, interlocking the hydration products. Studies have shown that the micro-mechanical

properties of cement-based materials, which govern the mesoscale behaviour, are influenced

by the chemical composition and microstructure of the graphene oxide [119], [120]. These

findings lead to the conclusion that GO, due to its nucleation and filling effect, can modify the

microstructure of the material, resulting in reduced porosity and preventing microcracking,

thereby significantly enhancing the mechanical properties of the concrete.

Though the positive effects of GO on porosity are evident, the exact mechanism and the effects

are unclear, with studies showing divergent results. Some studies [121] have reported that GO

does not significantly change the total porosity but reduces the critical pore diameter and the

35
volume of macropores. In contrast, other studies have reported increased porosity, with GO

even doubling pore volume in some cement matrices [44].

Likewise, Li et al. [122] observed minimal to no change in overall porosity, whereas

Mohammed et al. [123] reported an increase in the number of capillary pores in GO added

sample. In contrast, Gong et al. [44] found reductions in both total porosity and in the total

number of capillary pores, indicating inconsistencies. Additionally, Gong et al. [44], Long et

al. [44] and Kang et al. [124] reported an increase in gel pores with GO addition, possibly due

to enhanced C-S-H gel formation. However, Mohammed et al. [123] observed a reduction in

the gel pore area and volume, which may be attributed to nanoscale GO filling spaces within

the C-S-H interlayer. Further investigations are required to confirm this mechanism. These

inconsistencies highlight the influence of various factors such as water-to-cement ratio, curing

conditions, and characterization techniques [74].

Also, the current body of research predominantly relies on a qualitative method to assess crack

characteristics, such as length, width, and propagation patterns, often using electron

microscopy for visual comparisons [115]. However, this qualitative approach has inherent

limitations in definitively attributing crack inhibition to GO incorporation. To validate these

observations and to strengthen the evidence, more rigorous quantitative studies are essential.

4.4 Elastic Modulus and Deformation Behaviour

The intricate microstructure of concrete, composed of cementitious phases, aggregates,

and interfacial transition zones (ITZs), plays a crucial role in determining its mechanical

behaviour, particularly its toughness and impact resistance. Toughness refers to the concrete’s

ability to deform plastically and absorb energy under load, which mitigates crack propagation

and prevents brittle failure. Impact resistance, on the other hand, refers to the ability of concrete

to resist sudden dynamic loads, which is an important aspect for applications that involve heavy

36
loads, impact forces, and cyclic loading, such as pavements, bridges, and industrial flooring.

The toughness of concrete can be quantified by different measures, such as bending toughness,

impact toughness, and fracture toughness, which is calculated as the energy absorbed by

concrete during failure.

Both toughness and impact resistance are highly influenced by the ITZs, the regions between

aggregates and cement paste. The ITZ is typically weaker, with higher porosity and

microcracks, compared to the bulk cement paste [125]. By reducing ITZ thickness and volume,

the bond between aggregate particles and the cement paste strengthens, enhancing the

compressive, tensile, and flexural strength of the concrete[126], [127]. Several methods have

been developed to enhance the toughness and impact resistance of concrete, including

reinforced concrete, concrete-filled steel tubes, and prestressed concrete. These approaches

focus on modifying the internal structure of concrete by improving pore structure, enhancing

energy dissipation, optimizing ITZ performance, minimizing internal microcracks, and limiting

crack growth during failure. Transforming concrete from a brittle to a ductile material is

essentially a fundamental process of microstructural improvement [128], in which GO can play

a significant role.

37
Fig.10a SEM images of ITZ in GO added concrete at various scales (i, ii, iii), revealing
compact and refined structures [102]. Fig.10b(i) GO effect on reducing ITZ thickness;
Fig.10b(ii) GO effect on reducing ITZ volume fraction [108].

GO ability to influence the ITZ has been shown to improve overall concrete performance. For

instance, Liu et al. [102] demonstrated through the SEM images of the ITZ in GO added mortar

at different magnifications, as shown in Fig.10a (i-iii). The images reveal a compact ITZ with

minimal precipitated CH crystals and a network of fibrillar C–S–H and ettringite adhering to

the surface of a sand grain, indicating a denser structure. Additionally, Sun et al. [108]

investigated the impact of graphene oxide (GO) on the ITZ in cement mortar, finding that GO

38
addition reduced ITZ thickness by 11.64%, as shown in Fig.10b(i). The volume fraction of the

ITZ in the mortar also decreased by 11.77%, as illustrated in Fig.10b(ii), highlighting the

effectiveness of GO in refining the ITZ microstructure. Similarly, Yeke et al. [129] observed

that GO enhanced the mechanical performance of the ITZ between steel fibers and the

surrounding cement matrix. By incorporating 0.04% of GO, they observed significant changes

in the ITZ characteristics. The thickness decreased from 45 µm to 35 µm, while the average

elastic modulus increased from 19.63 GPa to 20.68 GPa. These alterations indicate a denser

and stiffer ITZ, potentially leading to improved crack bridging and energy dissipation within

the concrete matrix. Further supporting these findings, Nili et al. and Xu et al.[130], [131]

employed SEM, XRD, and EDS techniques to examine the microstructure of the ITZ. These

microscopic investigations further confirmed the positive effects of GO on mechanical strength

and structural integrity.

Beyond the ITZ, GO has shown potential for broader improvements in concrete's tensile

performance. Zhao et al. [132] reported a significant increase of 32.37% in Young's modulus

with the incorporation of just 0.02 wt% of GO. Similarly, Jamnam et al. [133] observed

enhanced toughness in a mortar containing 0.025 wt% GO through blast loading tests. Their

results demonstrated a lower permanent deflection and improved rebound after impact, which

can be attributed to a stronger bond and increased toughness. Also, these enhancements are

likely due to GO's ability to mitigate the intrinsic brittleness of cement paste caused by the

disorganised structure of its hydration products [134].

In addition, Chen et al. [135] investigated the impact of GO on the creep coefficient of concrete

and observed a reduction in them, particularly at higher dosages. They observed that with a

0.08 wt% GO addition, the creep coefficient decreased by 20.53%. As shown in Fig.11a, the

controlled sample (PSC) shows the highest creep coefficient across all time points. In contrast,

the GO-modified samples, GOSC-1 with 0.02 wt% of GO and GOSC-2 with 0.08 wt% of GO,

39
exhibited reduced creep coefficients, with GOSC-2 demonstrating the most significant

reduction. They proposed that GO acts as a brittleness regulation mechanism in cement matrix

by altering the morphology of hydration crystals within the cement matrix. The presence of

GO may induce stress around these rod-like crystals, causing them to elongate and form

columns oriented away from the GO surface. These columns then interlock within the pore

structure, enhancing the overall structural integrity and reduces creep. This finding suggests

that GO can act as a brittleness regulation mechanism within the cement matrix.

Fig.11a Change in Creep coefficient of PSC (controlled), GOSC-1 (0.02 wt% GO), GOSC-2
(0.08 wt% GO) concrete samples [135]; Fig.11b Improved elastic modulus of GO-enhanced
concrete at various dosages [107]; Fig.11c GO concrete demonstrating crack deflection(a,b,c)
vs straight fracture in controlled sample (a’, b’, c’) [102].

40
Similarly, Reddy et al. [92] studied the impact of GO on the elastic modulus of concrete and

observed a significant increase of 28.69% in their elastic modulus with the addition of 0.15wt%

of GO. They proposed that 0.15wt% is the optimal concentration beyond which reductions

were observed. Chen et al. [107] also observed an increase in the elastic modulus of concrete

with a 27.46% increase at 3 days and a 10.97% increase at 28 days with varying dosages of

0.02, 0.05, and 0.08 wt% as shown in Fig.11b. They noted that the elasticity of modulus

increases initially but then decreases with age. These findings highlight the potential of GO to

enhance the elasticity modulus of concrete.

However, further investigations are required to assess their long-term effects and optimal GO

concentration. For instance, Chen et al. [107] observed a decreasing trend in elastic modulus

with prolonged curing time, which is similar to the findings by Fonseka et al. [59], who reported

a 30% increase in elastic modulus after 28 days of curing and a subsequent decline after that.

The underlying reasons for this trend remain unclear, suggesting a need for further research

into its long-term effects.

Beyond its impact on elasticity, GO also offers significant advantages in mitigating fatigue

damage caused by repeated bending stresses. Cho et al. [136] developed a model to predict the

failure probability of concrete beams under different mixtures and bending fatigue conditions.

Their research suggests a fascinating mechanism by which GO improves concrete's fatigue

performance. They observed that the GO acts as a microscopic energy absorber during the

initial stages of crack formation, which reduces stress concentrations and hinders crack

propagation, leading to lower deformation. This improves the overall fatigue resistance. In

addition, GO toughens the concrete by deflecting, branching, and bridging cracks, enhancing

its resistance to failure [21], [137].

41
Liu et al. [102] in their study observed that GO deflects the crack propagation in concrete under

stress. Fig.11c illustrates the fracture patterns in GO modified (a,b,c) and plain cement (a’,b’,c’)

mortar beams. The plain mortar beams displayed a straight vertical crack and a relatively flat

fracture surface (Fig.11c a’,b’,c’), whereas the GO-incorporated mortar beams exhibited

inclined, tortuous cracks (Fig.11c a,b,c). Also, the GO beams exhibited rougher, more irregular

fracture surfaces than the brittle and flat surfaces in plain mortar beams. This crack pattern is

attributed to the presence of GO, which enhances mechanical resistance and deflects crack

propagation. Although GO has proven effective in mitigating the cracks, its long-term

performance under repetitive loading requires further investigation. To recognize the

limitations of the matrix, researchers have increasingly turned to advanced simulation

techniques such as Finite element modelling (FEM), Discrete element modelling (DEM) [138],

[139]. However, discrepancies between the simulation models and real-world performance still

exist. In addition to this, sophisticated experimental methodologies, such as advanced fatigue

testing equipment and microscopic and spectroscopic techniques [140], have emerged for a

better understanding of its elastic behaviour, as it serves as a critical component in the

infrastructure.

In summary, the potential of GO to enhance concrete performance is undeniable, but a clear

understanding of its mechanisms requires further investigation. On a positive note, GO strongly

demonstrated its capability to improve the mechanical strength of concrete, including its

compressive, flexural, and tensile properties. This improvement can be attributed to factors

such as enhanced hydration, improved interfacial bonding, a denser microstructure with

reduced porosity, reduced ITZ, microcrack filling, and crack deflection. These findings proves

the effectiveness of GO and its significant potential in improving the strength of concrete

structures. Table 1 summarizes the improvements in mechanical properties of concrete due to

the incorporation of graphene oxide. However, a gap remains between lab-scale experiments

42
and real-world applicability. Compared to the previous decade, the surge in both computational

and physical testing has significantly enhanced our knowledge, but concrete, with its inherent

complexities, demands further exploration. Challenges such as determining the optimal GO

dosage across diverse concrete formulations require further research. Additionally, a more

comprehensive understanding of how GO interacts with the hydration process and influences

the microstructure at various scales is crucial. Further research focused on optimizing GO's use

and standards, understanding its interaction mechanisms, and exploring its influence on long-

term durability is essential to unlocking its full potential and large-scale implementation

towards the future of concrete construction.

Table 1 Summary of GO enhanced mechanical strength of concrete.

Flexural Tensile
Additive Weigh Compressive w/c
Strength/ Strength/ Matrix Year Ref
material ratio Strength / days ratio
days days

GO 0.00% 11.67% / 90 - 26.39% / 0.42 Mortar 2023


[141]
days 90 days

GO 0.01% 12% / 28 days - - 0.475 Mortar 2023 [142]

GO 0.01% 19% / 28 days - 16.2%/ 0.475 Mortar 2023


[142]
28 days

GO 0.02% 33% /28 days 16.3% / - 0.43 Mortar 2023


[95]
28 days

GO 0.03% 39.4% / 28 17.7%/ 28 - 0.5 Mortar 2021


[26]
days days

43
GO 0.03% 41% /28days 44%/ 28 84%/ 28 0.6 Mortar 2022
[143]
days days

GO 0.03% 3% / 28 days 4% / 28 - 0.42 Mortar 2023


[144]
days

GO 0.03% 46.4% / 28 - - - Mortar 2021


[145]
days

GO + 0.03% 31%/28 days - - 0.485 Mortar 2021


[146]
SF

GO + 0.03% - 18% / 28 - 0.485 Mortar 2021


[129]
SF days

GO+ 0.03% 12.4% / 28 - - 0.4 Mortar 2021


[106]
SF+ FA days

GO 0.03% - 77.70% / - 0.36 Paste 2021


[105]
28 days

GO 0.03% 6.2% / 28days - - 0.5 Paste 2022 [147]

GO 0.04% 47.61% / 28 - - 0.36 Paste 2021


[105]
days

GO 0.04% 33.9% / - 44.3%/ 0.5 Paste 2021


[93]
28days 28 days

GO 0.05% 20.1% / 28 29.5%/ 28 26.2%/ 0.47 Concrete 2022


[148]
days days 28 days

GO 0.05% 38% / 28 days 44% / 28 - 0.38 Mortar 2022


[149]
days

GO 0.05% 20%/ 28 days 26%/ 28 - 0.45 Mortar 2022


[150]
days

44
G0 + FA 0.05% - 16.1% / - 0.4 Mortar 2021
[106]
+ SF 28days

GO 0.06% 14 %/ 28 days - - - Concrete 2023 [151]

GO+ 0.06 wt% 117 % / 28 60% / 28 42% / 28 0.45 Mortar 2022

Nano days days days [94]

silica

GO 0.06wt% 17.7 %/ 28 21.9% / - - paste 2023


[152]
days 28 days

GO 0.08% 12.7%/ 28 7.38% / - 0.35 Concrete 2020


[107]
days 28 days

GO 0.08% 21% / 28 days - 12%/ 28 - Concrete 2023


[59]
days

GO 0.08% 26% / 28 days - - - Concrete 2022 [153]

GO 0.08% 49%/ 90 days - 38% / 0.45 Concrete 2020


[66]
90days

GO 0.08% 25% / 28 days - - 0.42 Mortar 2023 [154]

GO 0.08% 22.25% / 28 38.77% / 10.25% / - Mortar 2021


[155]
days 28 days 28 days

GO 0.10% 38.4% / 28 12.07% / 14.87% / 0.45 Concrete 2022


[96]
days 28 days 28 days

GO 0.10% 38.46%/ 6.3% / 28 10.09%/ 0.3 Concrete 2022


[156]
28days days 28 days

GO+SF 0.10% 17% / 90days 28% / 40% / 0.34 Concrete 2021


[157]
90days 90days

45
GO 0.15% 24.3% / - - 0.3 Concrete 2023
[92]
28days

GO 0.15% 47.7% / 28 - - 0.45 Concrete 2023


[158]
days

GO 0.20% 17.6 /28 days 16.9%/ 28 - 0.4 Mortar 2022


[158]
days

GO 0.40% 4.7 / 28 days 14.23%/ - 0.5 Mortar 2020


[159]
28 days

*SF- silica fume; FA – Fly ash

5. Thermal and Electrical properties

The long-lasting performance of concrete depends not only on its durability but also on its

ability to withstand diverse environmental conditions and temperature fluctuations. Therefore,

understanding the material's thermal and electrical properties is crucial. Heat transfer plays an

important role in concrete structures because poor heat dissipation can result in thermal

cracking. During the hydration process, concrete undergoes an exothermic reaction, releasing

heat. If this heat isn't dissipated efficiently, a significant temperature difference can arise

between the concrete's inner core and its outer surface. This temperature gradient induces

tensile stress within the material, which can lead to cracking, compromising the structural

integrity. Similarly, the electrical properties are also important. With the growing adoption of

embedded sensors and the rise of smart infrastructure, concrete's ability to conduct electricity

has gained interest. Understanding and tailoring them specifically can open doors for new

innovative applications in areas like structural health monitoring and energy harvesting.

46
5.1 Thermo-Electrical Response and Heat Transfer

In concrete, the surface temperature is close to the ambient temperature, whereas its

core temperature is higher than the exothermic hydration process. This temperature difference

rises gradually from the core to the surface, which causes thermal cracking. To mitigate these

risks and to prevent thermal cracking, two main strategies are used, reducing heat generation

and enhancing heat dissipation. For instance, techniques such as incorporating cooling pipes

are adapted to lower the overall heat generated during hydration. Meanwhile, techniques like

using low-heat Portland cement or cement with additives and reinforcements focus on

improving the thermal conductivity of the cement matrix, thereby facilitating more efficient

heat transfer. This is where graphene oxide (GO) emerges as a potential solution. Its exceptional

thermal properties offer a promising avenue for enhancing concrete's thermal conductivity

[160]. The following sections will explore the mechanisms by which GO influences the thermal

and electrical properties of the concrete.

GO, when incorporated into concrete, can actively participate in heat transfer. By acting as a

conductive pathway within the concrete matrix, it helps in dissipating the heat generated during

hydration, reducing the temperature gradient and, consequently, the risk of internal cracking.

For instance, Shintani et al. [161] investigated the impact of graphene oxide (GO) on the

thermal properties of concrete. Their study demonstrated that incorporating GO into mortar led

to increased thermal conductivity, with notable improvements observed at a concentration of

0.02 wt%, as shown in Fig.12a. At this concentration, thermal conductivity increased by 8.6%

at 3 days, 4.0% at 7 days, and 4.9% at 28 days, after which the effect stabilized. This

improvement is attributed to the GO’s intrinsic conductive properties and its ability to reduce

porosity within the cementitious matrix. Additionally, GO facilitates the formation of an

internal heat transfer network that further enhances the material’s overall thermal conductivity.

47
Fig.12a Enhanced thermal conductivity of GO concrete at varying dosages and curing times
[161] ; Fig.12b Weight loss(%) of cement samples with varying GO concentrations under
elevated temperatures [160]

Xiang et al. [160] investigated the performance of GO-enhanced cement mortars under elevated

temperatures, with GO concentrations of 0.05 wt% (GO-1), 0.08 wt% (GO-2), 0.10 wt% (GO-

3), and 0.20wt% (GO-4) respectively. These specimens were subjected to varying temperatures

up to 800° C, respectively. Their findings revealed that GO-reinforced mortars demonstrated

significantly reduced weight loss compared to the control samples, as shown in Fig.12b. Even

at high temperatures of 600° C, the weight loss was approximately 3.15% at 0.08wt%

concentration, which is considerably lower when compared to that of the controlled sample.

They attributed this improvement to the GO’s ability to regulate hydration products and

enhance microstructure, thereby reducing water evaporation. These results align with the

observations of Mohammed et al. [162], who reported nearly a 50% less weight loss in GO-

enhanced samples than in control specimens at 800° C. This reduced weight loss and increased

thermal resilience highlights the potential of GO as an effective thermal regulator in

cementitious materials. Similarly, Han et al. [163] observed that the samples maintained their

strength even at elevated temperatures. They retained 92% of their initial strength at 600° C.

48
This thermal stability is attributed to the presence of GO, which delayed thermal

decomposition.

Studies have shown that GO can enhance the thermal diffusivity of concrete[164]. This

enhancement is achieved by lowering the thermal gradient within the concrete, resulting in a

more uniform temperature distribution and preserving the material's strength. For instance,

Sedaghat et al. reported that incorporating GO at 5wt% and 10wt% concentrations can improve

the thermal diffusivity of cement pastes by 25% and 75%, respectively. This increase in thermal

diffusivity was found to reduce the temperature gradient by 30˚C - 90˚C during hydration [165],

[166].

Additionally, Abd et al. [167] found that GO improved the specific heat capacity of concrete

by 30%, allowing it to absorb more heat without a significant temperature rise. This, along with

the reduced thermal decomposition rate observed in their study, suggests that GO can play an

important role in enhancing concrete's thermal stability.

Though GO is added to improve thermal conductivity, it can also be employed to improve the

insulating behaviour in a few cases, such as in a refractory mortar. Refractory mortar is a type

of mortar designed to withstand very high temperatures. Unlike regular mortar, refractory

mortar can resist heat and thermal cycling without cracking, which makes it ideal for bonding

firebricks, insulating bricks, or other refractory materials. In a study by Janjaroen et al.[168],

the addition of GO to refractory mortar significantly enhanced its insulating properties. The

GO-modified mortar exhibited a higher temperature difference between the hot and cold sides

compared to the control sample, indicating improved thermal insulation. These findings align

with those of findings by Rao et al. [169] that GO-enhanced concrete exhibited increased

stability at elevated temperatures. Specifically, GO addition improves water retention and

49
reduces the decomposition rate of hydrated products, contributing to the development of

stronger and more durable concrete structures.

While these enhancements are promising, a comprehensive understanding of the thermal

behaviour of GO-modified materials requires intensive thermal analysis techniques. These

methods assess changes in material properties in response to temperature and heat flux

variations. The four main techniques include thermogravimetric analysis (TGA), derivative

thermogravimetric (DTG), differential thermal analysis (DTA), and differential scanning

calorimetry (DSC). The decomposition of cement hydrate products can be quantified by

thermogravimetry analysis (TGA), which measures the mass loss of a sample as it's subjected

to increasing temperatures. It occurs in three stages, each corresponding to a different

temperature range. The first stage happens between 25° C and 350° C, where the C-S-H

decomposes and water evaporates. The second stage takes place from 350° C to 550° C, where

dihydroxylation of calcium hydroxide happens. The final stage occurs from 550° C to 800° C,

where CaCO3 decarbonisation happens [170].

Rehman et al. [171] investigated the properties of GO-based cement using TGA/DTA analysis,

and their findings reveal a similar three-stage decomposition process but temperatures lower

than the normal OPC. Water loss from carboaluminate hydrate and C-S-H is observed at 180°

C - 300° C, portlandite dihydroxylation at 430 – 480° C, and calcium carbonate decarbonation

at 600 – 780° C as shown in Fig.13a. These observations highlight the GO's influence on

thermal properties. Additionally, the weight loss at the C-S-H stage is found to be 1.01%, which

is 20.8% lesser, when compared to the plain concrete. This can be attributed to the C-S-H gel

bonding with GO. In a TGA study by Quershi et al. [73], they observed that the CH mass loss

was higher in rate for 7 days and 28 days. They stated that the hydrophilic nature of GO

absorbed an increased amount of water through its surface and interlayer, which may release

them later in the long run, resulting in the increased content of CH in the subsequent days

50
(1,7,28 days). Which is similar to the study by Wang et al. [118] indicating higher mass loss of

hydration products in a controlled sample.

