1. Introduction
Concrete is the second most used material in the world after water, due to its high demand in the construction industry [
1]. Its global consumption reaches approximately 30 billion metric tons annually, highlighting its fundamental role in the development of infrastructure worldwide [
2].
Conventional concrete is widely used in construction due to its versatility and ease of production. However, it is susceptible to crack formation, especially in massive structures, due to thermal stresses generated by cement hydration, which can compromise its structural integrity [
3]. These cracks can originate both from internal stresses and external restraints during the hardening process. On the other hand, porosity is a determining factor in the mechanical behavior of concrete, since it influences its physical and structural properties [
4]. In pervious concrete, it has been shown that there is a relationship between the propagation velocity of ultrasonic waves and porosity, which allows its non-destructive evaluation and highlights the importance of controlling this parameter in the design and performance of the material [
5].
As a fundamental composite material in modern construction, it has been the subject of numerous investigations to improve its mechanical properties and durability; in the search for high-performance materials, nanotechnology has emerged as a promising solution, offering the possibility of modifying concrete at the micro- and nanoscale level to optimize its characteristics [
6,
7,
8]. In this regard, the search for building materials with improved properties has led to the exploration of various nano-additions in concrete [
9,
10,
11]. One of the most promising nanomaterials is graphene oxide (GO), widely studied due to its exceptional electrical, optical, chemical, and mechanical properties, which make it attractive for various scientific and technological applications [
12]. GO is a two-dimensional nanoplatelet derived from graphene, distinguished by the presence of various oxygen functional groups in its structure, which facilitates its dispersion in water and thus in the concrete matrix [
9,
10,
13,
14,
15]. Its high mechanical strength, large surface area, and good thermal conductivity [
9], as well as its high electrical conductivity and thermal stability, make it a promising material for the construction industry [
16].
Graphene oxide (GO) has proven to be a promising solution to optimize the properties of concrete, improving its densification at the nanometer level, which increases its homogeneity and quality, reflected in better mechanical and durability characteristics [
10,
17]. In particular, the addition of GO has been found to significantly improve the compressive strength, flexural strength, and durability of concrete [
8]. Within carbon-based nanomaterials, GO stands out as an effective alternative to enhance the mechanical performance of cementitious composites, contributing to the increase in their toughness and strength [
18]. Furthermore, it has been observed that the incorporation of nanomaterials such as graphene oxide (GO) has shown significant improvements in the microstructure of concrete. The addition of GO contributes to the inhibition of crack propagation and acts as a nucleation agent in the formation of calcium silicate hydrates (CSHs), which favor the densification of the cementitious matrix and improve its mechanical and durability properties [
19].
Porosity reduction is another key factor in the performance of GO-modified concrete. It has been observed that the increase in GO concentration decreases the number of pores in the cementitious matrix, promoting a more compact and less permeable material. This effect is due to the acceleration in the nucleation and growth of cement hydration products, which improves the mechanical strength and reduces the permeability of the structure [
10,
20,
21]. These findings show that the addition of GO in cementitious composites not only optimizes the mechanical strength but also improves the microstructure of the material, increasing its durability and performance against external agents. In addition, the maintenance of concrete structures represents a considerable economic challenge, so the incorporation of reinforcing materials, such as GO, is a viable strategy to improve their properties and reduce costs in the long term [
22,
23].
The incorporation of GO in concrete has shown significant improvements in several of its mechanical properties, such as compressive strength [
9,
10,
24], flexural strength [
9,
24,
25], fracture toughness [
26], modulus of elasticity [
9,
24,
27], and durability [
24,
28]. These improvements are attributed to several mechanisms such as the filling effect of the nanopores by GO, which densifies the microstructure [
7,
24,
26]; the acceleration of cement hydration, which increases the formation of hydrated products such as calcium silicate hydrate (C-S-H) [
7,
9,
24,
27,
29,
30,
31]; and the ability of GO to act as a bridge that inhibits the propagation of microcracks [
7,
9,
26].
