Next Article in Journal
Influence of Friction Forces and Gravity on the Quality of Movement in the MWR Rehabilitation Supporting Device
Previous Article in Journal
A Digitally Enhanced Ethnography for Craft Action and Process Understanding
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Harnessing Secondary Resources for Sustainable and Circular Practices in the Construction Sector: A Scoping Review

CARTIF Technology Centre, Circular Economy Area, 47151 Valladolid, Spain
*
Author to whom correspondence should be addressed.
Appl. Sci. 2025, 15(10), 5410; https://doi.org/10.3390/app15105410
Submission received: 16 April 2025 / Revised: 8 May 2025 / Accepted: 9 May 2025 / Published: 12 May 2025

Abstract

:
The construction sector is heading towards more sustainable practices by focusing on the utilization of secondary materials previously considered waste. These include by-products generated in other industrial sectors, which are considered to be secondary resources for construction applications. Among them, cellulose, mining tailings, and metallurgy slags are particularly significant due to their prevalence and the volume of waste they represent. These materials are generated in vast quantities globally, posing environmental challenges, but also presenting substantial opportunities for reuse in construction, thereby reducing landfill use and promoting resource efficiency. Additionally, the reuse of cellulose, tailings, and slag has a substantial environmental impact. Transforming them into valuable construction resources can significantly reduce the environmental footprint associated with extracting and processing new raw materials. The growing use of cellulose in products like insulation and fiberboards, and the incorporation of tailings and slags into concrete and road materials, underscores the feasibility of these practices. Despite these promising developments, challenges such as economic feasibility, technological limitations, and regulatory issues continue to limit their full potential. This article presents a scoping literature review that systematically identifies and synthesizes the existing research on the reuse of these three industrial by-product streams in construction. A structured search was carried out in scientific databases, and relevant publications were analyzed in terms of reuse strategies, technical performance, environmental benefits, and barriers to adoption. The review concludes by highlighting the key innovations and strategic pathways needed to overcome current barriers, thus supporting the transition toward a more sustainable and circular construction industry.

1. Introduction

The construction industry plays a pivotal role in the global economy, accounting for 10–12% of the global economic output and providing substantial employment opportunities [1]. This sector not only drives significant economic activity but consumes about 50% of the total use of raw materials and 36% of the global final energy use, highlighting its massive resource demand [2]. The substantial consumption of resources by the construction industry is coupled with high waste production, making it one of the primary contributors to environmental degradation. Specifically, construction and demolition activities are responsible for approximately one-third of the total waste generated in the EU, much of which ends up in landfills, creating serious environmental problems [3]. This high level of resource consumption and waste production underscores the industry’s profound environmental impact, necessitating a shift towards sustainability. In response, recent international agreements and policies are pressing for the adoption of ‘net zero’ and sustainability-focused practices, aiming to align the industry with global environmental targets [4]. These initiatives emphasize the adoption of innovative, green building techniques and materials, which are essential for transforming construction into a more environmentally responsible sector. Figure 1 shows key figures related to the construction sector, emphasizing its global economic importance and environmental impact.
The concept of a circular economy is particularly relevant in this context, offering a model where resource input, waste, emissions, and energy leakage are minimized by slowing, closing, and narrowing material and energy loops [5]. This can be achieved through long-lasting design, maintenance, repair, reuse, remanufacturing, refurbishing, and recycling. In the construction sector, applying these principles means rethinking how building materials are sourced, processed, and utilized. Particularly noteworthy are the materials derived from construction and demolition processes. Recycled concrete, for example, demonstrates not only environmental benefits but economic viability, potentially offering a net benefit of approximately 31 million US dollars per year through resource conservation and energy reductions [6]. Other recycled materials, such as ceramics, are also being integrated to enhance the properties of construction materials, further promoting sustainability [7]. The inclusion of recycled glass, known for its thermal insulation properties, contributes significantly to energy savings in buildings [8]. Additionally, the use of recycled PET (polyethylene terephthalate) and marble waste in concrete has shown promise in enhancing compressive and flexural strengths while reducing water absorption, illustrating the diverse possibilities of using recycled materials in construction [9]. These practices not only support resource efficiency but contribute to the sector’s shift towards environmental conservation [10].
Beyond the specific challenges associated with construction and demolition waste (CDW), the escalating impacts of climate change and resource depletion amplify the urgency to adopt more sustainable practices across all sectors. This transition focuses particularly on the utilization of secondary materials from a variety of sectors—resources that were once considered waste but are now recognized for their potential to reduce the industry’s environmental footprint. Among these materials, cellulose waste, mining tailings, and metallurgy slags are particularly noteworthy due to their abundance and the significant volume of waste they represent. In 2022, approximately 265.7 million tons of cellulose pulp were produced globally, which were transformed into 430.5 million tons of paper and cardboard. China, the European Union, and the United States together accounted for about 66% of the global paper and paperboard production. Specifically, China led with 27.3%, followed by the United States (16.3%), and the EU-27 (20.5%), within which Germany, Italy, Sweden, Finland, and France were the top producers. In Latin America, Brazil and Uruguay also stood out, with Brazil’s production exceeding 10 million tons, supported by its rapidly growing eucalyptus-based pulp industry [11]. This production reflects the widespread availability of cellulose as a secondary material, which can be recycled and used in new manufacturing contexts. Mining tailings, which are the waste materials remaining after the processing and separation of the valuable fractions of an ore, are generated in the range of 5–7 billion tons per year throughout the world. This vast quantity of tailings presents significant environmental challenges and opportunities for reuse. In China alone, over 4.5 billion tons of industrial solid waste were produced in 2018, with a significant share corresponding to mining tailings. Additionally, Brazil has been a focus of international attention due to the Brumadinho disaster in 2019, which underscored the environmental risks associated with poor tailings management [12]. In the case of slags, according to statistics, just from steel slag alone, the world produces about 190 to 290 million tons every year. China stands out as the largest producer, with around 80 to 90 million tons annually, followed by India, which generated nearly 40 million tons between 2017 and 2018 [13]. In the European Union, countries such as Germany and the United Kingdom are actively promoting slag valorization; for example, in the UK, steel slag is already used as an aggregate in concrete and asphalt mixtures. However, in many regions the reuse rate remains low, highlighting the need for improved recovery strategies to enhance circularity in the construction sector, since much of this by-product is either left in open-air stockpiles or disposed of in landfills. These data clearly demonstrate that these materials are generated in vast quantities globally, with millions of tons produced annually, posing serious environmental challenges but also presenting substantial opportunities for reuse within the construction industry. Their reuse can significantly reduce the reliance on virgin materials, decrease landfill use, and promote greater resource efficiency.
Sourced from segments like paper manufacturing, agricultural production, the textile industry, and wastewater treatment plants, cellulose waste offers unique characteristics that are ideal for various building materials [14,15]. In the paper industry, cellulose waste mainly originates from the sludge produced during papermaking. This sludge is rich in high-quality cellulose fibers, which are increasingly being repurposed to manufacture eco-cement [16], insulation panels, and fiberboards [17]. These products are prized for their excellent thermal and acoustic properties, aligning well with modern building standards that prioritize energy efficiency. Agricultural residue, such as straw, hemp, and bagasse, also serves as valuable sources of cellulose waste. These materials can be transformed into composites used for wall panels, roofing, and flooring solutions, leveraging their lightweight and durable nature for eco-friendly construction projects [18]. For example, hempcrete, a composite made from hemp fibers and lime, is renowned for its sustainability, excellent thermal insulation, and moisture-regulating properties, making it a superior alternative to conventional building materials [19]. Additionally, textile waste, particularly from cotton and linen, provides another substantial source of cellulose. Recycled fabrics can be processed into fibers that enhance the mechanical properties of concrete, such as tensile strength and durability, thereby extending the lifespan of concrete structures and contributing to more sustainable construction practices [20].
In addition to the applications mentioned for cellulose waste, mining tailings are also finding a valuable place in the construction industry. Traditionally seen as an environmental issue due to their management and storage, these residues are being transformed into useful resources for building infrastructure. For instance, tailings can be processed to create lightweight aggregates used in concrete blocks and tiles, offering a sustainable alternative to conventional materials, and reducing the extraction of virgin natural resources [21]. Their ability to enhance the thermal and acoustic insulation properties of construction materials also makes them an attractive option for projects aimed at improving energy efficiency in buildings [22]. Furthermore, soil stabilization with tailings has proven effective for constructing roads and other infrastructure, providing a solid and durable base that can withstand adverse weather conditions and heavy use [23].
Similarly, slags from the mining and metallurgy industry, previously regarded only as a by-product to be disposed of, are now being explored for their potential in producing concrete, road foundations, and other structural materials [24]. These applications not only help in reducing the environmental impact associated with the extraction and processing of virgin materials but support global sustainability goals by lowering emissions and reducing resource depletion [25]. Originating from the intensive activities of the mining sector and the metal smelting processes in the metallurgy industry, slag is now recognized for its robust properties, such as durability and thermal resistance, making it highly suitable for construction uses [26]. For example, copper slag, a by-product of the copper smelting process, has been effectively utilized as a partial replacement for sand in concrete and mortar mixtures. This substitution not only helps conserve natural sand reserves but enhances the strength and durability of the concrete due to the higher hardness of copper slag compared to traditional sand [27]. Another innovative application is the use of slag from nickel and platinum processing for road construction. The high density and abrasiveness of this slag make it an ideal material for high-traffic road surfaces, offering better longevity and less maintenance than conventional materials [28]. These progressive uses of mining slag demonstrate the broad potential of these materials to contribute to more sustainable construction practices and resource management.
These practices enhance both the environmental and economic efficiency of the sectors involved by turning waste into wealth and reducing the overall industrial footprint. However, the integration of these secondary materials into mainstream construction practices is not without challenges. Economic feasibility, technological limitations, and regulatory frameworks often act as barriers to the widespread adoption of these innovative practices [29]. This review will examine these challenges, exploring the technological innovations that facilitate the recovery and reuse of these materials, and discuss the strategies needed to overcome existing barriers. Despite the growing body of literature on the use of secondary materials in construction, this review aims to bridge the gaps by providing a novel synthesis of up-to-date technological innovations, comprehensive economic analyses, and current regulatory frameworks. As such, this review endeavors to shed light on the multifaceted benefits and the transformative potential of these materials, ultimately contributing to the evolution of a more sustainable and circular construction industry.
Figure 1. The construction industry in numbers ([30,31,32]). GHG: greenhouse gases; GDP: gross domestic product; CDW: construction and demolition waste.
Figure 1. The construction industry in numbers ([30,31,32]). GHG: greenhouse gases; GDP: gross domestic product; CDW: construction and demolition waste.
Applsci 15 05410 g001

Review Methodology

This study adopts a scoping literature review approach aimed at mapping the current state of research on the reuse of secondary materials—specifically cellulose residue, mining tailings, and metallurgy slags—in the construction sector. The objective is to identify the main applications, benefits, limitations, and knowledge gaps associated with the valorization of these industrial by-products in construction materials and processes. A structured literature search was conducted across major scientific databases, including Scopus and Web of Science, using combinations of keywords such as “cellulose waste”, “mining tailings”, “metallurgical slags”, “construction sector”, “sustainable construction”, “waste valorisation”, and “circular economy”. Boolean operators (AND/OR) were applied to refine the queries and ensure the relevant coverage of interdisciplinary studies.
The inclusion criteria prioritized peer-reviewed journal articles written in English, published between 2000 and 2024, and directly addressing the reuse of the selected secondary materials in construction applications. Studies were preferentially included if they met these criteria; however, a few relevant sources falling outside this scope were considered when they provided essential contextual or technical insights.
After the initial screening of titles and abstracts, the eligible articles were selected for full-text review. The selected documents were then analyzed thematically to extract information on the types of materials used, processing methods, construction applications, performance characteristics (mechanical, thermal, acoustic, etc.), environmental benefits, and regulatory considerations. The results of this review are presented in a structured manner, grouping the findings by material type (cellulose, tailings, slag) and application (e.g., composites, insulation, concrete, cement, road construction, ceramics), followed by a discussion of cross-cutting challenges and future opportunities. Rather than treating each waste stream in isolation, this review aims to highlight their respective strengths and challenges within a common framework of performance, applicability, and environmental relevance, thereby supporting a comparative understanding of their roles in sustainable construction.
To improve methodological transparency, Boolean logic was applied to refine the search strategy using combinations of keywords tailored to each material stream and its relevant construction applications. Searches were conducted across Scopus, Web of Science, and Google Scholar using queries such as:
-
(“cellulose fiber” OR “cellulose waste”) AND (“insulation” OR “concrete” OR “mortar” OR “composites” OR “bio-based materials”) AND (“construction” OR “building” OR “civil engineering”)
-
(“mine tailings” OR “mining waste” OR “iron tailings” OR “copper tailings” OR “phosphate tailings”) AND (“geopolymer” OR “binder” OR “concrete” OR “paving blocks” OR “ceramics”) AND (“construction” OR “building materials” OR “road base”)
-
(“steel slag” OR “blast furnace slag” OR “copper slag” OR “ferrochrome slag” OR “ferronickel slag”) AND (“cement” OR “aggregate” OR “asphalt” OR “bricks” OR “concrete”) AND (“construction” OR “infrastructure” OR “building sector”)
Articles were included if they contributed substantively to at least one of the following dimensions: mechanical or durability performance, environmental assessment (e.g., life cycle analysis), regulatory context, or practical implementation in construction. Data extraction focused on parameters such as compressive and tensile strength, water absorption, environmental metrics (e.g., CO2 reduction, embodied energy), and application-specific processing methods. Although no formal bibliometric analysis (e.g., citation mapping or network clustering) was undertaken, the structured selection process was designed to ensure thematic breadth, technical depth, and representation of both established practices and emerging innovations. The resulting synthesis draws upon 302 references spanning two decades of the literature on circular construction.

2. The Role of Cellulose Waste in Sustainable Construction

Cellulose waste is emerging as a critical resource in the field of sustainable construction practices, offering a renewable and abundant alternative to traditional building materials. Derived from various industrial processes, cellulose waste has the potential to be repurposed into a wide range of construction materials that contribute to reducing the environmental impact of the building industry [33]. By leveraging the inherent properties of cellulose, such as its strength, durability, and thermal insulation capabilities, innovative applications are being developed that not only improve material efficiency but support the principles of a circular economy. The utilization of cellulose waste in construction can be categorized into several key areas, each demonstrating unique advantages and applications. These include the development of cellulose-based composites and bio-based materials, the production of high-performance insulation products, and the creation of sustainable concrete alternatives, as summarized in Table 1. Each of these applications harnesses the potential of cellulose waste to enhance the sustainability of construction practices, reduce reliance on non-renewable resources, and contribute to a more eco-friendly-built environment. The following sub-sections will explore these applications in detail.

2.1. Cellulose-Based Composites and Bio-Based Materials

Cellulose waste serves as a fundamental building block in the development of advanced composites and bio-based materials that are gaining traction in the construction industry. These materials leverage the natural properties of cellulose, such as its high tensile strength, biodegradability, and renewability, to create alternatives to conventional, often synthetic, building materials. Agricultural residue, such as straw and hemp, are particularly valuable in this context. Straw-based panels, for example, offer a sustainable alternative to traditional wood products. These panels are lightweight, durable, and can be produced with minimal energy input, significantly reducing the environmental impact associated with conventional wood processing. Furthermore, these insulating panels exhibit a high specific heat capacity of over 1600 J/(kg·K), leading to a volumetric specific heat capacity of approximately 0.35 MJ/(m3·K), despite their low density of around 210 kg/m3, enhancing their thermal performance and making them highly efficient for energy conservation in buildings [76]. Hemp fibers, combined with lime, are used to create hempcrete, a biocomposite that is highly valued for its excellent thermal insulation, moisture regulation, and carbon-sequestration capabilities. Hempcrete not only serves as a sustainable alternative to conventional construction materials but contributes to reducing the carbon footprint of buildings, as the hemp plant absorbs significant amounts of CO2 during its growth, effectively capturing carbon that remains stored in the building material [77]. Furthermore, hempcrete is a carbon-negative material itself; for example, 1 m2 of a hemp–lime wall (260 mm thick) requires 370–394 MJ of energy for production and sequesters 14–35 kg of CO2 over its 100-year lifespan, compared to an equivalent Portland cement concrete wall that requires 560 MJ of energy and releases 52.3 kg of CO2 [78]. Additionally, hemp–silica panels have demonstrated a thermal conductivity of 0.101 W/mK and an excellent moisture buffering value of 3.49, further enhancing their suitability for energy-efficient and moisture-regulating construction applications [79]. Cellulose from textile waste, particularly recycled cotton and linen, is being repurposed into fibers that enhance the mechanical properties of construction materials [80]. These fibers are used in concrete and other composites to improve tensile strength and durability, offering a sustainable solution for managing textile waste while contributing to the development of more resilient and long-lasting building products. Nonwoven composites produced from textile waste have demonstrated excellent acoustic and thermal performance. With an air flow resistivity lower than 100 kN s/m4, these materials achieve sound absorption coefficients higher than 0.6 starting at 500 Hz for 2.5 cm thick panels. Their thermal conductivity, lower than 0.05 W/(m·K), also makes them an ideal choice for improving the energy efficiency of building envelopes, contributing to both indoor comfort and sustainability [81]. However, despite their promising mechanical and thermal performance, cellulose-based materials may present durability limitations under certain conditions. In particular, the long-term exposure of these bio-based composites to moisture or microbial activity can compromise their structural integrity over time. Ongoing research is addressing this issue by developing fiber surface treatments, protective coatings, or hybrid formulations that enhance biological and moisture resistance, thereby improving their long-term applicability in the construction sector.

2.2. Cellulose Insulation Products

Cellulose waste is increasingly being utilized to produce highly effective insulation products, which align with the growing demand for energy-efficient building solutions. Derived from recycled paper [82], agricultural residues [83], and even textile waste [84], cellulose-based insulation materials are treated and processed to create products that can be used in various insulation applications. The natural properties of cellulose fibers, when processed appropriately, result in insulation materials that significantly reduce heat transfer, thereby improving the energy efficiency of buildings [85]. This is especially critical in sustainable construction, where reducing the energy required for heating and cooling is a key objective. In addition to their thermal properties, cellulose insulation products derived from both paper and agricultural waste offer excellent soundproofing capabilities [86]. This makes them ideal for use in a wide range of building types, from residential homes to commercial spaces. Moreover, these materials can be treated with non-toxic fire retardants, making them a safer and more environmentally friendly alternative to conventional insulation products that often rely on harmful chemicals [87].
A study carried out by Wang and Wang [88], set in the context of the UK’s climate and landfill crises, explored the potential of cellulose derived from recycled paper as an insulating material to address the country’s inadequate recycling rates. In a case study of the retrofit and new-build project in London, their research examined the performance of cellulose insulation compared to other conventional materials. The findings showed that, although cellulose insulation from recycled paper may not offer the highest thermal performance, it meets the requirements for most applications. Furthermore, this type of cellulose insulation emits significantly less carbon dioxide over its life cycle compared to other insulation materials, making it a highly sustainable option. While traditional paper recycling can sometimes result in lower carbon emissions, converting waste paper into cellulose insulation presents a viable alternative that not only improves the recycling rate but provides sufficient insulation for energy-efficient buildings. In exploring the use of agricultural waste for construction insulation, several innovative cases demonstrate both environmental and functional viability. For instance, insulation boards made from pineapple leaves have shown thermal conductivity values between 0.043 and 0.057 W/mK, highlighting their potential as effective and sustainable insulation materials [89]. Similarly, boards crafted from coconut husk and bagasse—by-products of coconut and sugar production—employ no chemical binders and exhibit thermal conductivity values ranging from 0.046 to 0.068 W/mK, underscoring their eco-friendliness [90]. Additionally, innovative boards combining wheat bran and banana peels offer promising properties suitable for building applications, with thermal conductivity ranging between 0.050 and 0.066 W/m·K, and mechanical properties comparable to those of commercial insulation boards [91]. Agricultural cellulose sources, such as straw and hemp, are also particularly effective in creating insulation products that offer superior thermal performance. For example, hemp-based lightweight concrete exhibits a thermal conductivity of 0.09 W/m·K, while straw-based concrete shows values as low as 0.078 W/m·K, making both materials excellent options for thermal insulation in sustainable construction [92]. These examples showcase how agricultural residue can be transformed into valuable, sustainable building materials, contributing to reduced environmental impact and enhanced building performance. Textile waste, which includes discarded garments, industrial scraps, and end-of-life textile products, provides a rich source of cellulose fibers that can be repurposed into high-quality insulation. Textile-based cellulose insulation is particularly valued for its excellent thermal properties; moreover, these materials can be treated to offer fire resistance and reduce noise pollution, making them suitable for a wide range of applications in both residential and commercial construction [93]. One notable example is the initiative carried out by Samardzioska et al. [94], which effectively demonstrated that insulation material made from recycled textile waste can achieve a thermal conductivity as low as 0.039 W/m·K. This level of performance makes it highly competitive with traditional insulation options, emphasizing its potential to improve energy efficiency in buildings significantly.