Fig.13a Thermal analysis (TGA and DTA) of GO-enhanced cement mortar vs controlled mortar
[171]; Fig.13b DTG analysis of controlled concrete(C) and GO enhanced concrete(GC) [169];
Fig.13c DSC plots comparing GO-added concrete and controlled samples at varying GO
dosages [118];

While TGA measures the change in mass of a sample as it's subjected to a controlled

temperature program. Derivative Thermogravimetric Analysis (DTG) measures the rate of

mass change with respect to temperature. This essentially provides the derivative of the TGA

curve, highlighting the points where the mass loss occurs most rapidly. By analysing the DTG

curves, critical temperatures corresponding to the peak mass loss rates can be identified. This

offers additional insights into the specific processes occurring during the thermal analysis,

51
complementing the data obtained from TGA. In DTG analysis by Rao et al. [169] three

endothermic peaks related to ettringite, Ca(OH)2, carbonate phases, and the corresponding

weight loss are clearly observed in Fig.13b. The shifting of peaks to higher temperatures

indicates improved water retention due to GO's high surface area and interaction with water

molecules. Additionally, the delayed decomposition of hydrated products implies that GO not

only enhances the concrete’s thermal resistance but also helps preserve its structural integrity

under extreme heat conditions.

While TGA focuses on changes in mass, Differential Scanning Calorimetry (DSC) takes a

different approach. Instead of measuring temperature differences, DSC measures the heat flow

required to maintain a constant temperature between a sample and an inert reference material

as they are subjected to a controlled heating program. By analysing the heat flow, DSC can

reveal various thermal events occurring within the sample, such as crystallisation and

oxidation.

Wang et al. [118] employed DSC to investigate the thermal behaviour of both pristine and

graphene oxide (GO) nanosheets-modified cement samples. The DSC plots for the control

sample (GOC0) and a high-GO-content sample (GOC4) containing 0.04wt% of GO are shown

in Fig.13c. The endothermic peak observed at 457° C corresponds to the decomposition of

portlandite Ca(OH)2. By analysing the area under this endothermic peak, the decomposition

temperature of Ca(OH)2 decreased from 459.30° C in GOC0 to 446.82° C in GOC4, suggesting

that GO nanosheets lowered the decomposition temperature of Ca(OH)2 in cement. The inset

in Fig.13c highlights a trend where the characteristic decomposition peak of Ca(OH)2

diminishes with increasing GO content, indicating that GO nanosheets reduce the amount of

Ca(OH)2 in the modified cement. The ΔH of GOC1 (118.55 J/g) decreases significantly to

96.76 J/g in GOC5, confirming that higher GO content correlates with a reduced quantity of

Ca(OH)2 in the cement matrix. This reduction in portlandite content suggests that GO

52
nanosheets influence the hydration process and enhance the thermal stability of the modified

cement structure.

While these findings emphasize the role of GO in improving thermal stability, similar studies

reveal additional mechanisms through which GO impacts the thermal properties of concrete.

For instance, findings by Yan et al. [160] suggests that GO actively reduces thermal stress by

heat transfer, with factors like the level of oxidation in GO influencing its heat transfer

efficiency. Studies indicate that GO with a higher oxidation level of 0.35 exhibits a lower

thermal conductivity of 72 W/mK, while with a low oxidation level of 0.05 results in a

significantly higher thermal conductivity of 670 W/mK [165], [172]. These highlights the

importance of optimizing the oxidation level of GO to achieve the desired thermal properties

in concrete. Furthermore, a computational study by Yang et al. [173] suggests that the size and

defect concentration of GO sheets can influence their thermal properties. Their findings

indicate that thermal conductivity decreases with the increasing defect concentration within the

GO sheets.

Despite promising signs, the influence of GO on concrete's thermal properties remains an area

of active research with some conflicting observations. For instance, Wang et al. [118] reported

a potential alteration in hydration due to GO, indicated by a lower decomposition temperature

of calcium hydroxide (Ca(OH)2). However, Wang et al. [174], found that GO had limited

effects on the hydration process based on TG-DSC analysis. Similarly, Han et al. [163] suggests

that the initial increase in hydration caused by GO might not translate into long-term benefits.

Their analysis showed minimal mass loss after 56 days and less intense carbonation peaks,

suggesting that long-term impact might be limited. These differences are likely due to the

concrete composite preparation, GO concentration, and dispersion within the matrix. Further

research is needed to reconcile these findings.

53
5.2 Dimensional Stability and Shrinkage

Shrinkage is another important factor affecting the thermal performance of concrete.

Concrete undergoes two main types of shrinkage, drying shrinkage and thermal shrinkage.

Drying shrinkage occurs due to the loss of moisture during the curing process, while thermal

shrinkage happens due to a reduction in volume as the temperature decreases. Both types of

shrinkage can lead to cracking, compromising the structure. Particularly, thermal shrinkage

affects the thermal properties of concrete by creating internal stresses and microcracks, which

disrupts the heat transfer pathways within the material. This disruption reduces the thermal

conductivity and diffusivity of the concrete, leading to uneven temperature distributions.

Therefore, reducing this thermal shrinkage is important.

Studies have explored the potential of GO in influencing the concrete's shrinkage behaviour,

particularly its thermal shrinkage at high temperatures. Mohammed et al. [175] investigated

the influence of GO on the thermal shrinkage behaviour of normal-strength concrete (NSC) at

high temperatures. Their study revealed that NSC specimens without GO exhibited significant

thermal contraction starting at 78° C due to the loss of free water, which continued up to 118°

C. This contraction resulted in a length reduction of 0.35% that led to the development of cracks

at early exposure stages, particularly around 110° C. Whereas GO-containing specimens

demonstrated a reduced thermal contraction compared to NSC. This positive effect is attributed

to the GO’s ability to restrict the movement of free water molecules within the concrete matrix.

The low capillary pores by GO hinders the escape of free water at elevated temperatures,

thereby reducing thermal shrinkage and potentially mitigating the risk of thermal cracking.

54
Fig.14a Ordinary vs GO concrete Specimens after exposure to elevated temperatures [175];
Fig.14b Spalling contour lines of ordinary and GO-added concrete specimens [175]

In addition, the GO-added specimens (GHS) shows better stability at high temperatures (800°

C), whereas the normal specimens (HS) crack into pieces, as shown in Fig.14a. This indicates

the thermal stability of GO-enhanced concrete at elevated temperatures. Also, the spalling

contour lines for both HS and GHS are shown in Fig.14b. Specimens containing GO (GHS)

exhibited much less spalling compared to the ordinary specimen (HS). While HS had irregular

spalling with large fragments, measuring up to 4×2×1cm, GHS exhibited only minor spalling,

mostly at the edges, with smaller fragments. The refined pore structure in GHS allowed for

vapor pressure to be released more effectively, thereby reducing severe spalling and improving

the overall durability of the concrete.

55
Similarly, Xiang et al. [160], investigated the performance of GO-enhanced cement mortars

under elevated temperatures up to 800° C, with varying GO concentrations Their findings

revealed that, even under high-temperature conditions, GO-reinforced mortars exhibited

significantly higher tensile and compressive strengths compared to that of control samples. The

controlled sample started cracking at 600° C, whereas the GO-reinforced mortars remained

structurally stable up to 800° C. This improvement is attributed to the GO’s bonding effect and

its ability to reinforce the C-S-H phase, which prevents crack growth under thermal stress.

Also, the CaO and CaCO3 produced from the decomposition of Ca(OH)2 at high temperatures

can fill cracks, which further enhances the material's strength. Likewise, Mohammed et al.

[162] studied the effect of high temperatures on the bond strength of GO-reinforced concrete

and observed significant retention of bond strength even at temperatures as high as 800° C,

indicating the effectiveness of GO in enhancing thermal stability.

In addition to its Stability, GO positively influences the shrinkage behaviour of the concrete.

Xu et al. [176] reported that incorporating 0.3wt% of GO reduced the chemical shrinkage of

cement paste, and this reduction is attributed to GO's ability to refine pore sizes within the

material. By regulating the formation of calcium hydroxide (Ca(OH)2), GO limits the water-

holding capacity of the pores, resulting in lower overall chemical shrinkage at the macroscopic

level. Similarly, studies by Guo et al. [177] observed a significant reduction in concrete's

shrinkage with the addition of graphene oxide (GO), achieving up to a 54.75% reduction in

shrinkage at 14 days with 0.06 wt% GO (rc06), as shown in Fig.15a. Furthermore, GO

demonstrates potential in mitigating early-age thermal contraction by hindering water loss at

high temperatures and preventing the escape of free water molecules This can potentially

reduce thermal shrinkage and the risk associated with it [175].

56
Fig.15a Shrinkage rate of concrete with varying GO dosages [177]; Fig.15b Electrical
resistivity of concrete specimens at varying GO dosages [178] ;

Although GO shows potential in improving the thermal behaviour of concrete, some

inconsistencies in their results remain. For instance, Abd et al. [167] observed a decrease in

thermal conductivity with increasing GO content. Using TG and FTIR analysis, they found that

with the addition of 1.2wt% of GO resulted in a reduced decomposition rate, which was

attributed to the insulating properties of GO sheets. These findings are similar to the

observations by Janjaroen et al. [168] and Maglad et al. [179]. However, GO's dual role in

enhancing electrical resistance and insulation behaviour remains unresolved, necessitating the

need for further investigation. Additionally, differences exist regarding the impact of GO on

strain rate over a period of time. While some studies report an initial increase in strain rate

[107], others suggest it eventually falls below that of ordinary concrete. Lu et al. [180] observed

a slightly higher initial shrinkage rate in GO-incorporated cement paste, attributing it to the

accelerated hydration process triggered by GO. Interestingly, they suggested that the water

absorbed by GO might be released back into the matrix later, contributing to a self-curing effect

57
that reduces overall shrinkage. These findings emphasize the need for optimizing GO

concentration and dispersion methods to achieve consistent performance.

5.3 Electrical properties

Concrete is traditionally considered an electrical insulator, and its need for enhanced electrical

properties has often been overlooked in the past. However, in recent years, developments in

nanomaterials have opened up the possibilities of next-generation structures with enhanced

monitoring capabilities that require conductivity. This has enabled the way for innovative

applications such as self-sensing, electromagnetic shielding, and energy storage-based

structures. However, among the conductive nanomaterials, graphene-based materials have

gained significant attention. Graphene oxide offers promising properties for concrete.

However, its electrical insulating nature due to oxygen functional groups limits its potential in

electrical applications [43]. This has led researchers to explore other graphene family materials

like reduced graphene oxide, graphene nanoplatelets, and few-layer graphene for their superior

electrical conductivity in cement composites [181], [182], [183].

The ability of graphene to transform non-conductive materials, like concrete, into conductive

ones opens up exciting possibilities. Studies are ongoing to explore the use of conductive

concrete for real-time monitoring and sensing of structures [184], [185]. Additionally,

researchers are investigating the potential application of conductive concrete in cold

environments, such as temperature-controlled roads and pavements that utilize electrical

resistance heating to melt snow and ice, improving winter safety [186].

However, achieving these electrical functionalities in concrete depends on exceeding the

percolation threshold of the added conductive material. This means there needs to be a

sufficient amount of conductive elements, like graphene, dispersed throughout the concrete to

58
create a network for electricity to flow. Interestingly, research suggests that even electrically

insulating GO can contribute to conductivity in concrete under certain conditions [187].

Particularly, when the GO dosage surpasses the percolation threshold and interacts with the

cementitious environment's high alkalinity and thermal effects of hydration, some of its oxygen

functionalities may be reduced, potentially allowing for a conductive path for electron flow,

thereby increasing the overall conductivity of the matrix [187].

Li et al. [178] studied the electrical properties of GO in cement paste and found that the

electrical resistivity increased with GO contents of 0.1wt% and 0.2wt%. However, when the

GO content reached 0.4 wt%, the resistivity of the sample decreased below the controlled

sample, as shown in Fig.15b. Similar findings were reported by Bagheri et al. [58] where

electrical resistivity initially increased with the addition of GO but began to decrease beyond

the threshold of 0.05wt%. This phenomenon can be attributed to the mechanisms governing

electrical resistivity in cementitious composites. The electrical resistivity in cementitious

composites is primarily influenced by two factors, the ion diffusion rate from cement particles

to the aqueous phase and the transport of ions within the cement matrix. Incorporating GO

sheets into cement mortar influences these mechanisms in a concentration-dependent manner.

At lower concentrations, GO accelerates the precipitation of hydration products, as hydration

progresses, it forms thick barriers of hydrates. These barriers restrict ion diffusion by reducing

the exposure of the solution to unhydrated cement particles, leading to an increase in electrical

resistivity. However, at higher GO concentrations beyond 0.04 wt%, the entrapped water

molecules associated with the GO sheets facilitate ion transport, counteracting the resistivity

increase. This is in agreement with the findings by Li et al. [178], where the addition of 0.06

wt% and 0.08 wt% GO led to a reduction in electrical resistivity. These results highlight the

dual role of GO in modulating electrical resistivity through its effects on hydration and ion

transport.

59
Similar results were observed by Quereshi et al. [73], where resistance increased during early

stages due to GO's water adsorption, which limited free water availability for conduction. At

later stages, as the GO network matures, it facilitates electron transport, resulting in decreased

resistivity.

In investigating the electrical properties of conductive concrete, a careful selection of

appropriate measurement techniques is required. The four-probe method is a widely recognized

standard for such evaluations. Additionally, electrical testing devices are also employed to

assess the properties of conductive concrete. These devices typically measure a combination

of factors like inductance, capacitance, and resistance[188]. A key concept in understanding

electrical conductivity within composites is the percolation theory [189].

This theory predicts a critical percolation threshold, beyond which the addition of conductive

materials creates a connected network throughout the material. This network allows for

efficient current flow, significantly increasing the material's overall conductivity [190].

However, the impact of GO on concrete's electrical resistivity is more nuanced than a simple

increase or decrease. Achieving and maintaining electrical conductivity in concrete composites

depends on various factors, including dosage, water content, curing time, etc [191]. Studies by

Balandin et al. [192] and Goracci et al. [193] have demonstrated the influence of water content

and curing time on conductivity. The presence of water content enhances the conductivity by

facilitating ion movement while curing age influences the development of C-S-H (calcium-

silicate-hydrate) gel formed during hydration, which influences water diffusion and contributes

to increased conductivity. However, maintaining consistent conductivity poses challenges.

Studies suggests that conductivity can be affected by the formation of cracks and pressure

variations within the concrete that can disrupt the conductivity network, impeding the flow of

electricity [190] [194].

60
These findings highlight the importance of optimizing GO content and dispersion methods to

achieve the desired electrical response for specific applications. While GO can initially increase

resistivity, its potential to enhance conductivity at later stages holds significant promise. This

paves the way for exciting possibilities such as self-sensing concrete, where changes in

electrical resistance could provide real-time information into concrete's structural health,

paving the way for next-generation intelligent concrete materials.

6. Durability Performance

Durability is a crucial aspect of any building material, determining its ability to

withstand various environmental challenges. The long-term durability and sustainability of

concrete structures remains a persistent challenge for the construction industry. While concrete

offers remarkable strength and versatility, its susceptibility to degradation from environmental

factors such as water ingress, chemical attack, and freeze-thaw cycles significantly impacts its

lifespan and structural integrity. Hence, a multifaceted solution is required to overcome these

issues. Recently, graphene oxide has emerged as a promising candidate for enhancing the

durability of concrete. This section explores the performance and long-term durability of GO-

modified concrete under these conditions.

6.1 Water sorptivity and Permeability

Water sorptivity is a critical challenge in concrete structures, referring to the material's

tendency to absorb and transport water into its internal matrix through capillary action. This

absorption can significantly weaken the concrete over time, as water ingress promotes chemical

degradation, freeze-thaw damage, and corrosion of embedded steel reinforcements, leading to

cracks and compromised structural integrity. Traditional methods, such as the use of

61
hydrophobic coatings, chemical admixtures, and supplementary cementitious materials such as

fly ash or silica fume, are commonly employed to overcome this issue. However, these

approaches often come with their own limitations. In this context, incorporating graphene oxide

has emerged as a promising solution for mitigating water sorptivity and creating more durable,

long-lasting construction materials.

Water absorption in concrete is strongly influenced by the size and distribution of pores within

the material. Capillary pores, in particular, play a critical role in the rate of water uptake. The

water absorption process typically involves two distinct phases, an initial phase and a

secondary phase. The initial phase represents how quickly water fills larger pores, while the

secondary phase represents the rate of water filling smaller air voids. Studies suggest that GO

can enhance the water resistance of concrete by modifying its pore structure. By potentially

reducing the volume and connectivity of capillary pores, GO may significantly impede water

ingress, thereby improving the durability of the material.

Bagheri et al. [58] studied the effect of GO on concrete and observed that the addition of

0.03wt% GO reduced the initial (Within the first six hours) and final (after 7 days) water

sorptivity by 88% and 66%, respectively, compared to the controlled sample. Fig.16a illustrates

the difference in initial and secondary water absorption rates of cement composites as a

function of GO content. Both rates exhibit similar trends, fluctuating based on the percentage

of GO. The study found that GO generally decreased water absorption in cement composites,

with the reduction reaching optimal levels at specific GO concentrations of 0.03% for the initial

absorption stage and 0.01% for the secondary stage.

62
Fig.16a Initial and secondary water absorption rates vs. GO concentrations [58]; Fig.16b
Water penetration comparison between GO-modified and controlled concrete specimens [111];
Fig.16c SEM images of concrete surfaces with no(i), low(ii) and high(iii) concentrations of GO
[195] ;

Similarly, Li et al. [196] observed that the addition of 0.4 wt% GO reduced the initial water

sorptivity by 4.6% and had a more significant impact on the secondary sorptivity, achieving a

44% reduction. This improvement is attributed to the GO’s ability to refine the microstructure

of the concrete. The enhanced performance of GO in secondary absorption highlights its

effectiveness in reducing long-term water absorption, furthermore, in a study by Zeng et al.

[111], GO specimens were tested for permeability under water pressure of 1.5MPa for 1 hour.

The results revealed that specimens containing 0.04wt% of GO exhibited a 55.5% reduction in

penetration depth as compared to the controlled sample, as shown in Fig.16b. Additionally, the

penetration depth decreased progressively with increasing GO content, indicating its

effectiveness in mitigating water ingress.

63
Similarly, Qureshi et al. [73] reported a significant 24.8% reduction in water absorption

coefficient with the addition of just 0.06 wt% GO in cement paste. This reduction is attributed

to the GO’s ability to act as a physical barrier, impeding water movement through the pore

network. Mohammed et al. [197] provided further insight, suggesting that GO increases the

tortuosity of pores within the concrete matrix. GO, along with other matrix components, forms

a more complex pathway for water movement, thereby hindering its absorption. These findings

highlight the potential of GO to enhance the water resistance of cement composites by

modifying the pore structure and impeding water transport, ultimately improving durability.

In addition to incorporating GO into the cement matrix, surface modification techniques with

GO have also proven effective in improving water permeability. Korayem et al. [195]

demonstrated that directly coating the concrete surface with GO reduced water absorption and

capillary absorption by approximately 40% and 57%, respectively. Fig.16c presents the SEM

images comparing samples with no GO coating (a), coating with low GO concentration (b),

and high concentration (c). These images reveal that the surface of the uncoated specimen

shows exposed pores and microcracks, leaving the hydrated cement vulnerable to water

penetration. Whereas a low GO concentration forms a thin, uniform layer of graphene oxide

on the concrete surface, providing partial coverage while maintaining surface visibility. At

higher GO concentrations, a thick protective layer entirely covers the concrete surface,

effectively creating a stronger barrier against water and ions ingress.

Similarly, He et al. [198] demonstrated that applying a GO aqueous solution as a surface sealant

can effectively reduce water absorption and air permeability in cement mortars, emphasizing

the importance of timing surface treatments for optimal effectiveness. Yu et al. [199] further

developed a novel GO-modified epoxy resin coating that significantly reduced water

penetration depth in C-S-H gel by 71.58%. These findings present additional possibilities for

enhancing water resistance through surface treatments.

64
The incorporation of GO, whether within the concrete matrix or as a surface treatment, presents

a comprehensive strategy to mitigate water sorptivity, which is a major factor impacting

concrete's durability. By potentially increasing pore tortuosity, acting as a physical barrier, and

enabling the development of effective surface sealants and coatings, GO offers a multifaceted

approach to enhancing concrete's water resistance. However, optimizing the distribution and

application of GO within the concrete matrix remains crucial for maximizing these benefits

and ensuring consistent performance. For instance, Bagheri et al. [58] observed a greater

impact on initial sorptivity, while Li et al. [196] observed a stronger impact on secondary

sorptivity. These differences highlight the influence of GO concentration and preparation

methods, underlining the need for further research and optimization to ensure consistent and

enhanced performance.

6.2 Chloride Ion Penetration and Corrosion Resistance

In addition to water ingress, chloride ion intrusion is one of the most significant threats

to the long-term performance of concrete structures, particularly in coastal and marine

environments. These ions penetrate into the matrix, promoting the corrosion of steel

reinforcement embedded within concrete. This rust formation expands the steel volume,

leading to cracking, spalling, and, ultimately, structural failure. This issue is particularly critical

in bridges, tunnels, and waterfront structures, where prolonged exposure to chlorides is

inevitable. To address this issue, graphene oxide has emerged as a promising solution. This

section explores the potential of graphene oxide to enhance the chloride ion resistance of

concrete.

Mohammed et al. [123] investigated the effect of GO on chloride ion intrusion in concrete and

found that incorporating 0.01wt% GO significantly reduced the chloride ion penetration depth

65
by 80%. As shown in Fig.17a, the GO-enhanced specimen exhibited a penetration depth of

only 5 mm (Fig.17a(i)) compared to 26 mm (Fig.17a(ii)) in the control specimen. This

improvement is attributed to the GO's layered and interconnected structures, which form a

sponge-like network within the cement matrix. This network acts as a physical barrier,

effectively capturing chloride ions and hindering their deeper penetration into the concrete.

Fig.17a Chloride Ion intrusion in GO added matrix (a) vs control mix (b) [200] ; Fig.17b
Chloride ion penetration depth in concrete with varying GO dosages [102]; Fig.17c
Mechanism of chloride ion binding with Friedel's salts in GO-added cement [200];

Similarly, Zeng et al. [115] observed that even a small amount of 0.03 wt% GO significantly

refined the pore structure of concrete, effectively reducing chloride ion intrusion. The addition

of GO dramatically reduced the volume of macropores and decreased the diameter of smaller

micropores, resulting in a denser matrix. This densification, driven by interactions between

GO's functional groups and hydrating crystals, creates a more formidable barrier for chloride

66
ion movement. Their study reported a significant 33.2% decrease in the chloride migration

coefficient within the mortar, aligning with the findings by Liu et al. [102]. Liu et al. observed

a 21.45% reduction in chloride ion intrusion with GO incorporation, as shown in Fig.17b.

Additionally, they found that penetration depth decreased with increasing GO concentrations,

further demonstrating its effectiveness.