Although GO exhibits excellent properties, its high surface energy makes it prone to particle agglomeration during both preparation and application. Additionally, the oxygen-containing functional groups in GO readily cross-link with Ca
2+ ions in the cement matrix, leading to granulation and aggregation, which significantly limits its effectiveness [
32]. To solve this problem, a variety of methods have been proposed to improve GO dispersion, including ultrasonic dispersion [
33,
34] and adding surfactants [
25,
35,
36,
37,
38,
39,
40].
Previous studies have confirmed that the cracking of mesoscale concrete is controlled by its micromechanical properties, which are directly dependent on the chemical composition and microstructure of the material [
26]. In this context, graphene oxide (GO) has captured the attention of the scientific community due to its exceptional properties, including its high mechanical strength, large specific surface area, and ability to enhance cement hydration [
24,
41,
42,
43,
44,
45,
46,
47,
48,
49].
This study investigates the influence of graphene oxide (GO) on the mechanical and microstructural properties of concrete with a compressive strength of f’c = 280 kg/cm2. Unlike previous studies that primarily focus on mechanical improvements, this research integrates advanced analytical techniques, including scanning electron microscopy (SEM) and energy-dispersive X-ray spectroscopy (EDS), X-ray diffraction (XRD), Fourier Transform Infrared Spectroscopy (FTIR), and thermogravimetric analysis (TGA), to establish a direct correlation between mechanical performance and microstructural modifications induced by GO. Additionally, it determines the optimal GO dosage to enhance mechanical properties while maintaining workability, a crucial factor for practical implementation. These findings provide new insights into the role of GO in cementitious materials, contributing to the development of durable and sustainable high-performance concrete.
2. Materials and Methods
The present study is framed within quantitative research, according to its approach, and applied, according to its purpose. The research level is explanatory since it seeks to analyze and describe the influence of the incorporation of graphene oxide on the mechanical properties and microstructural characteristics of concrete with f’c = 280 kg/cm2.
The methodology used is based on the Direct Observation Technique, allowing the recording and detailed analysis of the behavior and characteristics of the samples. For data collection, an Observation Guide was used, designed with specific criteria that ensure objectivity and accuracy in the collection of relevant information for the study.
The study population consists of 155 samples, including cylindrical specimens and prismatic beams, made with concrete compressive strength of f’c = 280 kg/cm
2 and with the addition of graphene oxide (GO) in different proportions with respect to the weight of cement. The aggregates used come from quarries located in the department of Ica, Peru. The samples are distributed in five groups according to the percentage of GO addition: 0.00% (reference concrete), 0.05%, 0.10%, 0.15%, and 0.20%. The curing process was carried out at 7, 14, and 28 days in order to evaluate the evolution of the material properties. Furthermore,
Figure 1 presents the schematic diagram of the experimental study, illustrating the key stages of the research methodology. It outlines the preparation of materials, including GO dispersion, the mixing and curing of concrete specimens, and the mechanical and microstructural characterization techniques employed. This diagram provides a clear overview of the experimental approach, ensuring clarity and reproducibility.
2.1. Characterization of Graphene Oxide: Diameter Size, Thickness, and Chemical Composition
The graphene oxide used in this research was purchased from Jiangsu Xfnano Materials Tech Co., Ltd., Nanjing, China, a well-known Chinese supplier of nanomaterials. According to the product data sheet, GO presents specific physical and chemical properties, which will be complemented with thermogravimetric analysis (TGA) for its detailed characterization.
Table 1 compares the properties of the GO used in this study with those reported in previous research, allowing us to contextualize its characteristics according to existing studies.
After receipt of the material, the four percentages of graphene oxide (GO) selected for the research were subjected to a prior dispersion process, with the aim of ensuring a homogeneous distribution within the cementitious matrix and optimizing its interaction with the concrete components.
2.2. Graphene Oxide Dispersion Process
The graphene oxide acquired is composed of sheets or layers, so its adequate dispersion is essential for its incorporation into the concrete matrix. For this purpose, it was initially subjected to a magnetic stirring process, with the purpose of obtaining a homogeneous aqueous solution, thus ensuring a uniform distribution before its addition to the cementitious mixture in a mixer [
21]. The correct dispersion of GO depends on key factors such as the amount added, the degree of oxidation, the particle size, and the method of dispersion [
50], which directly influences the distribution of nanoparticles within the concrete. Since GO contains functional groups that can affect the interaction between its layers, its dispersion was performed in water and optimized by applying ultrasound waves, reducing the Van Der Waals force and improving its integration into the mixture [
6]. In this context,
Table 2 presents the bulk density of the GO used in this research in comparison with previous studies, while
Table 3 details the degree of purity of the GO acquired and its relation to other similar investigations.