2.3. Sustainable Concrete Alternatives

The construction industry is also exploring the use of cellulose waste in the development of sustainable concrete alternatives. Traditional concrete, while essential for modern construction, has a significant environmental footprint, particularly due to the high carbon emissions associated with cement production. By incorporating cellulose fibers, including those derived from waste streams, into concrete, researchers and engineers aim to develop materials that maintain the strength and durability of conventional concrete while reducing its environmental impact. Kamaruzzaman et al. [95] used cellulose fibers as additives in concrete blocks to enhance their insulation properties, specifically improving their ability to limit heat transfer. Unlike other fillers, such as rubber particles, which can reduce the thermal insulating performance of cementitious composites, the use of cellulose fibers offers an ecological and environmentally sustainable alternative. The experimental results demonstrated that the incorporation of bio-cellulose significantly impacts the porous matrix structure, effectively reducing heat transfer and thermal conductivity. Cellulose-based composites can achieve thermal conductivities as low as 0.20 W/mK, which can be further optimized to 0.15 W/mK under specific conditions, such as temperature and the use of condensed phosphates. Similarly, Salahuddin et al. [96] explored the potential of waste paper concrete by incorporating waste paper into concrete mixtures, substituting fine aggregate, coarse aggregate, and cement in varying proportions (0%, 5%, 10%, 15%, and 20%). The study found that the fresh, structural, and durability properties of concrete improved with the addition of waste paper at 5% and 10%, while a decline was observed beyond 10%. At 10% incorporation, the concrete showed an increase in strength due to the presence of hydrated cement particles, outperforming standard concrete without paper. Cellulose-reinforced concrete exhibits improved tensile strength and reduced brittleness, which are critical factors in prolonging the lifespan of concrete structures. The incorporation of cellulose fibers helps to distribute stresses more evenly within the concrete, reducing the likelihood of cracking and other forms of degradation. This not only enhances the structural integrity of buildings but reduces the need for frequent repairs and maintenance, leading to lower lifecycle costs [97]. However, the majority of researchers have reported that the mechanical performance of concrete increased with the addition of fibers only up to a 1.0% concentration, while further additions of fibers decreased the mechanical performance of concrete due to reduced workability [98]. Building on the innovative uses of cellulose waste in concrete, another example of its application comes from a project at a hydropower station in China, where cellulose fiber-reinforced concrete was employed for its durability and enhanced mechanical properties. The study utilized four different fiber contents (0.6, 0.9, 1.2, and 1.5 kg/m3) to determine the optimal mix for concrete slabs exposed to extreme conditions. Particularly, the concrete with a fiber content of 0.9 kg/m3 demonstrated superior compressive, splitting tensile, and flexural strengths compared to plain concrete. This optimized cellulose fiber content not only improved the concrete’s structural integrity but its energy absorption capacity and toughness, making it highly suitable for the dam’s face slab in challenging environments. The implementation of this modified concrete in this project showcased significant benefits, as the dam face slab maintained a stable working condition through varying temperatures and even during seismic activities. These results underline the potential of cellulose fibers to enhance the resilience and longevity of concrete structures, particularly in hydroelectric projects where durability and performance under stress are paramount [42]. Additionally, cellulose-reinforced concrete often requires less energy to produce than traditional concrete, further contributing to its sustainability [14]. Nonetheless, the inclusion of organic components, such as waste paper, in concrete raises concerns regarding long-term durability, particularly in relation to water absorption and potential fiber degradation. Several studies have reported that prolonged exposure to moisture can weaken the fiber–matrix interface and reduce mechanical stability over time. Compared to other secondary resources, cellulose-based additions are particularly suited for enhancing tensile strength and reducing brittleness in cementitious materials, although their long-term performance is more sensitive to environmental exposure and fiber degradation. To mitigate these effects, strategies such as pre-treatment of paper fibers, incorporation of pozzolanic materials, or the use of hydrophobic admixtures are being explored to improve water resistance and prolong the service life of such concrete formulations.
Over the past decade, there has been a focused effort to explore the incorporation of agricultural waste products as alternative aggregates in both traditional concrete and bio-aggregate concrete. A wide range of these materials, including hemp, coconut shells, oil palm shells, corncob, straw, bamboo, and cork, have been explored for this purpose [99]. Kilani et al. [100], analyzed the specific mechanical properties of concrete integrated with hemp fibers, noting improvements in tensile strength and reductions in density, which offered lightweight construction benefits. They found that adding 5% hemp fiber to the concrete mixture increased the tensile strength by 10% while reducing the overall weight of the concrete by 15%. Rice husks are another agricultural by-product being utilized in concrete. Each kilogram of milled white rice generates roughly 0.28 kg of rice husks, which are typically used in energy generation or composting. However, studies have shown that rice husks can be incorporated into concrete for their insulating properties. Chabannes et al. [101] compared rice husk concrete with hemp concrete and found that, while rice husk concrete showed lower mechanical performance, it offered superior thermal insulation, with thermal conductivity values ranging between 0.10 and 0.14 W/mK, making it a viable eco-friendly material for filling walls. Additionally, Marques et al. [102] also demonstrated the suitability of rice husk concrete for use in acoustic barriers and thermal insulation in multilayer systems, highlighting its versatility and environmental benefits. Hemp shiv, another agricultural by-product, is widely recognized for its ecological benefits in construction, especially in the form of hempcrete. Hemp shiv, the woody core of the hemp stalk, is mixed with a binder, such as lime, to form hempcrete, which is valued for its excellent insulation properties and carbon sequestration capabilities. Hempcrete has been used in both prefabricated masonry blocks and as an in situ cast material, offering advantages such as lower thermal conductivity and improved sustainability compared to traditional concrete. The use of lime-based binders further enhances hempcrete’s environmental impact by reducing its carbon footprint [92]. Compared to mineral-based reinforcements, such as slags or tailings, cellulose fibers are particularly advantageous in enhancing tensile performance and reducing brittleness in lightweight applications, although their susceptibility to moisture-induced degradation remains a technical limitation.

3. The Role of Mining Tailings in Advancing Sustainable Construction Practices

Mining tailings, the by-products of mineral processing operations, are increasingly being recognized not merely as waste, but as valuable resources for sustainable construction [103]. The existence of critical minerals and rare earth elements in discarded mine tailings also suggests that reprocessing these materials could yield both economic and technological benefits [104]. Historically, these materials have been relegated to storage facilities or waste dumps, posing environmental and safety risks. However, with advances in technology and a shift towards more sustainable practices, mining tailings are now being repurposed into a variety of construction materials. This approach not only mitigates the environmental hazards associated with tailings storage but contributes to resource conservation by reducing the demand for virgin raw materials. This section explores the diverse applications of mining tailings in the construction industry, ranging from common uses, such as aggregates in concrete, to less common but innovative applications, like ingredients in glass manufacturing. Table 2 provides a summary of various construction applications for mining tailings, detailing the next sections their technical benefits and environmental impacts, and highlighting the versatility and potential of these materials in the construction sector.

3.1. Mining Tailings as Aggregates in Concrete

Mining tailings are commonly used as aggregates in concrete production but their application varies significantly depending on the type of mine from which these tailings are sourced. This distinction is crucial as the chemical composition, particle size, and overall suitability for concrete production can differ dramatically, influencing both the environmental impact and the mechanical properties of the resulting concrete. For instance, tailings from iron ore mines have been successfully used in concrete as they typically possess high density and few deleterious materials, which can enhance the strength and weight of the final product [153]. This makes iron ore tailings particularly suitable for heavy-duty applications, such as foundations and load-bearing structures [154]. A case study from an iron ore processing facility in China demonstrated that replacing up to 30% of traditional aggregates with iron tailings could significantly increase the compressive strength and durability of dam concrete for hydraulic engineering applications. Furthermore, the concrete prepared with iron ore tailings exhibited superior durability compared to that prepared with normal aggregate, with the abrasion resistance being higher by 1.32–2.90 h m2/kg. Due also to the higher compressive strength, lower saturated-surface-dry water absorption, and angular shape of the iron ore tailing aggregates of raw rock, the modified concrete showed overall durability performance. Significant economic and environmental benefits were also noted, including a cost savings of approximately 8.98 million USD in two practical hydraulic engineering applications. Additionally, the use of the modified aggregate resulted in a reduction of occupied land space, CO2 emissions, and water consumption by 151,900 m2, 8157 tons, and 942,000 m3, respectively, with relative reductions in CO2 emissions and water consumption of 15.7% and 27.4% compared to the normal aggregate [154]. Conversely, tailings from copper and zinc mines often contain higher levels of sulfides, which can lead to durability issues such as acid generation and expansion when exposed to moisture and air [155]. This makes them less suitable for use in environments exposed to water or in marine construction. However, these tailings have been effectively utilized in controlled low-strength materials for backfill and non-structural applications, where such limitations are less critical [150]. Research from a copper mine in Chile explored the use of these tailings for paving paths and lightweight concrete blocks, where they performed well within the acceptable limits for non-structural use [156]. In the study conducted by Adiguzel et al. [157], the use of process tailings in the concrete industry was closely examined with several key outcomes identified. It was reported that the compressive and flexural strength of concrete decreases as the proportion of tailings increases, with optimal substitution levels found to be between 10% and 40% for fine aggregates and between 5% and 20% for cement. Tailings were shown to enhance the mechanical and environmental properties of concrete when used in place of fine aggregates. Additionally, due to their filler effect, tailings positively affect the permeability and freeze–thaw resistance of concrete. However, if tailings exceed 10% of the mix, water permeability issues arise. High SiO2 content tailings (≥75%) proved beneficial as cement substitutes, particularly when the combined content of SiO2, Al2O3, and Fe2O3 in the tailings exceeded 75%. The particle size of the tailings also plays a crucial role; those greater than 1 mm can enhance the compressive strength of the concrete. Despite these advantages, concrete containing tailings exhibited reduced carbonation resistance compared to normal concrete, and could present environmental risks such as potential heavy metal mobilization and acid mine drainage. Nevertheless, according to Adiguzel et al., employing tailings in concrete production can lead to significant reductions in CO2 emissions from cement production and lower concrete production costs. The study emphasized the lack of extensive research on the environmental impacts of concrete with tailings, suggesting the need for more thorough investigations to widen the application of tailings in concrete production while considering both mechanical and environmental factors. Unlike cellulose, which is typically used for its reinforcement potential, mining tailings excel in bulk replacement applications and offer significant cost-savings opportunities when sourced and processed correctly. Their performance is strongly dependent on origin and treatment, making their integration more complex but potentially impactful when conditions allow.

3.2. Mining Tailings in Cement Production

Mining tailings, typically composed of silica (SiO2), alumina (Al2O3), iron oxide (Fe2O3), and lime (CaO), offer a valuable resource for the cement industry [158]. These chemical components are crucial for the formation of cement clinker, which is the core ingredient in cement production. By substituting or supplementing traditional raw materials with tailings, cement manufacturers can utilize a waste product beneficially while maintaining the essential chemical reactions required for high-quality cement. One of the primary advantages of using mining tailings in cement production is the potential to lower the sintering temperature needed for clinker formation [159]. Typically, the process of sintering cement clinker requires temperatures up to 1450 °C. However, when tailings rich in iron oxide and alumina are introduced, they can act as fluxing agents, reducing the temperature required to between 1300 °C and 1350 °C [160]. This reduction is particularly effective when tailings are used in conjunction with medium calcium limestone, which is a common raw material in cement production [161]. By decreasing the sintering temperature, the cement production process becomes significantly more energy-efficient, which not only reduces operational costs but lessens carbon emissions—both critical factors in the industry’s push towards sustainability. Additionally, the incorporation of tailings has been shown to enhance the mechanical properties of cement. Specifically, tailings can improve the burnability of the raw mix [159], which refers to the ease with which the raw materials combine to form clinker minerals under high temperatures. Improved burnability leads to a more consistent and complete reaction, enhancing the overall quality of the clinker. Furthermore, tailings can promote the formation of tricalcium silicate (C3S), the mineral primarily responsible for the strength developed in the early stages of cement setting [162]. The specific mineral compositions in tailings can vary depending on the source and type of mining activity, providing unique benefits in cement production. The presence of certain trace elements in the tailings, such as manganese [163], or vanadium or titanium [164], can act as mineralizers, facilitating the formation of C3S and improving the speed of clinker formation. The tailings from iron ore mining, which are high in Fe2O3, can contribute to early strength development in cement (Tai et al., 2024) [165]. By contrast, the tailings from aluminum ore processing might offer advantages in terms of durability and resistance to chemical attack [166].

3.3. Mining Tailings in Ceramic and Brick Manufacturing

Mining tailings, utilized in ceramic and brick manufacturing, serve as substitutes or supplements for conventional raw materials, such as clay and sand, which are typically extracted through environmentally costly mining processes [167]. These tailings are primarily composed of silica, alumina, and iron oxide. These components are particularly valued in the ceramics and brick industries for their ability to enhance the physical and aesthetic properties of the final products. Silica, for example, plays a critical role in the formation of the glass phase during the firing process in ceramics production [168]. This phase is essential for creating the desired hardness and durability in ceramic products. Alumina, also abundant in tailings, contributes to the structural integrity of ceramics, enhancing their resistance to wear and tear and their ability to withstand high temperatures without deforming [169]. Iron oxide, another significant component of many mining tailings, imparts a natural pigment to bricks, producing a range of earthy tones from reds to browns, which are highly sought after in building aesthetics [170]. Additionally, iron oxide influences the thermal properties of bricks, improving their insulative qualities and overall durability, which is crucial for structural applications where strength and longevity are required [171].
In the ceramics industry, the unique chemical makeup of mining tailings facilitates diverse applications, enhancing both the performance and aesthetic qualities of ceramic products. Solid tailings, especially those that are inert and non-sulfidic, are becoming increasingly popular as primary materials in ceramics. The similarity in chemical properties allows for the tailings to seamlessly replace clay in various ceramic manufacturing processes. Furthermore, the granular size of these tailings plays a crucial role in the ceramic production, affecting the key characteristics, such as plasticity, density, sintering rates, and porosity [172]. The use of finely sized tailings can significantly influence the manufacturing of porcelain, improving the end-product’s quality and durability [143]. This substitution is not only a matter of maintaining quality but enhancing it, as the physiochemical interactions during the sintering process can lead to improved hardness and structural integrity of the ceramic items. Hua et al. [173] utilized iron tailings as the primary raw material to develop a type of lightweight and high-strength ceramsite. By combining these iron tailings with dolomite and a small amount of clay, and processing them at 1150 °C in a nitrogen atmosphere, a new material was synthesized that exhibited enhanced mechanical strength, thermal stability, and significant adsorption capacity, making it ideal for diverse engineering applications. Ceramic floors and tiles are exemplary materials where industrial wastes, like tailings, can be utilized effectively. These products benefit from the enhanced properties provided by tailings, meeting the high standards required for modern construction and architectural designs. Paiva et al. [145] investigated using sulfidic mine tailings in ceramic roof tiles, finding that incorporating 20% of these tailings not only improved the tiles’ properties but reduced the firing temperatures, thus saving energy and costs. The study also proposed a method to manage sulfate emissions, enhancing sustainability in construction material production. Additionally, the high-temperature and pH-resistant nature of the ceramics made with tailings is crucial for applications in challenging industrial environments. The inherent chemical bonds in ceramics, predominantly covalent or ionic, are less reactive under extreme conditions, preserving the integrity and functionality of the ceramic products in oxidative and corrosive settings [174].
Brick has long been an essential material in the construction and building industry, with its usage expanding continuously alongside the growing global population and increased structuring. This surge in demand is rapidly depleting the natural raw materials traditionally required for brick production. In response, researchers are actively seeking alternative materials for brick manufacturing. Kang et al. [175] and Li et al., [176] showed that iron ore tailings, due to their high density and mechanical strength, are effective for making bricks. Thejas and Hossiney [177] reported the use of fine-grained, low-silica iron ore tailings along with a non-cement curing system to create environmentally friendly bricks, incorporating a triethanolamine hardening accelerator and a stearic acid emulsion for waterproofing. Chen et al. (2020) [178] developed high-strength autoclaved bricks from hematite tailings and investigated a hematite tailings–lime–sand mixture to guide their production. Nikvar-Hassani et al. [179] found that copper tailings with low silica content are suitable for producing autoclaved sand–lime bricks that comply with building standards. Huang et al. [180] demonstrated the potential of feldspar, quartz, and kaolin tailings in the ceramics industry, particularly in the manufacture of building clay-bricks fired at 1200 °C. To meet the mechanical resistance requirements of bricks, a binding agent is crucial due to its ability to durably adhere the primary materials together. Traditionally, clay and shale serve as the natural binding agents in brick manufacturing, necessitating high-temperature firing between 800 °C and 1000 °C to produce the bricks. However, mine tailings, rich in aluminum and silicon oxide, can substitute for clay in the production of bricks through geopolymerization [181].

3.4. Tailings-Based Geopolymer

Geopolymers are synthesized from a complex three-dimensional network primarily composed of Si–O–Si and Al–O–Si bonds. These networks typically form through the reaction of aluminosilicate minerals (like fly ash or slag) with a concentrated alkali hydroxide or silicate salt solution. The geopolymerization process begins with the formation of Si(OH)4 or Al(OH)4 oligomers in the presence of an alkali solution enriched with Al and Si. This process progresses until the condensation of the metal–oxygen bonds occurs, ultimately forming the geopolymer structure. During alkali activation, it is also possible to form hydrated C-S-H or C-A-S-H cementing gels [182]. The geopolymerization process is influenced by numerous factors, including the type and concentration of the alkali activator, the curing conditions, and the mineralogical composition of the aluminosilicates used. Other variables that can impact the process include the SiO2/Al2O3 ratio, the concentration of the alkali solution, the curing time and temperature, the liquid-to-solid ratio, and the pH [183]. These variables allow for the utilization of geopolymers in various applications, such as in cement and concrete, fire-resistant materials, and toxic waste containment [184].
Given their silica/aluminosilicate-based mineral composition, mine tailings are considered potential candidates for the geopolymer industry [185]. However, unlike pozzolanic materials, like slag, which mainly consist of amorphous minerals (e.g., SiO2), mine tailings typically have a more crystalline mineral structure. This crystallinity is due to the fact that mine tailings are usually produced through wet chemical processes (like flotation), unlike other by-products formed at high temperatures (like slag). This makes mine tailings less chemically reactive in the geopolymerization process. For instance, silica (-SiO2), an inert mineral in mine tailings, reacts less with alkali solutions due to its ordered crystalline structure [186]. Therefore, pretreatment (such as thermal) or blending with other aluminosilicates might be necessary to enhance the reactivity of mine tailings in geopolymer applications. Nonetheless, the formation of a silica polymer (Si–O–Si derived from silica sol) around the tailings’ particles can help stabilize and build the tailings-based geopolymer structure [172]. Recent applications of tailings-based geopolymers have included iron tailings/fly ash-based geopolymers used as binder materials due to their high thermal conductivities, low porosity, and microcracks, as well as enhanced setting times and ultimate compressive strength [181]. Phosphate tailings-based geopolymers also serve as binder materials due to their characteristics [187]. Slag/metakaolin/gold tailings-based geopolymers are used for lightweight aggregates thanks to their high strength characteristics [188]. Copper tailings-based geopolymers find application in road base construction materials [179] and bricks [189], with copper tailings/cement kiln dust-based geopolymers noted for their stability and enhanced ultimate compressive strength. Silica-rich vanadium tailings/fly ash-based geopolymers are employed in fire-resistant materials due to their stability over a wide range of temperatures [172].

3.5. Mining Tailings: Other Uses

In addition to all the applications discussed in the previous sections, mining tailings are finding expanded uses in the construction industry. One novel application is in glass manufacturing, where tailings rich in silica provide a perfect raw material. Okereafor et al. [190] evaluated raw, water-treated, nitric acid-treated, and sulfuric acid-treated gold mine tailings to determine their effectiveness in glass production. The findings revealed that sulfuric acid-treated tailings yielded satisfactory quality for green glass, whereas raw, water-treated, and nitric acid-treated tailings exhibited white residues that compromised the glass quality. The research suggested further refining of sulfuric acid-treated tailings to decrease the iron oxide content, potentially extending their use beyond green and amber glass production. Alfonso et al. [191] investigated the use of tailings from fluorite mines, which are rich in calcium, to produce commercial glass by adding varying amounts of soda. The study revealed that glasses with a lower soda content displayed increased stability against crystallization but were more challenging to produce due to stricter temperature constraints during processing. These findings demonstrated that fluorite mine tailings could be transformed into commercial glass that effectively encapsulates potentially toxic elements, thereby mitigating the environmental risks associated with tailings disposal. The glass containing 20% Na2CO3 proved particularly effective in this capacity, though economic considerations must be taken into account to balance the cost against the environmental benefits. This method offers a promising avenue for repurposing hazardous mine waste into useful and environmentally friendly products. Fu et al. [192] explored the creation of lightweight glass ceramics using phosphorus tailings and coal gangue, focusing on developing eco-friendly building materials. The study analyzed the effects of different proportions of phosphorus tailings and heat-treatment temperatures on the ceramics’ properties and environmental safety. The findings indicated that increased phosphorus tailings led to larger pore sizes and improved insulation properties, while higher heat-treatment temperatures enhanced the material’s resistance to water and corrosion. Optimal conditions were identified as 80 wt% phosphorus tailings at a heat-treatment temperature of 1160 °C, producing ceramics with excellent porosity, strength, and durability.
Tailings are also utilized as versatile fill material in various construction and landscaping endeavors. Their utility ranges from serving as sub-base materials for roads to filling voids or being used to reclaim unused land areas. Employing tailings in these capacities effectively manages large volumes of these materials, which helps in diminishing the need for extracting and processing new raw materials. This method not only provides a sustainable solution to waste management but supports environmental conservation efforts. For instance, Sangiorgi et al. [193] highlighted the use of mine waste as raw materials for infrastructure, potentially impacting local communities and broader European mining activities positively. This initiative explored the economic and social benefits of recycling mine tailings into new materials that could replace commonly used substances in infrastructure projects. The focus was on developing a design method for alkali-activated materials that utilize mine and quarry wastes, aiming to create less porous and more durable materials than traditional road paving options. Further research has categorized the use of mining wastes in road construction into several forms: direct use without treatment, stabilization with hydraulic binders like cement or lime, stabilization through geopolymerization, and use with asphalt [194]. Untreated coarse coal wastes, for example, have been adapted for use as gravel in various applications, including pipeline ditch refills and parking lot surfaces. The fine tailings have found use as aggregates in cobbling and even as fillers in the cement industry due to their suitable particle sizes and compositions [195]. Moreover, tailings that contain gypsum are increasingly being processed for use in the production of plaster and drywall [196]. These diverse applications of mining tailings not only extend the lifecycle of mined materials but contribute to more sustainable and cost-effective resource management within the construction industry.

4. The Role of Metallurgy Slags in Sustainable Construction

Metallurgy slags, the by-products of the metal smelting process, are increasingly being viewed not just as waste but as valuable resources for sustainable construction practices [197]. These slags are composed of a mixture of oxides, silicates, and other metal derivatives, which have significant potential in various construction applications. Historically considered a disposal challenge, these materials are now recognized for their potential in reducing the construction industry’s reliance on virgin raw materials, thereby supporting environmental sustainability [198]. In the iron and steel industry, the production of metallurgy slags is particularly notable, as the sector generates large quantities of this material during the steelmaking process. In 2023, global production of iron slag was estimated to range between 330 million and 390 million tons, while steel slag production was estimated to be between 190 million and 290 million tons [199]. The rising production of these slags highlights the need for effective management and reuse, particularly in the construction sector, where they can serve as substitutes for traditional materials, reducing environmental impacts and enhancing the sustainability of construction practices. Table 3 provides a summary of various construction applications for metallurgy slags, while the next sections detail their technical benefits and environmental impacts, underscoring the versatility and potential of these materials in the construction sector.