In addition to the physical effects, chemical interactions between GO and the concrete matrix

play a significant role in mitigating chloride penetration. Long et al. [200] studied the chemical

interactions between GO and the concrete matrix and proposed that GO's surface functional

groups interact with divalent cations, such as calcium ions, promoting the formation of Friedel's

salts (F-salts). These F-salts possess a superior ability to bind chloride ions compared to the

standard cementitious materials, as shown in Fig.17c. GO's strong affinity for calcium ions

stabilizes the F-salts, preventing their breakdown and ensuring their long-term effectiveness in

capturing chloride ions. This chemical interaction further enhances concrete's resistance to

chloride-induced corrosion.

In addition, the advancements in modelling and simulations complement experimental

findings, offering deeper insights into the mechanisms by which GO enhances concrete's

resistance to chloride ion penetration. For instance, Tong et al. [201] employed simulations to

demonstrate how incorporating graphene materials remodels the microstructure of

cementitious composites, effectively restricting water migration through the pore network.

Since chloride ions are often transported by water movement, a reduction in water flow

translates to a decrease in the number of Chloride ions carried into the concrete. Similarly, Lai

et al. [202] emphasized the role of GO size and concentration in determining critical transport

properties like tortuosity and ion diffusion coefficients. Their proposed model provides a

framework to quantify the influence of GO on chloride transport, enabling precise optimization

67
of GO-modified concrete for specific applications. These models can assist in optimizing GO

concentrations to achieve maximal durability with minimal material usage.

Maglad et al. [179] demonstrated that incorporating GO improved chloride resistance in

geopolymer concrete by 35.3%. Similarly, Guo et al. [177] reported an 8.4% reduction in the

chloride permeability coefficient of recycled concrete with the addition of an optimal 0.06 wt%

GO. However, their findings also indicated that beyond a certain GO concentration, the

improvements plateau decreases, suggesting the need for further research to determine the

optimal dosage for maximum effectiveness.

These studies collectively highlight the effectiveness of GO as a multifaceted defence

mechanism against chloride ingress. By refining pore structures, acting as a physical barrier,

and enabling beneficial chemical interactions, GO enhances concrete's durability. Additionally,

by promoting the formation of Friedel's salts and densifying the pore structure, GO creates a

robust, multi-layered defence against chloride-induced damage. This combined effect

addresses both physical and chemical aspects of chloride intrusion, highlighting its potential as

a transformative additive for concrete.

6.3 Sulfate Ion Resistance

Beyond the challenges posed by chloride ions, concrete structures face another significant

threat from sulfate ions. Structures in sulfate-rich environments, water-prone areas, or soils

with high sulfate content are particularly vulnerable to this attack. Unlike chloride-induced

corrosion, which primarily targets steel reinforcement, sulfate attack directly deteriorates the

cementitious matrix itself, leading to long-term structural degradation. Incorporating graphene

oxide presents a promising approach to mitigate this issue. This section explores the potential

of GO to revolutionize sulfate resistance in modern construction.

68
Sulfate ions, commonly present in soils and groundwater, can initiate a complex deterioration

process in concrete through both chemical and physical mechanisms. Chemically, sulfate ions

react with hydration products like calcium hydroxide and tricalcium aluminate, forming

expansive compounds such as ettringite and gypsum. These minerals precipitate within the

concrete matrix, exerting internal pressure that leads to cracking and a gradual loss of structural

integrity [203], [204]. Physically, the crystallization of sulfate salts within the concrete's pore

structure further increases internal pressure, exacerbating the damage over time. Together,

these mechanisms contribute to the progressive and potentially catastrophic deterioration of the

concrete matrix [205]. Recent studies suggest that GO can counteract these damaging effects

by reinforcing the concrete matrix at multiple levels [45], [206]. GO's ability to refine the pore

structure, enhance matrix density, and chemical interaction with cement hydration products

could inhibit sulfate ion ingress and mitigate the formation of expansive compounds.

Lai et al. [202] demonstrated the effectiveness of GO in enhancing sulfate resistance in cement

mortar. Their study revealed that the incorporation of just 0.03% GO could extend the structural

life of the concrete by a factor of 2.3. They attributed this improvement to the GO’s ability to

refine the pore structure and resist sulfate-induced corrosion. The findings were verified

through experiments involving mortar specimens immersed in a sulfate solution for two years.

Fig.18 compares the controlled specimen M00 (a) and the GO-enhanced specimen M13 (b)

immersed in sulfate solution for 6, 12, and 24 months. While both the samples appear intact at

the beginning, over time, noticeable deterioration occurs. However, the GO-modified

specimens exhibited significantly less damage compared to the controlled specimen. The GO-

reinforced specimens consistently demonstrated smaller damaged surface areas, highlighting

their ability to mitigate sulfate attack, with a 40.9% reduction in damage.

69
Fig.18 Surface deterioration comparison in controlled cement mortar (a, M00) versus GO-
reinforced mortar (b, M13) after immersion in a sulfate solution for 6, 12, and 24 months [207].

Similarly, Wang et al. [207] investigated GO's role in mitigating sulfate attack through a dry-

wet cyclic test, simulating real-world environmental conditions. Their findings showed that

specimens with 0.05 wt% GO exhibited reduced mass loss of 3.073% compared to 4.839% in

the control specimens after 120 cycles. As shown in Fig.19a, the mass of all specimens initially

increased due to the penetration and accumulation of sulfate ions. However, after 60 cycles,

the mass of the specimens began to decline due to rapid deterioration from sulfate attack.

Despite this, the GO-modified specimens demonstrated superior resistance to mass loss

throughout the testing period. This improvement is attributed to the unique properties of GO,

including its ability to densify the microstructure, reduce porosity, and enhance bonding within

the cement matrix.

70
Fig.19a Mass changes in cement mortar with different GO concentrations over 120 sulfate
cycles [207]. Fig.19b Unaffected depth of sulphate attack in ordinary specimen (i) and GO
added specimen (ii) [208].

Chintalapudi et al. [208] provided further evidence of improved resistance to sulfate attack in

GO-modified Ordinary Portland Cement (OPC) specimens. As shown in Fig.19b, the

specimens containing 0.04 wt% GO exhibited an average unaffected depth of 34.08 mm

(Fig.19b(ii)), whereas the controlled specimens without GO had an unaffected depth of only

29.56 mm (Fig.19b(i)). This represents a significant improvement of 15.29% in resistance to

sulfate ion penetration. This improvement is attributed to the GO’s ability to enhance the

hydration process, leading to the formation of denser microstructure. This densification reduces

the porosity and inhibits the ingress of sulphate ions.

Similarly, Xu et al. [209] emphasized that an optimal GO concentration of 0.03 wt% resulted

in a denser pore structure within the concrete. These denser pores can act as a barrier against

sulphate ions ingress and potentially enhances the concrete's resistance to sulfate attack.

Further supporting this, Tong et al. [201] investigated the corrosion resistance of concrete

incorporating GO. Their deterioration tests indicated that GO reduced the decrease in strength

71
due to sulphate attack compared to the controlled sample. GO slows down chemical attacks,

potentially including those caused by sulfate ions.

Furthermore, Lai et al. [202] proposed a chemo-mechanical model to predict the transport

properties of GO-modified cement paste. This model accounts for factors like the size,

concentration, and arrangement of GO nanosheets within the concrete matrix, predicting their

influence on ion pathways taken by ions and their diffusion coefficient. The model suggests

that the reduction in chloride diffusion and migration coefficients can be attributed to the

increased tortuosity caused by GO.

However, a crucial limitation of this model is its dependency on the GO dispersion within the

cement matrix. Uneven dispersion can significantly impact the model's accuracy, limiting its

practical application in real-world engineering scenarios. While models for plain concrete

under sulfate attack exist, they may not adequately account for the complex interactions

between GO and the concrete matrix. Overall, research on GO's impact on sulfate resistance

presents promising results but requires further optimisation.

6.4 Freeze-thaw resistance

Concrete structures in cold regions face one of their most severe durability challenges

from repeated freeze-thaw cycles. Particularly, infrastructures like roads, bridges, and dams in

cold regions are often subjected to harsh environmental conditions, including extreme cold,

where repeated cycles of freezing and thawing are common. This poses a significant threat to

their durability [210], [211].

To address this issue, traditional methods such as hydrophobic coatings, air-entraining agents,

which create tiny air voids within the concrete to relieve the pressure from freezing water, and

supplementary cementitious materials are generally used [212], [213]. However, these

72
approaches may not fully address the long-term durability concerns. Graphene oxide has

emerged as a promising solution for enhancing freeze-thaw resistance [214].

The primary mechanism behind freeze-thaw damage is the expansion of water trapped within

the concrete's capillary pores. As the temperature drops, the water freezes and expands,

generating internal pressure that forces unfrozen water to migrate through the pores. This

resulting high internal pressure can lead to internal erosion and surface deterioration over time.

To ensure the durability in such environments, its performance at very low temperatures should

be thoroughly evaluated and improved [215].

Qian et al. [216] investigated the impact of GO on the freeze-thaw durability of concrete

specimens by analysing its uniaxial compressive strength (UCS) under various freezing cycles.

Their findings reveal that the compressive strength of the specimens initially increased with

freeze-thaw cycles, likely due to densification effects. However, after 60 cycles, a decline in

strength was observed due to the cumulative damage from freeze-thaw stress. Despite this

decline, the rate of deterioration varied significantly between GO-modified and control

specimens. The UCS results for specimens subjected to freeze-thaw cycles are illustrated in

Fig.20a. GO-modified specimens exhibited significantly reduced mass loss rates of 24.2% after

120 cycles, compared to 31.4% in the control specimens. Furthermore, the GO specimens

exhibited their maximum performance at an optimal concentration of 0.05 wt%, demonstrating

superior resistance to freeze-thaw cycles while minimizing mass loss and structural

degradation.

73
Fig.20a Weight loss percentage of concrete during freeze-thaw cycles with varying GO
concentrations [216]. Fig.20b Freeze-thaw damage in Controlled specimen (CM) vs. GO-
added specimen (G1, G3) [197].

Similarly, studies by Mohammed et al. [197] demonstrated a substantial improvement in

freeze-thaw resistance with the addition of graphene oxide. Their experiments showed that

concrete modified with 0.06 wt% GO experienced only a 0.25% weight loss after 540 freeze-

thaw cycles, compared to 0.8% in the control samples, indicating significantly less damage

from repeated freezing and thawing. Furthermore, visible edge damages were evident in the

control specimens, whereas GO-modified concrete remained intact, highlighting its protective

effects, as shown in Fig.20b. They attributed this improvement to the GO's ability to mimic the

behaviour of an internal air-entraining agent by forming a network of voids within the concrete.

At an optimal concentration of 0.06 wt%, GO promotes the formation of an air void network

within the concrete, which reduces internal stresses caused by freezing. This network, along

with the tortuosity of pores caused by GO, contributes to enhanced freeze-thaw resistance of

the specimens.

Zeng et al. [217] reported an 18.9% improvement in freeze-thaw resistance with the addition

of 0.03 wt% GO, even without the use of traditional air entrainment. While GO does not

directly create air bubbles during mixing, its ability to densify the pore structure and mitigate

74
internal stresses contributes significantly to frost resistance. Their findings also suggest that

increasing the GO content further improves freeze-thaw resilience, indicating the need for

further research in optimising GO concentrations. Similarly, Xu et al. [209] found that an

optimal dosage of 0.03 wt%, GO, improved both the freeze-thaw stability and the compressive

strength of concrete. After 200 freeze-thaw cycles, GO-modified concrete demonstrated a

remarkable 34.83% increase in compressive strength, suggesting that GO not only enhances

durability under freezing conditions but also improves the overall mechanical performance of

the material.

Tong et al. [201] emphasized that improvements in freeze-thaw resistance are not solely

attributable to decreased porosity, but also their impeding water transport can contribute to the

performance. They employed atomistic modelling analyses and compared the behaviour of

water molecules within the gel pores of pure C-S-H and C-S-H modified with GO during

freeze-thaw cycles. After simulating 200 million steps, they found the amount of water

absorbed during the thawing phase decreased from 286 molecules to 262 molecules for C-S-

H/GO, indicating a reduced uptake and redistribution of water within the GO-modified matrix

Further supporting these findings, Alyaa et al. [218] and Xiaoya et al. [219] observed that GO

can inhibit the formation and expansion of microcracks within cement mortar, improving frost

resistance even at low concentrations. Their studies align with Mohammed et al.’s [197]

observations that GO acts as a multifaceted defence mechanism against freeze-thaw damage.

These studies strongly suggest that graphene oxide incorporation is a promising strategy for

enhancing concrete's resistance to freeze-thaw damage. However, certain limitations highlight

the need for further research. The precise mechanism by which GO interacts with water and

ice at both macroscopic and microscopic levels remains only partially understood.

Additionally, the long-term performance of GO-modified concrete under extended freeze

75
cycles and varying environmental conditions is yet to be fully explored. Addressing these gaps

is crucial for optimizing GO's application in real-world construction scenarios.

7. Scalability

The integration of graphene oxide (GO) into concrete has shown significant promise,

particularly in enhancing mechanical properties, durability, and resistance to environmental

factors. However, scaling up GO production from laboratory research to industrial levels

presents certain challenges. These challenges include identifying efficient methods to scale up,

ensuring the availability of raw materials, improving production efficiency, and addressing

environmental sustainability and cost-effectiveness. Systematically addressing these

challenges is essential for enabling the widespread adoption of GO-based concrete in the

construction industry. This section will delve into these aspects in detail.

Graphene oxide production methods have undergone substantial evolution from experimental

laboratory methods to industrial-scale synthesis. The GO’s journey began in the mid-19th

century with Brodie’s synthesis using potassium chlorate and nitric acid in 1859. While this

method was groundbreaking, it posed significant safety risks due to the use of potassium

chlorate and was highly labour-intensive [220]. In 1898, Staudenmaier introduced a

modification to the method that used sulfuric acid, which enhanced both safety and operational

efficiency [221]. However, these early methods were unsuitable for large-scale applications

due to their prolonged reaction times and hazardous by-products.

A major breakthrough was achieved in 1958 with the development of the Hummers’ Method.

By employing potassium permanganate and sulfuric acid, this method significantly reduced

reaction times and offered higher yields [222]. Though this process is scalable, it has its own

76
limitations, including the emission of toxic gases (e.g., NO₂ and N₂O₄) and the management of

hazardous waste. Over time, Modified Hummers’ Methods (MHMs) addressed these

limitations, optimizing the oxidation process, increasing oxidant efficiency, and incorporating

environmentally friendly reagents such as potassium ferrate while eliminating the need for

sodium nitrate. Since then, the MHM has become the cornerstone of industrial GO production

due to its easy scalability, reliability, and cost-effectiveness, as shown in Fig.21 [223], [224],

[225], [226], [227].

Fig.21 (a) Scaled-up laboratory-scale production of 500g of GO: (i) 50 L reactor setup, (ii)
crude reaction mixture upon completion, (iii) 10 L continuous-flow centrifuge, (iv) over 700 g
of freeze-dried GO; (b) Bench-scale facility at NSC Co. Ltd., for 10 kg-scale GO production
[227].

While the Modified Hummers' Method (MHM) is widely adopted for the industrial production

of graphene oxide (GO), there are several critical factors that must be addressed to ensure its

successful scaling for industrial applications. These factors include raw material availability,

consistency in product quality, process efficiency, and the ability to meet the demands of large-

scale production.

77
Another major challenge in scaling the Modified Hummers' Method (MHM) is achieving

complete oxidation of graphite. In conventional processes, oxidation often remains incomplete,

leaving unoxidized or partially oxidized graphite in the reactor. Overcoming this limitation is

essential for maximizing material utilization, reducing waste, and eliminating additional

purification steps.

One effective strategy involves reducing the size of the input graphite particles. Chen et al.

[228] demonstrated that using flake graphite smaller than 20 µm can achieve full oxidation

within 30 minutes during the MHM process. In contrast, larger graphite flakes (e.g., 100 µm)

require longer reaction times due to mass-transfer limitations, as the oxidant cannot fully

penetrate the graphite. To overcome this, optimizing mass transfer is critical. For instance, Park

et al. [229] reactor designs featuring a rotating inner cylinder that generates a Couette–Taylor

flow pattern have proven highly effective. These designs enable turbulent mixing, which allows

for the complete oxidation of 50 µm graphite flakes within 30 minutes, producing highly

oxidized GO.

Another key considerations for scaling the MHM includes the quality of Go produced. The

quality of the GO produced through MHM is highly dependent on reaction parameters such as

oxidant concentration, reaction time, and temperature. These factors directly influence the

degree of oxidation, the number of exfoliable layers, and the structural integrity of the resulting

GO sheets. For instance, excessive oxidation can introduce hole defects in the GO lattice, where

carbon atoms are removed as carbon dioxide. Such defects can compromise material

performance, highlighting the need for precise control over reaction conditions [230].

The drying process is a crucial step for GO storage and subsequent use. Typically, freeze-drying

is employed to produce a dry powder. However, vacuum freeze-drying is prohibitively

expensive at an industrial scale. Peng et al. [231] reported that commercially freeze-dried GO

78
can suffer from restacking, where GO sheets aggregate, limiting their dispersion in downstream

applications. This aggregation makes it challenging to re-disperse the material effectively.

Spray drying has emerged as a more cost-effective alternative for industrial applications.

Unlike freeze-drying, spray drying minimizes the restacking of GO sheets, facilitating easier

re-dispersion and enhancing the usability of the final product. Industrial-scale spray dryers are

readily available and offer an efficient solution for large-scale GO production. By addressing

these factors, the Modified Hummers Method can be optimized for scalable and cost-effective

graphene oxide production.

Despite the success of MHMs, alternative methods have been developed to overcome specific

challenges, including environmental sustainability, production speed, and resource utilization.

Methods such as electrochemical exfoliation and flash joule heating is in the verge of rise.

Particularly electrochemical exfoliation has emerged as a promising technique for large-scale

production of GO due to its simplicity, cost-effectiveness, and scalability. Unlike the traditional

Hummers method, which involves toxic chemicals and extended processing times, the

electrochemical method offers a rapid and environmentally friendly alternative. This process

typically employs graphite as an anode and involves the oxidation and intercalation of graphite

in an electrolyte under applied voltage. The choice of electrolyte plays a crucial role in

determining the yield and quality of GO. For instance, sulfuric acid is often used to produce

sheets with a high oxidation level due to its strong oxidizing nature. A comparative study by

Norazman et al. [232] demonstrated that H2SO4 electrolytes showed higher efficiency

compared to CuSO4 and Na2SO4 electrolytes. Furthermore, the process can be modified using

eco-friendly alternative electrolytes, such as whey or green reduction techniques with plant

extracts, to further enhance the sustainability of GO synthesis [233], [234]. Comparative

studies highlight the superior performance of electrochemical methods in producing GO with

fewer defects, improved morphological control, and reduced environmental impact [235].

79
Innovations in cell design and optimization of exfoliation parameters, such as varying

electrolyte concentrations, continue to refine the quality and yield of GO sheets.

Continuous flow synthesis represents a modern approach to electrochemical exfoliation

that integrates advanced engineering principles for large-scale GO production. This method

excels in scalability, reproducibility, and process optimization, addressing critical challenges

in industrial applications as shown in Fig.22a and Fig.22b. A notable application of this

approach is the production of holey graphene oxides (hGOs) [236]. Additionally, machine

learning algorithms can be employed to optimize reaction parameters, enabling consistent

quality across batches. Compared to traditional batch methods, continuous flow synthesis

offers enhanced energy efficiency, reduced waste generation, and improved cost-effectiveness,

making it a viable solution for industrial adoption of GO production [237].

80
Fig.22a 100g/day Continuous-flow electrochemical oxidation equipment; Fig.22b 1kg/day
advanced continuous flow equipment [227], [237]; Fig.22c Automated Flash Joule system
[238].

Flash Joule heating (FJH) is another groundbreaking method that utilizes rapid

electrical heating to exfoliate graphite or convert carbon-rich waste into graphene. This process

is distinguished by its simplicity, efficiency, and environmental sustainability, making it a

promising alternative to conventional methods. FJH involves applying a high-energy electrical

pulse through a large capacitor bank, generating ultra-high temperatures, often exceeding 3000

81
K in just a few milliseconds. This rapid heating induces graphitization and exfoliation,

converting amorphous or disordered carbon structures into highly ordered graphene layers.

Preliminary studies highlight its potential for high-speed production of flash graphene

at a fraction of the cost of traditional methods. For instance, Zhu et al. [238] developed the

automated FJH system with a production rate of 21.6 grams per hour, as illustrated in Fig.22c.

Similarly, Wu et al. [239] reported that their scaled-up system has a production capacity of 5

tons per year at an energy consumption of 5 kWh per kilogram. With further optimization, flash

Joule heating could become a cornerstone of graphene-concrete applications, offering an

economical and scalable pathway for large-scale implementation.

Unlike pristine Graphene, Graphene oxide (GO) offers a scalable and economically viable

alternative with simpler synthesis processes that are well-suited for large-scale production.

While ongoing research focuses on improving scalability through innovative methods such as

microwave-assisted synthesis and biological techniques. Additionally, the use of diverse

precursor materials introduces a sustainable dimension to GO production [240], [241], [242].

Industrial and agricultural wastes, including biomass, polystyrene, graphite rods, polymers, and

soot, can serve as cost-effective feedstocks for GO synthesis [243].

This waste-to-value approach not only advances sustainability but also makes GO

manufacturing accessible to a wide range of industries. Businesses can leverage these diverse

precursors to establish GO production as a supplementary venture, enabling value addition to

existing operations while promoting adaptability across markets. By integrating GO production

into their portfolios, companies can capitalize on its growing demand, contribute to a circular

economy, and unlock new opportunities for commercialization.

82
8. Case Studies

Graphene oxide (GO) has emerged as a promising material with the potential to revolutionize

the construction industry. However, to realise its full benefits in construction, it is important to

develop scalable and cost-effective production methods. Several companies are actively

exploring its potential across a wide range of industries and paving the way for large-scale

adoption. Particularly, companies like LayerOne, First Graphene Limited, and William Blythe

Limited stand out for their significant effort, investing heavily in research and development to

increase production capacity and reduce costs.

Layerone, formerly known as Abalonyx, has made huge progress in scaling up its production

capabilities. With a current production capacity of 2.2 kilograms per day, the company is

working toward expanding this to 30 kilograms per day. This expansion is expected to reduce

production costs by a remarkable 75%, making GO more accessible to a wider range of

industries, including construction [244]. Similarly, First Graphene Limited, a publicly listed

Australian company (ASX: FGR), on March 18, 2024, announced that it had successfully

produced multi-kilogram quantities of graphene oxide at its Henderson facility in Western

Australia. This demonstrates the industry’s interest in meeting the growing demand for GO

[245].

William Blythe Limited, a UK-based GO production company, has focused on high-impact

commercial applications of graphene oxide, particularly in filtration and polymer additives.