Likewise, after the dispersion of the GO to keep the nanoparticles separated, a surfactant must be added, which, in the case of concrete, are plasticizers and superplasticizers, because several studies showed that the GO decreases the workability of concrete, as shown in
Table 4; in this vein, other studies determined suitable percentages of plasticizer for concrete with GO, as shown in
Table 5, so that through various field tests, the percentages of the addition of plasticizer Sika-Cem by % GO were determined.
The incorporation of a plasticizer in the graphene oxide (GO) dispersion process serves two complementary mechanisms that are fundamental to the successful modification of concrete. First, the plasticizer acts as a surfactant, preventing the re-agglomeration of GO nanosheets after ultrasonic dispersion by maintaining particle separation through steric and electrostatic repulsion mechanisms. Second, it counteracts the adverse effect of GO on the workability of fresh concrete. As demonstrated in the preliminary tests (
Table 6), concrete with 0.20% GO without a plasticizer exhibited a slump of only 1 cm, whereas, with the addition of 1.10% plasticizer, the slump increased to 7.8 cm, restoring the necessary workability for proper placement and compaction. This phenomenon is explained by the hydrophilic nature of GO, whose numerous oxygen functional groups (hydroxyl, carboxyl, carbonyl, and epoxy) exhibit a high affinity for water molecules, significantly reducing the free water available for mixture lubrication [
29]. By lowering the surface tension of water and improving cement dispersion, the plasticizer mitigates this negative effect, enabling the benefits of GO on mechanical properties to be fully harnessed without compromising workability.
The following procedure describes the process of GO dispersion in water, following the guidelines recommended by Jiangsu Xfnano Materials Tech Co., Ltd., as well as references from several studies.
- -
Preparation of the materials:
Initially, the GO was weighed according to the weight of the cement, and the volume of water corresponding to the mix design was measured, leaving a margin for the prior wetting of the materials in the mixer before the incorporation of the aqueous solution of GO. In this study, we worked with a concentration of 5 mg/mL, recommended by the manufacturer.
Table 7 shows the concentrations of GO in water used in different investigations.
- -
Initial mixture:
GO and water were combined in a container and subjected to agitation in a magnetic stirrer at 2500 rpm for 2.5 min, reaching a total of 6250 revolutions. This procedure aligned with previous studies, as shown in
Table 8, and seeks to ensure a homogeneous predispersion of the solution, as shown in
Figure 2a.
- -
Ultrasonication:
Subsequently, the aqueous GO solution was subjected to an ultrasonication process with a power of 240 W, as seen in
Figure 2b. The power recommended by the manufacturer is 700 W, while the literature reports sonication times of approximately 30 min, as shown in
Table 9. However, considering the findings of [
25], we chose to apply only 30% of the maximum power to avoid damage to the GO nanoparticles.
The total power required was calculated as follows:
Based on the power of the equipment purchased (240 W), the sonication time was determined:
Since the ultrasound equipment only allows adjustments in whole minutes, the time was rounded to 27 min.
- -
Addition of plasticizer:
Finally, once the dispersion of graphene oxide in water was completed, the Sika-Cem plasticizer was incorporated, as shown in
Figure 2c, in order to avoid the agglomeration of the GO nanoparticles and improve the workability of the concrete. According to [
25], the addition of the plasticizer at the end of the dispersion process, i.e., after 30 min, favors better stability of the aqueous GO solution compared to its incorporation in intermediate stages (15 min) or at the beginning of the process.
This procedure guarantees an efficient dispersion of GO in the mixture, optimizing its incorporation into the concrete and maximizing its effect on the mechanical and rheological properties of the material.