4.1. Metallurgy Slags in Concrete Production

Slags have versatile applications across different industries, with concrete production being one of the most common [241]. Their reuse as replacements for binders and fine or coarse aggregates depends on the slags’ physical, chemical, and mechanical properties, as well as the specific treatments they undergo before being utilized. Among these, the reuse of iron and steel slags in concrete production stands out for its significant potential to enhance the properties of concrete mixtures while also offering environmental benefits. These slags, including blast furnace slag, basic oxygen furnace slag, electric arc furnace slag, and ladle furnace slag, have unique physical, chemical, and mechanical properties that make them suitable for various applications as substitutes for cement and fine or coarse aggregates. Specifically, blast furnace slag is commonly used as a partial replacement for cement in concrete due to its hydraulic properties, which improve the strength and durability of the concrete mix. Studies have shown that these slags can replace between 20% and 70% of cement, leading to reduced CO2 emissions and cost savings [242]. Optimal performance is observed with replacement levels between 20% and 30% [243], enhancing compressive and tensile strengths while maintaining good workability. However, challenges such as increased setting times and higher autogenous shrinkage need further investigation, particularly concerning long-term durability and resistance to chloride and sulfate attacks [244]. On the other hand, basic oxygen furnace slag exhibits good mechanical properties, but its use in concrete is limited due to its high volumetric expansion caused by free-CaO and MgO content. Treatments such as carbonation and scrubbing processes have been developed to reduce these components, making these slags a viable option for replacing natural aggregates or as a binder [245]. However, due to the lower hydraulic activity compared to cement clinker, basic oxygen furnace slag is more suitable for unbound applications, such as road construction, and further research is needed to optimize its use in concrete [246]. Similarly, electric arc furnace slag is widely used as a replacement for natural aggregates in concrete production, improving the mechanical properties of full-scale elements, like beams and columns [247]. However, when used as a cement replacement, issues such as slower strength development, increased setting times, and greater autogenous shrinkage have been observed [248]. While these slags show promise for enhancing the density and durability of concrete, further studies are needed to address their performance under freeze/thaw cycles, water permeability, and behavior in aggressive environments. Finally, ladle furnace slag has low hydraulic properties and is prone to volumetric instability, which complicates its reuse in concrete. Some studies have explored using these slags as a partial cement replacement in low-strength civil engineering applications, often requiring activation with alkali treatments. However, concerns regarding its workability, shrinkage, and long-term durability persist, necessitating more comprehensive research before real-world applications can be realized [249].
In addition to iron and steel slags, other slags from various metallurgical processes also have significant applications in concrete production. Copper slag, for example, is generated during the smelting process of copper, and contains high levels of iron, silicon, and aluminum oxides, making it suitable as a substitute for aggregates in concrete. These slags enhance abrasion resistance and the density of concrete, and have been successfully used in applications such as pavements and building blocks. However, their use is limited by the potential leaching of heavy metals, which requires strict environmental controls [250]. Copper slag specifically has demonstrated credibility for partial utilization in concrete, either as a binding material or as sand. Studies have shown that the optimal percentage of these slags in concrete can significantly impact its strength, with typical substitution rates ranging from 50 to 60% by weight of fine aggregates [251]. This variability in optimal use is often due to differences in the slag’s source. Despite its ability to enhance the mechanical properties of concrete, copper slag usage faces limitations in tensile strength. Nickel slag is another valuable by-product that can be used in construction. Rich in silicate and magnesium, nickel slag has been employed as a partial replacement for fine aggregates in concrete, improving wear resistance and thermal stability [252]. Additionally, nickel slag has been shown to increase the concrete’s resistance to aggressive environments, such as marine conditions, due to its dense microstructure and reduced permeability [253]. Shooshtarian et al. [254] demonstrated that the highest flexural and compressive strength for concrete using nickel slag sand was achieved at an optimal dosage of 30%. Phosphorous slag, a by-product of phosphoric acid production, also shows promise in concrete applications. Due to its high glass content and latent hydraulic properties, it can be used as a supplementary cementitious material, similar to ground granulated blast furnace slag. Phosphorous slag is currently utilized in concrete industries with a low cement replacement rate of 20–30%, due to its low pozzolanic activity [255]. But this material has been found to improve the durability of concrete, particularly its resistance to sulfate attack and alkali–silica reaction, making it a suitable material for infrastructures exposed to harsh chemical conditions [256]. These alternative slags expand the range of sustainable materials available for concrete production, helping to reduce the environmental footprint of construction while enhancing the performance of concrete in various applications. In contrast to cellulose and tailings, slags show the highest potential for widespread use due to their relatively uniform composition and established history in cement and asphalt applications. However, issues like expansion (due to free lime) require careful pre-treatment to ensure structural integrity.

4.2. Slags in Road Construction

Slags have found extensive use in road construction, primarily as a base material for roads and as an aggregate for asphalt. Their rough texture and high density provide excellent mechanical properties, such as stability and durability, essential for heavy-duty applications [257]. Steel slag, for example, offers numerous benefits as a material for road construction, thanks to its advantageous physical and chemical properties [258]. Steel slag exhibits superior performance compared to natural aggregates, with abrasion values ranging from 15% to 25%, significantly lower than those of typical natural aggregates, which can exceed 30% [202]. Its crushing strength is higher, often above 200 MPa, compared to around 150 MPa for natural aggregates, and its soundness index is less than 12%, indicating better resistance to weathering [259]. In addition, its rough, angular texture provides excellent interlocking, improving skid resistance by up to 50% compared to conventional aggregates [260]. Steel slag also enhances resistance to moisture damage, with tensile strength ratios often exceeding 90%, compared to 70–80% for natural aggregates, and it shows reduced rut depth in asphalt mixtures, often measuring less than 2.5 mm in repeated load tests [257]. Studies have consistently shown that incorporating steel slag into road pavements significantly improves durability and fatigue resistance, with fatigue life increasing by up to 30% compared to mixes without slag [261]. Additionally, its versatility allows it to serve as both a coarse and fine aggregate, optimizing the performance of asphalt and concrete mixtures.
Similarly, phosphorus slag has also been proven both theoretically and experimentally to be a feasible additive in asphalt concrete mixtures for road construction, enhancing the operational properties of road surfaces. Yang et al. [262] have shown that the significant porosity of phosphorus slag improves asphalt concrete’s structure, increasing the strength, density, and waterproofness by up to 20%, while reducing bitumen consumption by 15%. The authors also found that the slag’s interaction with bitumen is similar to limestone, and it has comparable effects on bitumen aging. Additionally, the coarse texture of slag enhances the internal friction angle, stability, and resistance of asphalt concrete, particularly in high temperatures, making it an excellent alternative to conventional materials like limestone. Extending the exploration of alternative slag materials, Kong et al. [263] successfully applied iron tailings slag in road subbase construction, demonstrating the material’s potential when integrated with optimal additives. Their research found that a mixture containing 6:1 slag to fly ash ratio, 4% calcium oxide, 5% cement, and 0.02% water-resistant base stabilizer achieved an impressive 7-day unconfined compressive strength of 1.97 MPa and an elastic modulus of 286 MPa. Copper slag, another promising alternative material, has also been explored for its potential in road construction, especially as a substitute for fine aggregates in asphalt and structural fill applications. Soni et al. [264] investigated the use of fine copper slag in road embankments, subgrades, and subbases, demonstrating its potential to enhance the mechanical properties of road layers. The study revealed that copper slag, due to its high specific gravity and angular particle shape, significantly improves the compaction characteristics and stability of road materials. The incorporation of 60% fine copper slag as a replacement for sand resulted in increased California Bearing Ratio (CBR, a parameter that measures the shear strength and load-bearing capacity of a soil or granular material) values, reaching up to 15%, compared to traditional soil-sand mixtures with CBR values of 8–10%. This improvement not only enhances the load-bearing capacity of subgrade layers but reduces swelling and moisture susceptibility, making copper slag a cost-effective and environmentally sustainable alternative to natural aggregates in road construction.

4.3. Innovative Uses in Bricks and Ceramics

Beyond conventional uses, metallurgy slags are being explored as primary materials in the manufacture of bricks and ceramics. The high silica and alumina content in these slags make them suitable for the production of heavy clay products. Incorporating slag into brick manufacturing can reduce the firing temperature needed, thus saving energy and reducing emissions. Furthermore, bricks made with slag exhibit enhanced durability and better thermal insulation properties compared to traditional clay bricks [265]. Several types of slag are currently utilized in brick. Liu et al. [266] explored the performance and reaction mechanisms of alkali-activated clay bricks incorporating steel slag and fly ash. Steel slag enhanced early compressive strength, while clay brick provided silicon and aluminum, aiding the reaction process. Fly ash dissolved quickly at higher temperatures, supplying additional silicon and aluminum, and together with clay brick, promoted calcium dissolution from steel slag. The optimal mix, with 30% steel slag, formed calcium-alumino-silicate-hydrate (C-A-S-H) and sodium-alumino-silicate-hydrate (N-A-S-H) gels that strengthened the system by wrapping partially hydrated clay brick. In another study, Ren et al. [267] conducted a comprehensive evaluation of aluminum ash and calcium carbide slag for brick manufacturing under ultra-high pressure conditions. The optimal conditions were a 1:9 ratio of aluminum ash to calcium carbide slag, 300 MPa pressure, 5-day curing time, and natural curing. The resulting bricks showed high compressive (75 MPa) and flexural strength (3.6 MPa), excellent waterproof properties, and met environmental standards for leaching toxicity, demonstrating a viable approach to recycling these industrial wastes into high-performance building materials. Hou et al. [268] examined the microstructure and mechanical properties of CO2-cured steel slag bricks at a pilot scale, while Wang et al. [97] investigated how moisture content affects bio-carbonated steel slag bricks. Both investigations demonstrated innovative uses of CO2 and microbial carbonation in steel slag brick production. Hou et al. [268] found that bricks with 25% steel slag achieved the highest CO2 uptake (7.5%) and a compressive strength of 27.7 MPa, meeting standards for load-bearing bricks, with excellent volume stability and freeze–thaw resistance. Wang et al. [97] showed that moisture content significantly influences the carbonation process, with optimal water levels enhancing CO2 diffusion and bacterial activity, improving carbonation efficiency, while too much moisture reduced these effects, impacting the final carbonation depth. Chen et al. [269] studied the impact of electric arc furnace slag on enhancing the quality and environmental safety of fired bricks that include municipal solid waste incineration fly ash. The slag increased density and reduced porosity and water absorption by over 50%, enhanced compressive strength by 50%, and lowered thermal conductivity. It also promoted chemical reactions and significantly reduced heavy metal leaching, making the bricks safer and more durable, addressing the drawbacks of using fly ash in brick manufacturing. Maierdan et al. [270] focused on recycling waste river sludge into unfired green bricks stabilized with phosphogypsum, slag, and cement. The study used hemihydrate phosphogypsum to dehydrate the sludge, and combined it with cement, sodium metasilicate, and slag. The optimal mix achieved a compressive strength of 31 MPa, meeting quality standards, with 96% waste utilization. The microstructural analysis confirmed the formation of calcium silicate hydrate and ettringite, enhancing brick strength. Fuchs et al. [271] developed CO2-cured, carbon-reducing slag-based masonry bricks at a pre-industrial scale, replacing conventional hydraulic binders with steel industry slags. The CO2-curing process at 15 bar, 50 °C, and 6 h resulted in compressive strengths of 24 to 49 MPa, surpassing those of traditional sand–lime bricks. The lifecycle assessment revealed a CO2-saving potential of 337 kg CO2 per cubic meter of brick compared to conventional methods, highlighting the significant environmental benefits and industrial feasibility of CO2-curing slag-based bricks. Guo et al. [272] analyzed the corrosion behavior of CaO–Al2O3–SiO2–MgO–Cr2O3 slag on MgO refractory bricks used in chromite smelting. Their study investigated the reaction mechanisms between the slag and the refractory, revealing that the main slag phases were pyroxene, chromium spinel, and magnesium olivine. An SEM-EDS analysis showed that, at the contact surface, calcium magnesium olivine and magnesium olivine were predominant, with calcium magnesium olivine increasing as slag basicity rose. Thermodynamic calculations indicated that higher slag basicity lowered the melting temperature and viscosity, allowing molten slag to infiltrate the brick pores, accelerating corrosion. Chinyama et al. [273] assessed the mechanical performance and contaminant leaching behavior of cement-based and fired bricks containing ferrochrome slag. The study compared these slag-containing bricks with conventional ones, revealing that bricks with 20–30% ferrochrome slag had higher compressive strength and lower water absorption. Contaminant leaching of trace metals and chloride was similar or lower than controls, except for higher manganese and chromium leaching in slag-containing debris due to their elevated concentrations in the slag. Cement-based bricks provided the highest strength, while fired bricks showed the least contaminant leaching. However, not all studies have yielded positive results. Gao et al. [274] examined the effect of slag composition on the corrosion resistance of high-chromia refractory bricks used in industrial entrained-flow gasifiers. Their study found that slag types influenced corrosion severity, with high-Ca/Na slag causing the most damage, followed by high-Fe slag, and high-Si/Al slag being the least corrosive. High-Ca/Na slag reacted with the refractory, forming low melting point phases that disrupted the brick’s matrix. High-Fe slag led to iron phase precipitation and spinel layer formation at the interface. The interaction of ZrO2 in the refractory with CaO from the slag formed calcium zirconate, which weakened the refractory’s structure and increased corrosion. After this analysis, it is observed that, although most studies highlight the benefits of using slag in bricks and ceramics manufacturing, such as improved mechanical properties and environmental advantages, it is essential to carefully assess the slag composition to mitigate potential negative effects, such as increased corrosion or contaminant leaching.

4.4. Slags in Cement Production

The use of slag in cement production is well-established, with its ability to substitute a portion of the clinker in Portland cement, known as slag cement. This substitution not only conserves limestone and reduces the energy required for clinker production but significantly diminishes CO2 emissions. The resulting slag cement generally shows improved performance characteristics, such as lower heat of hydration, better workability, and increased resistance to aggressive environments, making it highly desirable for marine and hydraulic constructions [275]. Several examples illustrate how different types of slag, with varying compositions and processing techniques, contribute to the performance and sustainability of eco-cements. Kurniati et al. [24] investigated the use of steel slag, ferronickel slag, and copper mining waste in cement production, highlighting their potential to partially replace traditional cement materials and enhance sustainability. Steel slag, with its high calcium oxide content and chemical properties similar to Portland cement, replaced up to 30% of traditional cement, significantly reducing CO2 emissions and enhancing mechanical properties, such as compressive strength and durability. Ferronickel slag, rich in silicon dioxide (up to 40%), exhibited pozzolanic activity, contributing to the formation of secondary calcium silicate hydrates and improving long-term strength in blended cements. Copper slag, characterized by high iron oxide (around 40%) and low calcium content, also showed promise as a pozzolanic material, particularly when rapidly cooled to maintain an amorphous silica state, allowing effective participation in the hydration process. When used at replacement levels of about 15–20%, copper slag contributed similar compressive strength gains as traditional pozzolans, demonstrating its value in cement formulations. Ryu et al. [276] successfully demonstrated that steelmaking slag, enriched with noncarbonated CaO from the decarbonization of limestone during steel production, can effectively replace a portion of natural limestone in cement clinker production. Their findings revealed that this substitution not only contributes to a significant reduction in CO2 emissions but enhances the overall material properties of the cement. The slag’s rich calcium content proved beneficial in maintaining the necessary lime saturation factor for optimal clinker formation. Through their detailed analysis, the authors also confirmed that the adjusted slag, when processed with targeted grinding and chemical stabilization, integrates well into cement mixtures, providing improved performance characteristics, such as increased durability and strength of the final cement product. Demarco et al. [277] investigated eco-sustainable cements made from iron-making slag and gypsum waste, with formulations containing 80–90% blast furnace slag, 10–20% gypsum, and 10–15% basic oxygen furnace slag. The best-performing eco-cement achieved compressive strengths greater than 50 MPa at 28 days and over 80 MPa at 180 days. This high performance was attributed to its composition, which included basic oxygen furnace slag with high basicity (3.1), a significant content of brownmillerite (19%), and dicalcium silicate (21%), along with blast furnace slag rich in aluminum (15%). These factors contributed to enhanced hydration, the formation of ettringite, and consistent strength growth throughout the curing process. Xie et al. [278] explored a large-scale industrial method for using modified magnesium slag in cement production to address the disposal challenges of magnesium slag waste. The study found that cement produced with modified magnesium slag met the standards for general Portland cement. The hydration process revealed a unique interaction between magnesium slag and clinker, enhancing early hydration and contributing to the cement’s performance. Additionally, using modified magnesium slag reduced carbon emissions by 7.95% and production costs by over 10% per ton of cement, demonstrating significant environmental and economic benefits.

5. Mechanical and Durability Performance of Construction Materials Incorporating Secondary Raw Materials

Beyond the environmental and economic advantages discussed in earlier sections, the incorporation of secondary resources, namely cellulose fibers, mining tailings, and metallurgical slags, also leads to measurable changes in key technical properties of construction materials. While previous parts of this manuscript have addressed material types, usage contexts, and sustainability benefits, this section focuses specifically on their influence on mechanical strength, dimensional stability, and long-term durability.
The studies already cited demonstrate that cellulose fibers, particularly those derived from waste paper or industrial by-products, enhance tensile strength and crack resistance when added in appropriate dosages to cementitious materials. For example, reinforced concrete prepared with 0.9 kg/m3 of cellulose fiber has shown notable improvements in compressive, tensile, and flexural strength under demanding structural conditions [42]. Similar benefits have been observed at lower incorporation levels (5–10%), where compressive strength increased without compromising workability [96]. However, higher dosages may induce durability issues due to increased water absorption and potential fiber degradation, especially under humid conditions [98]. Strategies such as fiber pre-treatment or combining with pozzolanic additives have been proposed to mitigate these effects and maintain matrix integrity [14].
In the case of mining tailings, several applications reflect a positive impact on the mechanical behavior of concrete, especially when using iron tailings as aggregate substitutes. As previously presented, replacing up to 30% of natural aggregate with iron tailings in hydraulic concrete led to increased compressive strength and abrasion resistance, as well as extended service life under abrasive and wet environments [154]. Optimal results were obtained when substitution ratios remained within 10–40% for fine aggregates and 5–20% for cement, providing improvements in permeability and freeze–thaw resistance due to the filler effect of finely ground tailings [157]. Nonetheless, the type of tailing remains a determining factor: materials from copper or zinc mines tend to be limited to non-structural uses due to risks of leaching or chemical instability [150,156]. Reactivity and durability can be optimized by selecting tailings with appropriate mineralogical compositions, particularly those with a combined content of silica, alumina, and iron oxides above 75% [157].
Similarly, metallurgical slags, such as blast furnace slag, steel slag, and copper slag, play a significant role in improving the mechanical and durability properties of cementitious materials. Ground granulated blast furnace slag, when used as a cement substitute in proportions between 20% and 30%, has been shown to increase both early and long-term compressive and tensile strengths while enhancing resistance to sulfate and chloride attacks [242,243]. Steel slags from electric arc furnaces serve effectively as aggregates, increasing density and strength in concrete structural elements [247], while copper slag, typically used as a partial fine aggregate replacement (50–60%), improves abrasion resistance and bulk density [251]. In asphalt applications, steel slag has demonstrated exceptional performance in enhancing moisture resistance, fatigue life, and rutting resistance in road surfaces [257,261]. Moreover, slags, such as phosphorus and nickel slag, also contribute positively to sulfate resistance, thermal stability, and impermeability, though their use requires careful control to prevent the leaching of heavy metals or volumetric instability [253,256].
To support and illustrate these findings, Table 4 compiles quantitative data from the literature analyzed in this review, highlighting how the use of these secondary raw materials influences compressive strength, tensile strength, water absorption, shrinkage, and other relevant parameters across different construction applications.

6. Overcoming Barriers to Integration

6.1. Economic Challenges

As deduced from the previous sections, the integration of secondary resources like cellulose waste, mining tailings, and metallurgy slags into the construction sector offers substantial environmental and economic benefits. However, significant barriers impede their widespread use, including economic challenges, technological limitations, restrictive regulations, and the need for innovative business models. Among them, economic challenges are a major obstacle to integrating secondary resources in construction [279]. The high costs associated with processing these materials, such as mining tailings and metallurgy slags, often involve specialized treatments, like grinding, chemical stabilization, or thermal activation, which require significant investments in technology and infrastructure. According to recent studies, despite the low cost or availability of these secondary materials, the processing expenses can be prohibitive, especially for small and medium-sized companies that lack access to the necessary capital and technological resources [280]. The inconsistent supply and variable quality of these materials further complicate cost–benefit analyses, making their integration less economically attractive [281]. Specific examples highlight the economic barriers associated with metallurgy slags, where steel slags need stabilization to reduce free lime content and prevent volumetric expansion, while copper slags must be rapidly cooled to retain their pozzolanic properties. The costs associated with these treatments can significantly impact the economic viability of using such materials. Similarly, cellulose derived from industrial processes, like paper manufacturing, requires extensive cleaning and refinement, which can outweigh the benefits if not optimized. Economic incentives, such as carbon pricing mechanisms like carbon taxes or emissions trading systems, can further encourage the use of secondary resources by making traditional, carbon-intensive construction materials comparatively more expensive. Government subsidies or tax breaks targeted at offsetting these initial processing costs could be instrumental in making the use of secondary resources more feasible economically [282].

6.2. Technological Limitations

Technological limitations also present substantial barriers. The varied composition of secondary materials with potential applications in the construction sector requires different processing techniques to ensure they meet the standards necessary for construction applications. Tailings, for instance, may contain hazardous elements, like heavy metals, which necessitate complex management strategies, including innovative treatments, such as thermal activation or blending with other materials, to mitigate environmental risks. Similarly, cellulose waste, sourced from paper mills or agricultural residue, requires tailored treatments, like refining or enzymatic processing, to optimize fiber quality for use in construction, but these processes can be costly and resource-intensive. Additionally, metallurgy slags used in cement production face challenges due to variable compositions that can affect hydration and stability, requiring further refinement of the activation methods. However, these advanced techniques are not yet widely adopted due to their complexity, cost, and the need for further research and development [283]. The lack of sufficient data and standardized processes further inhibits the broader use of these materials [284]. Innovative business models are also key to overcoming economic and logistical barriers. Industrial symbiosis, where companies collaborate to utilize each other’s by-products, can optimize resource use and supply chains for secondary materials. For instance, direct partnerships between steel manufacturers and cement producers can streamline the incorporation of steel slag into cement production, reducing transportation and processing costs while ensuring a consistent supply of high-quality materials [285].

6.3. Regulatory Barriers

The regulatory environment is another significant barrier, often rooted in outdated classifications and standards that fail to accommodate the innovative use of secondary materials [286]. For example, the classification of industrial by-products, like slags and tailings, as waste imposes strict handling and disposal regulations, increasing compliance costs and discouraging their reuse in construction [287]. There is a pressing need for the modernization of regulatory frameworks, which should recognize secondary materials as valuable resources rather than mere waste. By establishing comprehensive guidelines and standards, such as defining precise minimum quality requirements and acceptable levels of contamination, governments can significantly facilitate their broader adoption. This role is crucial in refining these frameworks to mirror the advancements in technology and adapt to the evolving market needs. Furthermore, policy adjustments that reclassify certain industrial by-products from waste to resources could lead to substantial reductions in compliance costs, thereby incentivizing companies to explore their use. In addition, the introduction of quality standards specifically designed for secondary materials could bridge the gap caused by market hesitancy, ensuring that these materials comply with the necessary performance and safety standards required for construction applications.
Some jurisdictions have already developed progressive strategies in this direction. For instance, Germany and the Netherlands have applied flexible interpretations of the EU Waste Framework Directive (2008/98/EC), enabling industrial by-products to achieve “end-of-waste” status when quality thresholds are met. Similarly, the Italian case of the “End of Waste Decree” for aggregates has helped streamline the valorization of certain mineral residues. These examples highlight how updated regulatory instruments can unlock SRM potential by providing legal certainty and market access. Another relevant challenge is the fragmentation of regulatory criteria across countries, which can hinder the circulation and transnational use of SRMs. Improving harmonization, or at least mutual recognition of quality standards, would strengthen regional markets and facilitate the use of SRMs in large-scale public projects or cross-border supply chains.