The company is scaling up its production capabilities to 50 tonnes per year, driven by the rising

demand for GO-based solutions. It has various collaborations with academia and partners,

actively scaling up its production further. [246]

Graphenea, established in 2010, has emerged as a pioneer in developing graphene for next-

generation applications, including G-FETs (Graphene Field Effect Transistors). The company

83
has already achieved a production capacity of 1 tonne of graphene oxide per annum, as shown

in Fig.23a. and is now planning to establish a larger industrial plant with a capacity of 500

tonnes per year [247]

Fig.23a Graphenea’s GO production plant with 1 tonne per annum production capacity [247];
Fig.23b The Sixth Element (Changzhou) Materials Technology Co., Ltd. GO production plant
with a production capacity of 1100 tonnes per annum [248]; Fig.23c Graphenemex GO-based
additive for concrete [249]; Fig.23d Concretene's carbon-saving concrete admixture in
parking bays outside the University of Manchester [250]; (All the images have been
reproduced with permission from the respective source company)

In parallel, several other manufacturers have also advanced large-scale production capabilities

for graphene oxide. According to Zhu et al. [251] multiple producers have established

production lines operating at a scale of several tons per annum. Among them, Sixth Element

Materials Technology Co. Ltd, based in Changzhou, China, stands out with a large production

capacity of 1,100 tonnes of graphene oxide per year, as shown in Fig.23b, making it one of the

84
largest producers globally [248]. Sixth Element has further strengthened its market position by

obtaining relevant qualifications for its graphene oxide products across multiple industries,

showcasing its ability to meet diverse application needs. These advancements highlight the

rapid progress in scaling graphene oxide production and the drive among industry leaders to

transform graphene oxide from a laboratory innovation into a scalable and impactful material

to be applied in various applications. While the scalable production of graphene oxide is

essential for meeting the growing demand in various industries, its practical application in

concrete has been increasingly validated through numerous real-world case studies. These case

studies highlight the large-scale implementation of graphene oxide-enhanced concrete for real-

world impact, bridging the gap between laboratory findings and large-scale implementation in

diverse scenarios.

Graphenemex, a leading company in Latin America, introduced the world’s first graphene

oxide water-based additive for concrete called Graphenergy Construcción, in 2018, as shown

in Fig.23c. This innovative product enhances the compressive and tensile strength of concrete

while significantly improving its waterproofing capabilities over 3.5 times the standard

performance [249]. The company has evaluated the effects of graphene oxide (GO) on

concrete, focusing on key properties such as thermal conductivity, thermal diffusivity, and

compressive strength. They found that their admixture improves thermal insulation, reducing

internal temperatures of concrete structures by up to 3° C, which can lower energy consumption

for air conditioning and heating. Furthermore, the field test results demonstrated significant

improvements in workability, impermeability, and heat dissipation. It also improved its

mechanical properties, making GO an ideal additive for rigid pavements, highways, and

bridges [252].

Another innovation comes from Concretene, a startup founded at the University of Manchester.

It is fully focused on developing new-age concrete by incorporating Graphene with the goal of

85
reducing its carbon footprint. With £3 million in venture capital funding, the company is

bringing its carbon-saving concrete admixture to market. A large-scale trial was conducted

using its graphene-modified concrete for the parking bays outside the University of

Manchester’s Graphene Engineering Innovation Centre (GEIC), as shown in Fig.23d. The trial

demonstrated a significant 37% reduction in slab thickness, along with a 21% increase in

compressive strength over the design specifications. Over a three-year period, the concrete

showed no signs of shrinkage or cracking. Additionally, the development resulted in an

estimated CO2 savings of 3.8 tonnes, highlighting the environmental benefits of graphene-

enhanced concrete [250], [253].

Similarly, an experimental study by Cho et al. [254] demonstrated the material efficiency of

0.1 wt% graphene oxide (GO) added concrete slabs. The slab thickness was reduced by 14.6%

and 12.6% for design service lives of 25 and 50 years, respectively, showcasing GO's potential

for improving durability and resource optimization in rigid pavements.

The Sixth Element Inc. conducted extensive testing on concrete materials using its graphene

oxide-based additive, SE2430. Results showed significant improvements in mechanical

strength with the addition of just 0.025 wt% of the additive [255]. Additionally, the company's

investigations into chloride penetration resistance revealed that incorporating as little as 0.005

wt% of graphene oxide particles enhanced resistance by 40%, as shown in Fig.24a. These

findings highlight the potential of industrial GO additives to significantly enhance the

durability and structural integrity of concrete.

86
Fig.24a The Sixth Element company’s GO additive, demonstrating its effectiveness in
increasing chloride resistance [255]; Fig.24b 3D printed GO-enhanced concrete, along with
the engineers of RMIT Credit: Jonathan Tran [256]; (All the images have been reproduced
with permission from the respective source)

Beyond enhancing performance, graphene oxide-enhanced concrete also offers innovative

solutions for sustainable construction. LayerOne Advanced Materials has developed a

graphene oxide-based additive capable of reducing concrete usage by up to 50%, leading to

significant cost savings and reduced carbon emissions [257].

To further enhance sustainability, AlterBiota, a startup specializing in sustainable solutions has

developed a graphene oxide admixture from wood called Hydrous Bio Graphene Oxide

(hBGO). This innovative bio-derived material reduces the consumption of ordinary Portland

cement while simultaneously enhancing the compressive strength of concrete, addressing both

performance and environmental concerns [258].

Similarly, Bio Graphene Solutions, a sustainability-focused company, has explored the use of

bio-based graphene additives in precast concrete applications. It successfully completed a trial

87
of precast concrete in collaboration with a manufacturer. The trial demonstrated significant

improvements, including a 43% increase in strength, a 10% reduction in cement usage, and a

47% decrease in overall admixture costs [259], [260]. These results highlight the potential of

bio-based graphene additives for sustainable and cost-effective solutions for the construction

industry.

The effect of GO is not just limited to traditional concrete; it also has a significant effect on the

future of the construction industry, such as 3D printed concrete. Civil engineers from RMIT

University and the University of Melbourne have incorporated graphene oxide into the cement

matrix to enhance the properties of 3D-printed concrete, as shown in Fig.24b. The

incorporation of 0.015wt% of GO not only increased the material's strength by 10% but also

introduced electrical conductivity, enabling the concrete to monitor cracking and structural

health in real-time. This forms a significant milestone in the development of graphene oxide-

enhanced concrete. By improving bonding within the matrix and addressing common printing

issues, graphene oxide has proven to be a game-changer for the emerging field of smart

concrete [256], [261]. This advancement highlights the potential for integrating nanomaterials

and electronics into construction materials, paving the way for intelligent infrastructure

systems.

In parallel, companies such as Nanoinova Technologies, Graphenea, Global Graphene Group,

Ceylon graphene, and Universal Matter are further pushing the boundaries of scalability and

adaptation. These case studies, along with the ongoing efforts, demonstrate the significant

potential of graphene oxide-enhanced concrete to revolutionize the construction industry. By

improving mechanical properties durability and reducing the environmental impact of concrete

production, GO-based technologies offer a promising pathway toward more sustainable and

efficient construction practices. As research and development continue to progress, we can

anticipate further breakthroughs in application and adaptability.

88
9. Economic Viability and Environmental Impacts

One of the important considerations for the widespread adoption of graphene oxide-

enhanced concrete is its cost-effectiveness. While graphene oxide is relatively expensive

compared to traditional cement, its exceptional properties and the minimal quantities

requirement in concrete present viable opportunities for its economic and environmental

adaptation. This section explores the cost implications and environmental impacts of

incorporating graphene oxide in concrete.

The cost of graphene oxide currently ranges between ₹100 to ₹250 per gram ($2.95; taking $1

= ₹84.7), which is substantially higher than the cost of cement, priced at ₹0.008 per gram (₹8

per kilogram). For instance, Devi et al. [66] analysed the cost impact of incorporating 0.06 wt%

graphene oxide (GO) into concrete. Their findings showed that the cost of GO-enhanced

concrete is higher than that of conventional concrete when evaluated using the economic index

of concrete, defined as the ratio of compressive strength to cost per cubic meter. However, the

incorporation of graphene oxide in concrete typically requires only 0.1 – 1 wt% relative to the

concrete aggregate. This results in a minimal addition to the overall costs [262], [263]. Also,

its ability to enhance the properties provides an economic advantage by its ability to reduce

cement consumption. For instance, Companies like Concretene and LayerOne Advanced

Materials claim that GO-enhanced concrete can reduce cement usage by up to 50%. This

reduction in cement consumption can directly contribute to the reduction in the overall cost of

the construction, which not only offsets the cost of graphene oxide but also makes them

economically viable. Concretene estimates that their GO-enhanced concrete can deliver overall

cost savings of 10–20% compared to conventional concrete while achieving equivalent or

superior performance. Similarly, LayerOne estimates to cut down the graphene oxide

production costs by up to 75%, aiming to reduce prices to as low as 20 euros per kilogram

89
(₹1,780/kg; taking €1 = ₹89) [244]. These cost reductions, combined with enhanced durability

and mechanical properties, make GO a promising and sustainable alternative for modern

construction practices.

Adding to this, the emergence of innovative production methods such as electrochemical

oxidation and flash graphene can potentially reduce the cost of graphene oxide production. For

instance, Flash graphene can currently be produced at an exceptionally low cost of $0.16/kg,

which is the most economical method ever reported. While currently used for the production

of turbostratic graphene, research indicates that this method can be adapted to produce

graphene with oxygen content, further broadening its applicability [264], [265].

In addition to its production efficiencies, studies suggest that the high strength and durability

of graphene-enhanced concrete contribute to long-term economic benefits by significantly

extending the lifespan of infrastructure and reducing maintenance costs compared to

conventional concrete. Furthermore, the use of AI-driven models to optimize GO concentration

for specific cement formulations can help refine production costs and improve resource

efficiency. Also, as the graphene oxide production scales up, prices are expected to decline

further, enhancing its cost-competitiveness with traditional construction materials. Larger-scale

manufacturing, automated production lines, and localized production facilities can make GO

more accessible and affordable [266].

These cost-saving advantages are further amplified by efficient integration methods. The most

common method of incorporating graphene oxide into concrete involves direct mixing of it

with cement paste, this can be done by adding graphene oxide powder to the cement at the

construction site. Alternatively, cement manufacturers can integrate graphene oxide during the

production stages, which offers distinct advantages. This approach not only reduces the

implementation cost but also utilizes existing transportation and distribution networks, which

90
further lowers the logistical expenses and the carbon emissions associated with it. However,

determining the optimal concentration of graphene oxide is necessary. Extensive research is

needed to standardize the optimal dosage for various concrete applications, considering factors

such as the type of cement, aggregate, and desired application performance. By carefully

optimizing the incorporation process and dosage, graphene oxide can be effectively integrated

into concrete, leading to substantial improvements in strength, durability, and other properties

while maintaining cost-effectiveness.

The growing potential of graphene oxide is evident in its expanding market size. According to

a 2024 Market Research report, the graphene oxide market size was valued at USD 179.93

million in 2023 and is projected to reach USD 232.24 million in 2024. Growing at a compound

annual growth rate (CAGR) of 32.3%, it is expected to reach USD 6.85 billion by 2036 due to

increasing demand from sectors such as electric vehicles and batteries [267]. Similarly, the

market for graphene-based concrete additives is projected to grow at a CAGR of 26.7% from

USD 18 million in 2022 to USD 151.4 million by 2031 [268] . Within this growth, graphene

oxide is expected to capture a substantial share due to its enhanced properties and rapid

scalability. Key opportunities for graphene oxide-enhanced concrete lie in addressing specific

regional challenges, such as saltwater corrosion resistance in coastal areas and improved

freeze-thaw durability in colder regions.

From an environmental perspective, in this current situation, sustainability has become a

crucial consideration in the production and use of materials, particularly in construction.

Cement manufacturing alone accounts for over 70% of concrete’s carbon emissions. Globally,

the cement industry contributes nearly 8% of total greenhouse gas emissions, and in India, it's

about 7%. [269], [270]. These alarming statistics stress the urgent need to reduce the carbon

footprint of concrete production. Graphene oxide (GO) offers a promising solution for reducing

the carbon footprint associated with the cement industry. By actively reducing the cement

91
consumption in the structures, GO can directly reduce the carbon emissions associated with it.

Based on the current estimates, GO-enhanced concrete has the potential to reduce cement usage

by 20–50%, which lowers the emissions by 18–20% associated with it.

Beyond reducing cement usage and emissions, GO contributes to sustainability in other ways,

too. By enhancing the durability and extending the lifecycle of concrete structures, GO can

reduce the need for frequent repairs and replacements, thereby conserving resources and

minimizing construction waste. Furthermore, the improved thermal insulation properties of

GO-enhanced concrete can lead to energy savings in buildings, which further contributes to

sustainability. In addition to this, studies reveal that GO offers a significant potential for carbon

sequestration in cement systems. Mishra et al. [271] demonstrated that incorporating just 0.5

wt% of GO into concrete can improve CO₂ intake by 30%, effectively capturing and storing

the greenhouse gas within the material. This capability of GO to act as a dual-purpose additive

while maintaining the alkalinity and mechanical integrity of cement systems opens up new

opportunities for sustainable and green construction technologies.

Additionally, a life cycle assessment was conducted by Long et al. [272] demonstrated that

GO-added concrete can reduce Greenhouse gas (GHG) emissions by 6.7% and the Primary

Energy Demand (PEG) by 2.2% compared to ordinary concrete with equivalent mechanical

strength. These findings highlight the potential of GO in revolutionizing traditional

construction practices. While GO has demonstrated potential in reducing the emissions of

concrete, the carbon footprint of its own production process must also be considered. A study

by Cossutta et al. [273] evaluated the global warming potential (GWP) of various GO synthesis

methods. The analysis includes various methods, including the Modified Hummers method

using Fugetsu, Bangal, and Jeong variants (GO1–GO3) and Staudenmaier and Brodie methods

(GO4 & GO5), as shown in Fig.25a.

92
Their findings revealed that the methods using potassium chlorate (GO4) and sodium chlorate

(GO5) as oxidizing agents exhibited significantly higher GWP compared to the methods using

potassium permanganate-based oxidizing agents such as Modified Hummers methods (GO1–

GO3). Based on the results, this study suggests that for industrial scaling, it is preferable to

adopt potassium permanganate-based synthesis methods due to their comparatively lower

environmental impact.

Additionally, they simulated GO production on a commercial scale and assessed its

environmental effects. They compared GWP values across lab-scale, commercial-scale, and

decarbonized production scenarios with KOH as the oxidizing agent, as shown in Fig.25b. The

results showed that the GWP of commercial-scale production is slightly lesser than the

laboratory-scale production due to its optimised process. Also, these results highlight that the

GWP of commercial scale GO production was much lower (0.284 kg) compared to that of the

GWP of Portland cement, 0.86 kg CO₂ equivalent per kilogram [274]. This indicates that while

GO production has an environmental cost, it remains a more sustainable alternative to

conventional materials such as cement, particularly when adopting optimized and scalable

production methods.

93
Fig.25a Global warming potential (GWP) of GO synthesis methods at lab scale by chemical
oxidation; Fig.25b Simulated GWP at various production scales and conditions [273].

Similarly, Arvidsson et al. [275] evaluated the environmental impacts of the Hummers method

and the ultrasonication method for synthesizing graphene oxide and found that the Hummers

method tends to have a high energy demand and a significant blue water footprint, while the

ultrasonication method has a low footprint but exhibited higher human toxicity. The study

highlighted that multiple factors influence the overall environmental impact, such as solvent

recovery and recycling. They suggest that solvent recycling is a critical step in minimizing the

environmental footprint of graphene oxide production. By incorporating efficient solvent

recovery processes, it is possible to mitigate the environmental impacts associated with the

production.

Another aspect of increasing the sustainability of concrete is developing new methods for

graphene oxide synthesis. Conventional methods of producing graphene oxide from graphite

involve a significant carbon footprint. New and innovative methods are essential to address the

need. For instance, Pei et al. [276] introduced an innovative electrochemical exfoliation

process that is highly scalable and environmentally friendly compared to traditional methods.

GO is synthesized by ultrafast electrochemical oxidation of graphite that is not only 100 times

faster but also significantly greener than the conventional Hummers method. Similarly, flash

graphene derived from biomass presents another sustainable approach. This method not only

produces graphene in seconds but also reduces its carbon emissions and water usage by over

90% compared to that of conventional production techniques. These techniques align with the

global efforts to minimize the environmental footprint of concrete materials [277].

Another effective way to reduce the environmental effect is the integration of GO into cement

composites, especially when combined with industrial byproducts and construction or

94
demolition waste. This potential stems from the low dosage of GO required, making it cost-

effective and environmentally advantageous [278]. Additionally, ongoing efforts by researchers

and bio-companies to synthesize and scale up GO production from bio-sources are accelerating

its transition toward sustainable large-scale applications [258]. However, achieving this goal

requires comprehensive lifecycle assessments (LCAs) of GO-enhanced cement composites to

quantify their full environmental implications accurately.

Beyond its environmental benefits, the impact of GO on human health and safety must also be

carefully evaluated. Studies by Wang et al. [279] have demonstrated that GO exhibits dose-

dependent toxicity in cells and animals, leading to issues such as cell apoptosis, lung granuloma

formation, and accumulation in the body, as it cannot be efficiently cleared by the kidneys.

These findings highlight the importance of mitigating risks associated with GO exposure,

particularly at construction sites [280].

To mitigate these risks, effective measures, including controlling the dispersion of GO in the

construction site, implementing recycling protocols, and ensuring its safe handling and storage

to minimize environmental and health risks, should be taken. Despite these concerns, recent

advancements suggest progress in developing safer forms of GO. For instance, Graphenea’s

non-toxic graphene oxide has been declared safe for use, highlighting its potential for safer

alternatives to conventional GO [281].

However, investigations into the sustainability, health implications, and safety practices

surrounding graphene oxide-reinforced cement composites remain insufficient.

Comprehensive studies are essential to address these gaps, ensuring that the adoption of GO in

construction is not only environmentally beneficial but also safe for workers and communities.

95
10. Summary and Future Prospects

Graphene oxide (GO) has emerged as a transformative material in the construction

industry due to its unique two-dimensional structure and exceptional mechanical properties.

When incorporated into concrete, GO not only enhances its strength and durability but also

contributes to improved long-term performance by addressing critical issues such as porosity

and degradation. This section summarizes the key findings on the role of GO in concrete and

explores its potential to redefine construction materials for sustainable development.

o GO significantly improves the mechanical, thermal, and durability properties of

concrete. However, its tendency to agglomerate in the mix can be mitigated using

effective additives such as polycarboxylate ether (PCE).

o While GO reduces the workability of concrete due to its high specific surface area, this

challenge can be effectively managed through the incorporation of additives such as

polycarboxylate ether (PCE). These additives act as dispersing agents, preventing the

agglomeration of GO sheets and ensuring uniform distribution within the concrete

matrix.

o The addition of graphene oxide (GO) significantly enhances the mechanical properties

of concrete. It improves both compressive and flexural strengths due to enhanced

interfacial bonding between GO and hydration products.

o GO's influence extends beyond mechanical strength. It also significantly impacts the

pore structure of concrete at the nanoscale, leading to a denser microstructure with

reduced porosity.

o GO prevents the development of cracks at the nano level through its crack bridging

properties, which prevents the development of cracks and also redirects the crack

96
propagation. This contributes to enhanced stability and increased longevity of concrete

structures.

o Incorporation of GO can enhance the thermal conductivity of concrete. By acting as a

conductive pathway, it helps in dissipating the heat generated during hydration,

reducing temperature gradients, and thereby lowering the risk of internal cracking. This

leads to a stronger and more durable concrete structure across various environmental

conditions.

o GO is inherently electrically insulating due to the presence of oxygen functional groups,

which limits its direct applications in electrical applications. However, under specific

conditions, GO can contribute to conductivity in concrete under certain conditions.

When the GO concentration surpasses the percolation threshold, interactions with the

highly alkaline cementitious environment and the thermal effects of hydration can

reduce some of GO's oxygen functional groups. This reduction may create a conductive

path for electrons, thereby increasing the electrical conductivity of the concrete matrix.

o The incorporation of GO significantly enhances the durability of concrete by increasing

its resistance to water sorptivity. This improvement is attributed to the potential increase

in pore tortuosity and reduced porosity that acts as a physical barrier. This structural

densification hinders the ingress of harmful ions, thereby slowing down the degradation

of concrete.

o GO's unique ability to form a sponge-like network within the cement mortar acts as a

physical barrier to chloride and sulfate ions, thus reducing harmful ions penetration and

enhancing resistance to ions-induced corrosion. This contributes to the long-term

durability and resilience of concrete structures.

97
o GO has been suggested to mimic the behaviour of air-entraining agents, which improve

the concrete's freezing and thawing. By promoting the formation of an air void network

within the concrete, GO enhances its freeze-thaw durability.

o GO can effectively absorb electromagnetic (EM) radiation, making it a promising

material for enhancing the EM shielding effectiveness of concrete. The oxygen-

containing functional groups within GO trap electrons, forming electric dipoles that

further enhance the absorption of EM waves, providing an additional functional benefit

in specialized applications.

o GO offers wide scalability through advanced production methods that can be easily

scaled up to meet market demands. Its potential to reduce cement usage and lower cost

makes GO an economically viable option, contributing to reduced construction

expenses.

o Synthesising GO from biowaste through innovative methods can significantly reduce

the cost and carbon footprint associated with the construction industry, This positions

GO as a green alternative to traditional cement, aligning with sustainable development

goals.

o Real-world testing of GO has shown significant promise under various conditions.

However, more extensive experimentation, iteration testing, and in-depth analysis are

required to fully analyse its long-term performance.

o A major challenge in adopting GO is determining the optimal dosage of GO for

consistent improvements across diverse concrete formulations. Further research is

required to establish the best practices for incorporating GO into concrete mixtures.

98
o Understanding the interactions between GO and the hydration process is essential. GO’s

effect on hydration kinetics and microstructure development at various scales needs a

more comprehensive investigation to optimize its usage.

The future prospects of graphene oxide (GO)-added concrete lies in addressing key challenges

across multiple dimensions to unlock its full potential. In terms of mechanical and structural

properties, optimizing factors such as loading, orientation, and alignment, alongside rheology

and interfacial adhesion, is critical to enhance mechanical reinforcement and mitigate issues

like agglomeration, curing, and aging. Compatibility with various cement types, including

blended specialty cement, requires a deeper understanding of hydration kinematics, admixture

interactions, and consistency to ensure reproducibility and prevent localized weaknesses and

autogenous shrinkage.

Enhancing the durability and long-term performance of GO-enhanced concrete must account

for factors like alkali-silica reactions, delayed ettringite formation, and the need for robust

testing and standardization protocols. Health and environmental considerations are equally

important, including the risks of inhalation exposure, skin contact, and broader toxicological

concerns. These issues must be managed through improved supply chain infrastructure and

safety measures to minimize potential health and safety implications.