2.3. Aggregates, Cement, and Additions
Experimental laboratory tests were carried out to characterize the behavior of the reference concrete with a compressive strength of f’c = 280 kg/cm
2. Additionally, tests were carried out with the addition of (GO) in proportions of 0.05%, 0.10%, 0.15%, and 0.20% with respect to the weight of cement. The selection of GO concentrations (0.00%, 0.05%, 0.10%, 0.15%, and 0.20%) is based on a comprehensive review of previous studies that have reported improvements in compressive strength within this range, with optimal values generally found between 0.05% and 0.15%, depending on the cementitious matrix used, as evidenced in
Table 10. The inclusion of 0.00% serves as a reference without GO, while the incorporation of progressive concentrations allows for the evaluation of trends in mechanical property development and the identification of potential optimum content. Furthermore, this selection enhances comparability with previous research, strengthening the analysis and discussion of the results within a broader scientific context.
Mechanical tests were carried out in the laboratory in accordance with the corresponding ASTM standards, using standardized specimens distributed at different curing ages to evaluate the performance of concrete with graphene oxide (GO) addition. The compressive strength (
Figure 3a) was determined in 50 cylindrical specimens of 10 cm × 20 cm, organized in a series of 3 specimens tested at 7 days, 3 at 14 days, and 4 at 28 days for each percentage of GO, following ASTM C39/C39M-18 [
66]. The indirect tensile strength by diametral (
Figure 3b) compression was evaluated in 35 cylindrical specimens of 10 cm × 20 cm, distributed in a series of 3 specimens tested at 7 days and 4 at 28 days for each GO dosage, according to ASTM C496/C496M-17 [
67]. For flexural strength (
Figure 3c), 35 prismatic beams of 15 cm × 15 cm × 51 cm were used, divided into 3 specimens tested at 7 days and 4 at 28 days for each percentage of GO, in accordance with ASTM C293/C293M-16 [
68]. Finally, the evaluation of the modulus of elasticity (
Figure 3d) was carried out on 35 cylindrical specimens of 10 cm × 30 cm, with a series of 3 specimens tested at 7 days and 4 at 28 days for each GO proportion, following ASTM C469/C469M [
69].
Aggregates were obtained from the Palomino-Parcona quarry, located in the department of Ica (
Table 11). The cement used was Sol Type I.
Table 12 presents the specific amounts of GO, aggregates, cement, water, and additions used in the study.
From the identification of the optimum %, samples of the hardened concrete were taken for testing and determination of the microstructural characteristics of the reference concrete contrasted with the concrete with the optimum % of graphene oxide.
4. Discussion
Several studies have investigated the effect of graphene oxide (GO) addition, relative to cement weight, on the mechanical and microstructural properties of concrete. Ref. [
32] reported increases of 20.1%, 29.8%, and 26.2% in compressive, flexural, and tensile strength, respectively, with a 0.06% GO dosage. Additionally, the enhanced hydration of C
2S and C
2S, reduced pore size, and the formation of denser hydration products were observed through SEM analysis. Similarly, Ref. [
51] recorded improvements of 16.04%, 30.50%, 23.40%, and 12.95% in compressive, tensile, and flexural strength, as well as in the elastic modulus, with 0.06% GO, along with a 12.99% reduction in porosity. Likewise, Ref. [
24] reported increases of 22.36%, 27.44%, 23.88%, and 22.47% in the same properties, accompanied by a reduction in microcracks and increased C-S-H gel formation.
GO is a nanomaterial with a high specific surface area, which promotes agglomeration and affects its dispersion in fluid media [
75]. Its dispersion is influenced by the Peclet number (Pe), which improves as the GO sheet diameter increases, as considered in this study (20 µm) [
76]. GO contains oxygen functional groups (hydroxyl, epoxy, carboxyl, and carbonyl), which facilitate hydration and the formation of cementitious products [
53,
77]. Furthermore, the mass loss of GO occurs in three stages: water evaporation (110–130 °C), the pyrolysis of oxygen-containing groups (~450 °C), and carbon decomposition (>600 °C) [
78,
79].