6.4. Policy Recommendations and Stakeholder Collaboration

Policy recommendations to enhance the use of secondary resources should aim to create an enabling environment that balances safety, performance, and sustainability. For example, mandatory requirements for incorporating a minimum percentage of recycled content in construction materials, alongside incentives for companies that exceed these targets, could significantly boost the use of secondary resources [288]. Clear labeling and certification programs could also play a vital role in building consumer confidence and driving market demand for sustainable construction products [289]. Collaboration among industry, academia, and government is essential for overcoming the barriers to integrating secondary resources. Industry can offer practical insights into the feasibility and challenges of using secondary materials, while academic researchers contribute advanced knowledge and innovations in processing and application technologies. Government support, through funding and regulatory adjustments, can bridge the gap between research and practical implementation, ensuring that new technologies reach the market effectively. Public–private partnerships, in particular, provide a valuable platform for piloting the new uses of secondary materials, allowing for the refinement of processes and the validation of economic and environmental benefits under real-world conditions.
In several countries, green public procurement (GPP) frameworks are already being used to promote SRMs by prioritizing materials with verified environmental benefits in public works. For example, France mandates minimum recycled content thresholds in state-funded building projects, and South Korea has implemented financial incentives and pilot schemes for the use of industrial by-products in concrete and asphalt. Moreover, policy instruments, such as tax incentives, dedicated SRM certification schemes, and national registries of valorized materials have proven useful in reducing market risk and encouraging innovation. Integrating these tools within broader circular economy strategies can help align construction policies with climate and waste reduction targets.
In conclusion, overcoming the barriers to integrating secondary resources into construction requires a multi-faceted approach that addresses economic, technological, and regulatory challenges. By investing in research and development, updating regulations, and fostering innovative business models and economic incentives, the construction sector can make significant strides towards sustainability. Collaborative efforts between industry, academia, and government will be pivotal in unlocking the full potential of secondary resources, driving the transition to more circular and resource-efficient construction practices.

7. Future Perspectives and Sustainability Goals

7.1. Technological Innovations in Waste Valorization

The future of sustainable construction is increasingly focused on the innovative reuse of secondary resources, such as mining wastes, industrial by-products, and other recyclable materials. Technological advancements are playing a crucial role in transforming these waste materials into valuable construction components, aligning with broader circular economy goals and sustainability targets [290]. As the industry strives to reduce its environmental footprint, integrating these secondary resources offers a pathway to significantly lower impacts while creating new economic opportunities. Advanced processing technologies—including mechanical, thermal, and chemical processes—are at the forefront of converting mining wastes into high-value construction materials. These technologies enable the transformation of materials like mining tailings into lightweight aggregates, paving blocks, and bricks, providing sustainable alternatives to traditional, resource-intensive materials [291]. This not only diverts waste from landfills but conserves natural resources, contributing to lower emissions and a reduced environmental impact throughout the construction lifecycle. Geopolymer technology represents a significant opportunity for the construction sector by using mining wastes to produce geopolymer binders. These binders can replace traditional cement, significantly reducing CO2 emissions linked to conventional construction practices. By adopting geopolymer technology, the industry can turn problematic waste streams into durable, environmentally friendly construction materials with enhanced chemical resistance [292]. Nanotechnology is another promising area for the reuse of secondary resources. For example, nano-silica extracted from mining tailings can replace portions of traditional binders in construction materials, improving strength and durability while reducing environmental impact [293]. This approach aligns with sustainability goals by minimizing the use of conventional raw materials and enhancing the overall performance of construction products.
However, while these innovations are promising, their implementation at industrial scale requires a realistic assessment of technological readiness and commercial viability. For instance, some solutions, such as the use of cellulose-based insulation panels or fiber-reinforced concrete blocks derived from textile waste, have reached high technology readiness levels (TRL 8–9), being already available on the market [42,94]. Conversely, other developments, such as the use of nano-silica from mining tailings or CO2-cured slag-based bricks, remain at lower TRLs (5–6), as they are still undergoing pilot-scale validation or early pre-industrial implementation [93,268]. Geopolymer technologies using mining tailings have also demonstrated laboratory feasibility and mechanical performance but often face scalability issues due to the need for pretreatment, alkali activation, and durability under real exposure conditions [183,185]. Therefore, although the innovation pipeline is rich and diverse, bridging the gap between research and market adoption still requires investment in demonstration-scale trials, standardization efforts, and supportive policy frameworks to overcome current barriers.

7.2. Digital Technologies for Optimization

The integration of digital technologies, such as artificial intelligence, machine learning, and 3D printing, is set to revolutionize how secondary resources are used in construction [294]. These technologies can optimize material formulations, enhance production efficiency, and allow for precise material application, reducing waste and energy consumption. In particular, 3D printing offers the potential to create customized, efficient structures using materials derived from waste, thus minimizing environmental impact. Emerging digital technologies, like blockchain, the internet of things (IoT), and digital twins, are further enhancing the integration of secondary resources into construction. Blockchain can create secure and transparent records of material quality and origin, increasing confidence in the use of secondary materials. The IoT connects physical devices and systems, enabling better data exchange and automation in material processing and construction, while digital twins simulate construction processes to optimize the use of secondary materials by predicting their performance under different conditions. These future perspectives highlight the significant potential of secondary resources to contribute to a more sustainable construction industry.

7.3. Circular Business Models and Environmental Assessment

Circular business models are also essential for advancing the reuse of secondary resources in construction. By creating new value chains, these models turn waste materials into economic assets, providing new revenue streams for industries, like mining and metallurgy, while reducing their environmental footprint. Establishing partnerships across sectors ensures a steady supply of quality secondary materials, enhancing both the economic and the environmental performance [295]. In this frame, and to ensure that the use of secondary materials meets sustainability goals, life cycle assessment (LCA) is an invaluable tool. An LCA evaluates the environmental impacts of using mining waste and industrial by-products compared to traditional raw materials, helping to identify the most sustainable options. This approach allows for optimizing material use, ensuring that the environmental benefits are maximized throughout the lifecycle of construction products [296]. On the other hand, a material flow analysis (MFA) complements the LCA by assessing the movement of secondary resources within the construction sector. An MFA helps identify opportunities for increased recycling and reuse, guiding the efficient allocation of mining and industrial wastes to applications where they provide the greatest environmental benefit. For instance, an MFA can help determine the most effective uses for slags or tailings, optimizing resource use and reducing waste [297]. Adopting these strategies not only addresses the environmental challenges associated with construction but opens new avenues for economic growth and sustainable development.
Recent LCA studies support the environmental advantages of using secondary materials such as cellulose, mining tailings, and metallurgical slags in construction. For example, insulation materials made from recycled cellulose fibers consistently show lower environmental impacts across categories, such as global warming potential and embodied energy, when compared to conventional insulation, like expanded polystyrene or mineral wool [298]. One study even reported a net carbon benefit of −210% for cellulose-insulated walls, due to biogenic carbon storage [299]. Similarly, concrete incorporating mining tailings has demonstrated CO2 emissions reductions of over 50% when used in geopolymeric binders, while a partial cement substitution (5–20%) with tailings can avoid up to 72 kg of CO2 per m3 of concrete produced [300]. Regarding slags, replacing 70% of Portland clinker with ground granulated blast furnace slag was found to reduce greenhouse gas emissions by approximately 47.5% [301]. Copper slag used as a supplementary cementitious material in concrete achieved a ~12.8% reduction in CO2 emissions compared to conventional concrete [302]. In addition, steel slag used as aggregate does not increase lifecycle emissions and can help preserve natural resources without compromising structural performance. These findings confirm that integrating waste-derived materials not only diverts waste from landfills but significantly lowers the environmental footprint of buildings and infrastructure.

8. Conclusions

The construction sector is progressing toward greater sustainability by harnessing secondary resources, such as cellulose waste, mining tailings, and metallurgy slags. Traditionally viewed as waste, these materials are now being repurposed into valuable components for construction, significantly reducing the need for virgin raw materials and minimizing environmental impacts. By integrating these secondary resources, the industry can achieve notable reductions in greenhouse gas emissions, enhance the performance of construction materials, and align with global sustainability goals. The reuse of cellulose waste, for example, contributes to the development of eco-friendly insulation and composite materials, enhancing thermal efficiency in buildings. Mining tailings, when used as aggregates in concrete and road construction, improve durability and reduce the demand for natural aggregates, while metallurgy slags offer robust properties for cement and concrete production, lowering the sector’s carbon footprint. These practices underscore the potential of secondary resources to transform the construction industry, promoting a circular economy that emphasizes resource efficiency and waste reduction. However, the widespread adoption of these sustainable practices faces significant challenges, including the economic costs of processing and refining secondary materials, technological limitations related to material quality and performance, and restrictive regulatory frameworks that often classify these by-products as waste. To overcome these barriers, it is crucial to invest in research and development to improve material processing technologies, develop innovative treatments, and optimize the performance of these resources. Regulatory frameworks also need to be updated to facilitate the safe and effective use of secondary materials, encouraging their integration into mainstream construction practices. Innovative business models, such as industrial symbiosis, where companies collaborate to utilize each other’s by-products, can play a pivotal role in overcoming logistical and economic challenges. These models, along with public–private partnerships, can help scale up the use of secondary resources, demonstrating their viability and driving market confidence. The future of construction hinges on the ability to scale these sustainable practices and integrate them into standard industry operations. By embracing circular economy principles, the construction sector can achieve significant environmental and economic benefits, reducing waste, conserving natural resources, and enhancing the resilience of the built environment. This approach not only addresses current environmental challenges but sets the foundation for a more efficient, sustainable, and forward-looking construction industry aligned with global environmental objectives.
Among the three material streams reviewed, metallurgical slags currently exhibit the most mature and standardized integration in construction due to their chemical consistency and availability. Mining tailings offer a versatile but highly site-specific potential, whereas cellulose-based materials provide unique advantages in lightweight and insulating applications, though their durability requires careful design considerations. These distinctions underline the need for tailored valorization strategies aligned with the specific demands and constraints of each construction context. These findings reinforce the notion that secondary raw materials are not interchangeable but complementary. Their incorporation should be guided by performance-based criteria rather than merely availability or environmental appeal. Future efforts should therefore prioritize the development of adaptive standards, targeted material conditioning techniques, and application-specific performance benchmarks. This will ensure that circular construction strategies move beyond material substitution toward truly optimized, resilient, and scalable solutions.

Comparative Synthesis and Evaluation Framework

To complement the preceding conclusions and support more informed decision-making, a comparative framework has been developed to assess the relative strengths and limitations of the three SRM streams analyzed in this review, cellulose fibers, mining tailings, and metallurgical slags, across multiple dimensions. While previous sections have explored their specific applications, technical performance, and environmental potential, this synthesis aims to articulate how these materials differ in practical deployment, maturity of use, and the trade-offs involved in their adoption.
The framework presented in Table 5 is structured around four key evaluation criteria: (i) technical performance (e.g., strength, durability, workability), (ii) environmental impact (e.g., emission reduction potential, resource conservation), (iii) economic viability (e.g., cost of processing, scalability, market value), and (iv) regulatory and standardization context (e.g., availability of technical standards, acceptance in codes and regulations). This multi-criteria perspective enables a more integrated understanding of how each SRM aligns with different construction needs and sustainability goals.
This analysis highlights that metallurgical slags generally score highest in terms of technical performance and standardization, benefiting from decades of research, codification, and widespread industrial integration. Mining tailings, though offering high potential for volume replacement and localized sourcing, are more heterogeneous in composition, and face greater regulatory and environmental scrutiny, requiring site-specific adaptation. Cellulose-based materials, on the other hand, stand out for their low environmental footprint and suitability in thermal and acoustic insulation, but raise questions of long-term durability in structural applications, especially under moisture exposure.
Ultimately, this comparative evaluation reinforces the need for material-specific valorization strategies guided not only by sustainability metrics, but by technical feasibility, life-cycle performance, and regulatory readiness. It also suggests that hybrid approaches, combining different SRMs according to context-specific priorities, may offer a pathway toward more robust and scalable circular construction solutions.
To complement the comparative summary presented in Table 5, Figure 2 provides a visual synthesis of the relative strengths and limitations of each material stream across key performance and sustainability dimensions. This radar chart highlights the distinct profiles of the three material types: cellulose excels in insulation and environmental metrics, while metallurgy slags demonstrate superior mechanical performance and industrial maturity. Mining tailings, although versatile, show high variability and require tailored approaches depending on source and processing conditions. The values assigned to each criterion in the radar chart (ranging from 1 to 5) are qualitative and comparative, not absolute. They were derived directly from the cross-sectional analysis conducted in this section Table 5. This evaluation is based on the synthesis of the published literature included in the review.
Based on this comparative synthesis, a staggered implementation outlook can be envisioned: metallurgical slags, due to their maturity and standardization, are ready for immediate scale-up; mining tailings require site-specific pilot projects and regulatory streamlining in the medium term; while cellulose-based materials, with their strong environmental profile, should be prioritized for non-structural and insulation uses in parallel with long-term durability research.

Author Contributions

All authors contributed to the study conception and design. Material preparation, data collection and analysis were performed by D.H. and F.V. The first draft of the manuscript was written by D.H. and all authors commented on previous versions of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