From a manufacturing and quality control perspective, innovations in controlling oxygen

ratios, structural defects, and functionalization are necessary to improve the scalability,

uniformity, and cost-effectiveness of GO production. Balancing these technical advancements

with environmental impact assessments is essential to ensure that GO-added concrete emerges

as a viable, sustainable, and safe material for widespread adoption in the construction industry.

99
It's important to acknowledge that GO's multi-functionality might be sensitive to real-world

conditions like temperature, humidity, and applied loads. Innovations in manufacturing

scalable, non-toxic GO variants will further accelerate its adoption in the construction industry.

Additionally, the integration of computational tools and advanced imaging techniques can

provide deeper insights into GO's interactions within cement matrices, paving the way for

customized solutions tailored to specific applications. With the current experimental findings,

companies can readily adapt and incorporate GO into their applications, provided the

concentration challenges are effectively addressed. Ultimately, GO holds the potential to

redefine sustainable construction by balancing superior performance with environmental

stewardship. By overcoming these challenges, GO-added concrete could transform

infrastructure development, offering durable, efficient, and eco-friendly solutions for the

future.

11. Conclusion

The potential of graphene oxide (GO) to enhance concrete performance is undeniable, yet a

clear understanding of the underlying mechanisms driving these improvements requires further

exploration. This study examined the multifaceted influence of GO on various aspects of

concrete, highlighting both its promise and emphasizing the need for continued investigation.

On the positive side, GO demonstrated a strong capability to enhance the mechanical properties

of concrete, including compressive and flexural strengths. This improvement is attributed to

enhanced interfacial bonding between GO and hydration products, a denser microstructure with

reduced porosity, and, potentially, the ability to redirect crack propagation. These factors

collectively indicate significant potential for substantially enhancing the strength of existing

methods. GO influences the concrete's pore structure at the nanoscale, leading to a denser

matrix with fewer and smaller pores. This not only translates to improved mechanical

100
properties but also hinders the ingress of harmful substances, ultimately slowing down the

degradation process of the concrete. However, its journey from the lab to real-world

applications still evolves, and challenges remain. Determining the optimal GO dosage for

consistent improvements across diverse concrete formulations necessitates further research.

Additionally, a more comprehensive understanding of how GO interacts with the hydration

process and influences the microstructure at various scales is crucial. The future of GO-

enhanced concrete lies in advancing its integration through interdisciplinary innovations.

Future research should focus on optimizing the large-scale production of GO through cost-

effective and environmentally sustainable methods. Innovations in scalable synthesis

techniques, improved dispersion strategies, and advanced computational modelling will be

essential in bridging the gap between laboratory research and industrial adoption. Additionally,

comprehensive life cycle assessments (LCAs) must be conducted to evaluate GO’s economic

feasibility and environmental footprint. In conclusion, GO presents a promising avenue for

developing a new generation of concrete with superior performance, its integration into

mainstream construction could redefine sustainable engineering, paving the way for resilient

and high-performance concrete solutions for the future.

Conflict of interest

The authors declare no conflicts of interest.

101
Acknowledgments

This work is financially supported by SERB, India, under Core Research Grant (CRG),

Reference No. CRG/2022/004381. The authors extend their appreciation to the Deanship of

Scientific Research and Graduate Studies at King Khalid University for funding this work

through the Large Project number, R.G.P. 2/627/45 and the authors acknowledge the Research

Center for Advanced Materials (RCAMS) at King Khalid University, Saudi Arabia for their

valuable technical support.

References

[1] N. Tkachenko et al., “Global database of cement production assets and upstream suppliers,”
Scientific Data 2023 10:1, vol. 10, no. 1, pp. 1–9, Oct. 2023, doi: 10.1038/s41597-023-02599-
w.

[2] D. Gardner, R. Lark, T. Jefferson, and R. Davies, “A survey on problems encountered in current
concrete construction and the potential benefits of self-healing cementitious materials,” Case
Studies in Construction Materials, vol. 8, pp. 238–247, Jun. 2018, doi:
10.1016/J.CSCM.2018.02.002.

[3] S. Bao, Y. Zhang, H. Liu, Y. Niu, W. Zhang, and K. Zeng, “Study on crack propagation of
functionally graded ultra-high performance cementitious composite,” Eng Fract Mech, vol.
282, p. 109143, Apr. 2023, doi: 10.1016/J.ENGFRACMECH.2023.109143.

[4] Y. Niu, H. Huang, J. Wei, C. Jiao, and Q. Miao, “Investigation of fatigue crack propagation
behavior in steel fiber-reinforced ultra-high-performance concrete (UHPC) under cyclic
flexural loading,” Compos Struct, vol. 282, p. 115126, Feb. 2022, doi:
10.1016/J.COMPSTRUCT.2021.115126.

[5] A. S. Chahar and P. Pal, “Study on various properties of reinforced concrete – A review,” Mater
Today Proc, vol. 65, pp. 597–602, Jan. 2022, doi: 10.1016/J.MATPR.2022.03.193.

[6] A. Trebukhin et al., “Influence of Nanoparticles on the Mechanical and Durability Properties of
Concrete: A Microstructural Analysis,” E3S Web of Conferences, vol. 588, p. 03005, Oct. 2024,
doi: 10.1051/E3SCONF/202458803005.

[7] Z. Najeeb and M. A. Mosaberpanah, “Mechanical and durability properties of modified High-
Performance mortar by using cenospheres and Nano-Silica,” Constr Build Mater, vol. 362, p.
129782, Jan. 2023, doi: 10.1016/J.CONBUILDMAT.2022.129782.

102
[8] “DRAFT MALAWI STANDARD (COMESA AND SADC HARMINIZED) Cement-Part 1: Composition,
specifications and conformity criteria for common cements”, Accessed: Jul. 02, 2024. [Online].
Available: www.mbsmw.org

[9] R. Si and Y.-C. Choi, “Degree of Hydration, Microstructure, and Mechanical Properties of
Cement-Modified TiO2 Nanoparticles,” Materials 2024, Vol. 17, Page 4541, vol. 17, no. 18, p.
4541, Sep. 2024, doi: 10.3390/MA17184541.

[10] R. Sridhar, P. Aosai, T. Imjai, M. Setkit, A. Shirkol, and I. Laory, “Influence of Nanoparticles and
PVA Fibers on Concrete and Mortar on Microstructural and Durability Properties,” Fibers
2024, Vol. 12, Page 54, vol. 12, no. 7, p. 54, Jun. 2024, doi: 10.3390/FIB12070054.

[11] H. Yang, D. Zheng, W. Tang, X. Bao, and H. Cui, “Application of graphene and its derivatives in
cementitious materials: An overview,” Journal of Building Engineering, vol. 65, p. 105721, Apr.
2023, doi: 10.1016/J.JOBE.2022.105721.

[12] X. Zhang, S. Wan, J. Pu, L. Wang, and X. Liu, “Highly hydrophobic and adhesive performance of
graphene films,” J Mater Chem, vol. 21, no. 33, pp. 12251–12258, Aug. 2011, doi:
10.1039/C1JM12087E.

[13] K. Chintalapudi and R. M. R. Pannem, “An intense review on the performance of Graphene
Oxide and reduced Graphene Oxide in an admixed cement system,” Constr Build Mater, vol.
259, p. 120598, Oct. 2020, doi: 10.1016/J.CONBUILDMAT.2020.120598.

[14] P. Zhang, Y. Sun, J. Wei, and T. Zhang, “Research progress on properties of cement-based
composites incorporating graphene oxide,” Reviews on Advanced Materials Science, vol. 62,
no. 1, Jan. 2023, doi: 10.1515/RAMS-2022-0329/ASSET/GRAPHIC/J_RAMS-2022-
0329_FIG_019.JPG.

[15] A. Talaat, A. Emad, A. Tarek, M. Masbouba, A. Essam, and M. Kohail, “Factors affecting the
results of concrete compression testing: A review,” Ain Shams Engineering Journal, vol. 12, no.
1, pp. 205–221, Mar. 2021, doi: 10.1016/J.ASEJ.2020.07.015.

[16] P. M. Walunjkar and M. N. Bajad, “Review on effect of graphene oxide reinforcement on


properties of cement concrete,” Mater Today Proc, Sep. 2023, doi:
10.1016/J.MATPR.2023.09.085.

[17] M. Åvec et al., “Van der Waals interactions mediating the cohesion of fullerenes on
graphene,” Phys Rev B Condens Matter Mater Phys, vol. 86, no. 12, p. 121407, Sep. 2012, doi:
10.1103/PHYSREVB.86.121407/FIGURES/3/MEDIUM.

[18] X. Li, L. Wang, Y. Liu, W. Li, B. Dong, and W. H. Duan, “Dispersion of graphene oxide
agglomerates in cement paste and its effects on electrical resistivity and flexural strength,”
Cem Concr Compos, vol. 92, pp. 145–154, Sep. 2018, doi:
10.1016/J.CEMCONCOMP.2018.06.008.

[19] G. Xu, S. Du, J. He, and X. Shi, “The role of admixed graphene oxide in a cement hydration
system,” Carbon N Y, vol. 148, pp. 141–150, Jul. 2019, doi: 10.1016/J.CARBON.2019.03.072.

[20] A. H. Korayem, N. Tourani, M. Zakertabrizi, A. M. Sabziparvar, and W. H. Duan, “A review of


dispersion of nanoparticles in cementitious matrices: Nanoparticle geometry perspective,”
Constr Build Mater, vol. 153, pp. 346–357, Oct. 2017, doi:
10.1016/J.CONBUILDMAT.2017.06.164.

103
[21] Z. Pan et al., “Mechanical properties and microstructure of a graphene oxide–cement
composite,” Cem Concr Compos, vol. 58, pp. 140–147, Apr. 2015, doi:
10.1016/J.CEMCONCOMP.2015.02.001.

[22] A. M. Sabziparvar, E. Hosseini, V. Chiniforush, and A. H. Korayem, “Barriers to achieving highly


dispersed graphene oxide in cementitious composites: An experimental and computational
study,” Constr Build Mater, vol. 199, pp. 269–278, Feb. 2019, doi:
10.1016/J.CONBUILDMAT.2018.12.030.

[23] R. Kaur and N. C. Kothiyal, “Positive synergistic effect of superplasticizer stabilized graphene
oxide and functionalized carbon nanotubes as a 3-D hybrid reinforcing phase on the
mechanical properties and pore structure refinement of cement nanocomposites,” Constr
Build Mater, vol. 222, pp. 358–370, Oct. 2019, doi: 10.1016/J.CONBUILDMAT.2019.06.152.

[24] C. S. R. Indukuri, R. Nerella, and S. R. C. Madduru, “Workability, microstructure, strength


properties and durability properties of graphene oxide reinforced cement paste,” Australian
Journal of Civil Engineering, vol. 18, no. 1, pp. 73–81, Jan. 2020, doi:
10.1080/14488353.2020.1721952.

[25] L. Zhao et al., “Deep research about the mechanisms of graphene oxide (GO) aggregation in
alkaline cement pore solution,” Constr Build Mater, vol. 247, p. 118446, Jun. 2020, doi:
10.1016/J.CONBUILDMAT.2020.118446.

[26] W. Qin, Q. Guodong, Z. Dafu, W. Yue, and Z. Haiyu, “Influence of the molecular structure of a
polycarboxylate superplasticiser on the dispersion of graphene oxide in cement pore solutions
and cement-based composites,” Constr Build Mater, vol. 272, p. 121969, Feb. 2021, doi:
10.1016/J.CONBUILDMAT.2020.121969.

[27] Z. Lu, B. Chen, C. K. Y. Leung, Z. Li, and G. Sun, “Aggregation size effect of graphene oxide on
its reinforcing efficiency to cement-based materials,” Cem Concr Compos, vol. 100, pp. 85–91,
Jul. 2019, doi: 10.1016/J.CEMCONCOMP.2019.04.005.

[28] W. J. Long, Y. cun Gu, B. X. Xiao, Q. ming Zhang, and F. Xing, “Micro-mechanical properties and
multi-scaled pore structure of graphene oxide cement paste: Synergistic application of
nanoindentation, X-ray computed tomography, and SEM-EDS analysis,” Constr Build Mater,
vol. 179, pp. 661–674, Aug. 2018, doi: 10.1016/J.CONBUILDMAT.2018.05.229.

[29] G. Xiong et al., “Effect of power ultrasound assisted mixing on graphene oxide in cement
paste: Dispersion, microstructure and mechanical properties,” Journal of Building
Engineering, vol. 69, p. 106321, Jun. 2023, doi: 10.1016/J.JOBE.2023.106321.

[30] S. Chuah, W. Li, S. J. Chen, J. G. Sanjayan, and W. H. Duan, “Investigation on dispersion of


graphene oxide in cement composite using different surfactant treatments,” Constr Build
Mater, vol. 161, pp. 519–527, Feb. 2018, doi: 10.1016/J.CONBUILDMAT.2017.11.154.

[31] D. Lu, Z. Sheng, B. Yan, and Z. Jiang, “Co-effects of Graphene Oxide and Silica Fume on the
Rheological Properties of Cement Paste,” Lecture Notes in Civil Engineering, vol. 356 LNCE, pp.
251–258, 2023, doi: 10.1007/978-981-99-3330-3_26/FIGURES/4.

[32] Z. Lu, D. Hou, A. Hanif, W. Hao, G. Sun, and Z. Li, “Comparative evaluation on the dispersion
and stability of graphene oxide in water and cement pore solution by incorporating silica
fume,” Cem Concr Compos, vol. 94, pp. 33–42, Nov. 2018, doi:
10.1016/J.CEMCONCOMP.2018.08.011.

104
[33] Z. Lu, A. Hanif, G. Sun, R. Liang, P. Parthasarathy, and Z. Li, “Highly dispersed graphene oxide
electrodeposited carbon fiber reinforced cement-based materials with enhanced mechanical
properties,” Cem Concr Compos, vol. 87, pp. 220–228, Mar. 2018, doi:
10.1016/J.CEMCONCOMP.2018.01.006.

[34] L. Zhao et al., “Investigation of dispersion behavior of GO modified by different water


reducing agents in cement pore solution,” Carbon N Y, vol. 127, pp. 255–269, Feb. 2018, doi:
10.1016/J.CARBON.2017.11.016.

[35] H. Qin, W. Wei, and Y. Hang Hu, “Synergistic effect of graphene-oxide-doping and microwave-
curing on mechanical strength of cement,” Journal of Physics and Chemistry of Solids, vol.
103, pp. 67–72, Apr. 2017, doi: 10.1016/J.JPCS.2016.12.009.

[36] G. Jing et al., “The non-uniform spatial dispersion of graphene oxide: A step forward to
understand the inconsistent properties of cement composites,” Constr Build Mater, vol. 264,
p. 120729, Dec. 2020, doi: 10.1016/J.CONBUILDMAT.2020.120729.

[37] J. Li and Q. Zheng, “The first experimental evidence for improved nanomechanical properties
of calcium silicate hydrate by polycarboxylate ether and graphene oxide,” Cem Concr Res, vol.
156, p. 106787, Jun. 2022, doi: 10.1016/J.CEMCONRES.2022.106787.

[38] L. Zhao et al., “Synergistic effects of silica nanoparticles/polycarboxylate superplasticizer


modified graphene oxide on mechanical behavior and hydration process of cement
composites,” RSC Adv, vol. 7, no. 27, pp. 16688–16702, Mar. 2017, doi: 10.1039/C7RA01716B.

[39] F. Babak, H. Abolfazl, R. Alimorad, and G. Parviz, “Preparation and Mechanical Properties of
Graphene Oxide: Cement Nanocomposites,” The Scientific World Journal, vol. 2014, no. 1, p.
276323, Jan. 2014, doi: 10.1155/2014/276323.

[40] S. Ghazizadeh, P. Duffour, N. T. Skipper, and Y. Bai, “Understanding the behaviour of graphene
oxide in Portland cement paste,” Cem Concr Res, vol. 111, pp. 169–182, Sep. 2018, doi:
10.1016/J.CEMCONRES.2018.05.016.

[41] W. Zhu, Q. Feng, Q. Luo, X. Bai, X. Lin, and Z. Zhang, “Effects of PCE on the Dispersion of
Cement Particles and Initial Hydration,” Materials 2021, Vol. 14, Page 3195, vol. 14, no. 12, p.
3195, Jun. 2021, doi: 10.3390/MA14123195.

[42] X. Yan, D. Zheng, H. Yang, H. Cui, M. Monasterio, and Y. Lo, “Study of optimizing graphene
oxide dispersion and properties of the resulting cement mortars,” Constr Build Mater, vol.
257, p. 119477, Oct. 2020, doi: 10.1016/J.CONBUILDMAT.2020.119477.

[43] D. I. Petukhov, O. O. Kapitanova, E. A. Eremina, and E. A. Goodilin, “Preparation, chemical


features, structure and applications of membrane materials based on graphene oxide,”
Mendeleev Communications, vol. 31, no. 2, pp. 137–148, Mar. 2021, doi:
10.1016/J.MENCOM.2021.03.001.

[44] K. Gong et al., “Reinforcing Effects of Graphene Oxide on Portland Cement Paste,” Journal of
Materials in Civil Engineering, vol. 27, no. 2, p. A4014010, Jul. 2014, doi:
10.1061/(ASCE)MT.1943-5533.0001125.

[45] P. K. Akarsh, S. Marathe, and A. K. Bhat, “Influence of graphene oxide on properties of


concrete in the presence of silica fumes and M-sand,” Constr Build Mater, vol. 268, p. 121093,
Jan. 2021, doi: 10.1016/J.CONBUILDMAT.2020.121093.

105
[46] Q. Wang, J. Wang, C. X. Lu, X. Y. Cui, S. Y. Li, and X. Wang, “Rheological behavior of fresh
cement pastes with a graphene oxide additive,” New Carbon Materials, vol. 31, no. 6, pp.
574–584, Dec. 2016, doi: 10.1016/S1872-5805(16)60033-1.

[47] K. Vallurupalli, W. Meng, J. Liu, and K. H. Khayat, “Effect of graphene oxide on rheology,
hydration and strength development of cement paste,” Constr Build Mater, vol. 265, p.
120311, Dec. 2020, doi: 10.1016/J.CONBUILDMAT.2020.120311.

[48] Q. Wang, S. Y. Li, S. Pan, and Z. W. Guo, “Synthesis and properties of a silane and copolymer-
modified graphene oxide for use as a water-reducing agent in cement pastes,” New Carbon
Materials, vol. 33, no. 2, pp. 131–139, Apr. 2018, doi: 10.1016/S1872-5805(18)60330-0.

[49] R. Roy, A. Mitra, A. T. Ganesh, and V. Sairam, “Effect of Graphene Oxide Nanosheets dispersion
in cement mortar composites incorporating Metakaolin and Silica Fume,” Constr Build Mater,
vol. 186, pp. 514–524, Oct. 2018, doi: 10.1016/J.CONBUILDMAT.2018.07.135.

[50] O. FABER, “PLASTIC YIELD, SHRINKAGE, AND OTHER PROBLEMS OF CONCRETE, AND THEIR
EFFECT ON DESIGN.,” https://doi.org/10.1680/imotp.1928.14232, vol. 225, no. 1928, pp. 27–
73, Jun. 2015, doi: 10.1680/IMOTP.1928.14232.

[51] D. Lange-Kornbak and B. L. Karihaloo, “Design of concrete mixes for minimum brittleness,”
Advanced Cement Based Materials, vol. 3, no. 3–4, pp. 124–132, Apr. 1996, doi:
10.1016/S1065-7355(96)90044-9.

[52] A. F. Al Fuhaid and A. Niaz, “Carbonation and Corrosion Problems in Reinforced Concrete
Structures,” Buildings, vol. 12, no. 5, pp. 586–586, May 2022, doi:
10.3390/BUILDINGS12050586.

[53] H. Azari and S. Nazarian, “Assessment of Concrete Structures Affected by Alkali–Silica


Reaction and Freeze–Thaw by Use of Ultrasonic Surface Wave Method,” Transp Res Rec, vol.
2592, no. 2592, pp. 117–125, Oct. 2016, doi: 10.3141/2592-13.

[54] J. J. Biernacki et al., “Cements in the 21st century: Challenges, perspectives, and
opportunities,” Journal of the American Ceramic Society, vol. 100, no. 7, pp. 2746–2773, Jul.
2017, doi: 10.1111/JACE.14948.

[55] S. Paruthi, I. Rahman, A. Husain, A. H. Khan, A. M. Manea-Saghin, and E. Sabi, “A


comprehensive review of nano materials in geopolymer concrete: Impact on properties and
performance,” Developments in the Built Environment, vol. 16, p. 100287, Dec. 2023, doi:
10.1016/J.DIBE.2023.100287.

[56] D. Y. Yoo, T. Oh, and N. Banthia, “Nanomaterials in ultra-high-performance concrete (UHPC) –


A review,” Cem Concr Compos, vol. 134, p. 104730, Nov. 2022, doi:
10.1016/J.CEMCONCOMP.2022.104730.

[57] S. Kumar, N. Bheel, S. Zardari, A. S. Alraeeini, A. H. Almaliki, and O. Benjeddou, “Effect of


graphene oxide on mechanical, deformation and drying shrinkage properties of concrete
reinforced with fly ash as cementitious material by using RSM modelling,” Scientific Reports
2024 14:1, vol. 14, no. 1, pp. 1–23, Aug. 2024, doi: 10.1038/s41598-024-69601-2.

[58] A. Bagheri, E. Negahban, A. Asad, H. A. Abbasi, and S. M. Raza, “Graphene oxide-incorporated


cementitious composites: a thorough investigation,” Mater Adv, vol. 3, no. 24, pp. 9040–
9051, Dec. 2022, doi: 10.1039/D2MA00169A.

106
[59] I. Fonseka, D. Mohotti, K. Wijesooriya, C. K. Lee, and P. Mendis, “Influence of Graphene oxide
on abrasion resistance and strength of concrete,” Constr Build Mater, vol. 404, p. 133280,
Nov. 2023, doi: 10.1016/J.CONBUILDMAT.2023.133280.

[60] M. M. Mokhtar, S. A. Abo-El-Enein, M. Y. Hassaan, M. S. Morsy, and M. H. Khalil, “Mechanical


performance, pore structure and micro-structural characteristics of graphene oxide nano
platelets reinforced cement,” Constr Build Mater, vol. 138, pp. 333–339, May 2017, doi:
10.1016/J.CONBUILDMAT.2017.02.021.

[61] S. Lv, Y. Ma, C. Qiu, T. Sun, J. Liu, and Q. Zhou, “Effect of graphene oxide nanosheets of
microstructure and mechanical properties of cement composites,” Constr Build Mater, vol.
49, pp. 121–127, Dec. 2013, doi: 10.1016/J.CONBUILDMAT.2013.08.022.