GO enhances the mechanical performance of concrete, but its effect is dosage-dependent. Ref. [
32] reported a 20.1% increase in compressive strength with 0.05% GO but a reduction at 0.07%. Ref. [
7] observed improvements of 23%, 48%, and 57% with 0.01%, 0.03%, and 0.05% GO, respectively. In this study, a 14.61% increase was recorded with 0.05% GO. Regarding indirect tensile strength, some studies [
32,
51,
57] have shown improvements ranging from 22.66% to 27.44%, whereas this study reported a 12.33% increase. For flexural strength, Refs. [
7,
64] documented enhancements of 41% and 48%, respectively, while this study achieved a 6.09% increase. Finally, the elastic modulus increased by 27.38%, surpassing the values reported in [
24,
51].
FTIR analysis confirmed the formation of calcium carbonate (CaCO
3) at bands between 1402 and 1425 cm
−1 [
11,
80], as well as silicates at bands between 943 and 997 cm
−1 [
29,
81]. The observed pore size reduction in this study (16.27%) was lower compared to the values reported in [
32,
51], which reached up to 39.88%. The Ca/Si ratio increased in the GO-modified sample, consistent with the findings in [
10,
25]. SEM analysis in [
32,
75] revealed denser and more crystalline microstructures in GO-modified concrete, aligning with the results obtained in this study. Lastly, the thermal degradation behavior of GO-modified concrete exhibited consistency with previous studies, indicating mass losses in three stages: water evaporation (<200 °C), portlandite decomposition (~444 °C), and decarbonation (>700 °C) [
7,
82,
83].
Microstructural analyses (SEM, EDS, FTIR, TGA, and XRD) in this study revealed that incorporating 0.05% GO promotes a denser and more homogeneous microstructure, as evidenced by a 16.28% reduction in average pore and crack size. XRD analysis showed a 26.31% increase in silicon compounds, indicating greater C-S-H formation, while portlandite content decreased by 2.06%, confirming a more complete and efficient hydration process. These findings were supported by FTIR, which showed increased absorption in silicate- and carbonate-related bands, highlighting the higher reactivity of the cementitious matrix with GO. Consequently, mechanical properties improved significantly, with 14.61% higher compressive strength, 12.33% higher indirect tensile strength, 6.09% higher flexural strength, and a 27.38% increase in elastic modulus, suggesting that GO not only enhances late-stage hydration but also reinforces the strength and durability of concrete. In agreement with these results, Ref. [
29] analyzed early hydration (1 to 3 days) using TGA, DTG, and SEM, reporting that at 24 h, GO-modified samples exhibited a slight reduction in portlandite (Ca(OH)
2) content compared to the control, attributed to the initial absorption of water and Ca
2+ by GO, temporarily limiting their availability for hydration product formation. However, this effect was reversed as GO acted as a water reservoir, facilitating a more efficient hydration process. By day 3, compressive strength significantly increased by 24.7%, 30.7%, and 42.3% for GO contents of 0.1%, 0.15%, and 0.20%, respectively, a trend correlated with greater C-S-H formation, confirmed by thermal analysis and electron microscopy. Furthermore, EDS and XRD results confirm that GO primarily acts as a nucleation site for C-S-H formation, as evidenced by the increase in silicon compounds and the reduction in portlandite, suggesting a more efficient conversion of Ca(OH)
2 into hydrated products. This effect has been previously reported in studies such as [
29], where GO was found to promote the nucleation and growth of C-S-H without altering its chemical structure. Additionally, TGA and DTG analyses demonstrated that GO modifies hydration dynamics by influencing the availability of water and Ca
2+, which can temporarily reduce portlandite content in the first 24 h but subsequently accelerate its transformation into C-S-H. Collectively, these findings indicate that GO not only acts as a nucleation center but also modulates the chemical reactions involved in portlandite conversion, optimizing hydration and microstructural densification, which translates into significant improvements in concrete strength and durability. The thermogravimetric analysis (TGA) in this study revealed that GO undergoes a 93.9% mass loss, occurring in three main stages: elimination of volatile compounds (up to 280 °C), decomposition of organic functional groups (280–600 °C), and structural carbon degradation (600–850 °C). In GO-modified concrete, the total mass loss was significantly lower (14.54%), indicating that the cementitious matrix significantly mitigates GO’s thermal degradation. Similarly, the study in [
80] reported that TGA showed a significant mass loss between 400 °C and 800 °C, primarily attributed to the progressive decomposition of hydrated compounds, such as C-S-H and portlandite (Ca(OH)
2), with reductions of 1.75–1.84% between 400 and 475 °C and 8.90–11.35% between 600 and 800 °C. This thermal degradation directly impacts the mechanical strength of concrete, with losses of up to 28% at 400 °C and 70% at 800 °C, due to the conversion of Ca(OH)
2 into CaO and the decomposition of C-S-H. Additionally, an increase in crack and fissure formation was observed from 600 °C onward, which compromises structural stability and increases porosity, reducing the material’s performance under high-temperature exposure.