Co-funded by the European Union under grant agreement No 101138646 (Increasing circularity in process industries by upcycling secondary resources–ICARUS project). Views and opinions are those of the author(s) only and do not necessarily reflect those of the European Union or the European Health and Digital Executive Agency (HADEA). Neither the European Union nor HADEA can be held responsible for them.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Domljan, I.; Domljan, V. Major challenges in the construction industry. In New Technologies, Development and Application VII; Karabegovic, I., Kovačević, A., Mandzuka, S., Eds.; Lecture Notes in Networks and Systems; Springer: Cham, Switzerland, 2024; Volume 1070. [Google Scholar] [CrossRef]
  2. Norouzi, M.; Chàfer, M.; Cabeza, L.F.; Jiménez, L.; Boer, D. Circular economy in the building and construction sector: A scientific evolution analysis. J. Build. Eng. 2021, 44, 102704. [Google Scholar] [CrossRef]
  3. Elhegazy, H.; Zhang, J.; Amoudi, O.; Zaki, J.N.; Yahia, M.; Eid, M.; Mahdi, I. An exploratory study on the impact of the construction industry on climate change. J. Ind. Integr. Manag. 2023, 9, 397–418. [Google Scholar] [CrossRef]
  4. Bulkeley, H.; Chan, S.; Fransen, A.; Landry, J.; Seddon, N.; Deprez, A.; Kok, M. Building Synergies Between Climate Biodiversity Governance: A Primer for COP28. 2023. Available online: https://shorturl.at/tCNZ0 (accessed on 28 August 2024).
  5. Husgafvel, R.; Sakaguchi, D. Circular economy development in the wood construction sector in Finland. Sustainability 2023, 15, 7871. [Google Scholar] [CrossRef]
  6. Purchase, C.K.; Al Zulayq, D.M.; O’Brien, B.T.; Kowalewski, M.J.; Berenjian, A.; Tarighaleslami, A.H.; Seifan, M. Circular economy of construction and demolition waste: A literature review on lessons, challenges, and benefits. Materials 2022, 15, 76. [Google Scholar] [CrossRef]
  7. Xu, F.; Lin, X.; Zhou, A. Effect of recycled ceramic aggregate on hydration heat and permeability of high performance concrete. Cement Concr. Compos. 2023, 137, 104930. [Google Scholar] [CrossRef]
  8. Cozzarini, L.; De Lorenzi, L.; Barago, N.; Sbaizero, O.; Bevilacqua, P. Expanded glass for thermal and acoustic insulation from recycled post-consumer glass and textile industry process waste. Materials 2023, 16, 1721. [Google Scholar] [CrossRef]
  9. Bostanci, S.C. Use of waste marble dust and recycled glass for sustainable concrete production. J. Clean. Prod. 2020, 251, 119785. [Google Scholar] [CrossRef]
  10. Papamichael, I.; Voukkali, I.; Loizia, P.; Zorpas, A.A. Construction and demolition waste framework of circular economy: A mini review. Waste Manag. Res. 2023, 41, 1728–1740. [Google Scholar] [CrossRef]
  11. Valois, M. UNECE/FAO Data Brief 2023–Pulp, Paper and Paperboard. 2024. Available online: https://unece.org/forestry-timber/documents/2023/11/informal-documents/2023-data-brief-pulp-paper-and-paperboard (accessed on 28 August 2024).
  12. Bamigboye, G.O.; Bassey, D.E.; Olukanni, D.O.; Ngene, B.U.; Adegoke, D.; Odetoyan, A.O.; Kareem, M.A.; Enabulele, D.O.; Nworgu, A.T. Waste materials in highway applications: An overview on generation and utilization implications on sustainability. J. Clean. Prod. 2021, 283, 124581. [Google Scholar] [CrossRef]
  13. Ji, R.; Liu, T.J.; Kang, L.L.; Wang, Y.T.; Li, J.G.; Wang, F.P.; Yu, Q.; Wang, X.M.; Liu, H.; Guo, H.W.; et al. A review of metallurgical slag for efficient wastewater treatment: Pretreatment, performance and mechanism. J. Clean. Prod. 2022, 380, 135076. [Google Scholar] [CrossRef]
  14. Rbihi, S.; Neggaoui, H.; Laallam, L.; Jouaiti, A. Cellulose waste for making sustainable cellulose/concrete biocomposite: Functional and environmental value for the thermal insulation. Biomass Convers. Biorefinery 2024, 14, 2333–2340. [Google Scholar] [CrossRef]
  15. Wang, J.; Wang, L.; Gardner, D.J.; Shaler, S.M.; Cai, Z. Towards a cellulose-based society: Opportunities and challenges. Cellulose 2021, 28, 4511–4543. [Google Scholar] [CrossRef]
  16. Kumar, V.; Verma, P. Pulp-paper industry sludge waste biorefinery for sustainable energy and value-added products development: A systematic valorization towards waste management. J. Environ. Manag. 2024, 352, 120052. [Google Scholar] [CrossRef] [PubMed]
  17. Kethiri, M.A.; Chikhi, M.; Belghar, N.; Chekired, F.; Guerira, B. A thermophysical investigation of the potential for the use of wastepaper for insulation building materials. Int. Wood Prod. J. 2024, 15, 63–71. [Google Scholar] [CrossRef]
  18. Khalife, E.; Sabouri, M.; Kaveh, M.; Szymanek, M. Recent advances in the application of agricultural waste in construction. Appl. Sci. 2024, 14, 2355. [Google Scholar] [CrossRef]
  19. Sharma, R.; Jha, P.; Bansal, T.; Sharma, D.; Pathak, A. Sustainable construction: Unveiling the potential of hempcrete in the modern era. Asian J. Civ. Eng. 2024, 25, 4157–4167. [Google Scholar] [CrossRef]
  20. Fode, T.A.; Jande, Y.A.C.; Kivevele, T. Physical, mechanical, and durability properties of concrete containing different waste synthetic fibers for green environment—A critical review. Heliyon 2024, 10, e32950. [Google Scholar] [CrossRef]
  21. Yang, L.; Liu, J.; Ma, X.; Hu, X.; Wu, Z.; Shi, C. Use of bauxite tailing for the production of fine lightweight aggregates. J. Clean. Prod. 2022, 372, 133603. [Google Scholar] [CrossRef]
  22. Deng, H.; Ran, C.; Liu, Y. Properties of cement thermal insulation materials containing tailing waste for connecting mines assessed using the orthogonal method with the response surface method. Processes 2023, 11, 2652. [Google Scholar] [CrossRef]
  23. Lage, G.T.; Bessa, S.A. Soil stabilization with Brazilian iron ore tailings to produce rammed earth. In International Conference on Earthen Construction; Springer Nature: Cham, Switzerland, 2024; pp. 206–215. [Google Scholar] [CrossRef]
  24. Kurniati, E.O.; Pederson, F.; Kim, H.J. Application of steel slags, ferronickel slags, and copper mining waste as construction materials: A review. Resour. Conserv. Recycl. 2023, 198, 107175. [Google Scholar] [CrossRef]
  25. Meshram, S.; Raut, S.P.; Ansari, K.; Madurwar, M.; Daniyal, M.; Khan, M.A.; Katare, V.; Khan, A.H.; Khan, N.A.; Hasan, M.A. Waste slags as sustainable construction materials: A compressive review on physico mechanical properties. J. Mater. Res. Technol. 2023, 23, 5821–5845. [Google Scholar] [CrossRef]
  26. Ramón-Álvarez, I.; Sánchez-Delgado, S.; Torres-Carrasco, M. Viability of hybrid and alkali-activated slag materials for thermal energy storage: Analysis of the evolution of mechanical and thermal properties. J. Build. Eng. 2024, 95, 110073. [Google Scholar] [CrossRef]
  27. Wang, Y.; Hu, Y.; Yang, J.; He, Y.; He, X.; Su, Y.; Strnadel, B. Performance of cement-based materials incorporating ultra-fine copper slag. Constr. Build. Mater. 2023, 402, 132949. [Google Scholar] [CrossRef]
  28. Gaus, A.; Rauf, I.; Siregar, F.; Heryanto, H. Proximity porosity and crystallinity analysis as clay/nickel slag characteristics for material stabilization application. Trends Sci. 2024, 21, 7970. [Google Scholar] [CrossRef]
  29. Ababio, B.K.; Lu, W. Barriers and enablers of circular economy in construction: A multi-system perspective towards the development of a practical framework. Constr. Manag. Econ. 2023, 41, 3–21. [Google Scholar] [CrossRef]
  30. Caro, D.; Lodato, C.; Damgaard, A.; Cristóbal, J.; Foster, G.; Flachenecker, F.; Tonini, D. Environmental and socio-economic effects of construction and demolition waste recycling in the European Union. Sci. Total Environ. 2024, 908, 168295. [Google Scholar] [CrossRef]
  31. European Commission. Construction Sector. Available online: https://ec.europa.eu/eurostat/cache/digpub/housing/bloc-3a.html?lang=en (accessed on 8 May 2025).
  32. European Commission. Internal Market, Industry, Entrepreneurship and SMEs. Available online: https://single-market-economy.ec.europa.eu/industry/sustainability/buildings-and-construction_en (accessed on 8 May 2025).
  33. Jing, L.; Shi, T.; Chang, Y.; Meng, X.; He, S.; Xu, H.; Yang, S.; Liu, J. Cellulose-based materials in environmental protection: A scientometric and visual analysis review. Sci. Total Environ. 2024, 929, 172576. [Google Scholar] [CrossRef]
  34. Zhang, Z.; Scherer, G.W. Measuring chemical shrinkage of ordinary Portland cement pastes with high water-to-cement ratios by adding cellulose nanofibrils. Cem. Concr. Compos. 2020, 111, 103625. [Google Scholar] [CrossRef]
  35. Gorzelańczyk, T.; Schabowicz, K.; Szymków, M. Tests of fiber cement materials containing recycled cellulose fibers. Materials 2020, 13, 2758. [Google Scholar] [CrossRef]
  36. Feng, H.; Zheng, Z.; Guo, A.; Sun, Z.; Yu, Z.; Yao, C.; Du, Y. Mechanical properties and microstructure of cellulose nanocrystal modified cement pastes subject to chloride erosion. J. Build. Eng. 2025, 112670. [Google Scholar] [CrossRef]
  37. Ghahari, S.; Assi, L.N.; Alsalman, A.; Alyamaç, K.E. Fracture properties evaluation of cellulose nanocrystals cement paste. Materials 2020, 13, 2507. [Google Scholar] [CrossRef] [PubMed]
  38. Xu, H.; Wang, Z.; Shao, Z.; Cai, L.; Jin, H.; Zhang, Z.; Chen, T. Experimental study on durability of fiber reinforced concrete: Effect of cellulose fiber, polyvinyl alcohol fiber and polyolefin fiber. Constr. Build. Mater. 2021, 306, 124867. [Google Scholar] [CrossRef]
  39. Arslan, C.; Gencel, O.; Borazan, I.; Sutcu, M.; Erdogmus, E. Effect of waste-based micro cellulose fiber as pore maker on characteristics of fired clay bricks. Constr. Build. Mater. 2021, 300, 124298. [Google Scholar] [CrossRef]
  40. Chopperla, K.S.T.; Ramanathan, S.; Moon, R.J.; Isgor, O.B.; Weiss, W.J. Field demonstrations using cellulose nanomaterials as concrete additives. Concr. Int. 2023, 45, 33–38. Available online: https://www.fpl.fs.usda.gov/documnts/pdf2023/fpl_2023_chopperla001.pdf (accessed on 8 May 2025).
  41. Barabanshchikov, Y.; Pham, H.; Usanova, K. Influence of microfibrillated cellulose additive on strength, elastic modulus, heat release, and shrinkage of mortar and concrete. Materials 2021, 14, 6933. [Google Scholar] [CrossRef]
  42. Ma, W.; Qin, Y.; Li, Y.; Chai, J.; Zhang, X.; Ma, Y.; Liu, H. Mechanical properties and engineering application of cellulose fiber-reinforced concrete. Mater. Today Commun. 2020, 22, 100818. [Google Scholar] [CrossRef]
  43. Nassiri, S.; Chen, Z.; Jian, G.; Zhong, T.; Haider, M.M.; Li, H.; Wolcott, M. Comparison of unique effects of two contrasting types of cellulose nanomaterials on setting time, rheology, and compressive strength of cement paste. Cem. Concr. Compos. 2021, 123, 104201. [Google Scholar] [CrossRef]
  44. de Souza, L.O.; Cordazzo, M.; de Souza, L.M.S.; Tonoli, G.; de Andrade Silva, F.; Mechtcherine, V. Investigation of dispersion methodologies of microcrystalline and nano-fibrillated cellulose on cement pastes. Cem. Concr. Compos. 2022, 126, 104351. [Google Scholar] [CrossRef]
  45. Aziz, M.A.; Zubair, M.; Saleem, M. Development and testing of cellulose nanocrystal-based concrete. Case Stud. Constr. Mater. 2021, 15, e00761. [Google Scholar] [CrossRef]
  46. de Souza, L.O.; Liebscher, M.; de Souza, L.M.S.; de Andrade Silva, F.; Mechtcherine, V. Effect of microcrystalline and nano-fibrillated cellulose on the mechanical behaviour and microstructure of cement pastes. Constr. Build. Mater. 2023, 408, 133812. [Google Scholar] [CrossRef]
  47. Choi, H.; Choi, Y.C. Setting characteristics of natural cellulose fiber reinforced cement composite. Constr. Build. Mater. 2021, 271, 121910. [Google Scholar] [CrossRef]
  48. Singh, H.; Gupta, R. Influence of cellulose fiber addition on self-healing and water permeability of concrete. Case Stud. Constr. Mater. 2020, 12, e00324. [Google Scholar] [CrossRef]
  49. Zhang, Z.; Angst, U. Microstructure and moisture transport in carbonated cement-based materials incorporating cellulose nanofibrils. Cem. Concr. Res. 2022, 162, 106990. [Google Scholar] [CrossRef]
  50. Goncalves, J.; Boluk, Y.; Bindiganavile, V. Cellulose nanofibers mitigate chloride ion ingress in cement-based systems. Cem. Concr. Compos. 2020, 114, 103780. [Google Scholar] [CrossRef]
  51. Gwon, S.; Shin, M. Rheological properties of cement pastes with cellulose microfibers. J. Mater. Res. Technol. 2021, 10, 808–818. [Google Scholar] [CrossRef]
  52. Onuaguluchi, O.; Banthia, N. Sulfate resistance of cement composites containing Nano-Fibrillated Cellulose (NFC). Cem. Concr. Compos. 2023, 135, 104831. [Google Scholar] [CrossRef]
  53. Terrones-Saeta, J.M.; Suárez-Macías, J.; Iglesias-Godino, F.J.; Corpas-Iglesias, F.A. Development of porous asphalt with bitumen emulsion, electric arc furnace slag and cellulose fibers for medium traffic roads. Minerals 2020, 10, 872. [Google Scholar] [CrossRef]
  54. Castro, M.P.P.; Effects of Cellulose Nano-Fiber as an Additive on Performance of Asphalt Binders and Mixes. South Dakota State University. 2020. Available online: https://www.proquest.com/openview/9d7cb9eca3df40f17be867250d12623f/1?pq-origsite=gscholar&cbl=18750&diss=y (accessed on 8 May 2025).
  55. Guha, A.H.; Assaf, G.J. Effects of different fillers on pavement deformation of hot mix asphalt in hot climates. Constr. Build. Mater. 2020, 261, 119898. [Google Scholar] [CrossRef]
  56. Li, M.; Zheng, F.; Xiao, Y.; Guan, Y.; Wang, J.; Zhen, Q.; Yu, Y. Utilization of residual heat to prepare high performance foamed glass-ceramic from blast furnace slag and its reinforce mechanism. Process Saf. Environ. Prot. 2021, 156, 391–404. [Google Scholar] [CrossRef]
  57. Aljubory, A.; Teama, Z.T.; Salman, H.T.; Abd Alkareem, H.M. Effects of cellulose fibers on the properties of asphalt mixtures. Mater. Today Proc. 2021, 42, 2941–2947. [Google Scholar] [CrossRef]
  58. Ghabchi, R.; Castro, M.P.P. Effect of laboratory-produced cellulose nanofiber as an additive on performance of asphalt binders and mixes. Constr. Build. Mater. 2021, 286, 122922. [Google Scholar] [CrossRef]
  59. Nambiar, R.A.; Haridharan, M.K. Mechanical and durability study of high performance concrete with addition of natural fiber (jute). Mater. Today Proc. 2021, 46, 4941–4947. [Google Scholar] [CrossRef]
  60. Zhou, C.; Cai, L.; Chen, Z.; Li, J. Effect of kenaf fiber on mechanical properties of high-strength cement composites. Constr. Build. Mater. 2020, 263, 121007. [Google Scholar] [CrossRef]
  61. Guo, A.; Sun, Z.; Satyavolu, J. Impact of modified kenaf fibers on shrinkage and cracking of cement pastes. Constr. Build. Mater. 2020, 264, 120230. [Google Scholar] [CrossRef]
  62. Candamano, S.; Crea, F.; Coppola, L.; De Luca, P.; Coffetti, D. Influence of acrylic latex and pre-treated hemp fibers on cement based mortar properties. Constr. Build. Mater. 2021, 273, 121720. [Google Scholar] [CrossRef]
  63. Affan, M.; Ali, M. Experimental investigation on mechanical properties of jute fiber reinforced concrete under freeze-thaw conditions for pavement applications. Constr. Build. Mater. 2022, 323, 126599. [Google Scholar] [CrossRef]
  64. Sáez-Pérez, M.P.; Brümmer, M.; Durán-Suárez, J.A. Effect of the state of conservation of the hemp used in geopolymer and hydraulic lime concretes. Constr. Build. Mater. 2021, 285, 122853. [Google Scholar] [CrossRef]
  65. Gencel, O.; Bayraktar, O.Y.; Kaplan, G.; Benli, A.; Martinez-Barrera, G.; Brostow, W.; Bodur, B. Characteristics of hemp fiber reinforced foam concretes with fly ash and Taguchi optimization. Constr. Build. Mater. 2021, 294, 123607. [Google Scholar] [CrossRef]
  66. Poletanovic, B.; Dragas, J.; Ignjatovic, I.; Komljenovic, M.; Merta, I. Physical and mechanical properties of hemp fiber reinforced alkali-activated fly ash and fly ash/slag mortars. Constr. Build. Mater. 2020, 259, 119677. [Google Scholar] [CrossRef]
  67. Juradin, S.; Boko, I.; Grubeša, I.N.; Jozić, D.; Mrakovčić, S. Influence of different treatment and amount of Spanish broom and hemp fibers on the mechanical properties of reinforced cement mortars. Constr. Build. Mater. 2021, 273, 121702. [Google Scholar] [CrossRef]
  68. Poletanovic, B.; Janotka, I.; Janek, M.; Bacuvcik, M.; Merta, I. Influence of the NaOH-treated hemp fibers on the properties of fly-ash based alkali-activated mortars prior and after wet/dry cycles. Constr. Build. Mater. 2021, 309, 125072. [Google Scholar] [CrossRef]
  69. Kouta, N.; Saliba, J.; Saiyouri, N. Effect of flax fibers on early age shrinkage and cracking of earth concrete. Constr. Build. Mater. 2020, 254, 119315. [Google Scholar] [CrossRef]
  70. Rahimi, M.; Hisseine, O.A.; Tagnit-Hamou, A. Effectiveness of treated flax fibers in improving the early age behavior of high-performance concrete. J. Build. Eng. 2022, 45, 103448. [Google Scholar] [CrossRef]
  71. Page, J.; Amziane, S.; Gomina, M.; Djelal, C.; Audonnet, F. Using linseed oil as flax fiber coating for fiber-reinforced cementitious composite. Ind. Crops Prod. 2021, 161, 113168. [Google Scholar] [CrossRef]
  72. Rahimi, M.; Omran, A.; Tagnit-Hamou, A. Role of homogenization and surface treatment of flax fiber on performance of cement-based composites. Clean. Mater. 2022, 3, 100037. [Google Scholar] [CrossRef]
  73. Zhang, D.; Tan, G.Y.; Tan, K.H. Combined effect of flax fibers and steel fibers on spalling resistance of ultra-high-performance concrete at high temperature. Cem. Concr. Compos. 2021, 121, 104067. [Google Scholar] [CrossRef]
  74. Kouta, N.; Saliba, J.; Saiyouri, N. Fracture behavior of flax fibers reinforced earth concrete. Eng. Fract. Mech. 2021, 241, 107378. [Google Scholar] [CrossRef]
  75. Page, J.; Khadraoui, F.; Gomina, M.; Boutouil, M. Enhancement of the long-term mechanical performance of flax fiber-reinforced cementitious composites by using alternative binders. J. Build. Eng. 2021, 40, 102323. [Google Scholar] [CrossRef]
  76. Czajkowski, Ł.; Kocewicz, R.; Weres, J.; Olek, W. Estimation of thermal properties of straw-based insulating panels. Materials 2022, 15, 1073. [Google Scholar] [CrossRef]
  77. Parmar, N.; Vats, A.; Devi, V.; Kwatra, S.; Srivastava, R.K.; Singh, S.B. Assessing sustainability and performance of hempcrete. Int. J. Environ. Clim. Change 2024, 14, 609–622. [Google Scholar] [CrossRef]
  78. Walker, R.; Pavia, S. Moisture transfer and thermal properties of hemp–lime concretes. Constr. Build. Mater. 2014, 64, 270–276. [Google Scholar] [CrossRef]
  79. Ayati, B.; Gutierrez, A.; Chandler, A. Mechanical and hygrothermal properties of hemp-silica bio-composites. Constr. Build. Mater. 2024, 425, 136077. [Google Scholar] [CrossRef]
  80. Fernando, S.; Gunasekara, C.; Shahpasandi, A.; Nguyen, K.; Sofi, M.; Setunge, S.; Mendis, P.; Rahman, M.T. Sustainable cement composite integrating waste cellulose fibre: A comprehensive review. Polymers 2023, 15, 520. [Google Scholar] [CrossRef] [PubMed]
  81. Rubino, C.; Liuzzi, S.; Stefanizzi, P.; Martellotta, F. Characterization of sustainable building materials obtained from textile waste: From laboratory prototypes to real-world manufacturing processes. J. Clean. Prod. 2023, 390, 136098. [Google Scholar] [CrossRef]
  82. Abdulmunem, A.R.; Hussein, N.F.; Samin, P.M.; Sopian, K.; Hussien, H.A.; Ghazali, H. Integration of recycled waste paper with phase change material in building enclosure. J. Energy Storage 2023, 64, 107140. [Google Scholar] [CrossRef]
  83. Sangmesh, B.; Patil, N.; Jaiswal, K.K.; Gowrishankar, T.P.; Selvakumar, K.K.; Jyothi, M.S.; Jyothilakshmi, R.; Kumar, S. Development of sustainable alternative materials for the construction of green buildings using agricultural residues: A review. Constr. Build. Mater. 2023, 368, 130457. [Google Scholar] [CrossRef]
  84. Ayed, R.; Bouadila, S.; Skouri, S.; Boquera, L.; Cabeza, L.F.; Lazaar, M. Recycling textile waste to enhance building thermal insulation and reduce carbon emissions: Experimentation and model-based dynamic assessment. Buildings 2023, 13, 535. [Google Scholar] [CrossRef]
  85. Petcu, C.; Hegyi, A.; Stoian, V.; Dragomir, C.S.; Ciobanu, A.A.; Lăzărescu, A.V.; Florean, C. Research on thermal insulation performance and impact on indoor air quality of cellulose-based thermal insulation materials. Materials 2023, 16, 5458. [Google Scholar] [CrossRef]
  86. Song, M.; Ripoll, K. A comparative study: Exploring diverse paper-based fabrics for innovative soundproofing solutions. J. Stud. Res. 2023, 12, 1–8. [Google Scholar] [CrossRef]
  87. Jefferson Andrew, J.; Sain, M.; Ramakrishna, S.; Jawaid, M.; Dhakal, H.N. Environmentally friendly fire retardant natural fibre composites: A review. Int. Mater. Rev. 2024, 69, 267–308. [Google Scholar] [CrossRef]
  88. Wang, Z.; Wang, D. Can paper waste be utilised as an insulation material in response to the current crisis. Sustainability 2023, 15, 15939. [Google Scholar] [CrossRef]
  89. Ali, M.; Al-Suhaibani, Z.; Almuzaiqer, R.; Albahbooh, A.; Al-Salem, K.; Nuhait, A. New composites derived from the natural fiber polymers of discarded date palm surface and pineapple leaf fibers for thermal insulation and sound absorption. Polymers 2024, 16, 1002. [Google Scholar] [CrossRef] [PubMed]
  90. Rodríguez Neira, K.; Cárdenas-Ramírez, J.P.; Rojas-Herrera, C.J.; Haurie, L.; Lacasta, A.M.; Torres Ramo, J.; Sánchez-Ostiz, A. Assessment of elaboration and performance of rice husk-based thermal insulation material for building applications. Buildings 2024, 14, 1720. [Google Scholar] [CrossRef]