[62] C. Chhorn, S. J. Hong, and S. W. Lee, “Relationship between compressive and tensile strengths
of roller-compacted concrete,” Journal of Traffic and Transportation Engineering (English
Edition), vol. 5, no. 3, pp. 215–223, Jun. 2018, doi: 10.1016/j.jtte.2017.09.002.

[63] Y. Zhou et al., “Optimization of Graphene Nanoplatelets Dispersion and Its Performance in
Cement Mortars,” Materials 2022, Vol. 15, Page 7308, vol. 15, no. 20, p. 7308, Oct. 2022, doi:
10.3390/MA15207308.

[64] Y. Y. Wu, L. Que, Z. Cui, and P. Lambert, “Physical Properties of Concrete Containing Graphene
Oxide Nanosheets,” Materials 2019, Vol. 12, Page 1707, vol. 12, no. 10, p. 1707, May 2019,
doi: 10.3390/MA12101707.

[65] G. Jing et al., “From graphene oxide to reduced graphene oxide: Enhanced hydration and
compressive strength of cement composites,” Constr Build Mater, vol. 248, p. 118699, Jul.
2020, doi: 10.1016/J.CONBUILDMAT.2020.118699.

[66] S. C. Devi and R. A. Khan, “Effect of graphene oxide on mechanical and durability performance
of concrete,” Journal of Building Engineering, vol. 27, p. 101007, Jan. 2020, doi:
10.1016/J.JOBE.2019.101007.

[67] M. Degefe and A. Parashar, “Effect of non-bonded interactions on failure morphology of a


defective graphene sheet,” Mater Res Express, vol. 3, no. 4, p. 045009, Apr. 2016, doi:
10.1088/2053-1591/3/4/045009.

[68] K. L. Goh, “Fibre Debonding, Matrix Yielding and Cracks,” pp. 77–97, 2017, doi: 10.1007/978-
1-4471-7305-2_4.

[69] X. Li, H. Grassl, C. Hesse, and J. Dengler, “Unlocking the potential of ordinary Portland cement
with hydration control additive enabling low-carbon building materials,” Communications
Materials 2024 5:1, vol. 5, no. 1, pp. 1–9, Jan. 2024, doi: 10.1038/s43246-023-00441-9.

[70] S. K. U. Rehman, S. Kumarova, S. A. Memon, M. F. Javed, and M. Jameel, “A Review of


Microscale, Rheological, Mechanical, Thermoelectrical and Piezoresistive Properties of
Graphene Based Cement Composite,” Nanomaterials 2020, Vol. 10, Page 2076, vol. 10, no.
10, p. 2076, Oct. 2020, doi: 10.3390/NANO10102076.

[71] D. Hou, Z. Lu, X. Li, H. Ma, and Z. Li, “Reactive molecular dynamics and experimental study of
graphene-cement composites: Structure, dynamics and reinforcement mechanisms,” Carbon
N Y, vol. 115, pp. 188–208, May 2017, doi: 10.1016/J.CARBON.2017.01.013.

107
[72] S. Lv, H. Hu, J. Zhang, Y. Lei, L. Sun, and Y. Hou, “Structure, performances, and formation
mechanism of cement composites with large-scale regular microstructure by distributing
uniformly few-layered graphene oxide in cement matrix,” Structural Concrete, vol. 20, no. 1,
pp. 471–482, Feb. 2019, doi: 10.1002/SUCO.201800078.

[73] T. S. Qureshi and D. K. Panesar, “Impact of graphene oxide and highly reduced graphene oxide
on cement based composites,” Constr Build Mater, vol. 206, pp. 71–83, May 2019, doi:
10.1016/J.CONBUILDMAT.2019.01.176.

[74] L. Zhao et al., “Hydration kinetics, pore structure, 3D network calcium silicate hydrate, and
mechanical behavior of graphene oxide reinforced cement composites,” Constr Build Mater,
vol. 190, pp. 150–163, Nov. 2018, doi: 10.1016/J.CONBUILDMAT.2018.09.105.

[75] W. J. Long, T. H. Ye, Y. C. Gu, H. D. Li, and F. Xing, “Inhibited effect of graphene oxide on
calcium leaching of cement pastes,” Constr Build Mater, vol. 202, pp. 177–188, Mar. 2019,
doi: 10.1016/J.CONBUILDMAT.2018.12.194.

[76] S. Sharma and N. C. Kothiyal, “Comparative effects of pristine and ball-milled graphene oxide
on physico-chemical characteristics of cement mortar nanocomposites,” Constr Build Mater,
vol. 115, pp. 256–268, Jul. 2016, doi: 10.1016/J.CONBUILDMAT.2016.04.019.

[77] H. Peng, Y. Ge, C. S. Cai, Y. Zhang, and Z. Liu, “Mechanical properties and microstructure of
graphene oxide cement-based composites,” Constr Build Mater, vol. 194, pp. 102–109, Jan.
2019, doi: 10.1016/J.CONBUILDMAT.2018.10.234.

[78] S. Lv, S. Ting, J. Liu, and Q. Zhou, “Use of graphene oxide nanosheets to regulate the
microstructure of hardened cement paste to increase its strength and toughness,”
CrystEngComm, vol. 16, no. 36, pp. 8508–8516, Aug. 2014, doi: 10.1039/C4CE00684D.

[79] A. Anwar, X. Liu, and L. Zhang, “Nano-cementitious composites modified with Graphene Oxide
– a review,” Thin-Walled Structures, vol. 183, p. 110326, Feb. 2023, doi:
10.1016/J.TWS.2022.110326.

[80] R. Kaur, N. C. Kothiyal, and H. Arora, “Studies on combined effect of superplasticizer modified
graphene oxide and carbon nanotubes on the physico-mechanical strength and electrical
resistivity of fly ash blended cement mortar,” Journal of Building Engineering, vol. 30, p.
101304, Jul. 2020, doi: 10.1016/J.JOBE.2020.101304.

[81] E. Horszczaruk, E. Mijowska, R. J. Kalenczuk, M. Aleksandrzak, and S. Mijowska,


“Nanocomposite of cement/graphene oxide – Impact on hydration kinetics and Young’s
modulus,” Constr Build Mater, vol. 78, pp. 234–242, Mar. 2015, doi:
10.1016/J.CONBUILDMAT.2014.12.009.

[82] W. J. Long, J. J. Wei, F. Xing, and K. H. Khayat, “Enhanced dynamic mechanical properties of
cement paste modified with graphene oxide nanosheets and its reinforcing mechanism,” Cem
Concr Compos, vol. 93, pp. 127–139, Oct. 2018, doi: 10.1016/J.CEMCONCOMP.2018.07.001.

[83] L. Wang et al., “Effect of Graphene Oxide (GO) on the Morphology and Microstructure of
Cement Hydration Products,” Nanomaterials 2017, Vol. 7, Page 429, vol. 7, no. 12, p. 429,
Dec. 2017, doi: 10.3390/NANO7120429.

108
[84] H. Yang, M. Monasterio, H. Cui, and N. Han, “Experimental study of the effects of graphene
oxide on microstructure and properties of cement paste composite,” Compos Part A Appl Sci
Manuf, vol. 102, pp. 263–272, Nov. 2017, doi: 10.1016/J.COMPOSITESA.2017.07.022.

[85] H. Cui, X. Yan, L. Tang, and F. Xing, “Possible pitfall in sample preparation for SEM analysis - A
discussion of the paper ‘Fabrication of polycarboxylate/graphene oxide nanosheet composites
by copolymerization for reinforcing and toughening cement composites’ by Lv et al.,” Cem
Concr Compos, vol. 77, pp. 81–85, Mar. 2017, doi: 10.1016/J.CEMCONCOMP.2016.12.007.

[86] A. Gholampour, M. V. Kiamahalleh, D. N. H. Tran, T. Ozbakkaloglu, and D. Losic, “Revealing the


dependence of the physiochemical and mechanical properties of cement composites on
graphene oxide concentration,” RSC Adv, vol. 7, no. 87, pp. 55148–55156, Dec. 2017, doi:
10.1039/C7RA10066C.

[87] D. Kang, K. S. Seo, H. Y. Lee, and W. Chung, “Experimental study on mechanical strength of
GO-cement composites,” Constr Build Mater, vol. 131, pp. 303–308, Jan. 2017, doi:
10.1016/J.CONBUILDMAT.2016.11.083.

[88] C. Lu, Z. Lu, Z. Li, and C. K. Y. Leung, “Effect of graphene oxide on the mechanical behavior of
strain hardening cementitious composites,” Constr Build Mater, vol. 120, pp. 457–464, Sep.
2016, doi: 10.1016/J.CONBUILDMAT.2016.05.122.

[89] V. Romero, J. Antonio, C.-S. German, M. Maldonado, and E. Raymundo, “International Journal
of scientific research and management (IJSRM) ||Volume||4||Issue||06||Pages||4324-
4332||2016|| Website: www.ijsrm.in ISSN (e): 2321-3418 Optimizing content graphene oxide
in high strength concrete,” vol. 4, no. 6, 2016, Accessed: Jan. 10, 2024. [Online]. Available:
www.ijsrm.in

[90] K. R. M. Shareef, S. A. Rawoof, and K. Sowjanya, “A FEASIBILITY STUDY ON MECHANICAL


PROPERTIES OF CONCRETE WITH GRAPHENE OXIDE,” International Research Journal of
Engineering and Technology, vol. 9001, 2008, Accessed: Jan. 10, 2024. [Online]. Available:
www.irjet.net

[91] X. Yuan, L. Zhang, X. Chen, and F. Liu, “Study on the mechanical properties and frost resistance
of multiple modified concrete,” Mater Res Express, vol. 9, no. 4, p. 045013, Apr. 2022, doi:
10.1088/2053-1591/AC3951.

[92] P. V. R. K. Reddy and D. R. Prasad, “The role of graphene oxide in the strength and vibration
characteristics of standard and high-grade cement concrete,” Journal of Building Engineering,
vol. 63, p. 105481, Jan. 2023, doi: 10.1016/J.JOBE.2022.105481.

[93] C. Ruan, J. Lin, S. Chen, K. Sagoe-Crentsil, and W. Duan, “Effect of Graphene Oxide on the Pore
Structure of Cement Paste: Implications for Performance Enhancement,” ACS Appl Nano
Mater, vol. 4, no. 10, pp. 10623–10633, Oct. 2021, doi:
10.1021/ACSANM.1C02090/SUPPL_FILE/AN1C02090_SI_001.PDF.

[94] K. Ghouchani, H. Abbasi, and E. Najaf, “Some mechanical properties and microstructure of
cementitious nanocomposites containing nano-SiO2 and graphene oxide nanosheets,” Case
Studies in Construction Materials, vol. 17, p. e01482, Dec. 2022, doi:
10.1016/J.CSCM.2022.E01482.

[95] Y. Chen, G. Li, L. Li, W. Zhang, and K. Dong, “Molecular dynamics simulation and experimental
study on mechanical properties and microstructure of cement-based composites enhanced by

109
graphene oxide and graphene,” Mol Simul, vol. 49, no. 3, pp. 251–262, Feb. 2023, doi:
10.1080/08927022.2022.2156560.

[96] P. V. R. K. Reddy and D. Ravi Prasad, “Graphene oxide reinforced cement concrete – a study on
mechanical, durability and microstructure characteristics,” Fullerenes, Nanotubes and Carbon
Nanostructures, vol. 31, no. 3, pp. 255–265, 2023, doi: 10.1080/1536383X.2022.2141231.

[97] K. Samimi, M. Pakan, and J. Eslami, “Investigating the compressive strength and
microstructural analysis of mortar containing synthesized graphene and natural pozzolan in
the face of alkali-silica reactions,” Journal of Building Engineering, vol. 68, p. 106126, Jun.
2023, doi: 10.1016/J.JOBE.2023.106126.

[98] I. Merta and E. K. Tschegg, “Fracture energy of natural fibre reinforced concrete,” Constr Build
Mater, vol. 40, pp. 991–997, Mar. 2013, doi: 10.1016/J.CONBUILDMAT.2012.11.060.

[99] S. Zhang, B. Han, H. Xie, M. An, and S. Lyu, “Brittleness of Concrete under Different Curing
Conditions,” Materials 2021, Vol. 14, Page 7865, vol. 14, no. 24, p. 7865, Dec. 2021, doi:
10.3390/MA14247865.

[100] C. Chen, W. Zhang, H. Wu, C. Chen, and B. Ma, “Study on the damage resistance and
deterioration behavior of GO concrete under the harsh environment,” PLoS One, vol. 18, no.
4, p. e0284186, Apr. 2023, doi: 10.1371/JOURNAL.PONE.0284186.

[101] J. Wang, S. Dong, X. Yu, and B. Han, “Mechanical properties of graphene-reinforced reactive
powder concrete at different strain rates,” J Mater Sci, vol. 55, no. 8, pp. 3369–3387, Mar.
2020, doi: 10.1007/S10853-019-04246-5/TABLES/3.

[102] S. Liu, F. Lu, Y. Chen, B. Dong, H. Du, and X. Li, “Efficient Use of Graphene Oxide in Layered
Cement Mortar,” Materials 2022, Vol. 15, Page 2181, vol. 15, no. 6, p. 2181, Mar. 2022, doi:
10.3390/MA15062181.

[103] A. M. B. Chandima and S. P. Guluwita, “A Review on Mechanical Properties and Morphological


Properties of Concrete with Graphene Oxide,” Lecture Notes in Civil Engineering, vol. 94, pp.
129–140, 2021, doi: 10.1007/978-981-15-7222-7_12.

[104] A. Gholampour, M. Valizadeh Kiamahalleh, D. N. H. Tran, T. Ozbakkaloglu, and D. Losic, “From


Graphene Oxide to Reduced Graphene Oxide: Impact on the Physiochemical and Mechanical
Properties of Graphene-Cement Composites,” ACS Appl Mater Interfaces, vol. 9, no. 49, pp.
43275–43286, Dec. 2017, doi:
10.1021/ACSAMI.7B16736/SUPPL_FILE/AM7B16736_SI_001.PDF.

[105] C. S. R. Indukuri and R. Nerella, “Enhanced transport properties of graphene oxide based
cement composite material,” Journal of Building Engineering, vol. 37, p. 102174, May 2021,
doi: 10.1016/J.JOBE.2021.102174.

[106] C. Liu, X. Huang, Y. Y. Wu, X. Deng, and Z. Zheng, “The effect of graphene oxide on the
mechanical properties, impermeability and corrosion resistance of cement mortar containing
mineral admixtures,” Constr Build Mater, vol. 288, p. 123059, Jun. 2021, doi:
10.1016/J.CONBUILDMAT.2021.123059.

[107] Z. Chen, Y. Xu, J. Hua, X. Wang, L. Huang, and X. Zhou, “Mechanical Properties and Shrinkage
Behavior of Concrete-Containing Graphene-Oxide Nanosheets,” Materials 2020, Vol. 13, Page
590, vol. 13, no. 3, p. 590, Jan. 2020, doi: 10.3390/MA13030590.

110
[108] H. Sun et al., “Influence of Graphene Oxide on Interfacial Transition Zone of Mortar,” J
Nanomater, vol. 2020, no. 1, p. 8919681, Jan. 2020, doi: 10.1155/2020/8919681.

[109] C. Liu et al., “Research progress on individual effect of graphene oxide in cement-based
materials and its synergistic effect with other nanomaterials,” Nanotechnol Rev, vol. 10, no. 1,
pp. 1208–1235, Jan. 2021, doi: 10.1515/NTREV-2021-0080/ASSET/GRAPHIC/J_NTREV-2021-
0080_FIG_014.JPG.

[110] X. Li et al., “Improvement of mechanical properties by incorporating graphene oxide into


cement mortar,” Mechanics of Advanced Materials and Structures, vol. 25, no. 15–16, pp.
1313–1322, Dec. 2018, doi: 10.1080/15376494.2016.1218226.

[111] H. Zeng, Y. Lai, S. Qu, and F. Yu, “Effect of graphene oxide on permeability of cement
materials: An experimental and theoretical perspective,” Journal of Building Engineering, vol.
41, p. 102326, Sep. 2021, doi: 10.1016/J.JOBE.2021.102326.

[112] B. Ren, E. Bai, X. Luo, T. Wang, and Z. Wang, “Impact mechanical properties and pore
structure of graphene oxide concrete at high temperature,” Journal of Building Engineering,
vol. 85, p. 108593, May 2024, doi: 10.1016/J.JOBE.2024.108593.

[113] S. Chuah, Z. Pan, J. G. Sanjayan, C. M. Wang, and W. H. Duan, “Nano reinforced cement and
concrete composites and new perspective from graphene oxide,” Constr Build Mater, vol. 73,
pp. 113–124, Dec. 2014, doi: 10.1016/J.CONBUILDMAT.2014.09.040.

[114] Y. Suo, R. Guo, H. Xia, Y. Yang, B. Zhou, and Z. Zhao, “A review of graphene oxide/cement
composites: Performance, functionality, mechanisms, and prospects,” Journal of Building
Engineering, vol. 53, p. 104502, Aug. 2022, doi: 10.1016/J.JOBE.2022.104502.

[115] H. Zeng, S. Qu, and Y. Qin, “Microstructure and transport properties of cement-based material
enhanced by graphene oxide,” https://doi.org/10.1680/jmacr.19.00558, vol. 73, no. 19, pp.
1011–1024, Jan. 2021, doi: 10.1680/JMACR.19.00558.

[116] M. Sharma and S. Bishnoi, “The Interfacial Transition Zone: Microstructure, Properties, and Its
Modification,” Lecture Notes in Civil Engineering, vol. 12, pp. 745–754, 2019, doi:
10.1007/978-981-13-0365-4_63.

[117] S. Y. Chung and J. S. Kim, “Investigation of gradient properties for cement mortar interfacial
transition zones and their effects on mechanical characteristics using micro-CT,”
Developments in the Built Environment, vol. 18, p. 100390, Apr. 2024, doi:
10.1016/J.DIBE.2024.100390.

[118] M. Wang, R. Wang, H. Yao, S. Farhan, S. Zheng, and C. Du, “Study on the three dimensional
mechanism of graphene oxide nanosheets modified cement,” Constr Build Mater, vol. 126,
pp. 730–739, Nov. 2016, doi: 10.1016/J.CONBUILDMAT.2016.09.092.

[119] S. Li, O. Jensen, Z. Wang, Q. Y.-C. P. B. Engineering, and undefined 2021, “Influence of
micromechanical property on the rate-dependent flexural strength of ultra-high performance
concrete containing coarse aggregates (UHPC-CA),” Elsevier, Accessed: Feb. 14, 2024.
[Online]. Available: https://www.sciencedirect.com/science/article/pii/S1359836821007654

[120] A. Santos, M. do R. Veiga, … A. S.-C. and C., and undefined 2020, “Microstructure as a critical
factor of cement mortars’ behaviour: The effect of aggregates’ properties,” Elsevier, Accessed:

111
Feb. 14, 2024. [Online]. Available:
https://www.sciencedirect.com/science/article/pii/S0958946520301207

[121] S. Sharma and N. C. Kothiyal, “Influence of graphene oxide as dispersed phase in cement
mortar matrix in defining the crystal patterns of cement hydrates and its effect on mechanical,
microstructural and crystallization properties,” RSC Adv, vol. 5, no. 65, pp. 52642–52657, Jun.
2015, doi: 10.1039/C5RA08078A.

[122] X. Li et al., “Effects of graphene oxide agglomerates on workability, hydration, microstructure


and compressive strength of cement paste,” Constr Build Mater, vol. 145, pp. 402–410, Aug.
2017, doi: 10.1016/J.CONBUILDMAT.2017.04.058.

[123] A. Mohammed, J. G. Sanjayan, W. H. Duan, and A. Nazari, “Incorporating graphene oxide in


cement composites: A study of transport properties,” Constr Build Mater, vol. 84, pp. 341–
347, Jun. 2015, doi: 10.1016/J.CONBUILDMAT.2015.01.083.

[124] X. Kang, X. Zhu, J. Qian, J. Liu, and Y. Huang, “Effect of graphene oxide (GO) on hydration of
tricalcium silicate (C3S),” Constr Build Mater, vol. 203, pp. 514–524, Apr. 2019, doi:
10.1016/J.CONBUILDMAT.2019.01.117.

[125] S. Zhang, C. Zhang, L. Liao, and C. Wang, “Numerical study of the effect of ITZ on the failure
behaviour of concrete by using particle element modelling,” Constr Build Mater, vol. 170, pp.
776–789, May 2018, doi: 10.1016/J.CONBUILDMAT.2018.03.040.

[126] M. Maleki, I. Rasoolan, A. Khajehdezfuly, and A. P. Jivkov, “On the effect of ITZ thickness in
meso-scale models of concrete,” Constr Build Mater, vol. 258, p. 119639, Oct. 2020, doi:
10.1016/J.CONBUILDMAT.2020.119639.

[127] F. Xu, B. Tian, and G. Xu, “Influence of the ITZ Thickness on the Damage Performance of
Recycled Concrete,” Advances in Materials Science and Engineering, vol. 2021, no. 1, p.
6643956, Jan. 2021, doi: 10.1155/2021/6643956.

[128] C. E. Torrence, J. E. Trageser, R. E. Jones, and J. M. Rimsza, “Sensitivity of the strength and
toughness of concrete to the properties of the interfacial transition zone,” Constr Build Mater,
vol. 336, p. 126875, Jun. 2022, doi: 10.1016/J.CONBUILDMAT.2022.126875.

[129] L. Yeke and Z. Yu, “Effect of graphene oxide on mechanical properties of UHPC and analysis of
micro-control mechanism,” Mater Res Express, vol. 8, no. 9, p. 095001, Sep. 2021, doi:
10.1088/2053-1591/AC2015.

[130] J. Xu, B. Wang, and J. Zuo, “Modification effects of nanosilica on the interfacial transition zone
in concrete: A multiscale approach,” Cem Concr Compos, vol. 81, pp. 1–10, Aug. 2017, doi:
10.1016/J.CEMCONCOMP.2017.04.003.

[131] M. Nili and A. Ehsani, “Investigating the effect of the cement paste and transition zone on
strength development of concrete containing nanosilica and silica fume,” Mater Des, vol. 75,
pp. 174–183, Jun. 2015, doi: 10.1016/J.MATDES.2015.03.024.

[132] L. Zhao et al., “Mechanical behavior and toughening mechanism of polycarboxylate


superplasticizer modified graphene oxide reinforced cement composites,” Compos B Eng, vol.
113, pp. 308–316, Mar. 2017, doi: 10.1016/J.COMPOSITESB.2017.01.056.

112
[133] S. Jamnam et al., “Effect of graphene oxide nanoparticles on blast load resistance of steel
fiber reinforced concrete,” Constr Build Mater, vol. 343, p. 128139, Aug. 2022, doi:
10.1016/J.CONBUILDMAT.2022.128139.

[134] Y. Sui, S. Liu, C. Ou, Q. Liu, and G. Meng, “Experimental investigation for the influence of
graphene oxide on properties of the cement-waste concrete powder composite,” Constr Build
Mater, vol. 276, p. 122229, Mar. 2021, doi: 10.1016/J.CONBUILDMAT.2020.122229.

[135] Z. Chen, Y. Xu, J. Hua, X. Zhou, X. Wang, and L. Huang, “Modeling Shrinkage and Creep for
Concrete with Graphene Oxide Nanosheets,” Materials 2019, Vol. 12, Page 3153, vol. 12, no.
19, p. 3153, Sep. 2019, doi: 10.3390/MA12193153.

[136] B. H. Cho, B. H. Nam, and M. Khawaji, “Flexural fatigue behaviors and damage evolution
analysis of edge-oxidized graphene oxide (EOGO) reinforced concrete composites,” Cem Concr
Compos, vol. 122, p. 104082, Sep. 2021, doi: 10.1016/J.CEMCONCOMP.2021.104082.

[137] S. Wu, T. Qureshi, and G. Wang, “Application of Graphene in Fiber-Reinforced Cementitious


Composites: A Review,” Energies 2021, Vol. 14, Page 4614, vol. 14, no. 15, p. 4614, Jul. 2021,
doi: 10.3390/EN14154614.

[138] L. Shuguang and L. Qingbin, “Method of meshing ITZ structure in 3D meso-level finite element
analysis for concrete,” Finite Elements in Analysis and Design, vol. 93, no. C, pp. 96–106, Jan.
2015, doi: 10.1016/J.FINEL.2014.09.006.

[139] X. Zhao, Q. Dong, X. Chen, Y. Xiao, and D. Zheng, “Fatigue damage numerical simulation of
cement-treated base materials by discrete element method,” Constr Build Mater, vol. 276, p.
122142, Mar. 2021, doi: 10.1016/J.CONBUILDMAT.2020.122142.

[140] A. Habib, A. AL Houri, S. Al-Toubat, and M. T. Junaid, “Experimental Techniques for Testing the
Properties of Construction Materials,” May 2024, doi: 10.20944/PREPRINTS202405.1987.V1.

[141] K. Bhatrola, S. K. Maurya, R. Kaur, and N. C. Kothiyal, “Mechanical and electrical resistivity
performance of Pozzolana Portland cement mortar admixed graphene oxide,” Mater Today
Proc, vol. 78, pp. 830–838, Jan. 2023, doi: 10.1016/J.MATPR.2022.11.414.

[142] N. J. Edwards, Y. Lin, H. Du, and D. Ruan, “Effect of graphene oxide on cement mortar under
quasi-static and dynamic loading,” Journal of Building Engineering, vol. 74, p. 106783, Sep.
2023, doi: 10.1016/J.JOBE.2023.106783.

[143] S. Ganesh, C. Thambiliyagodage, S. V. T. J. Perera, and R. K. N. D. Rajapakse, “Influence of


Laboratory Synthesized Graphene Oxide on the Morphology and Properties of Cement
Mortar,” Nanomaterials, vol. 13, no. 1, p. 18, Jan. 2023, doi: 10.3390/NANO13010018/S1.

[144] A. Zanichelli, A. Carpinteri, C. Ronchei, D. Scorza, and S. Vantadori, “Fracture and mechanical
behaviour of a GO reinforced mortar,” Procedia Structural Integrity, vol. 47, pp. 37–42, Jan.
2023, doi: 10.1016/J.PROSTR.2023.06.038.

[145] K. Chintalapudi and R. M. R. Pannem, “Strength properties of graphene oxide cement


composites,” Mater Today Proc, vol. 45, pp. 3971–3975, Jan. 2021, doi:
10.1016/J.MATPR.2020.08.369.

[146] A. Abdullah, M. Taha, M. Rashwan, and M. Fahmy, “Efficient Use of Graphene Oxide and Silica
Fume in Cement-Based Composites,” Materials 2021, Vol. 14, Page 6541, vol. 14, no. 21, p.
6541, Oct. 2021, doi: 10.3390/MA14216541.

113
[147] F. Yang, J. Xie, W. Wang, W. Wang, and Z. Wang, “Effect of graphene oxide or triethanolamine-
modified graphene oxide on the hydration of calcium sulfoaluminate cement,” Constr Build
Mater, vol. 345, p. 128315, Aug. 2022, doi: 10.1016/J.CONBUILDMAT.2022.128315.

[148] C. Liu, X. Hunag, Y. Y. Wu, X. Deng, Z. Zheng, and B. Yang, “Studies on mechanical properties
and durability of steel fiber reinforced concrete incorporating graphene oxide,” Cem Concr
Compos, vol. 130, p. 104508, Jul. 2022, doi: 10.1016/J.CEMCONCOMP.2022.104508.

[149] D. Lu, X. Shi, and J. Zhong, “Nano-engineering the interfacial transition zone in cement
composites with graphene oxide,” Constr Build Mater, vol. 356, p. 129284, Nov. 2022, doi:
10.1016/J.CONBUILDMAT.2022.129284.

[150] S. Biswas and S. Mandal, “Mechanical and micro-structural study of cement mortar with
graphene oxide,” Journal of Building Pathology and Rehabilitation, vol. 7, no. 1, pp. 1–11, Dec.
2022, doi: 10.1007/S41024-022-00232-8/FIGURES/13.

[151] T. Susan Ja, M. Mathew, and S. C. George, “Experimental investigations on the impact of
graphene-based oxides in concrete,” Mater Today Proc, Aug. 2023, doi:
10.1016/J.MATPR.2023.08.071.

[152] J. Jiang, J. Qin, and H. Chu, “Improving mechanical properties and microstructure of ultra-
high-performance lightweight concrete via graphene oxide,” Journal of Building Engineering,
vol. 80, p. 108038, Dec. 2023, doi: 10.1016/J.JOBE.2023.108038.

[153] D. Mohotti, P. Mendis, K. Wijesooriya, I. Fonseka, D. Weerasinghe, and C. K. Lee, “Abrasion


and Strength of high percentage Graphene Oxide (GO) Incorporated Concrete,” Electronic
Journal of Structural Engineering, vol. 22, no. 01, pp. 37–43, May 2022, doi:
10.56748/ejse.2233001.

[154] R. Mohan, G. Selina Ruby, and I. Padmanaban, “Synthesis of Graphene Oxide and Study on
Strength Properties of Graphene Oxide in Cement Mortar,” AIP Conf Proc, vol. 2766, no. 1,
Jun. 2023, doi: 10.1063/5.0139351/2894995.

[155] K. Sreeja and T. Naresh Kumar, “Effect of graphene oxide on fresh, hardened and mechanical
properties of cement mortar,” Mater Today Proc, vol. 46, pp. 2235–2239, Jan. 2021, doi:
10.1016/J.MATPR.2021.03.574.

[156] P. V. R. K. Reddy and D. R. Prasad, “Investigation on the impact of graphene oxide on


microstructure and mechanical behaviour of concrete,” Journal of Building Pathology and
Rehabilitation, vol. 7, no. 1, pp. 1–10, Dec. 2022, doi: 10.1007/S41024-022-00166-
1/FIGURES/14.

[157] M. Somasri and B. Narendra Kumar, “Graphene oxide as Nano material in high strength self-
compacting concrete,” Mater Today Proc, vol. 43, pp. 2280–2289, Jan. 2021, doi:
10.1016/J.MATPR.2020.12.1085.

[158] B. Liu, L. Wang, G. Pan, and D. Li, “Dispersion of graphene oxide modified polycarboxylate
superplasticizer in cement alkali solution for improving cement composites,” Journal of
Building Engineering, vol. 57, p. 104860, Oct. 2022, doi: 10.1016/J.JOBE.2022.104860.

[159] G. Jing et al., “Enhanced Dispersion of Graphene Oxide in Cement Matrix with Isolated-
Dispersion Strategy,” Ind Eng Chem Res, vol. 59, no. 21, pp. 10221–10228, May 2020, doi:
10.1021/ACS.IECR.0C01230/SUPPL_FILE/IE0C01230_SI_001.PDF.

114
[160] J. Yan, J. Wang, H. Chen, and P. Xiang, “High Temperature Exposure Assessment of Graphene
Oxide Reinforced Cement,” Front Mater, vol. 9, p. 786260, Mar. 2022, doi:
10.3389/FMATS.2022.786260/BIBTEX.

[161] T. Shintani, S. Soeya, T. Saiki, W. Tian, J. Chai, and J. Cao, “Cement-based composites modified
by graphene oxide nano-materials: porosity and thermal conductivity,” J Phys Conf Ser, vol.
2553, no. 1, p. 012003, Aug. 2023, doi: 10.1088/1742-6596/2553/1/012003.

[162] A. Mohammed, N. T. K. Al-Saadi, and R. Al-Mahaidi, “Bond behaviour between NSM CFRP
strips and concrete at high temperature using innovative high-strength self-compacting
cementitious adhesive (IHSSC-CA) made with graphene oxide,” Constr Build Mater, vol. 127,
pp. 872–883, Nov. 2016, doi: 10.1016/J.CONBUILDMAT.2016.10.066.

[163] S. Han, M. S. Hossain, T. Ha, and K. K. Yun, “Graphene-oxide-reinforced cement composites


mechanical and microstructural characteristics at elevated temperatures,” Nanotechnol Rev,
vol. 11, no. 1, pp. 3174–3194, Jan. 2022, doi: 10.1515/NTREV-2022-
0495/ASSET/GRAPHIC/J_NTREV-2022-0495_FIG_012.JPG.

[164] H. yan Chu, J. yang Jiang, W. Sun, and M. Zhang, “Mechanical and thermal properties of
graphene sulfonate nanosheet reinforced sacrificial concrete at elevated temperatures,”
Constr Build Mater, vol. 153, pp. 682–694, Oct. 2017, doi:
10.1016/J.CONBUILDMAT.2017.07.157.

[165] E. Shamsaei, F. B. de Souza, X. Yao, E. Benhelal, A. Akbari, and W. Duan, “Graphene-based


nanosheets for stronger and more durable concrete: A review,” Constr Build Mater, vol. 183,
pp. 642–660, Sep. 2018, doi: 10.1016/J.CONBUILDMAT.2018.06.201.

[166] A. Sedaghat et al., “Investigation of Physical Properties of Graphene-Cement Composite for


Structural Applications,” Open Journal of Composite Materials, vol. 4, no. 1, pp. 12–21, Jan.
2014, doi: 10.4236/OJCM.2014.41002.

[167] Z. B. Abd, N. A. M. Habib, and A. Khammas, “Influence of Graphene Oxide on Thermal


Stability of Cement Mixture Nanocomposite,” AIP Conf Proc, vol. 2922, no. 1, Feb. 2024, doi:
10.1063/5.0183159/3265308.

[168] T. Janjaroen et al., “The Mechanical and Thermal Properties of Cement CAST
Mortar/Graphene Oxide Composites Materials,” Int J Concr Struct Mater, vol. 16, no. 1, pp. 1–
13, Dec. 2022, doi: 10.1186/S40069-022-00521-Z/TABLES/3.

[169] P. T. Rao, J. Prakash, R. Alexander, M. J. Shinde, and K. Dasgupta, “Role of graphene oxide
infusion in concrete to elevate strength and fire performance in construction concrete,” Diam
Relat Mater, vol. 147, p. 111269, Aug. 2024, doi: 10.1016/J.DIAMOND.2024.111269.

[170] L. Alarcon-Ruiz, G. Platret, E. Massieu, and A. Ehrlacher, “The use of thermal analysis in
assessing the effect of temperature on a cement paste,” Cem Concr Res, vol. 35, no. 3, pp.
609–613, Mar. 2005, doi: 10.1016/J.CEMCONRES.2004.06.015.

[171] S. K. Ur Rehman et al., “Influence of Graphene Nanosheets on Rheology, Microstructure,


Strength Development and Self-Sensing Properties of Cement Based Composites,”
Sustainability 2018, Vol. 10, Page 822, vol. 10, no. 3, p. 822, Mar. 2018, doi:
10.3390/SU10030822.

115
[172] J. Chen and L. Li, “Effect of oxidation degree on the thermal properties of graphene oxide,”
Journal of Materials Research and Technology, vol. 9, no. 6, pp. 13740–13748, Nov. 2020, doi:
10.1016/J.JMRT.2020.09.092.

[173] Y. Yang et al., “Thermal Conductivity of Defective Graphene Oxide: A Molecular Dynamic
Study,” Molecules 2019, Vol. 24, Page 1103, vol. 24, no. 6, p. 1103, Mar. 2019, doi:
10.3390/MOLECULES24061103.

[174] Y. Wang, J. Yang, and D. Ouyang, “Effect of Graphene Oxide on Mechanical Properties of
Cement Mortar and its Strengthening Mechanism,” Materials 2019, Vol. 12, Page 3753, vol.
12, no. 22, p. 3753, Nov. 2019, doi: 10.3390/MA12223753.

[175] A. Mohammed, J. G. Sanjayan, A. Nazari, and N. T. K. Al-Saadi, “Effects of graphene oxide in


enhancing the performance of concrete exposed to high-temperature,” Australian Journal of
Civil Engineering, vol. 15, no. 1, pp. 61–71, Jan. 2017, doi: 10.1080/14488353.2017.1372849.

[176] Y. Xu et al., “The role of graphene oxide on the hydration process and chemical shrinkage of
cement composites,” irsm.cas.czY Xu, B Li, J Zeng, R Jin, W Chen, D ZhuCeramics–Silikáty,
2020•irsm.cas.cz, vol. 64, no. 3, pp. 310–319, 2020, doi: 10.13168/cs.2020.0020.

[177] K. Guo, H. Miao, L. Liu, J. Zhou, and M. Liu, “Effect of graphene oxide on chloride penetration
resistance of recycled concrete,” Nanotechnol Rev, vol. 8, no. 1, pp. 681–689, May 2019, doi:
10.1515/NTREV-2019-0059/MACHINEREADABLECITATION/RIS.

[178] W. Li, X. Li, S. J. Chen, Y. M. Liu, W. H. Duan, and S. P. Shah, “Effects of graphene oxide on early-
age hydration and electrical resistivity of Portland cement paste,” Constr Build Mater, vol.
136, pp. 506–514, Apr. 2017, doi: 10.1016/J.CONBUILDMAT.2017.01.066.

[179] A. M. Maglad et al., “A Study on the Properties of Geopolymer Concrete Modified with Nano
Graphene Oxide,” Buildings 2022, Vol. 12, Page 1066, vol. 12, no. 8, p. 1066, Jul. 2022, doi:
10.3390/BUILDINGS12081066.

[180] Z. Lu et al., “Early-age interaction mechanism between the graphene oxide and cement
hydrates,” Constr Build Mater, vol. 152, pp. 232–239, Oct. 2017, doi:
10.1016/J.CONBUILDMAT.2017.06.176.

[181] S. Polverino, A. E. Del Rio Castillo, and F. Bonaccorso, “Few-Layer-Graphene Based Smart
Concrete: A New Paradigm in Construction Materials,” pp. 417–424, 2024, doi: 10.1007/978-
981-97-1972-3_45.

[182] J. L. Le, H. Du, and S. D. Pang, “Use of 2D Graphene Nanoplatelets (GNP) in cement
composites for structural health evaluation,” Compos B Eng, vol. 67, pp. 555–563, Dec. 2014,
doi: 10.1016/J.COMPOSITESB.2014.08.005.

[183] G. Qi, Q. Wang, R. Zhang, Z. Guo, D. Zhan, and S. Liu, “Effect of rGO/GNP on the electrical
conductivity and piezoresistance of cement-based composite subjected to dynamic loading,”
Constr Build Mater, vol. 368, p. 130340, Mar. 2023, doi:
10.1016/J.CONBUILDMAT.2023.130340.

[184] D. Lu, D. Wang, and J. Zhong, “Highly conductive and sensitive piezoresistive cement mortar
with graphene coated aggregates and carbon fiber,” Cem Concr Compos, vol. 134, p. 104731,
Nov. 2022, doi: 10.1016/J.CEMCONCOMP.2022.104731.

116
[185] W. Li, F. Qu, W. Dong, G. Mishra, and S. P. Shah, “A comprehensive review on self-sensing
graphene/cementitious composites: A pathway toward next-generation smart concrete,”
Constr Build Mater, vol. 331, p. 127284, May 2022, doi:
10.1016/J.CONBUILDMAT.2022.127284.

[186] X. Wang, Y. Wu, P. Zhu, and T. Ning, “Snow Melting Performance of Graphene Composite
Conductive Concrete in Severe Cold Environment,” Materials 2021, Vol. 14, Page 6715, vol.
14, no. 21, p. 6715, Nov. 2021, doi: 10.3390/MA14216715.

[187] V. Singh, D. Joung, L. Zhai, S. Das, S. I. Khondaker, and S. Seal, “Graphene based materials:
Past, present and future,” Prog Mater Sci, vol. 56, no. 8, pp. 1178–1271, Oct. 2011, doi:
10.1016/J.PMATSCI.2011.03.003.

[188] Y. Y. Wang, Y. Q. Tan, K. Liu, and H. N. Xu, “Preparation and electrical properties of conductive
asphalt concretes containing graphene and carbon fibers,” Constr Build Mater, vol. 318, p.
125875, Feb. 2022, doi: 10.1016/J.CONBUILDMAT.2021.125875.

[189] Y. Feng et al., “Conductivity Analysis of Carbon Black and Graphite Composites based on
Percolation Theory,” Proceedings of the 3rd IEEE International Conference on Micro/Nano
Sensors for AI, Healthcare and Robotics, NSENS 2024, pp. 120–123, 2024, doi:
10.1109/NSENS62142.2024.10561336.

[190] Z. Wu, J. Wei, R. Dong, and H. Chen, “Epoxy Composites with Reduced Graphene Oxide–
Cellulose Nanofiber Hybrid Filler and Their Application in Concrete Strain and Crack
Monitoring,” Sensors 2019, Vol. 19, Page 3963, vol. 19, no. 18, p. 3963, Sep. 2019, doi:
10.3390/S19183963.

[191] S. Bai, L. Jiang, Y. Jiang, M. Jin, S. Jiang, and D. Tao, “Research on electrical conductivity of
graphene/cement composites,” https://doi.org/10.1680/jadcr.16.00170, vol. 32, no. 2, pp.
45–52, Jan. 2020, doi: 10.1680/JADCR.16.00170.

[192] A. A. Balandin, “Thermal properties of graphene and nanostructured carbon materials,”


Nature Materials 2011 10:8, vol. 10, no. 8, pp. 569–581, Jul. 2011, doi: 10.1038/nmat3064.

[193] G. Goracci and J. S. Dolado, “Elucidation of Conduction Mechanism in Graphene


Nanoplatelets (GNPs)/Cement Composite Using Dielectric Spectroscopy,” Materials 2020, Vol.
13, Page 275, vol. 13, no. 2, p. 275, Jan. 2020, doi: 10.3390/MA13020275.

[194] H. Liu, A. Deshmukh, N. Salowitz, J. Zhao, and K. Sobolev, “Resistivity Signature of Graphene-
Based Fiber-Reinforced Composite Subjected to Mechanical Loading,” Front Mater, vol. 9, p.
818176, Jan. 2022, doi: 10.3389/FMATS.2022.818176/BIBTEX.

[195] A. Habibnejad Korayem, P. Ghoddousi, A. A. Shirzadi Javid, M. A. Oraie, and H. Ashegh,


“Graphene oxide for surface treatment of concrete: A novel method to protect concrete,”
Constr Build Mater, vol. 243, p. 118229, May 2020, doi:
10.1016/J.CONBUILDMAT.2020.118229.

[196] X. Li et al., “Effects of graphene oxide aggregates on hydration degree, sorptivity, and tensile
splitting strength of cement paste,” Compos Part A Appl Sci Manuf, vol. 100, pp. 1–8, Sep.
2017, doi: 10.1016/J.COMPOSITESA.2017.05.002.

[197] A. Mohammed, J. G. Sanjayan, W. H. Duan, and A. Nazari, “Graphene Oxide Impact on


Hardened Cement Expressed in Enhanced Freeze–Thaw Resistance,” Journal of Materials in

117
Civil Engineering, vol. 28, no. 9, p. 04016072, Mar. 2016, doi: 10.1061/(ASCE)MT.1943-
5533.0001586.

[198] J. He, S. Du, Z. Yang, and X. Shi, “Laboratory investigation of graphene oxide suspension as a
surface sealer for cementitious mortars,” Constr Build Mater, vol. 162, pp. 65–79, Feb. 2018,
doi: 10.1016/J.CONBUILDMAT.2017.12.022.

[199] J. Yu et al., “Insights on the capillary transport mechanism in the sustainable cement hydrate
impregnated with graphene oxide and epoxy composite,” Compos B Eng, vol. 173, p. 106907,
Sep. 2019, doi: 10.1016/J.COMPOSITESB.2019.106907.

[200] W. J. Long et al., “Investigation on chloride binding capacity and stability of Friedel’s salt in
graphene oxide reinforced cement paste,” Cem Concr Compos, vol. 132, p. 104603, Sep. 2022,
doi: 10.1016/J.CEMCONCOMP.2022.104603.

[201] T. Tong, Z. Fan, Q. Liu, S. Wang, S. Tan, and Q. Yu, “Investigation of the effects of graphene and
graphene oxide nanoplatelets on the micro- and macro-properties of cementitious materials,”
Constr Build Mater, vol. 106, pp. 102–114, Mar. 2016, doi:
10.1016/J.CONBUILDMAT.2015.12.092.

[202] Y. Lai, H. Zeng, S. Qu, and Y. Qin, “Graphene oxide-enhanced cementitious materials under
external sulfate attack: Implications for long structural life,” ACS Appl Nano Mater, vol. 3, no.
10, pp. 9784–9795, Oct. 2020, doi:
10.1021/ACSANM.0C01885/SUPPL_FILE/AN0C01885_SI_001.PDF.

[203] W. Piasta, “Analysis of carbonate and sulphate attack on concrete structures,” Eng Fail Anal,
vol. 79, pp. 606–614, Sep. 2017, doi: 10.1016/J.ENGFAILANAL.2017.05.008.

[204] K. G. Babu, K. V. Rao, and M. Duraisamy, “Sulphate attack on concrete,”


https://doi.org/10.1680/macr.1990.42.153.249, vol. 3, no. pt A, pp. 93–96, May 2015, doi:
10.1680/MACR.1990.42.153.249.

[205] C. Zhang, J. Li, M. Yu, Y. Lu, and S. Liu, “Mechanism and Performance Control Methods of
Sulfate Attack on Concrete: A Review,” Materials 2024, Vol. 17, Page 4836, vol. 17, no. 19, p.
4836, Sep. 2024, doi: 10.3390/MA17194836.

[206] L. Sabapathy, B. S. Mohammed, A. Al-Fakih, M. M. A. Wahab, M. S. Liew, and Y. H. M. Amran,


“Acid and Sulphate Attacks on a Rubberized Engineered Cementitious Composite Containing
Graphene Oxide,” Materials 2020, Vol. 13, Page 3125, vol. 13, no. 14, p. 3125, Jul. 2020, doi:
10.3390/MA13143125.

[207] X. Wang, G. Zhang, L. Liu, Y. Li, H. Kong, and C. Zhang, “Mechanical and fatigue properties of
graphene oxide concrete subjected to sulfate corrosion,” Front Mater, vol. 10, p. 1318366,
Dec. 2023, doi: 10.3389/FMATS.2023.1318366/BIBTEX.

[208] K. Chintalapudi and R. M. R. Pannem, “Enhanced chemical resistance to sulphuric acid attack
by reinforcing Graphene Oxide in Ordinary and Portland Pozzolana cement mortars,” Case
Studies in Construction Materials, vol. 17, p. e01452, Dec. 2022, doi:
10.1016/J.CSCM.2022.E01452.

[209] Y. Xu and Y. Fan, “Effects of Graphene Oxide Dispersion on Salt-Freezing Resistance of


Concrete,” Advances in Materials Science and Engineering, vol. 2020, no. 1, p. 4673739, Jan.
2020, doi: 10.1155/2020/4673739.

118
[210] R. Wang, Q. Zhang, and Y. Li, “Deterioration of concrete under the coupling effects of freeze–
thaw cycles and other actions: A review,” Constr Build Mater, vol. 319, p. 126045, Feb. 2022,
doi: 10.1016/J.CONBUILDMAT.2021.126045.

[211] R. Wang, Z. Hu, Y. Li, K. Wang, and H. Zhang, “Review on the deterioration and approaches to
enhance the durability of concrete in the freeze–thaw environment,” Constr Build Mater, vol.
321, p. 126371, Feb. 2022, doi: 10.1016/J.CONBUILDMAT.2022.126371.

[212] G. P. Millán Ramírez, H. Byliński, and M. Niedostatkiewicz, “Effectiveness of various types of


coating materials applied in reinforced concrete exposed to freeze–thaw cycles and
chlorides,” Scientific Reports 2023 13:1, vol. 13, no. 1, pp. 1–14, Aug. 2023, doi:
10.1038/s41598-023-40203-8.

[213] H.-S. Shang, T.-H. Yi, S. Chen, M. Jha, Q. Q. Liang, and E. Lui, “Freeze-Thaw Durability of Air-
Entrained Concrete,” The Scientific World Journal, vol. 2013, no. 1, p. 650791, Jan. 2013, doi:
10.1155/2013/650791.

[214] J. Gong, L. Lin, and S. Fan, “Modification of cementitious composites with graphene oxide and
carbon nanotubes,” SN Appl Sci, vol. 2, no. 9, pp. 1–16, Sep. 2020, doi: 10.1007/S42452-020-
03456-W/TABLES/8.

[215] H. Cai and X. Liu, “Freeze-thaw durability of concrete: ice formation process in pores,” Cem
Concr Res, vol. 28, no. 9, pp. 1281–1287, Sep. 1998, doi: 10.1016/S0008-8846(98)00103-3.

[216] J. Qian, L. Q. Zhou, X. Wang, and J. P. Yang, “Degradation of Mechanical Properties of


Graphene Oxide Concrete under Sulfate Attack and Freeze–Thaw Cycle Environment,”
Materials 2023, Vol. 16, Page 6949, vol. 16, no. 21, p. 6949, Oct. 2023, doi:
10.3390/MA16216949.

[217] H. Zeng, Y. Lai, S. Qu, and F. Yu, “Exploring the effect of graphene oxide on freeze–thaw
durability of air-entrained mortars,” Constr Build Mater, vol. 324, p. 126708, Mar. 2022, doi:
10.1016/J.CONBUILDMAT.2022.126708.

[218] A. Mohammed, J. Sanjayan, A. Nazari, N. T. K. Al-Saadi, and W. Duan, “Graphene Oxide as


Additive to Replace Using Air-Entraining Agents,” ACI Mater J, vol. 114, no. 6, Dec. 2017, doi:
10.14359/51700990.

[219] Y. Xiaoya, Z. Junjie, Zehai Q, and Jiawei N, “Effect of different water-reducing agents on
mechanical properties and microstructure of graphite oxide-blended cement mortar.,” Journal
of Functional Materials, vol. 49, no. 10, p. 10184, 2018, doi: 10.3969/j.issn.1001-
9371.2018.10.032.

[220] “XIII. On the atomic weight of graphite,” Philos Trans R Soc Lond, vol. 149, pp. 249–259, Dec.
1859, doi: 10.1098/RSTL.1859.0013.

[221] A. Sepehri and M. K. Ghassem Alaskari, “Simple Method of Exfoliation Multilayer Graphene
from Graphite Sheets,” SSRN Electronic Journal, Apr. 2018, doi: 10.2139/SSRN.3157265.

[222] W. S. Hummers and R. E. Offeman, “Preparation of Graphitic Oxide,” J Am Chem Soc, vol. 80,
no. 6, p. 1339, Mar. 1958, doi: 10.1021/JA01539A017/ASSET/JA01539A017.FP.PNG_V03.

[223] H. Yu, B. Zhang, C. Bulin, R. Li, and R. Xing, “High-efficient Synthesis of Graphene Oxide Based
on Improved Hummers Method,” Scientific Reports 2016 6:1, vol. 6, no. 1, pp. 1–7, Nov. 2016,
doi: 10.1038/srep36143.

119
[224] V. Purwandari, S. Gea, B. Wirjosentono, and A. Haryono, “Synthesis of graphene oxide from
the Sawahlunto-Sijunjung coal via modified hummers method,” AIP Conf Proc, vol. 2049, no.
1, Dec. 2018, doi: 10.1063/1.5082470/726403.

[225] A. M. Dimiev and J. M. Tour, “Mechanism of graphene oxide formation,” ACS Nano, vol. 8, no.
3, pp. 3060–3068, Mar. 2014, doi: 10.1021/NN500606A/SUPPL_FILE/NN500606A_SI_001.PDF.

[226] D. C. Marcano et al., “Improved synthesis of graphene oxide,” ACS Nano, vol. 4, no. 8, pp.
4806–4814, Aug. 2010, doi: 10.1021/NN1006368/ASSET/IMAGES/MEDIUM/NN-2010-
006368_0012.GIF.

[227] Y. Nishina, “Mass Production of Graphene Oxide Beyond the Laboratory: Bridging the Gap
Between Academic Research and Industry,” ACS Nano, Dec. 2024, doi:
10.1021/ACSNANO.4C13297/ASSET/IMAGES/LARGE/NN4C13297_0012.JPEG.

[228] J. Chen, Y. Li, L. Huang, C. Li, and G. Shi, “High-yield preparation of graphene oxide from small
graphite flakes via an improved Hummers method with a simple purification process,” Carbon
N Y, vol. 81, no. 1, pp. 826–834, Jan. 2015, doi: 10.1016/J.CARBON.2014.10.033.

[229] W. K. Park et al., “Facile synthesis of graphene oxide in a Couette–Taylor flow reactor,” Carbon
N Y, vol. 83, pp. 217–223, Mar. 2015, doi: 10.1016/J.CARBON.2014.11.024.

[230] S. Eigler et al., “Wet Chemical Synthesis of Graphene,” Advanced Materials, vol. 25, no. 26,
pp. 3583–3587, Jul. 2013, doi: 10.1002/adma.201300155.

[231] L. Peng et al., “An iron-based green approach to 1-h production of single-layer graphene
oxide,” Nature Communications 2015 6:1, vol. 6, no. 1, pp. 1–9, Jan. 2015, doi:
10.1038/ncomms6716.

[232] S. B. Norazman, W. L.-I. J. of, and undefined 2023, “Synthesis of Few Layers Graphene Sheets
from Graphite Rod via Electrochemical Exfoliation.,” ijneam.unimap.edu.mySNA Binti
Norazman, WW LiuInternational Journal of Nanoelectronics & Materials,
2023•ijneam.unimap.edu.my, Accessed: Dec. 14, 2024. [Online]. Available:
https://ijneam.unimap.edu.my/images/PDF/ijneam%20july%202023%20pdf/BOND21-NM-
023_Final%20CRC.pdf

[233] G. Bilgiç, “Reduced Graphene Synthesis via Eco-Friendly Electrochemical Exfoliation Method,”
Kastamonu University Journal of Engineering and Sciences, vol. 10, no. 1, pp. 38–43, Jun.
2024, doi: 10.55385/KASTAMONUJES.1477345.

[234] G. Konstantopoulos, E. Fotou, A. Ntziouni, K. Kordatos, and C. A. Charitidis, “A systematic


study of electrolyte effect on exfoliation efficiency and green synthesis of graphene oxide,”
Ceram Int, vol. 47, no. 22, pp. 32276–32289, Nov. 2021, doi:
10.1016/J.CERAMINT.2021.08.122.

[235] M. O. Ariff, W. W. Liu, and K. L. Foo, “Improvement Synthesis of Graphene Oxide Yield in Two
Steps of Intercalation and Oxidation of Flexible Graphite Foil by Electrochemical Exfoliation,”
International Journal of Nanoelectronics and Materials (IJNeaM) , vol. 17, no. 2, pp. 288–291,
May 2024, doi: 10.58915/IJNEAM.V17I2.719.

[236] W. Chen, M. J. Abedin, T. Barua, M. S. Mirshekarloo, S. El Meragawi, and M. Majumder,


“Customized Production of Holey Graphene Oxides via a Continuous Flow Process,” Small, vol.
20, no. 39, p. 2304227, Sep. 2024, doi: 10.1002/SMLL.202304227.

120
[237] B. D. L. Campéon, M. Akada, M. S. Ahmad, Y. Nishikawa, K. Gotoh, and Y. Nishina, “Non-
destructive, uniform, and scalable electrochemical functionalization and exfoliation of
graphite,” Carbon N Y, vol. 158, pp. 356–363, Mar. 2020, doi: 10.1016/J.CARBON.2019.10.085.

[238] X. Zhu et al., “Continuous and low-carbon production of biomass flash graphene,” Nature
Communications 2024 15:1, vol. 15, no. 1, pp. 1–11, Apr. 2024, doi: 10.1038/s41467-024-
47603-y.

[239] D.-N. Wu, J. Sheng, H.-G. Lu, S.-D. Li, and Y. Li, “100-Gram Batch Production of Graphene Using
High-Power Rapid Joule Heating Method,” Sep. 2024, doi: 10.26434/CHEMRXIV-2024-0C0LK.

[240] S. E. Lowe et al., “Scalable Production of Graphene Oxide Using a 3D-Printed Packed-Bed
Electrochemical Reactor with a Boron-Doped Diamond Electrode,” ACS Appl Nano Mater, vol.
2, no. 2, pp. 867–878, Feb. 2019, doi:
10.1021/ACSANM.8B02126/SUPPL_FILE/AN8B02126_SI_001.PDF.

[241] J. Zhang, Q. Liu, Y. Ruan, S. Lin, K. Wang, and H. Lu, “Monolithic Crystalline Swelling of
Graphite Oxide: A Bridge to Ultralarge Graphene Oxide with High Scalability,” Chemistry of
Materials, vol. 30, no. 6, pp. 1888–1897, Mar. 2018, doi:
10.1021/ACS.CHEMMATER.7B04458/SUPPL_FILE/CM7B04458_SI_001.PDF.

[242] P. Ranjan, S. Agrawal, A. Sinha, T. R. Rao, J. Balakrishnan, and A. D. Thakur, “A Low-Cost Non-
explosive Synthesis of Graphene Oxide for Scalable Applications,” Scientific Reports 2018 8:1,
vol. 8, no. 1, pp. 1–13, Aug. 2018, doi: 10.1038/s41598-018-30613-4.

[243] R. Ikram, B. M. Jan, and W. Ahmad, “An overview of industrial scalable production of
graphene oxide and analytical approaches for synthesis and characterization,” Journal of
Materials Research and Technology, vol. 9, no. 5, pp. 11587–11610, Sep. 2020, doi:
10.1016/J.JMRT.2020.08.050.

[244] “Large scale production of graphene oxide and its derivatives - The Graphene Council.”
Accessed: Dec. 14, 2024. [Online]. Available:
https://www.thegraphenecouncil.org/blogpost/1501180/381372/Large-scale-production-of-
graphene-oxide-and-its-
derivatives?hhSearchTerms=%22%22Concrete%22+and+%22Graphene+oxide%22%22&terms
=

[245] “Share Price & ASX Announcements - First Graphene.” Accessed: Dec. 14, 2024. [Online].
Available: https://firstgraphene.net/investors/share-price-asx-announcements/

[246] “Taking nanotechnology from lab to market: a graphene oxide case study - Chemical Industry
Journal.” Accessed: Dec. 14, 2024. [Online]. Available:
https://www.chemicalindustryjournal.co.uk/taking-nanotechnology-from-lab-to-market-a-
graphene-oxide-case-study

[247] “Graphenea opens new graphene oxide pilot plant.” Accessed: Dec. 14, 2024. [Online].
Available: https://www.graphenea.com/blogs/graphene-news/graphenea-opens-new-
graphene-oxide-pilot-plant

[248] “About Us - The Sixth Element (Changzhou) Materials Technology Co.,Ltd.” Accessed: Dec. 14,
2024. [Online]. Available: https://www.c6th.com/about-us

121
[249] “Towards a sustainable construction - Graphenemex.” Accessed: Dec. 14, 2024. [Online].
Available: https://www.graphenemex.com/en/solutions-with-graphene/graphene-
oxide/concrete-additive/construction-industry-go/towards-a-sustainable-construction/

[250] “Graphene@Manchester solves concrete’s big problem.” Accessed: Dec. 14, 2024. [Online].
Available: https://www.manchester.ac.uk/about/news/greener-and-cheaper-
graphenemanchester-solves-concretes-big-problem/

[251] Y. Zhu, H. Ji, H. M. Cheng, and R. S. Ruoff, “Mass production and industrial applications of
graphene materials,” Natl Sci Rev, vol. 5, no. 1, pp. 90–101, Jan. 2018, doi:
10.1093/NSR/NWX055.

[252] “The graphene additive for concrete - Graphenemex.” Accessed: Dec. 14, 2024. [Online].
Available: https://www.graphenemex.com/en/solutions-with-graphene/graphene-
oxide/concrete-additive/construction-industry-go/the-graphene-additive-for-concrete/

[253] “Concretene - Home.” Accessed: Dec. 14, 2024. [Online]. Available:


https://www.concretene.co.uk/concretene-home

[254] B. H. Cho and B. H. Nam, “Concrete composites reinforced with graphene oxide nanoflake
(GONF) and steel fiber for application in rigid pavement,” Case Studies in Construction
Materials, vol. 17, p. e01346, Dec. 2022, doi: 10.1016/J.CSCM.2022.E01346.

[255] “Utilizing Graphene in Concrete Production And Manufacture - Company news - News - The
Sixth Element (Changzhou) Materials Technology Co.,Ltd.” Accessed: Dec. 14, 2024. [Online].
Available: https://www.c6th.com/news/w-56283343.html

[256] “Graphene oxide study strengthens the case for smart concrete - RMIT University.” Accessed:
Dec. 14, 2024. [Online]. Available: https://www.rmit.edu.au/news/all-
news/2023/dec/graphene-oxide-concrete

[257] “Mechanically reinforced composites & concrete — LayerOne Advanced Materials.” Accessed:
Dec. 14, 2024. [Online]. Available: https://www.layeronematerials.com/mechanically-
reinforced-composites-and-concrete

[258] “Alterbiota – We make the construction industry greener.” Accessed: Dec. 14, 2024. [Online].
Available: https://alterbiota.com/

[259] “Graphene News and Updates - The Graphene Council.” Accessed: Dec. 14, 2024. [Online].
Available: https://www.thegraphenecouncil.org/blogpost/1501180/Graphene-News-and-
Updates?tag=Bio+Graphene+Solutions

[260] “HOME - Bio Graphene Solutions.” Accessed: Dec. 14, 2024. [Online]. Available:
https://biographenesolutions.com/

[261] J. Liu, P. Tran, T. Ginigaddara, and P. Mendis, “Exploration of using graphene oxide for strength
enhancement of 3D-printed cementitious mortar,” Additive Manufacturing Letters, vol. 7, p.
100157, Dec. 2023, doi: 10.1016/J.ADDLET.2023.100157.

[262] O. Zaid, S. R. Z. Hashmi, F. Aslam, Z. U. Abedin, and A. Ullah, “Experimental study on the
properties improvement of hybrid graphene oxide fiber-reinforced composite concrete,”
Diam Relat Mater, vol. 124, p. 108883, Apr. 2022, doi: 10.1016/J.DIAMOND.2022.108883.

122
[263] K. Rajesh, K. R. SEKAR, V. RAMASWAMY, K. KANNAN, and R. MANIKANDAN, “Demand and
Supply Analysis on Manufacturing of Cement for Monsoon Periods Using Dvorak Technique,”
International Journal of Mechanical and Production Engineering Research and Development
(IJMPERD) ISSN (P): 2249-6890; ISSN (E): 2249-8001 Vol, vol. 7, pp. 187–194, 2017.

[264] L. Eddy et al., “Automated Laboratory Kilogram-Scale Graphene Production from Coal,” Small
Methods, vol. 8, no. 3, p. 2301144, Mar. 2024, doi: 10.1002/SMTD.202301144.

[265] S. Mathan, M. Selvaraj, M. A. Assiri, K. Kandiah, and R. Rajendran, “Synthetic


nanoarchitectonics with ultrafast Joule heating of graphene-based electrodes for high energy
density supercapacitor application,” Surfaces and Interfaces, vol. 51, p. 104707, Aug. 2024,
doi: 10.1016/J.SURFIN.2024.104707.

[266] H. Lv, M. Du, Z. Li, L. Xiao, and S. Zhou, “Cost Optimization of Graphene Oxide-Modified Ultra-
High-Performance Concrete Based on Machine Learning Methods,” Inorganics 2024, Vol. 12,
Page 181, vol. 12, no. 7, p. 181, Jun. 2024, doi: 10.3390/INORGANICS12070181.

[267] “Graphene Oxide Market Size | Growth Analysis 2036.” Accessed: Dec. 14, 2024. [Online].
Available: https://www.researchnester.com/reports/graphene-oxide-market/3269

[268] “Graphene-based Concrete Additives Market Share & Size | 2031.” Accessed: Dec. 14, 2024.
[Online]. Available: https://www.transparencymarketresearch.com/graphene-based-concrete-
additives-market.html

[269] R. B. Meshram and S. Kumar, “Comparative life cycle assessment (LCA) of geopolymer cement
manufacturing with Portland cement in Indian context,” International Journal of
Environmental Science and Technology, vol. 19, no. 6, pp. 4791–4802, Jun. 2022, doi:
10.1007/S13762-021-03336-9/TABLES/5.

[270] G. Habert et al., “Environmental impacts and decarbonization strategies in the cement and
concrete industries,” Nature Reviews Earth & Environment 2020 1:11, vol. 1, no. 11, pp. 559–
573, Sep. 2020, doi: 10.1038/s43017-020-0093-3.

[271] G. Mishra, A. Warda, and S. P. Shah, “Carbon sequestration in graphene oxide modified
cementitious system,” Journal of Building Engineering, vol. 62, p. 105356, Dec. 2022, doi:
10.1016/J.JOBE.2022.105356.

[272] W. J. Long, D. Zheng, H. bo Duan, N. Han, and F. Xing, “Performance enhancement and
environmental impact of cement composites containing graphene oxide with recycled fine
aggregates,” J Clean Prod, vol. 194, pp. 193–202, Sep. 2018, doi:
10.1016/J.JCLEPRO.2018.05.108.

[273] M. Cossutta, J. McKechnie, and S. J. Pickering, “A comparative LCA of different graphene


production routes,” Green Chemistry, vol. 19, no. 24, pp. 5874–5884, Dec. 2017, doi:
10.1039/C7GC02444D.

[274] I. Papanikolaou, N. Arena, and A. Al-Tabbaa, “Graphene nanoplatelet reinforced concrete for
self-sensing structures – A lifecycle assessment perspective,” J Clean Prod, vol. 240, p.
118202, Dec. 2019, doi: 10.1016/J.JCLEPRO.2019.118202.

[275] R. Arvidsson, D. Kushnir, B. A. Sandén, and S. Molander, “Prospective life cycle assessment of
graphene production by ultrasonication and chemical reduction,” Environ Sci Technol, vol. 48,

123
no. 8, pp. 4529–4536, Apr. 2014, doi:
10.1021/ES405338K/SUPPL_FILE/ES405338K_SI_001.PDF.

[276] S. Pei, Q. Wei, K. Huang, H. M. Cheng, and W. Ren, “Green synthesis of graphene oxide
by seconds timescale water electrolytic oxidation,” Nature Communications 2018 9:1, vol. 9,
no. 1, pp. 1–9, Jan. 2018, doi: 10.1038/s41467-017-02479-z.

[277] C. Jia et al., “Graphene environmental footprint greatly reduced when derived from biomass
waste via flash Joule heating,” One Earth, vol. 5, no. 12, pp. 1394–1403, Dec. 2022, doi:
10.1016/J.ONEEAR.2022.11.006/ATTACHMENT/5BF9837E-61A6-4EEB-A7A5-
7593EA7627E3/MMC2.PDF.

[278] W. J. Long, D. Zheng, H. bo Duan, N. Han, and F. Xing, “Performance enhancement and
environmental impact of cement composites containing graphene oxide with recycled fine
aggregates,” J Clean Prod, vol. 194, pp. 193–202, Sep. 2018, doi:
10.1016/J.JCLEPRO.2018.05.108.

[279] K. Wang et al., “Biocompatibility of Graphene Oxide,” Nanoscale Res Lett, vol. 6, no. 1, pp. 1–
8, Aug. 2011, doi: 10.1007/S11671-010-9751-6/FIGURES/8.

[280] C. Liao, Y. Li, and S. C. Tjong, “Graphene Nanomaterials: Synthesis, Biocompatibility, and
Cytotoxicity,” International Journal of Molecular Sciences 2018, Vol. 19, Page 3564, vol. 19,
no. 11, p. 3564, Nov. 2018, doi: 10.3390/IJMS19113564.

[281] L. Ou et al., “Toxicity of graphene-family nanoparticles: A general review of the origins and
mechanisms,” Part Fibre Toxicol, vol. 13, no. 1, Oct. 2016, doi: 10.1186/S12989-016-0168-Y.

124

You might also like