5. Conclusions
The results of this study demonstrate that the addition of graphene oxide (GO) in low concentrations significantly enhances the mechanical and microstructural properties of concrete with a specified compressive strength of f’c = 280 kg/cm2.
From a physicochemical perspective, the GO used in this study exhibits suitable characteristics for its application in cementitious materials. Its average diameter of 20 µm facilitates homogeneous dispersion in water, optimizing its integration into the concrete matrix. Additionally, its chemical composition (63.40% carbon, 35.49% oxygen, and 1.11% sulfur) falls within the ranges reported in previous studies, ensuring the presence of functional groups that act as nucleation sites for hydration products, promoting a more compact and homogeneous microstructure.
Thermogravimetric analysis (TGA) identified three main stages of mass loss: (i) between 30 °C and 280 °C, associated with the removal of adsorbed water; (ii) between 280 °C and 600 °C, related to the decomposition of organic functional groups; and (iii) between 600 °C and 850 °C, corresponding to structural carbon degradation. The consistency of these results with previous studies validates the thermal stability of GO within the cementitious matrix.
Regarding mechanical properties, the incorporation of GO in proportions of 0.05%, 0.10%, 0.15%, and 0.20% by cement weight exhibited a positive effect, increasing compressive strength, indirect tensile strength, flexural strength, and elastic modulus. The optimal dosage was 0.05%, with respective increases of 14.61%, 12.33%, 6.09%, and 27.38% compared to the reference sample. However, at higher concentrations, mechanical performance showed a decreasing trend, suggesting that at GO dosages equal to or greater than 0.10%, this reduction is primarily due to the agglomeration of nanolayers caused by Van der Waals forces, leading to discontinuities in the cementitious matrix and stress concentrators that compromise its structural integrity, as evidenced in SEM micrographs. Additionally, the high specific surface area of GO increases water absorption, modifying the effective water-to-cement ratio and generating zones with non-uniform hydration, which affects the homogeneous formation of cementitious products, as confirmed by XRD analyses that reveal an imbalance in C-S-H and portlandite formation at higher GO concentrations. Furthermore, excessive GO content reduces concrete workability, requiring additional doses of superplasticizer to mitigate its effect, which may induce segregation and heterogeneity in the microstructure. These inter-related factors explain the decrease in mechanical strength observed in samples with GO concentrations above 0.05%, highlighting the importance of an optimal dosage to maximize benefits without compromising the material’s integrity.
Microstructural analysis confirmed substantial improvements in the sample with 0.05% GO. FTIR spectroscopy revealed shifts and increased intensity in absorption peaks associated with calcium carbonate (CaCO3), silicates, and Al(OH)3, indicating a higher degree of hydration and crystallinity. SEM images showed a denser structure with a 45.43% reduction in average pore and crack size. Additionally, a more uniform distribution of hydration products was observed, with CSH crystals in an expanded sponge-like form, Ca(OH)2 in hexagonal plates, and ettringite (AFt) in needle-like structures.
EDS analysis showed an increase in the Ca/Si ratio from 2.27 to 3.28 in the GO-modified sample, indicating a higher presence of hydrated cementitious products and suggesting improved strength of the material. XRD results confirmed a 26.31% increase in silicon-based compounds (CSH), while portlandite (Ca(OH)2) decreased by 2.06%, evidencing more efficient hydration and reduced susceptibility to carbonation.
Finally, TGA and DTG thermal analyses indicated that the GO-modified sample exhibited greater volatile and free water release at 101 °C, the more efficient decomposition of organic compounds at 150 °C, and portlandite degradation at lower temperatures compared to the reference sample. These findings suggest that GO addition enhances the thermal stability and hydration efficiency of cementitious compounds.