  91. Di Canto, J.A.; Malfait, W.J.; Wernery, J. Turning waste into insulation–a new sustainable thermal insulation board based on wheat bran and banana peels. Build. Environ. 2023, 244, 110740. [Google Scholar] [CrossRef]
  92. Babenko, M.; Estokova, A.; Unčik, S.; Savytskyi, M. Comparative study of lightweight concretes based on hemp and flax straw. Slovak. J. Civ. Eng. 2022, 30, 11–16. [Google Scholar] [CrossRef]
  93. Agnihotri, S.; Sheikh, J.N.; Singh, S.P.; Behera, B.K. Flame-retardant textile structural composites for construction application: A review. J. Mater. Sci. 2024, 59, 1788–1818. [Google Scholar] [CrossRef]
  94. Samardzioska, T.; Jovanoska, M.; Grujoska, V. Sustainable thermal insulation derived from recycled textile waste. Adv. Environ. Eng. Res. 2023, 4, 1–8. [Google Scholar] [CrossRef]
  95. Kamaruzzaman, A.S.; Jusoh, W.A.W.; Zin, F.S.; Roslan, N.H. A study on thermal insulation of cement bricks incorporate with cellulose fiber. Prog. Eng. Appl. Technol. 2023, 4, 701–710. [Google Scholar] [CrossRef]
  96. Salahuddin, B.; Sha, Z.; Aziz, S.; Faisal, S.N.; Islam, M.S. Durability against fatigue and moisture of natural fibre composite. Sustain. Nat. Fiber Compos. 2022, 122, 128–153. [Google Scholar] [CrossRef]
  97. Wang, R.; Jin, P.; Dong, H.; Liu, Y.; Ding, Z.; Zhang, W. Effect of moist content on the bio-carbonated steel slag bricks. Constr. Build. Mater. 2021, 269, 121313. [Google Scholar] [CrossRef]
  98. Ahmad, J.; Zhou, Z. Mechanical properties of natural as well as synthetic fiber reinforced concrete: A review. Constr. Build. Mater. 2022, 333, 127353. [Google Scholar] [CrossRef]
  99. Dias, S.; Almeida, J.; Tadeu, A.; de Brito, J. Alternative concrete aggregates—Review of physical and mechanical properties and successful applications. Cem. Concr. Compos. 2024, 152, 105663. [Google Scholar] [CrossRef]
  100. Kilani, A.; Fapohunda, C.; Adeleke, O.; Metiboba, C. Evaluating the effects of agricultural wastes on concrete and composite mechanical properties: A review. Res. Eng. Struct. Mater. 2021, 8, 307–336. [Google Scholar] [CrossRef]
  101. Chabannes, M.; Bénézet, J.C.; Clerc, L.; Garcia-Diaz, E. Use of raw rice husk as natural aggregate in a lightweight insulating concrete: An innovative application. Constr. Build. Mater. 2014, 70, 428–438. [Google Scholar] [CrossRef]
  102. Marques, B.; Tadeu, A.; Almeida, J.; António, J. Experimental characterisation of cement-based composites with rice husk. Int. J. Des. Nat. Ecodynamics 2019, 14, 147–153. [Google Scholar] [CrossRef]
  103. Maruthupandian, S.; Chaliasou, N.A.; Kanellopoulos, A. Recycling mine tailings for a sustainable future built environment. In Energy and Sustainable Futures: Proceedings of 2nd ICESF 2020; Springer International Publishing: Cham, Switzerland, 2021. [Google Scholar] [CrossRef]
  104. Araujo, F.S.; Taborda-Llano, I.; Nunes, E.B.; Santos, R.M. Recycling and reuse of mine tailings: A review of advancements and their implications. Geosciences 2022, 12, 319. [Google Scholar] [CrossRef]
  105. Amrani, M.; Taha, Y.; El Haloui, Y.; Benzaazoua, M.; Hakkou, R. Sustainable reuse of coal mine waste: Experimental and economic assessments for embankments and pavement layer applications in Morocco. Minerals 2020, 10, 851. [Google Scholar] [CrossRef]
  106. Susanto, I.; Fransisco, S.; Firdaus, Y.; Ronny, Y. Utilization of tailing waste as aggregate for road pavement. IOP Conf. Ser. Earth Environ. Sci. 2021, 802, 012020. [Google Scholar] [CrossRef]
  107. Tong, J.; Niu, X.; Wang, Y.; Lu, Y. Strength characteristics of iron tailings blended soil as a road base material. Appl. Sci. 2021, 11, 7587. [Google Scholar] [CrossRef]
  108. Mashifana, T.; Sithole, T. Clean production of sustainable backfill material from waste gold tailings and slag. J. Clean. Prod. 2021, 308, 127357. [Google Scholar] [CrossRef]
  109. Solouki, A.; Fathollahi, A.; Viscomi, G.; Tataranni, P.; Valdrè, G.; Coupe, S.J.; Sangiorgi, C. Thermally treated waste silt as filler in geopolymer cement. Materials 2021, 14, 5102. [Google Scholar] [CrossRef] [PubMed]
  110. Solouki, A.; Tataranni, P.; Sangiorgi, C. Mixture optimization of concrete paving blocks containing waste silt. Sustainability 2022, 14, 451. [Google Scholar] [CrossRef]
  111. Solouki, A.; Viscomi, G.; Tataranni, P.; Sangiorgi, C. Preliminary evaluation of cement mortars containing waste silt optimized with the design of experiments method. Materials 2021, 14, 528. [Google Scholar] [CrossRef]
  112. Solouki, A.; Tataranni, P.; Tarsi, G.; Sangiorgi, C. Waste silt as filler in hot mix asphalt: A laboratory characterization. Appl. Sci. 2023, 13, 3473. [Google Scholar] [CrossRef]
  113. Ipekyol, A.; Tortum, A.; Rasouli, R.; Yazdani, M. Evaluating fatigue and crack resistance of asphalt mixture containing zinc tailing aggregates. Case Stud. Constr. Mater. 2022, 17, e01384. [Google Scholar] [CrossRef]
  114. Ikotun, J.; Adeyeye, R.; Otieno, M. Application of mine tailings sand as construction material—A review. MATEC Web Conf. 2022, 364, 05008. [Google Scholar] [CrossRef]
  115. Cao, L.; Zhou, J.; Zhou, T.; Dong, Z.; Tian, Z. Utilization of iron tailings as aggregates in paving asphalt mixture: A sustainable and eco-friendly solution for mining waste. J. Clean. Prod. 2022, 375, 134126. [Google Scholar] [CrossRef]
  116. Baldo, N.; Rondinella, F.; Daneluz, F.; Pasetto, M. Foamed bitumen mixtures for road construction made with 100% waste materials: A laboratory study. Sustainability 2022, 14, 6056. [Google Scholar] [CrossRef]
  117. Calandra, P.; Quaranta, S.; Figueira, B.A.M.; Caputo, P.; Porto, M.; Rossi, C.O. Mining wastes to improve bitumen performances: An example of circular economy. J. Colloid. Interface Sci. 2022, 614, 277–287. [Google Scholar] [CrossRef]
  118. Sá, T.S.W.; Oda, S.; Machado, V.K.C.B.L.; Toledo Filho, R.D. Use of iron ore tailings and sediments on pavement structure. Constr. Build. Mater. 2022, 342, 128072. [Google Scholar] [CrossRef]
  119. Wei, Z.; Jia, Y.; Wang, S.; Li, Z.; Li, Y.; Wang, X.; Gao, Y. Utilization of iron ore tailing as an alternative mineral filler in asphalt mastic: High-temperature performance and environmental aspects. J. Clean. Prod. 2022, 335, 130318. [Google Scholar] [CrossRef]
  120. Xiao, Y.; Ju, X.; Li, C.; Wang, T.; Wu, R. Research on recycling of phosphorus tailings powder in open-graded friction course asphalt concrete. Materials 2023, 16, 2000. [Google Scholar] [CrossRef] [PubMed]
  121. Singh, S.; Kumar, A.; Sitharam, T.G. Experimental study on strength, durability, hydraulic and toxicity characteristics of soil treated with mine tailings based geopolymers for sustainable road subgrade application. Constr. Build. Mater. 2024, 414, 134894. [Google Scholar] [CrossRef]
  122. Lei, B.; Li, X.; Guo, Y.; Qu, F.; Zhao, C.; Tam, V.W.; Li, W. Recycling of copper tailing as filler material in asphalt paving mastic: A sustainable solution for mining waste recovery. Case Stud. Constr. Mater. 2024, 20, e03237. [Google Scholar] [CrossRef]
  123. Li, J.; Sha, A.; Wang, Z.; Song, R.; Cao, Y. Investigation of the self-healing, road performance and cost–benefit effects of an iron tailing/asphalt mixture in pavement. Constr. Build. Mater. 2024, 422, 135788. [Google Scholar] [CrossRef]
  124. Vargas, F.; Rigamonti, L. Environmental evaluation of treated tailing as supplementary cementitious material. Procedia CIRP 2020, 90, 280–284. [Google Scholar] [CrossRef]
  125. Yao, G.; Wang, Q.; Wang, Z.; Wang, J.; Lyu, X. Activation of hydration properties of iron ore tailings and their application as supplementary cementitious materials in cement. Powder Technol. 2020, 360, 863–871. [Google Scholar] [CrossRef]
  126. Magalhaes, L.F.; França, S.; dos Santos Oliveira, M.; Peixoto, R.A.F.; Bessa, S.A.L.; da Silva Bezerra, A.C. Iron ore tailings as a supplementary cementitious material in the production of pigmented cements. J. Clean. Prod. 2020, 274, 123260. [Google Scholar] [CrossRef]
  127. Bezerra, C.G.; Rocha, C.A.A.; de Siqueira, I.S.; Toledo Filho, R.D. Feasibility of iron-rich ore tailing as supplementary cementitious material in cement pastes. Constr. Build. Mater. 2021, 303, 124496. [Google Scholar] [CrossRef]
  128. Zhang, Y.; Liu, B.; Gu, X.; Nehdi, M.L.; Zhang, L.V. Mechanochemical activation of iron ore tailing-based ternary supplementary cementitious materials. Constr. Build. Mater. 2022, 346, 128420. [Google Scholar] [CrossRef]
  129. Zhang, Y.; Yang, D.; Gu, X.; Chen, H.; Li, Z. Application of iron tailings-based composite supplementary cementitious materials (SCMs) in green concrete. Materials 2022, 15, 3866. [Google Scholar] [CrossRef] [PubMed]
  130. Wang, Y.; Li, Z.; Jin, Q.; Zhang, M.; Zhou, Z. High-efficiency utilization of limestone tailings: Used as cementitious materials and fine aggregate to prepare karst structure filling material. Constr. Build. Mater. 2022, 316, 125841. [Google Scholar] [CrossRef]
  131. Ghazi, A.B.; Jamshidi-Zanjani, A.; Nejati, H. Utilization of copper mine tailings as a partial substitute for cement in concrete construction. Constr. Build. Mater. 2022, 317, 125921. [Google Scholar] [CrossRef]
  132. Yun-Hong, C.; Si-hui, Y.; Jing-yu, Z.; Xiao-hui, S. Test research on hydration process of cement-iron tailings powdercomposite cementitious materials. Powder Technol. 2022, 399, 117215. [Google Scholar] [CrossRef]
  133. Duarte, M.S.; Almada, B.S.; dos Santos, W.J.; Bessa, S.A.L.; da Silva Bezerra, A.C.; Aguilar, M.T.P. Influence of mechanical treatment and magnetic separation on the performance of iron ore tailings as supplementary cementitious material. J. Build. Eng. 2022, 59, 105099. [Google Scholar] [CrossRef]
  134. Zhang, K.; Xu, H.; Kong, X.X.; Zhang, C.Y.; Lu, H.F.; He, D. Study on the influence mechanism of micro-mechanical properties of heterogeneous geopolymer gels. J. Build. Eng. 2023, 76, 107164. [Google Scholar] [CrossRef]
  135. Singh, A.; Wang, Y.; Zhou, Y.; Sun, J.; Xu, X.; Li, Y.; Wang, X. Utilization of antimony tailings in fiber-reinforced 3D printed concrete: A sustainable approach for construction materials. Constr. Build. Mater. 2023, 408, 133689. [Google Scholar] [CrossRef]
  136. Chen, B.; Pang, L.; Zhou, Z.; Chang, Q.; Fu, P. Study on the hydration properties of a ternary cementitious material system containing activated gold tailings and granulated blast furnace slag. J. Build. Eng. 2023, 63, 105574. [Google Scholar] [CrossRef]
  137. Sepúlveda-Vásquez, C.; Carrasco-Astudillo, N.; Muñoz, L.; Molina, P.; Ringuedé, A.; Guerra, C.; Sancy, M. Electrochemical analysis of carbon steel embedded in mortars with pretreated copper tailings as supplementary cementitious material. Int. J. Electrochem. Sci. 2024, 19, 100584. [Google Scholar] [CrossRef]
  138. Hernández-Ramos, S.M.; Trejo-Arroyo, D.L.; Cholico-González, D.F.; Rodríguez-Torres, G.M.; Zárate-Medina, J.; Vega-Azamar, R.E.; Ortíz-Lara, N. Characterization and effect of mechanical and thermal activation in mining tailings for use as supplementary cementitious material. Case Stud. Constr. Mater. 2024, 20, e02770. [Google Scholar] [CrossRef]
  139. Hu, Z.; Gu, X.; Cheng, B.; Wang, Q.; Liu, J.; Ge, X.; Yin, S. The role of chemical activation in strengthening iron ore tailings supplementary cementitious materials. Buildings 2024, 14, 963. [Google Scholar] [CrossRef]
  140. Hu, Z.; Gu, X.; Liu, J.; Ge, X.; Wang, S.; Yin, S. Mechanical properties and hydration mechanism of iron tailings–cement-based supplementary cementitious materials. Buildings 2024, 14, 1044. [Google Scholar] [CrossRef]
  141. Fernandes, J.V.; Guedes, D.G.; da Costa, F.P.; Rodrigues, A.M.; Neves, G.D.A.; Menezes, R.R.; Santana, L.N.L.D.L. Sustainable ceramic materials manufactured from ceramic formulations containing quartzite and scheelite tailings. Sustainability 2020, 12, 9417. [Google Scholar] [CrossRef]
  142. Figueirêdo, J.M.R.D.; Costa, F.P.D.; Fernandes, J.V.; Rodrigues, A.M.; Neves, G.D.A.; Menezes, R.R.; Santana, L.N.L.D.L. Development of scheelite tailings-based ceramic formulations with the potential to manufacture porcelain tiles, semi-stoneware and stoneware. Materials 2020, 13, 5122. [Google Scholar] [CrossRef]
  143. Suvorova, O.V.; Selivanova, E.A.; Mikhailova, J.A.; Masloboev, V.A.; Makarov, D.V. Ceramic products from mining and metallurgical waste. Appl. Sci. 2020, 10, 3515. [Google Scholar] [CrossRef]
  144. Baziz, A.; Bouzidi, N.; Eliche-Quesada, D. Recycling of gold mining reject from Amesmessa mine as ceramic raw material: Microstructure and mechanical properties. Environ. Sci. Pollut. Res. 2021, 28, 46738–46747. [Google Scholar] [CrossRef]
  145. Paiva, H.; Simões, F.; Maljaee, H.; Yliniemi, J.; Illikainen, M.; Ferreira, V.M. Production of ceramic construction materials as an environmental management solution for sulfidic mine tailings. SN Appl. Sci. 2021, 3, 1–10. [Google Scholar] [CrossRef]
  146. Almeida, E.P.; Carreiro, M.E.A.; Rodrigues, A.M.; Ferreira, H.S.; Santana, L.N.L.D.L.; Menezes, R.R.; Neves, G.A. A new eco-friendly mass formulation based on industrial mining residues for the manufacture of ceramic tiles. Ceram. Int. 2021, 47, 11340–11348. [Google Scholar] [CrossRef]
  147. Simão, F.V.; Chambart, H.; Vandemeulebroeke, L.; Cappuyns, V. Incorporation of sulphidic mining waste material in ceramic roof tiles and blocks. J. Geochem. Explor. 2021, 225, 106741. [Google Scholar] [CrossRef]
  148. Simão, F.V.; Chambart, H.; Vandemeulebroeke, L.; Nielsen, P.; Cappuyns, V. Turning mine waste into a ceramic resource: Plombières tailing case. J. Sustain. Metall. 2021, 7, 1469–1482. [Google Scholar] [CrossRef]
  149. Ferreira, J.F.; Costa, F.P.D.; Borborema, L.F.D.; Arimateia, R.R.; Leite, R.S.; Apolinário, R.C.; Menezes, R.R. Incorporation of bentonite mining waste in ceramic formulations for the manufacturing of porcelain stoneware. Sustainability 2022, 14, 15973. [Google Scholar] [CrossRef]
  150. Lam, E.J.; Zetola, V.; Ramírez, Y.; Montofré, Í.L.; Pereira, F. Making paving stones from copper mine tailings as aggregates. Int. J. Environ. Res. Public Health 2020, 17, 2448. [Google Scholar] [CrossRef] [PubMed]
  151. Zhou, Y.; Duan, X.; Chen, T.; Yan, B.; Li, L. Mechanical properties and toxicity risks of lead-zinc sulfide tailing-based construction materials. Materials 2021, 14, 2940. [Google Scholar] [CrossRef] [PubMed]
  152. Cechin, L.; Mymrine, V.; Avanci, M.A.; Povaluk, A.E. Ceramics composites from iron ore tailings and blast furnace slag. Ceram. Int. 2022, 48, 10506–10515. [Google Scholar] [CrossRef]
  153. Huang, H.; Li, S.; Gou, H.; Zhang, N.; Liu, L. Efficient recovery of feldspar, quartz, and kaolin from weathered granite. Minerals 2024, 14, 300. [Google Scholar] [CrossRef]
  154. Lv, X.; Lin, Y.; Chen, X.; Shi, Y.; Liang, R.; Wang, R.; Peng, Z. Environmental impact, durability performance, and interfacial transition zone of iron ore tailings utilized as dam concrete aggregates. J. Clean. Prod. 2021, 292, 126068. [Google Scholar] [CrossRef]
  155. Vallejos, C.M.C.; Rodríguez, J.S.T.; Janet, L.; Reyes, C.; Serrano, J.C.E.; Lindo, M.A.S.; Giraldo, R.D.L.; Huerta, D.M.; González, C.G.P. Copper and zinc mining tailings as a partial replacement for cement to increase the compressive strength of concrete. Int. J. Membr. Sci. Technol. 2023, 10, 1543–1550. [Google Scholar] [CrossRef]
  156. Junior, P.; Franks, D.M. Construction materials and sustainable development. In Routledge Handbook of the Extractive Industries and Sustainable Development; Routledge: Abingdon-on-Thames, UK, 2022; pp. 228–246. [Google Scholar]
  157. Adiguzel, D.; Tuylu, S.; Eker, H. Utilization of tailings in concrete products: A review. Constr. Build. Mater. 2022, 360, 129574. [Google Scholar] [CrossRef]
  158. Gou, M.; Zhou, L.; Then, N.W.Y. Utilization of tailings in cement and concrete: A review. Sci. Eng. Compos. Mater. 2019, 26, 449–464. [Google Scholar] [CrossRef]
  159. Jian, S.; Gao, W.; Lv, Y.; Tan, H.; Li, X.; Li, B.; Huang, W. Potential utilization of copper tailings in the preparation of low heat cement clinker. Constr. Build. Mater. 2020, 252, 119130. [Google Scholar] [CrossRef]
  160. Faria, N.C.; Pereira, V.M.; Nobre, T.R.; Romano, R.C.D.O.; Coelho, A.C.V.; Angulo, S.C. Clinker Portland with iron ore tailing and its characterization by integrated laboratory methods. Constr. Build. Mater. 2023, 402, 132958. [Google Scholar] [CrossRef]
  161. Wang, Z.; Wang, Q.; Zhao, W.; Xia, C.; Tian, X.; Jiang, Y.; Zhou, X.; Chen, G.; Wang, L.; Chen, M. Influence of carlin-type gold mine tailings addition on the synthesis temperature, alkali-resistant performance, and hydration mechanism of Portland cement. Constr. Build. Mater. 2022, 359, 129458. [Google Scholar] [CrossRef]
  162. Da, Y.; He, T.; Shi, C. Unveiling the cooperation effects of fluorine and copper on tricalcium silicate (C3S) during cement kiln co-processing hazardous wastes containing Cu. Constr. Build. Mater. 2022, 337, 127612. [Google Scholar] [CrossRef]
  163. Barbosa, P.D.C.; Mau, J.P.M.; Cheloni, L.M.D.M.S.; Cota, T.G.; Reis, É.L. Briquetting of manganese ore fines with cement as binder. REM-Int. Eng. J. 2024, 77, e230047. [Google Scholar] [CrossRef]
  164. Tian, X.; Wang, C.; Asrah, H.; Han, L.C.; Zheng, Y. Investigating the grinding characteristics of vanadium-titanium iron ore tailings for sustainable utilization in cementitious material preparation. Alex. Eng. J. 2024, 101, 78–89. [Google Scholar] [CrossRef]
  165. Tai, W.; Tong, X.; Zhang, K.; Luo, Z. Effect of Iron Tailings Fine Powder as Admixture on the Properties of Cement Mortar and Concrete. In Advanced Construction Technology and Research of Deep-Sea Tunnels; Springer Nature: Singapore, 2024; pp. 283–300. [Google Scholar]
  166. Eid, M.S.; Saleh, H.M. Characterizations of cement and modern sustainable concrete incorporating different waste additives. In Sustainability of Concrete with Synthetic and Recycled Aggregates, 1st ed.; IntechOpen: London, UK, 2021; pp. 3–18. [Google Scholar]
  167. Pell, R.; Tijsseling, L.; Goodenough, K.; Wall, F.; Dehaine, Q.; Grant, A.; Deak, D.; Yan, X.; Whattoff, P. Towards sustainable extraction of technology materials through integrated approaches. Nat. Rev. Earth Environ. 2021, 2, 665–679. [Google Scholar] [CrossRef]
  168. DeCeanne, A.V.; Rodrigues, L.R.; Wilkinson, C.J.; Mauro, J.C.; Zanotto, E.D. Examining the role of nucleating agents within glass-ceramic systems. J. Non Cryst. Solids 2022, 591, 121714. [Google Scholar] [CrossRef]
  169. Smirnov, A.D.; Kholodkova, A.A.; Rybalchenko, V.V.; Tarasovskii, V.P. Porous alumina-bentonite ceramics: Effects of fillers and molding technique. Ceramics 2023, 6, 132–145. [Google Scholar] [CrossRef]
  170. Khanam, J.; Hasan, M.R.; Biswas, B.; Jahan, S.A.; Sharmin, N.; Ahmed, S.; Al-Reza, S.M. Development of ceramic grade red iron oxide pigment from waste iron source. Heliyon 2023, 9, e12854. [Google Scholar] [CrossRef]
  171. Ozturk, S. Optimization of thermal conductivity and lightweight properties of clay bricks. Eng. Sci. Technol. Int. J. 2023, 48, 101566. [Google Scholar] [CrossRef]
  172. Yilmaz, E.; Koohestani, B.; Cao, S. Recent practices in mine tailings’ recycling and reuse. In Managing Mining and Minerals Processing Wastes: Concepts, Design and Applications, 1st ed.; Elsevier: Amsterdam, The Netherlands, 2023; pp. 271–304. [Google Scholar] [CrossRef]
  173. Hua, S.; Wu, D.; Wu, J.; Li, S.; Liu, G.; Pei, D. Characterization of the physical chemistry properties of iron-tailing-based ceramsite. Molecules 2023, 28, 2258. [Google Scholar] [CrossRef] [PubMed]
  174. Kinnunen, P.; Ismailov, A.; Solismaa, S.; Sreenivasan, H.; Räisänen, M.L.; Levänen, E.; Illikainen, M. Recycling mine tailings in chemically bonded ceramics—A review. J. Clean. Prod. 2018, 174, 634–649. [Google Scholar] [CrossRef]
  175. Kang, X.; Li, Y.; Li, W.; Zhou, Y.; Cui, J.; Cai, B.; Zi, Y.; Fang, J.; Chen, Y.; Li, S.; et al. Synthesis and characterization of sustainable eco-friendly alkali-activated high-content iron ore tailing bricks. Buildings 2023, 13, 2743. [Google Scholar] [CrossRef]
  176. Li, W.; Lei, G.; Xu, Y.; Huang, Q. The properties and formation mechanisms of eco-friendly brick building materials fabricated from low-silicon iron ore tailings. J. Clean. Prod. 2018, 204, 685–692. [Google Scholar] [CrossRef]
  177. Thejas, H.K.; Hossiney, N. A short review on environmental impacts and application of iron ore tailings in development of sustainable eco-friendly bricks. Mater. Today Proc. 2022, 61, 327–331. [Google Scholar] [CrossRef]
  178. Chen, H.; Wang, Z.; Pei, L.; Xue, Z.; Yu, C.; Fan, C. Fabrication of baking-free bricks from iron ore tailings. Curr. Mater. Sci. 2020, 13, 97–110. [Google Scholar] [CrossRef]
  179. Nikvar-Hassani, A.; Hodges, R.; Zhang, L. Production of green bricks from low-reactive copper mine tailings: Durability and environmental aspects. Constr. Build. Mater. 2022, 337, 127571. [Google Scholar] [CrossRef]
  180. Huang, X.; Tian, Y.; Jiang, J.; Lu, X.; He, Z.; Jia, K. Mechanical properties and enhancement mechanism of iron ore tailings as aggregate for manufacturing ultra-high performance geopolymer concrete. Constr. Build. Mater. 2024, 439, 137362. [Google Scholar] [CrossRef]
  181. Das, D.; Gołąbiewska, A.; Rout, P.K. Geopolymer bricks: The next generation of construction materials for sustainable environment. Constr. Build. Mater. 2024, 445, 137876. [Google Scholar] [CrossRef]
  182. Zhang, Y.; Zhang, L.; Wang, Q.; Han, D.; Li, Z. Iron ore tailings, phosphate slags, and lithium slags as ternary supplementary cementitious materials for concrete: Study on compression strength and microstructure. Mater. Today Commun. 2023, 36, 106644. [Google Scholar] [CrossRef]
  183. Zhang, J.; Sun, N.; Huo, Z.; Chen, J. Understanding the synergistic geopolymerization mechanism of multiple solid wastes in ternary geopolymers. J. Build. Eng. 2024, 95, 110295. [Google Scholar] [CrossRef]
  184. Kriven, W.M.; Leonelli, C.; Provis, J.L.; Boccaccini, A.R.; Attwell, C.; Ducman, V.S.; Ferone, C.; Rossignol, S.; Luukkonen, T.; Van Deventer, J.S.; et al. Why geopolymers and alkali-activated materials are key components of a sustainable world: A perspective contribution. J. Am. Ceram. Soc. 2024, 107, 5159–5177. [Google Scholar] [CrossRef]
  185. Palma, G.; Bolaños, H.; Huamani, R.; Clements, C.; Hedayat, A. Optimization of geopolymers for sustainable management of mine tailings: Impact on mechanical, microstructural and toxicological properties. Minerals 2024, 14, 997. [Google Scholar] [CrossRef]
  186. Manaviparast, H.R.; Miranda, T.; Pereira, E.; Cristelo, N. A comprehensive review on mine tailings as a raw material in the alkali activation process. Appl. Sci. 2024, 14, 5127. [Google Scholar] [CrossRef]
  187. Wang, H.; Wang, L.; Ren, Z.; Liu, S. Optimizing raw material composition and alkali content in low silicon-aluminum tailings-based geopolymers. Constr. Build. Mater. 2023, 404, 133339. [Google Scholar] [CrossRef]
  188. Asadizadeh, M.; Hedayat, A.; Tunstall, L.; Gonzalez, J.A.V.; Alvarado, J.W.V.; Neira, M.T. The impact of slag on the process of geopolymerization and the mechanical performance of mine-tailings-based alkali-activated lightweight aggregates. Constr. Build. Mater. 2024, 411, 134347. [Google Scholar] [CrossRef]
  189. Das, D.; Neupane, K.; Gołąbiewska, A.; Rout, P.K. Utilization of iron tailings in alkali-activated binders: A review. Case Stud. Constr. Mater. 2023, 18, e02050. [Google Scholar] [CrossRef]
  190. Okereafor, U.; Makhatha, M.; Mekuto, L.; Mavumengwana, V. Gold mine tailings: A potential source of silica sand for glass making. Minerals 2020, 10, 448. [Google Scholar] [CrossRef]
  191. Alfonso, P.; Tomasa, O.; Domenech, L.M.; Garcia-Valles, M.; Martinez, S.; Roca, N. The use of tailings to make glass as an alternative for sustainable environmental remediation: The case of Osor, Catalonia, Spain. Minerals 2020, 10, 819. [Google Scholar] [CrossRef]
  192. Fu, F.; Hu, N.; Ye, Y.; Chen, G.; Guan, W.; Yang, S.; Li, Q. Production of high-value-added lightweight glass ceramics based on phosphorus tailings and coal gangue. Ceram. Int. 2024, 50, 16725–16735. [Google Scholar] [CrossRef]
  193. Sangiorgi, C.; Lantieri, C.; Tataranni, P.; Castro-Gomes, J.; Gabriel, M. Reuse of mining waste into innovative alkali-activated-based materials for road pavement applications. In Functional Pavement Design; CRC Press: Boca Raton, FL, USA, 2016. [Google Scholar]
  194. Segui, P.; Safhi, A.E.M.; Amrani, M.; Benzaazoua, M. Mining wastes as road construction material: A review. Minerals 2023, 13, 90. [Google Scholar] [CrossRef]
  195. Jahandari, S.; Tao, Z.; Chen, Z.; Osborne, D.; Rahme, M. Coal wastes: Handling, pollution, impacts, and utilization. In The Coal Handbook; Woodhead Publishing: Sawston, UK, 2023; pp. 97–163. [Google Scholar] [CrossRef]
  196. Prasad, M.N.V. Resource potential of natural and synthetic gypsum waste. In Environmental Materials and Waste; Elsevier: Amsterdam, The Netherlands, 2024; pp. 369–424. [Google Scholar] [CrossRef]
  197. Randhawa, N.S. An overview of sustainable solutions towards recycling of metallurgical slags. In iConFIRM: International Conference on Fundamental and Industrial Research in Materials; Springer Nature: Singapore, 2023; pp. 47–56. [Google Scholar] [CrossRef]
  198. Baalamurugan, J.; Ganesh Kumar, V.; Naveen Prasad, B.S.N.; Padmapriya, R.; Karthick, V.; Govindaraju, K. Recycling of induction furnace steel slag in concrete for marine environmental applications towards ocean acidification studies. Int. J. Environ. Sci. Technol. 2022, 19, 5039–5048. [Google Scholar] [CrossRef]
  199. USGS. US Geological Survey, Mineral Commodity Summaries. Iron and Steel Slag. 2024. Available online: https://www.usgs.gov/centers/national-minerals-information-center/mineral-commodity-summaries (accessed on 8 May 2025).
  200. Esther, L.A.; Pedro, L.G.; Irune, I.V.; Gerardo, F. Comprehensive analysis of the environmental impact of electric arc furnace steel slag on asphalt mixtures. J. Clean. Prod. 2020, 275, 123121. [Google Scholar] [CrossRef]
  201. Dondi, G.; Mazzotta, F.; Lantieri, C.; Cuppi, F.; Vignali, V.; Sangiovanni, C. Use of steel slag as an alternative to aggregate and filler in road pavements. Materials 2021, 14, 345. [Google Scholar] [CrossRef]
  202. Díaz-Piloneta, M.; Terrados-Cristos, M.; Álvarez-Cabal, J.V.; Vergara-González, E. Comprehensive analysis of steel slag as aggregate for road construction: Experimental testing and environmental impact assessment. Materials 2021, 14, 3587. [Google Scholar] [CrossRef]
  203. Georgiou, P.; Loizos, A. Characterization of sustainable asphalt mixtures containing high reclaimed asphalt and steel slag. Materials 2021, 14, 4938. [Google Scholar] [CrossRef]
  204. Chen, W.; Wei, J.; Xu, X.; Zhang, X.; Han, W.; Yan, X.; Lu, Z. Study on the optimum steel slag content of SMA-13 asphalt mixes based on road performance. Coatings 2021, 11, 1436. [Google Scholar] [CrossRef]
  205. Swathi, M.; Andiyappan, T.; Guduru, G.; Reddy, M.A.; Kuna, K.K. Design of asphalt mixes with steel slag aggregates using the Bailey method of gradation selection. Constr. Build. Mater. 2021, 279, 122426. [Google Scholar] [CrossRef]
  206. Zhao, X.; Sheng, Y.; Lv, H.; Jia, H.; Liu, Q.; Ji, X.; Meng, J. Laboratory investigation on road performances of asphalt mixtures using steel slag and granite as aggregate. Constr. Build. Mater. 2022, 315, 125655. [Google Scholar] [CrossRef]
  207. Goli, A. The study of the feasibility of using recycled steel slag aggregate in hot mix asphalt. Case Stud. Constr. Mater. 2022, 16, e00861. [Google Scholar] [CrossRef]
  208. Wang, Z. Quantitative assessment of road performance of recycled asphalt mixtures incorporated with steel slag. Materials 2022, 15, 5005. [Google Scholar] [CrossRef] [PubMed]
  209. Chen, H.; Liu, Y.; Cui, H.; Zhang, W.; Hu, L.; Mao, L. Effects of electric arc furnace slag on promoting quality and environmental safety of fired bricks incorporating municipal solid waste incineration fly ash. Constr. Build. Mater. 2022, 345, 128327. [Google Scholar] [CrossRef]
  210. Liu, J.; Jing, H.; Wang, Z.; Wang, X.; Zhang, L. Recycling of steel slag in sustainable bituminous mixtures: Self-healing performance, mechanism, environmental and economic analyses. J. Clean. Prod. 2023, 429, 139496. [Google Scholar] [CrossRef]
  211. Liu, J.; Wang, W.; Wang, Y.; Zhou, X.; Wang, S.; Liu, Q.; Yu, B. Towards the sustainable utilization of steel slag in asphalt pavements: A case study of moisture resistance and life cycle assessment. Case Stud. Constr. Mater. 2023, 18, e01722. [Google Scholar] [CrossRef]
  212. Sun, J.; Luo, S.; Wang, Y.; Dong, Q.; Zhang, Z. Pre-treatment of steel slag and its applicability in asphalt mixtures for sustainable pavements. Chem. Eng. J. 2023, 476, 146802. [Google Scholar] [CrossRef]
  213. Tangadagi, R.B.; Manjunatha, M.; Bharath, A.; Preethi, S. Utilization of steel slag as an eco-friendly material in concrete for construction. J. Green. Eng. 2020, 10, 2408–2419. [Google Scholar]
  214. Sha, F.; Liu, P.; Ding, Y. Application investigation of high-phosphorus steel slag in cementitious material and ordinary concrete. J. Mater. Res. Technol. 2021, 11, 2074–2091. [Google Scholar] [CrossRef]
  215. Lee, J.I.; Choi, S.J. Effect of replacement ratio of ferronickel slag aggregate on characteristics of cementitious mortars at different curing temperatures. Case Stud. Constr. Mater. 2023, 18, e01882. [Google Scholar] [CrossRef]
  216. Fares, A.I.; Sohel, K.M.A.; Al-Jabri, K.; Al-Mamun, A. Characteristics of ferrochrome slag aggregate and its uses as a green material in concrete—A review. Constr. Build. Mater. 2021, 294, 123552. [Google Scholar] [CrossRef]
  217. Wang, D.; Wang, Q.; Huang, Z. Reuse of copper slag as a supplementary cementitious material: Reactivity and safety. Resour. Conserv. Recycl. 2020, 162, 105037. [Google Scholar] [CrossRef]
  218. Chen, Y.; Ji, T.; Yang, Z.; Zhan, W.; Zhang, Y. Sustainable use of ferronickel slag in cementitious composites and the effect on chloride penetration resistance. Constr. Build. Mater. 2020, 240, 117969. [Google Scholar] [CrossRef]
  219. Gupta, N.; Siddique, R. Durability characteristics of self-compacting concrete made with copper slag. Constr. Build. Mater. 2020, 247, 118580. [Google Scholar] [CrossRef]
  220. Beaucour, A.L.; Pliya, P.; Faleschini, F.; Njinwoua, R.; Pellegrino, C.; Noumowé, A. Influence of elevated temperature on properties of radiation shielding concrete with electric arc furnace slag as coarse aggregate. Constr. Build. Mater. 2020, 256, 119385. [Google Scholar] [CrossRef]
  221. Luo, Z.; Ma, Y.; Mu, W.; Liu, J.; He, J.; Zhou, X. Magnesium phosphate cement prepared with electric furnace ferronickel slag: Properties and its hydration mechanism. Constr. Build. Mater. 2021, 300, 123991. [Google Scholar] [CrossRef]
  222. Nuruzzaman, M.; Kuri, J.C.; Sarker, P.K. Strength, permeability and microstructure of self-compacting concrete with the dual use of ferronickel slag as fine aggregate and supplementary binder. Constr. Build. Mater. 2022, 318, 125927. [Google Scholar] [CrossRef]
  223. Shoukry, H.; Perumal, P.; Abadel, A.; Alghamdi, H.; Alamri, M.; Abdel-Gawwad, H.A. Performance of limestone-calcined clay cement mortar incorporating high volume ferrochrome waste slag aggregate. Constr. Build. Mater. 2022, 350, 128928. [Google Scholar] [CrossRef]
  224. Liu, J.; Lin, C.; Liu, T.; Han, L.; Shen, X.; Li, C.; Lu, A. An eco-friendly permeable brick with excellent permeability and high strength derived from steel slag wastes. Int. J. Appl. Ceram. Technol. 2020, 17, 584–597. [Google Scholar] [CrossRef]
  225. Liu, Y.; Zhang, Z.; Hou, G.; Yan, P. Preparation of sustainable and green cement-based composite binders with high-volume steel slag powder and ultrafine blast furnace slag powder. J. Clean. Prod. 2021, 289, 125133. [Google Scholar] [CrossRef]
  226. Islam, M.Z.; Sohel, K.M.; Al-Jabri, K.; Al Harthy, A. Properties of concrete with ferrochrome slag as a fine aggregate at elevated temperatures. Case Stud. Constr. Mater. 2021, 15, e00599. [Google Scholar] [CrossRef]
  227. Yang, H.J.; Lee, C.H.; Shim, S.H.; Kim, J.H.J.; Lee, H.J.; Park, J. Performance evaluation of cement paste incorporating ferro-nickel slag powder under elevated temperatures. Case Stud. Constr. Mater. 2021, 15, e00727. [Google Scholar] [CrossRef]
  228. Jena, S.; Panigrahi, R. Feasibility study of the properties of geopolymer concrete with ferrochrome slag and silica fume. Mater. Today Proc. 2021, 38, 2476–2480. [Google Scholar] [CrossRef]
  229. Maharishi, A.; Singh, S.P.; Gupta, L.K. Strength and durability studies on slag cement concrete made with copper slag as fine aggregates. Mater. Today Proc. 2021, 38, 2639–2648. [Google Scholar] [CrossRef]
  230. Roslan, N.H.; Ismail, M.; Khalid, N.H.A.; Muhammad, B. Properties of concrete containing electric arc furnace steel slag and steel sludge. J. Build. Eng. 2020, 28, 101060. [Google Scholar] [CrossRef]
  231. Ruan, W.; Ma, Y.; Liao, J.; Ma, T.; Zhu, Y.; Zhou, A. Effects of steel slag on the microstructure and mechanical properties of magnesium phosphate cement. J. Build. Eng. 2022, 49, 104120. [Google Scholar] [CrossRef]
  232. Pang, L.; Liao, S.; Wang, D.; An, M. Influence of steel slag fineness on the hydration of cement-steel slag composite pastes. J. Build. Eng. 2022, 57, 104866. [Google Scholar] [CrossRef]
  233. Fang, M.; Fang, G.; Xia, Y.; Wang, H. Study on compressive strength of concrete mixed by steel slag powder and fly ash. IOP Conf. Ser. Earth Environ. Sci. 2020, 508, 012183. [Google Scholar] [CrossRef]
  234. Wu, Q.; Huang, Z. Preparation and performance of lightweight porous ceramics using metallurgical steel slag. Ceram. Int. 2021, 47, 25169–25176. [Google Scholar] [CrossRef]
  235. Rahou, J.; Rezqi, H.; El Ouahabi, M.; Fagel, N. Characterization of Moroccan steel slag waste: The potential green resource for ceramic production. Constr. Build. Mater. 2022, 314, 125663. [Google Scholar] [CrossRef]
  236. Ji, Y.; Li, E.; Zhu, G.; Wang, R.; Sha, Q. Preparation and performance of ceramic tiles with steel slag and waste clay bricks. Materials 2024, 17, 1755. [Google Scholar] [CrossRef]
  237. Liu, J.; Yu, B.; Wang, Q. Application of steel slag in cement treated aggregate base course. J. Clean. Prod. 2020, 269, 121733. [Google Scholar] [CrossRef]
  238. Weng, Y.; Liu, Y.; Liu, J. Study on mathematical model of hydration expansion of steel slag-cement composite cementitious material. Environ. Technol. 2021, 42, 2776–2783. [Google Scholar] [CrossRef] [PubMed]
  239. Li, N.; Zhan, H.; Yu, X.; Tang, W.; Xue, Q. Investigation of the aging behaviour of cellulose fiber in reclaimed asphalt pavement. Constr. Build. Mater. 2021, 271, 121559. [Google Scholar] [CrossRef]
  240. Mikhailenko, P.; Piao, Z.; Poulikakos, L.D. Electric arc furnace slag as aggregates in semi-dense asphalt. Case Stud. Constr. Mater. 2023, 18, e02049. [Google Scholar] [CrossRef]
  241. Piemonti, A.; Conforti, A.; Cominoli, L.; Sorlini, S.; Luciano, A.; Plizzari, G. Use of iron and steel slags in concrete: State of the art and future perspectives. Sustainability 2021, 13, 556. [Google Scholar] [CrossRef]
  242. Shumuye, E.D.; Zhao, J.; Wang, Z. Effect of the curing condition and high-temperature exposure on ground-granulated blast-furnace slag cement concrete. Int. J. Concr. Struct. Mater. 2021, 15, 1–20. [Google Scholar] [CrossRef]
  243. Ahmad, J.; Kontoleon, K.J.; Majdi, A.; Naqash, M.T.; Deifalla, A.F.; Ben Kahla, N.; Isleem, H.F.; Qaidi, S.M. A comprehensive review on the ground granulated blast furnace slag (GGBS) in concrete production. Sustainability 2022, 14, 8783. [Google Scholar] [CrossRef]
  244. Asaad, M.A.; Huseien, G.F.; Memon, R.P.; Ghoshal, S.K.; Mohammadhosseini, H.; Alyousef, R. Enduring performance of alkali-activated mortars with metakaolin as granulated blast furnace slag replacement. Case Stud. Constr. Mater. 2022, 16, e00845. [Google Scholar] [CrossRef]
  245. Omur, T.; Miyan, N.; Kabay, N.; Özkan, H. Utilization and optimization of unweathered and weathered basic oxygen furnace slag aggregates in cement based mortar. J. Build. Eng. 2023, 64, 105634. [Google Scholar] [CrossRef]
  246. Zago, S.C.; Vernilli, F.; Cascudo, O. The reuse of basic oxygen furnace slag as concrete aggregate to achieve sustainable development: Characteristics and limitations. Buildings 2023, 13, 1193. [Google Scholar] [CrossRef]
  247. Diotti, A.; Cominoli, L.; Galvin, A.P.; Sorlini, S.; Plizzari, G. Sustainable recycling of electric arc furnace steel slag as aggregate in concrete: Effects on the environmental and technical performance. Sustainability 2021, 13, 521. [Google Scholar] [CrossRef]
  248. Kurecki, M.; Meena, N.; Shyrokykh, T.; Korobeinikov, Y.; Jarnerud Örell, T.; Voss, Z.; Sridhar, S. Recycling Perspectives of Electric Arc Furnace Slag in the United States: A Review. Steel Res. Int. 2024, 2300854. [Google Scholar] [CrossRef]
  249. Araos, P.; Aponte, D.; Barra, M. Ladle furnace slag as a new source of supplementary cementitious material: Evaluating long-term performance and environmental impact in Portland cement systems. J. Build. Eng. 2024, 95, 109995. [Google Scholar] [CrossRef]
  250. Ahmad, J.; Majdi, A.; Deifalla, A.F.; Isleem, H.F.; Rahmawati, C. Concrete made with partially substitutions of copper slag (CPS): State of the art review. Materials 2022, 15, 5196. [Google Scholar] [CrossRef] [PubMed]
  251. Sahu, A.; Kumar, S.; Srivastava, A.K.L.; Pratap, B. Performance of recycled aggregate concrete using copper slag as fine aggregate. J. Build. Eng. 2024, 82, 108364. [Google Scholar] [CrossRef]
  252. Yanning, S.; Qiao, H.; Qiong, F.; Chao, W.; Jianghua, Z. Application of metallurgical ferronickel slag in building materials: A review. J. Build. Eng. 2024, 96, 110632. [Google Scholar] [CrossRef]
  253. Nguyen, Q.D.; Castel, A. Developing geopolymer concrete by using ferronickel slag and ground-granulated blast-furnace slag. Ceramics 2023, 6, 1861–1878. [Google Scholar] [CrossRef]
  254. Shooshtarian, S.; Maqsood, T.; Wong, P.S.; Caldera, S.; Ryley, T.; Zaman, A.; Cáceres Ruiz, A.M. Circular economy in action: The application of products with recycled content in construction projects—A multiple case study approach. Smart Sustain. Built Environ. 2024, 13, 370–394. [Google Scholar] [CrossRef]
  255. He, Y.; Liu, Y.; Liu, X.; Lan, M.; Lei, B.; Chen, Q.; Xue, X. Mechanical properties of eco-cement mortar containing MgO-modified phosphorous slag. Constr. Build. Mater. 2024, 428, 136223. [Google Scholar] [CrossRef]
  256. Jia, R.; Wang, Q.; Luo, T. Mechanisms and differences between sodium and magnesium sulfate attacks on alkali-activated phosphorus slag. Constr. Build. Mater. 2023, 403, 133117. [Google Scholar] [CrossRef]
  257. Wang, H.; Qian, J.; Zhang, H.; Nan, X.; Chen, G.; Li, X. Exploring Skid Resistance Over Time: Steel Slag as a Pavement Aggregate—Comparative Study and Morphological Analysis. J. Clean. Prod. 2024, 464, 142779. [Google Scholar] [CrossRef]
  258. Aquib, A.R.; Probha, Z.T.; Haque, M.A. An overview on utilization of steel slag as road construction materials. In International Conference on Civil Engineering, Lecture Notes in Civil Engineering; Springer Nature: Singapore, 2023; pp. 51–68. [Google Scholar]
  259. Kim, K.; Haeng Jo, S.; Kim, N.; Kim, H. Characteristics of hot mix asphalt containing steel slag aggregate according to temperature and void percentage. Constr. Build. Mater. 2018, 188, 1128–1136. [Google Scholar] [CrossRef]
  260. Pasetto, M.; Baliello, A.; Giacomello, G.; Pasquini, E. The use of steel slags in asphalt pavements: A state-of-the-art review. Sustainability 2023, 15, 8817. [Google Scholar] [CrossRef]
  261. Liu, J.; Xu, J.; Liu, Q.; Wang, S.; Yu, B. Steel slag for roadway construction: A review of material characteristics and application mechanisms. J. Mater. Civ. Eng. 2022, 34, 03122001. [Google Scholar] [CrossRef]
  262. Yang, S.; Krayushkina, K.; Akmaldinova, O.; Ji, J.; Bieliatynskyi, A. Determination of the possibility of using phosphorus slag in the road industry. Mater. Sci. 2023, 41, 44–61. [Google Scholar] [CrossRef]
  263. Kong, Y.; Zhang, X.; Zhang, L.; Xu, J.; Ji, W.; Pan, L.; Lu, R.; Zuo, J.; Ma, X.; Ma, S. Investigation on utilization and microstructure of fine iron tailing slag in road subbase construction. Constr. Build. Mater. 2024, 447, 138019. [Google Scholar] [CrossRef]
  264. Soni, J.; Chokshi, T.; Sharma, R.; Gujar, R.; Jariwala, N.; Pitroda, J.R. Assessing the applicability of fine copper slag in road and structural fill application. Mater. Today Proc. 2022, 62, 7040–7043. [Google Scholar] [CrossRef]
  265. Zakaria, S.K.; Sharif, N.M.; Abu Seman, A.; Taib, M.A.A.; Mohamed, J.J.; Yusoff, M.; Yusoff, A.H.; Masri, M.N.; Ali, A.; Ter Teo, P. Recycling of metallurgical steel slag waste as partial replacement of raw materials for sustainable clay-based ceramics: A brief review. AIP Conf. Proc. 2024, 2750, 090004. [Google Scholar] [CrossRef]
  266. Liu, Y.; Gu, F.; Zhou, H.; Li, Q.; Shang, S. Study on the performance and reaction mechanism of alkali-activated clay brick with steel slag and fly ash. Constr. Build. Mater. 2024, 411, 134406. [Google Scholar] [CrossRef]
  267. Ren, Y.; Wang, Z.; Qu, G.; Kuang, L.; Ning, P.; Ren, N.; Zhang, C.; Chen, X.; Wu, F.; Hu, Y.; et al. Comprehensive performance study of aluminum ash and calcium carbide slag for brick making under ultra-high pressure. Constr. Build. Mater. 2022, 359, 129526. [Google Scholar] [CrossRef]
  268. Hou, G.; Yan, Z.; Sun, J.; Naguib, H.M.; Lu, B.; Zhang, Z. Microstructure and mechanical properties of CO2-cured steel slag brick in pilot-scale. Constr. Build. Mater. 2021, 271, 121581. [Google Scholar] [CrossRef]
  269. Chen, Z.; Leng, Z.; Jiao, Y.; Xu, F.; Lin, J.; Wang, H.; Zhang, Y. Innovative use of industrially produced steel slag powders in asphalt mixture to replace mineral fillers. J. Clean. Prod. 2022, 344, 131124. [Google Scholar] [CrossRef]
  270. Maierdan, Y.; Haque, M.A.; Chen, B.; Maimaitiyiming, M.; Ahmad, M.R. Recycling of waste river sludge into unfired green bricks stabilized by a combination of phosphogypsum, slag, and cement. Constr. Build. Mater. 2020, 260, 120666. [Google Scholar] [CrossRef]
  271. Fuchs, K.; Gaikwad, A.D.; Prokein, M.; Schieferstein, E.; Nühlen, J.; Mölders, N.; Renner, M. CO2-curing of carbon-reducing slag-based masonry bricks at elevated pressures in pre-industrial scale. J. Supercrit. Fluids 2023, 203, 106064. [Google Scholar] [CrossRef]
  272. Guo, Y.; Li, Z.; Liu, F.; Wang, S.; Chen, F.; Yang, L.; Li, Z. Corrosion behavior of CaO–Al2O3–SiO2–MgO-Cr2O3 slag on MgO refractory brick for smelting of chromite. J. Eur. Ceram. Soc. 2024, 44, 544–556. [Google Scholar] [CrossRef]
  273. Chinyama, S.R.; Wenga, T.; Gwenzi, W. Mechanical and contaminant leaching behaviour of cement-based and fired bricks incorporating ferrochrome slag. J. Build. Eng. 2023, 72, 106572. [Google Scholar] [CrossRef]
  274. Gao, J.; Su, W.; Song, X.; Lv, P.; Yang, J.; Yu, G. Effect of slag composition on corrosion resistance of high chromia refractory bricks for industrial entrained-flow gasifier. Chin. J. Chem. Eng. 2023, 62, 139–149. [Google Scholar] [CrossRef]
  275. Amran, M.; Murali, G.; Khalid, N.H.A.; Fediuk, R.; Ozbakkaloglu, T.; Lee, Y.H.; Haruna, S.; Lee, Y.Y. Slag uses in making an ecofriendly and sustainable concrete: A review. Constr. Build. Mater. 2021, 272, 121942. [Google Scholar] [CrossRef]
  276. Ryu, G.U.; Kim, H.J.; Yu, H.J.; Pyo, S. Utilization of steelmaking slag in cement clinker production: A review. J. CO2 Util. 2024, 84, 102842. [Google Scholar] [CrossRef]
  277. Demarco, M.; Vernilli, F.; Zago, S.C. Clinker-Free Cement Manufactured with Metallurgical Slags. Buildings 2024, 14, 1739. [Google Scholar] [CrossRef]
  278. Xie, G.; Liu, L.; Suo, Y.; Zhu, M.; Yang, P.; Sun, W. High-value utilization of modified magnesium slag solid waste and its application as a low-carbon cement admixture. J. Environ. Manag. 2024, 349, 119551. [Google Scholar] [CrossRef]
  279. Amarasinghe, I.; Hong, Y.; Stewart, R.A. Visualising a framework for enhancing material circularity in building construction projects: Drivers, barriers, and strategies. Build. Environ. 2024, 253, 111359. [Google Scholar] [CrossRef]
  280. Sajid, Z.W.; Aftab, U.; Ullah, F. Barriers to Adopting Circular Procurement in the Construction Industry: The Way Forward. Sustain. Futures 2024, 8, 100244. [Google Scholar] [CrossRef]
  281. Shooshtarian, S.; Maqsood, T.; Wong, P.S.; Zaman, A.; Caldera, S.; Ryley, T. Utilisation of certification schemes for recycled products in the Australian building and construction sector. Bus. Strateg. Environ. 2024, 33, 1759–1777. [Google Scholar] [CrossRef]
  282. Jayalath, C. Policy Instruments Fostering Site Level Application of Circular Practices in Construction. Preprints 2024, 2024, 2024061646. [Google Scholar] [CrossRef]
  283. Giorgi, S.; Lavagna, M.; Wang, K.; Osmani, M.; Liu, G.; Campioli, A. Drivers and barriers towards circular economy in the building sector: Stakeholder interviews and analysis of five European countries policies and practices. J. Clean. Prod. 2022, 336, 130395. [Google Scholar] [CrossRef]
  284. Munaro, M.R.; Tavares, S.F. A review on barriers, drivers, and stakeholders towards the circular economy: The construction sector perspective. Clean. Responsible Consum. 2023, 8, 100107. [Google Scholar] [CrossRef]
  285. Chen, P.; Zhang, S.; Meng, J.; Lei, T.; Li, B.; Coffman, D.M. Technological solutions to China’s carbon neutrality in the steel and cement sectors. Earths Future 2023, 11, e2022EF003255. [Google Scholar] [CrossRef]
  286. Babalola, A.; Harinarain, N. Policy barriers to sustainable construction practice in the Nigerian construction industry: An exploratory factor analysis. J. Eng. Des. Technol. 2024, 22, 214–234. [Google Scholar] [CrossRef]
  287. Sorvari, J.; Wahlström, M. Industrial by-products. In Handbook of Recycling; Elsevier: Amsterdam, The Netherlands, 2024; pp. 259–285. [Google Scholar] [CrossRef]
  288. Assaad, R.H. Examining low-carbon material initiatives: Existing policies, impacts on the procurement of projects, current challenges, and potential solutions to reduce embodied carbon in the construction industry. J. Leg. Aff. Disput. Resolut. Eng. Constr. 2024, 16, 04523060. [Google Scholar] [CrossRef]
  289. Feijão, D.; Reis, C.; Marques, M.C. Comparative analysis of sustainable building certification processes. J. Build. Eng. 2024, 96, 110401. [Google Scholar] [CrossRef]
  290. Statsenko, L.; Samaraweera, A.; Bakhshi, J.; Chileshe, N. Construction 4.0 technologies and applications: A systematic literature review of trends and potential areas for development. Constr. Innov. 2023, 23, 961–993. [Google Scholar] [CrossRef]
  291. Thakur, A.; Kumar, A. Development of Novel Processes and Products from Mining Residues. In Sustainable Management of Mining Waste and Tailings; CRC Press: Boca Raton, FL, USA, 2024; pp. 301–330. [Google Scholar]
  292. Devarajan, P.; Kumar, R.S.; King, J.A.F. A short review on substantial role of geopolymer in the sustainable construction industry. IOP Conf. Ser. Earth Environ. Sci. 2023, 1130, 012002. [Google Scholar] [CrossRef]
  293. Barraza, F.P.; Thiyagarajan, D.; Ramadoss, A.; Manikandan, V.S.; Dhanabalan, S.S.; Abarzúa, C.V.; Soloaga, P.S.; Nazer, J.C.; Morel, M.J.; Thirumurugan, A. Unlocking the potential: Mining tailings as a source of sustainable nanomaterials. Renew. Sustain. Energy Rev. 2024, 202, 114665. [Google Scholar] [CrossRef]
  294. Geng, S.; Luo, Q.; Liu, K.; Li, Y.; Hou, Y.; Long, W. Research status and prospect of machine learning in construction 3D printing. Case Stud. Constr. Mater. 2023, 18, e01952. [Google Scholar] [CrossRef]
  295. Osei-Tutu, S.; Ayarkwa, J.; Osei-Asibey, D.; Nani, G.; Afful, A.E. Barriers impeding circular economy (CE) uptake in the construction industry. Smart Sustain. Built Environ. 2023, 12, 892–918. [Google Scholar] [CrossRef]
  296. Adrianto, L.R.; Ciacci, L.; Pfister, S.; Hellweg, S. Toward sustainable reprocessing valorization of sulfidic copper tailings: Scenarios prospective, LCA. Sci. Total Environ. 2023, 871, 162038. [Google Scholar] [CrossRef]
  297. Meglin, R.; Kytzia, S.; Habert, G. Regional circular economy of building materials: Environmental and economic assessment combining material flow analysis, input-output analyses, and life cycle assessment. J. Ind. Ecol. 2022, 26, 562–576. [Google Scholar] [CrossRef]
  298. Füchsl, S.; Rheude, F.; Röder, H. Life cycle assessment (LCA) of thermal insulation materials: A critical review. Clean. Mater. 2022, 5, 100119. [Google Scholar] [CrossRef]
  299. Greenfiber. Carbon Impact of Insulation: Putting Cellulose to the Test. Available online: https://www.greenfiber.com/blog/carbon-impact-of-insulation-putting-cellulose-to-the-test (accessed on 3 May 2025).
  300. Brekke, S.; O’Born, R. Life Cycle Assessment of Geopolymer Concrete Made with Tailings from Ilmenite Mining. In The International Conference on Net-Zero Civil Infrastructures: Innovations in Materials, Structures, and Management Practices (NTZR); Springer Nature: Cham, Switzerland, 2024; pp. 821–829. [Google Scholar]
  301. Crossin, E. The greenhouse gas implications of using ground granulated blast furnace slag as a cement substitute. J. Clean. Prod. 2015, 95, 101–108. [Google Scholar] [CrossRef]
  302. de Pedro, J.P.Q.; Lagao, J.A.T.; Ongpeng, J.M.C. Life cycle assessment of concrete using copper slag as a partial cement substitute in reinforced concrete buildings. Buildings 2023, 13, 746. [Google Scholar] [CrossRef]
Figure 2. Comparative summary of reviewed SRM.
Figure 2. Comparative summary of reviewed SRM.
Applsci 15 05410 g002
Table 1. Applications of cellulose waste in sustainable construction.
Table 1. Applications of cellulose waste in sustainable construction.
SRMApplicationDosageReference
CNFPortland cement paste1.5 wt% of the cement weight[34]
RCFFiber cement boardFrom 10% to 50% of the total cellulose fiber content[35]
RCFCellulose nanocrystals cement paste0.2% cellulose nanocrystals[36]
CNCCement paste0.2 vol–1.0 vol%[37]
CFConcrete1.0 kg/m3 (0.09%)–1.2 kg/m3 (0.11%)–1.5 kg/m3 (0.14%)–1.8 kg/ m3 (0.16%)–2.0 kg/m3 (0.18%)[38]
CFClay brick2.5–15 wt%[39]
CNCConcrete0.1 vol–0.2 vol%[40]
MFCMortar and concrete0.4 to 4.5% by weight of the cement[41]
CFReinforced concrete0.6–0.9–1.2–1.5 kg/cm3[42]
CNC/CNFCementCNF 0.020–0.035–0.045–0.050–0.065–0.085–0.10–0.15
CNC 0.020–0.035–0.045–0.050–0.065–0.085–0.10–0.15–0.20–0.40–0.60 wt%
[43]
CFConcreteFrom 10 to 90 wt%[14]
MCC/NFCCement pastesMCC: 0.05–0.5–0.75–1.0%
NFC: 0.05–0.5–1.0%
[44]
CNCMortar and concrete mixes0.25–0.50–0.75–1.00 wt% relative to the cement[45]
MCC/NFCCement pastes0.050–0.075–0.100% by cement weight[46]
HF/JFCement composite (mortar and paste)0.25–0.5–1–2% of the cement weight[47]
CFConcrete0.5 vol%[48]
CNFCement pastes and mortars0.3 wt–0.8 wt% of cement mass[49]
CNFCement based systems0.1–0.2–0.3–0.4–0.5 (vol%)[50]
MFCCement pastes0.3–0.6–1–2 wt.% of the cement[51]
CNFCement composites1.5 kg/m3[52]
CFAsphalt0.5%[53]
CNFAsphalt binders and mixes0.2–0.3–0.5–0.7% by the weight of binder[54]
CFAs filler in hot mix asphalt0.25–0.5–0.75%[55]
CFSMA pavement1.0% by weight of asphalt[56]
CFAsphalt mixtures2.5–5.0–7.5–10.0% by the weight of asphalt cement[57]
CNFAdditive in asphalt pavements.0.2–0.3–0.5–0.7% by the weight of binder[58]
CFConcrete1–2% by the weight of cement[59]
KFCement composites1–1.5–2% by mass[60]
KFCement pastes0.25–0.5% by weight of cement[61]
HFCement based mortar0.5 wt–1 wt–1.5 wt% on cement[62]
JFReinforced concrete5% by cement mass (0.8% content by volume of concrete)[63]
HFConcrete51.45–60.60–64.45 by % volume[64]
HFReinforced foam concretes0.75–1.5–3% by weight of cement[65]
HFMortars0.5 vol–1.0 vol%[66]
BF/HFReinforced cement mortars0.34–0.5–0.68–1 vol%[67]
HFMortars1 vol%[68]
FFConcrete0.3–0.6% mass percentages[69]
FFHigh-performance concrete0.15–0.30–0.50%[70]
FFReinforced cementitious composites1.0 vol% relative to the total batch volume[71]
FFCement-based composites0.3–0.6–0.9–1.2–1.5–1.8 by weight of cement[72]
FFUltra high-performance concrete2–4–6 (kg/m3)[73]
FFConcrete0.3–0.6% (5.7–11.4 kg/m3)[74]
FFCementitious composite1.0–2.0–3.0% vol[75]
BF: broom fibers; CF: cellulose fibers; CNC: cellulose nanocrystal; CNF: cellulose nanofibrils; FF: flax fiber; HF: hemp fiber; JF: jute fiber; KF: kenaf fibers; MCC: microcrystalline cellulose; MFC: microfibrillated cellulose; NFC: nanofibrillated cellulose; RCF: recycled cellulose fiber; SRM: secondary raw material; wt%: weight percent.
Table 2. Comprehensive applications of mining tailings in construction.
Table 2. Comprehensive applications of mining tailings in construction.
SRMApplicationDosageReference
CMWRRoad embankment or pavement layers80%[105]
GTRoad base layer and asphalt baseby 60% tailing and 7% asphalt[106]
ITRoad base material50%[107]
GT/GBFSRoad pavement materialGBFS:GT ratios: 100–10:90–20:80–30:70–40:60–50:50[108]
SWCement mortarsFrom 2 to 20% wt%[109]
SWConcrete paving blocksFrom 2 to 40% wt%[110]
SWGeopolymer cement mixtureFrom 53% to 81.6%[111]
SWFiller in asphalt pavement7 wt%[112]
ZTAsphalt mixturesUnspecified[113]
IT/GTConcrete and asphalt mixturesUnspecified[114]
ITAggregates in paving asphalt mixtureCoarse aggregates: 74%
Fine aggregates: 12% of the total mass of aggregates
[115]
SS/LFSAggregates in foamed bitumen-stabilized mixtures for road foundation layers20/10 wt% of aggregates[116]
MTFiller in bitumen1–3–5–10 (%w/w)[117]
ITMaterial in soil–cement mixtures10–20%[118]
ITFiller in asphalt mixturesFiller/asphalt weight ratios: 0.6–0.8–1.0–1.2[119]
PTModifier for the asphalt binder and
replacement for traditional limestone mineral powder
4–7–10–12–15%.[120]
RM/IT/ZTGeopolymer binders to stabilize the soil for road subgrade application10–20–30%[121]
CTFiller material in asphalt paving mastic30–60–90–120 wt% by asphalt weight[122]
ITFiller in an asphalt mixture20–40–60–80%[123]
CTSupplementary cementitious material10–14.4–20–27.8–40–43.9%[124]
ITSupplementary cementitious materials in cement10–20–30–40 wt%[125]
ITSupplementary cementitious material10–20–30 wt% by cement weight[126]
ITReplacing cement in pastes5–10–15% cement in weight[127]
IT/PS/SSConcreteIT: 6–8.5–17–25–44–63–126 (kg/m3)
PS/SS: 17–32–34–41–50–60 (kg/m3)
[128]
ITCementitious materials30% of cement by weight replaced[129]
LMCementitious materialsFrom 30% to 70%[130]
CTSubstitute for cement in concreteFrom 10 to 70%[131]
ITSubstitute for cement in cementitious materials10–30–50%[132]
ITSupplementary cementitious
material as partial replacement of cement
15%[133]
IT/PS/LSTernary supplementary cementitious materials for concreteIT: 30–15–10% PS: 0–15–10% LS: 0–15–10.
Total IOT + PS + LS = 30% replacement
[134]
ATPartially replace the silica sand in the mix25–50–75–100%[135]
GT/GBFSTernary cementitious
material
40.5–67.5–94.5–135 kg/m3[136]
CTSupplementary cementitious materialPartial cement replacement, between 5 and 15 wt%,[137]
GT/STSupplementary cementitious material10–15–20 wt%[138]
IT/SS/GBFSSupplementary cementitious materialIT: 2.2 wt%/SS: 1.5 wt%/GBFS: 1 wt%[139]
IT/SS/GBFSSupplementary cementitious materialIT: 2.2 wt%
SS/GBFS: from 0.5 to 2 wt%
[140]
QT/SchTCeramic tilesQT: 11 wt%/SchT: from 0 to 8 wt%[141]
SchTFeldspar substitute in porcelain tiles, semi-stoneware and stoneware formulationsFrom 0 to 8 wt%[142]
ANT/FQT/CNSCeramic formulations(A) 60% CNS, 20% ANT, and 20% FQT
(B) 50% CNS, 30% ANT, and 20% FQT
(C) 40% CNS, 40% ANT, and 20% FQT
[143]
GTSubstitute of feldspar in ceramic material5–10–30 wt%[144]
MTCeramic roof tiles5–10–20%[145]
SchT/KTCeramic tiles2–5–10 wt %.[146]
MT/WR Roof
tiles and blocks
Roof tiles: 5 wt%
Ceramic blocks: 10 wt%
[147]
MTCeramic products (roof tiles, blocks, and pavers)Roof tiles: 5–10–20 wt%
Block and paver: 10–20 wt%
[148]
BMWPorcelain stoneware2.5–5–10–15–20–25–40 wt.%[149]
CTAggregates in paving stones (cement Portland mixtures)25–50% for a total weight of aggregates[150]
MT Construction materialsFrom 0 to 58% w/w[151]
ITCeramics composites25–40–45%[152]
ANT: apatite–nepheline tailings; AT: antimony tailings; BMW: bentonite mining waste; CMWR: coal mine waste rocks; CNS: copper–nickel slag; CT: copper tailings; FQT: ferruginous quartzite tailings; GBFS: granulated blast furnace slag; GT: gold tailings; IT: iron tailings; KT: kaolin tiles; LFS: ladle furnace; LS: lithium slag; MT: mine tailings; PS: phosphate slag; PT: phosphorus tailing; QT: quartzite tailing; RM: red mud; SchT: scheelite tailings; SS: steel slag; ST: silver tailings; SW: silt waste; WR: waste rock; wt%: weight percent; ZT: zinc tailings.
Table 3. Applications of metallurgy slags in construction.
Table 3. Applications of metallurgy slags in construction.
SRMApplicationDosageReference
SS (EAF)Coarse aggregates in asphalt mixtures66.21–67.30%[200]
SSFine aggregates and filler in bituminous mixtures10%[201]
SS (BOF)Coarse aggregates in road surface layers and asphalt pavements.15%[202]
SS (EAF)For wearing courses of asphalt pavements(Not clear)[203]
SSAggregates in the production of asphalt mixes25–50–75–100%[204]
SSFine or coarse aggregate fractions in asphalt mixesUp to 100%[205]
SSFine aggregates in asphalt concrete and stone matrix asphalt37 vol% (AC)–22 vol% (SMA)[206]
SSAggregates (unspecified) in conventional an asphalt mixture25–50–75 wt%[207]
SSCoarse aggregates in recycled asphalt pavementSS: Not specified/10–20–30–40–50% RAP + 6% (rejuvenator in RAM) + limestone (fine aggregates: not specified)[208]
SSPMineral filler in asphalt mixtureThe proportions of the coarse aggregate, fine aggregate, and filler in the mineral mixture were 46%, 50% and 4%, respectively.[209]
SSTo partially replace limestone to prepare sustainable self-healing steel slag bituminous mixtures25–50–75–100%[210]
SSAggregates in asphalt mixtures50–100%[211]
SSAggregates in asphalt mixtures50 vol% of the initial BA[212]
SSConcrete as replacement of fine aggregatesFrom to 50% with an increment 10%[213]
HPSSCementitious material and ordinary concrete35%[214]
SS-IFCoarse aggregates in concrete20–40% replaced concrete[198]
FNSFine aggregates (sand) in cementitious mortars50%[215]
FCSAggregates in concreteCoarse aggregates (10–100%)
Fine aggregates (20–100%)
Fine aggregates (mortar) 20%
[216]
CSSupplementary cementitious material30% cement[217]
FNSPortland cement in cementitious composites10–30–50–70% by mass[218]
CSNatural sand in self-compacting concrete10–20–30–40–50–60%[219]
SS(EAF)Coarse aggregates in concrete100%[220]
FNSMagnesium phosphate cement66–80%[221]
FNSFine aggregates and supplementary binder in self-compacting concrete30% cement replacement[222]
FCSAggregates in mortar as a substitute for natural sand25% to 100% with a stride of 25 wt%.[223]
SSAggregates in cementitious material30–50–70 vol%[224]
SS/UFSCementitious material50% SS–40% SS + 10% UFS (wt%)[225]
FCSFine aggregates (sand) in concrete production.25–50–75–100%[226]
FNS/GNBSReplacement of cement pastes30% (FNS)–60% (GNBS)[227]
FCSCoarse aggregates in geopolymer concrete mixesFrom 0 to 40%[228]
CSFine aggregates in cement concrete20–40–60–80–100 vol%[229]
SS(EAF)Substitute of cement in concrete5%, 10%, 15%,–20% by weight.[230]
SSMagnesium phosphate cement5–10–15–20–25 wt%[231]
SSReplacement of cement30% by mass[232]
SSPConcreteSteel slag powder 20–25–30–40%[233]
SSLightweight porous ceramics31.5–35%[234]
SS(EAF)Clay bricksFrom 5 to 25% by weight[235]
SSCeramic tiles35–40–45%[236]
SSBricks90 wt%[237]
SSCementitious material10–20–30–40–50%[238]
SS(BOF)Ceramics composites20–25–45%[152]
SS(BOF)Foamed glass–ceramicSS(BOF)/WG ratios: 30/70–35/65–40/60–45/55–50/50[239]
SS(EAF)Aggregates in semi-dense asphalt13 vol% of the aggregates[240]
AC: asphalt concrete; BA: basalt aggregates; BOF: blast oxygen furnace; CS: copper slag; EAF: electric arc furnace; FCS: ferrochrome slag; FNS: ferronickel slag; GNBS: granulated blast-furnace slag; HPSS: high-phosphorus steel slag; IF: induction furnace; IS: iron slag; RAM: recycled asphalt mixtures; RAP: reclaimed asphalt pavement; SMA: stone matrix asphalt; SS: steel slag; SSP: steel slag powder; UFS: ultrafine ground granulated blast furnace; WG: waste glass.
Table 4. Performance of construction materials incorporating SRM (↑ increase; ↓decrease).
Table 4. Performance of construction materials incorporating SRM (↑ increase; ↓decrease).
SRMApplicationDosageEffect on PropertiesReference
Cellulose fibersConcrete 0.9 kg/m3↑ Compressive, tensile, and flexural strength[42]
Waste paperConcrete 5–10%↑ Compressive strength, ↓ beyond 10%[96]
General cellulose useConcrete≤1.0% (typical)↑ Tensile strength, ↓ Brittleness, ↑ Durability (if treated)[14,98]
Iron tailingsConcrete Up to 30% aggregate replacement↑ Compressive strength, ↑ Abrasion resistance[154]
Various tailingsFine aggregate replacement10–40%↑ Freeze–thaw resistance, ↓ Permeability[157]
Tailings as cement substituteConcrete5–20%↑ Durability, depends on SiO2 + Al2O₃ + Fe2O₃ content ≥75%[157]
Blast furnace slagCement replacement20–30%↑ Compressive and tensile strength, ↑ Sulfate resistance[242,243]
Steel slag (EAF)Aggregates in concrete100%↑ Density, ↑ Mechanical strength[247]
Copper slagFine aggregates in concrete50–60%↑ Abrasion resistance, ↑ Density, ↓ Tensile strength[251]
Steel slagAsphalt mixturesVarious↑ Moisture resistance, ↑ Fatigue life, ↓ Rutting[257,261]
Nickel slagConcrete30%↑ Flexural and compressive strength, ↑ Durability[254]
Phosphorus slagConcrete20–30%↑ Sulfate resistance, ↑ Chemical durability[256]
Table 5. Comparative summary of reviewed SRM.
Table 5. Comparative summary of reviewed SRM.
Evaluation CriteriaCellulose-Based MaterialsMining TailingsMetallurgical Slags
Technical performanceModerate; enhances tensile strength and crack control; durability depends on moisture protectionVariable; good performance in aggregates with proper processing; site-dependentHigh; proven mechanical performance, good durability as binder or aggregate
Environmental impactVery high; renewable source, low CO2 footprint, promotes insulation and lightweight designHigh; reduces demand for virgin aggregates, helps in land recoveryHigh; reduces clinker use and emissions in cement, valorizes industrial waste
Economic viabilityVariable; low-cost raw material but processing (e.g., fiber treatment) may add costModerate to high; abundant but processing and environmental safeguards increase costsHigh; often already integrated into industrial supply chains
Regulatory and standardizationLimited; lacks universal standards, acceptance varies by region and applicationEmerging; some applications standardized, but others face regulatory hurdlesStrong; widely accepted in standards and technical codes
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Hidalgo, D.; Verdugo, F. Harnessing Secondary Resources for Sustainable and Circular Practices in the Construction Sector: A Scoping Review. Appl. Sci. 2025, 15, 5410. https://doi.org/10.3390/app15105410

AMA Style

Hidalgo D, Verdugo F. Harnessing Secondary Resources for Sustainable and Circular Practices in the Construction Sector: A Scoping Review. Applied Sciences. 2025; 15(10):5410. https://doi.org/10.3390/app15105410

Chicago/Turabian Style

Hidalgo, Dolores, and Francisco Verdugo. 2025. "Harnessing Secondary Resources for Sustainable and Circular Practices in the Construction Sector: A Scoping Review" Applied Sciences 15, no. 10: 5410. https://doi.org/10.3390/app15105410

APA Style

Hidalgo, D., & Verdugo, F. (2025). Harnessing Secondary Resources for Sustainable and Circular Practices in the Construction Sector: A Scoping Review. Applied Sciences, 15(10), 5410. https://doi.org/10.3390/app15105410

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop