[go: up one dir, main page]

Next Article in Journal
Safety Risk Assessment Using a BP Neural Network of High Cutting Slope Construction in High-Speed Railway
Next Article in Special Issue
Metal-Nails Waste and Steel Slag Aggregate as Alternative and Eco-Friendly Radiation Shielding Composites
Previous Article in Journal
Effect of the Progressive Failure of Shear Connectors on the Behavior of Steel-Reinforced Concrete Composite Girders
Previous Article in Special Issue
Influence of Water-Binder Ratio on the Mechanical Strength and Microstructure of Arch Shell Interface Transition Zone
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Bibliometric-Statistical Review of Organic Residues as Cementitious Building Materials

by
Sergio Yanez
,
Constanza Márquez
,
Benjamín Valenzuela
and
Cristina Alejandra Villamar-Ayala
*
Civil Engineering Department, University of Santiago of Chile, Av. Victor Jara 3659, Santiago 9170124, Chile
*
Author to whom correspondence should be addressed.
Buildings 2022, 12(5), 597; https://doi.org/10.3390/buildings12050597
Submission received: 1 April 2022 / Revised: 22 April 2022 / Accepted: 24 April 2022 / Published: 5 May 2022
(This article belongs to the Special Issue Advanced Eco-Friendly Cementitious Materials)
Figure 1
<p>Generation of agricultural residues (1980–2019): (<b>a</b>) worldwide; and (<b>b</b>) by continent (Source: Figures created by the authors based on [<a href="#B21-buildings-12-00597" class="html-bibr">21</a>]).</p> ">
Figure 2
<p>Bibliometric analysis based on the specific keyword “organic residues as building materials”: (<b>a</b>) VOSviewer map displaying the most prevalent keywords for <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>654</mn> </mrow> </semantics></math>; and (<b>b</b>) temporal evolution of studies related to organic residues in the construction industry between 1992 and 2020 for <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>1416</mn> </mrow> </semantics></math>.</p> ">
Figure 3
<p>Categorization of organic residues (<math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>158</mn> </mrow> </semantics></math>) by: (<b>a</b>) constructive use; (<b>b</b>) origin; and (<b>c</b>) type of agricultural residue.</p> ">
Figure 4
<p>Compressive strength behaviour by (<b>a</b>) constructive use; (<b>b</b>) industry; and (<b>c</b>) agricultural industry. (<span class="html-italic">n</span> = 557). (*) there are significant differences (<math display="inline"><semantics> <mrow> <mi>p</mi> <mo>&lt;</mo> <mn>0.05</mn> </mrow> </semantics></math>).</p> ">
Figure 5
<p>Physicochemical and mechanical properties correlations. (<b>a</b>) Residue density-compressive strength; (<b>b</b>) Residue density-tensile strength; and, (<b>c</b>) Lignin content-flexural strength.</p> ">
Figure 6
<p>Physicochemical and thermal properties correlations. (<b>a</b>) Residue density-thermal conductivity; and (<b>b</b>) Extractives content-thermal conductivity.</p> ">
Review Reports Versions Notes

Abstract

:
Climate deterioration and environmental pollution has been widely studied by a wide scientific community. The effects of the ecosystem deterioration impacts directly to human activities. In this scenario, the building industry has increased the pressure on proposing new materials to replace the cementicious component and natural resources (water, sand, gravel, and limestone) on mortar and concrete to reverse this trend. To this end, organic residues can offer opportunities as an available alternative for construction applications. Therefore, this paper aims to broaden the scope of research in this field by investigating the potential use of organic residues as cementicious building material based on bibliometric-statistical analysis using scientific information. A preliminary bibliometric analysis using VOSviewer was carried out to define the keywords co-ocurrence from Scopus database. Type of organic material, constructive use, and its properties (physicochemical, mechanical, and thermal) were extracted from scientific publications. Then, a systematic analysis criteria was defined to limit the scope of the study. Finally, statistical variance analysis and multiple correlation for identifying constructive application were applied. From the co-ocurrence analysis of keywords, we determined that 54% of the selected scientific publications were closely related to the scope of this study. State-of-the-art study established that related researches grew exponentially at a rate of about 30 % /year. Moreover, scientific publications reported the use of a wide variety of organic residues, such as wheat, paper, hemp, rice, wood, molluscs, olive, coconut, among others. Mainly, agricultural residues ( 82 % ) with building applications related to structural concrete, mortar, bricks, and blocks, had been evaluated. Physicochemical properties from organic residues (extractives content, lignin content, and density) were correlated to mechanical (compressive, flexural and tensile strength) and thermal properties (thermal conductivity). The identification of the physicochemical properties of the organic residues allow us to predict the mechanical and thermal behavior of the material with residues. In summary, agricultural residues are the most promising organic building material due to their abundance and lignin content, exhibiting better mechanic and thermal properties than any other organic residues.

1. Introduction

The construction industry plays a fundamental role in the economic growth of each country, and its performance can achieve vital advances in the socio-environmental development [1]. The construction industry is expected to grow at an annual average of 3.9% until 2030, increasing the Growth Development Production (GDP) worldwide by more than one percentage point (14.7% of GDP world) [2]. However, this industry generates 19% of the total greenhouse gas emissions (CO2−eq), and it is responsible for the consumption of 32% of the global energy related with buildings heating [3]. Therefore, there is an urgent need to refocus of research and policy efforts in light of the climate crisis [4].
In the construction industry, cementicious materials, and specially concrete, are the most used building materials [5]. Moreover, concrete is the second most produced and consumed material, only after water, reaching a global consumption of around 25 billion ton/year [6,7]. From a mechanical point of view, concrete has relatively high mechanical performance under compressive strength (5–60 MPa) when it is compared to other materials. Also, concrete is one of the most versatile materials used in buildings, allowing structural engineers to optimise application requirements with many possible structural configurations. Even more, fresh concrete is easy to mix, place and finish, increasing the workability. In spite of that, the increasing demand of concrete also generates pressure on the aggregates extraction, water and cement production [8,9].
The aggregate/cement/water ratio in the concrete production borders around 3/1/0.5 (volume/volume or v/v) depending on the construction application [10]. Fine aggregates 0.08–5 mm and coarse aggregates (>5 mm) that give stiffness to concrete are a natural resource, which is scarce, finite, and highly demanded. In the last 15 years, aggregates extraction have gradually increased, up to 17.5 Gton/year [11,12,13,14]. Meanwhile, Portland cement (95% [3CaO · SiO2 + 2CaO · SiO2 + 3CaO · Al2O3 + 3CaO · Al2O3 · Fe2O3] + 5% CaSO4), which is chemically synthesized from clay/calcareous rocks, has reached production demand close to 4.1 Gton/year [15]. Despite the high mechanical performance and the technological advances in the concrete/cement industry, the production industry of cement is one of the most polluting industries [16,17].
As a building material, concrete generates up to 5% of CO2−eq [18]. Also, concrete production is responsible for 9% (2 Gton/year) of industrial water withdrawals, corresponding to 1.7% of total world water extraction [12]. As a consequence, it is highly recommended to pursue in the innovation, testing, and analysis of the whole life cyclic assessment of novel construction materials [19]. Reuse of waste materials, the use of fly-ash to replace the cement content in the binder, and the use of organic residues are some of the solutions that can increase the sustainability of the construction industry. In this paper, we focused the bibliometric analysis in organic residues as a potential alternative in cementicious building materials considering mechanical and thermal properties.

2. The Production of Organic Residues

Worldwide, nearly 2010 million ton/year of residues are produced, of which 63% (1260 million ton/year) comes from organic residues [20]. One of the main sources of organic residues generation is the agricultural industry, whose generation has been growing at rates of 1.4%/year, reaching values of up to 37.5 million tons in 2019, value that could increase to 39.5 million ton/year by 2050 [21].
The evolution and generation of organic residues from 1980 to 2019 (by year) and by continent are shown in Figure 1. Here, the organic residues considered were barley, soy, rice husk, corn, wheat, and others. A clear incremental trend in the generation of agricultural residues can be seen during the last 40 years from Figure 1a. When it is separated by continent (Figure 1b), Asia and America show a generation of 45.8 % and 29% of the total agricultural residues worldwide, respectively. Indeed, wheat production generates 9.9 million tons of organic residues, while maize, rice, and soy generates 9.3, 9.2 and 4.5 million tons, respectively [21].
Several authors have proposed novel techniques and materials to improve the buildings’ thermal comfort [22,23,24]. Typical thermal barriers are placed in walls to prevent the heat to dissipate. Fiberglass, mineral (rock or slag) wool, plastic fibers, or natural fibers are used to insulate stud and joist spacing that is relatively free from obstructions. This solution, however, tends to be relatively inefficient if it is done by non-trained personal. Foam boards, to be placed on outside of wall (usually new construction) or inside of wall (existing homes) are used to insulate concrete walls, but requires specialized skills.
Despite the many techniques available in the market, the thermal performance of insulation is very dependent on proper installation and the skills of the insulation company. Therefore, the thermal performance of the material needs to be improved to avoid additional solutions. On the other hand, thermal improvements using non-degradable materials could give non-inert (toxic) properties to construction wastes (i.e., concrete) that are considered inert during the demolition [25]. Multiple studies have focused on evaluating the alternative use of organic residues as a total or partial replacement, or as an additive to conventional building materials. Indeed, various organic residues, such as wheat, paper, hemp, rice, wood, mollusks, olives, walnuts, coconut, among others have been studied [26,27,28,29,30,31,32,33,34]. In addition, different constructive manufacturing or replacement of materials (aggregates or cement) using organic residues have been reported, such as concrete, mortar, blocks, bricks, panels and asphalt and also in replacement of materials [35,36,37,38].
There is robust bibliography available that deals with the use of organic residues as a total or partial replacement of cement, fine or coarse aggregates in the cement paste and concrete (>100 journals). In general, studies have established reuse of some of these organic materials, towards construction applications with a focus on thermal improvement with a large amount of incorporated organic material (65%, v/v). Meanwhile, structural application reports use a lower amount of organic residues in its composition (1%, v/v) [39,40]. However, no systematic or statistical analysis on the use of organic residues as an alternative building material has been reported. In this review paper, the use of organic residues as building material is analyzed bibliometrically through a meta-analysis of the state-of-the-art (systematic) and statistical analysis of existing scientific information. Conclusions are presented to correlate the mechanical and thermal properties of the construction material with physicochemical properties of the organic residue. Comparisons on the origin of the organic residue as a potential replacement of cement or aggregate is also presented.

3. Methodology

This section describes the procedure used in the identification of the most suitable database as well as methods of database analysis.

3.1. Database Collection

A preliminary bibliometric analysis was carried out using the scientific journal repository Scopus from Elsevier. The selected repository offers a more comprehensive coverage when compared to Web of Science in fields outside of medicine and physical sciences [41,42]. In order to narrow the potential keywords to be used in the search engine, specific criteria were defined in terms of start date and applications. The keyword used for this preliminary analysis was “organic residue as building material”. This search generated 654 results (scientific publications). Additional documents were excluded from this search based on the documentation type. For instance, short surveys, notes, reports, and erratas were not considered.
After the bibliometric analysis using Scopus database, a more specific analysis was carried out using Science Direct. The analysis consisted in three different phases. First, an exploring phase was performed to inspect journals while using the same keyword as input in the search engine. This phase is called exploring phase. The final searching process gave 62,893 results, including scientific publications from different topics and knowledge areas.
A second phase, named identification phase, was implemented by using the most compatible journal. This analysis determined that the three journals with the most compatible research topic were Construction and Building Materials (19.8%), Journal of Cleaner Production (3%), and Journal of Science of the Total Environment (1.4%). The percentages given here are the amount of available scientific publications in each selected journal with respect to the total search. The keyword used for this analysis was “organic residue”. This phase gave 1448 results (scientific publications). Finally, a third phase called confirmation phase, considered the inclusion and exclusion criteria described in Table 1, reaching 158 results (scientific publications).

3.2. Processing Information

In this section, general and specific information were obtained from the confirmation phase described before (158 scientific publications). The general information analysis, as well as the specific information are described in Table 2. For each analysis, the information was processed by objective, factor, variable, and analysis type. On the one hand, preliminary bibliometric analysis used VOSviewer (version 1.6.16, Centre for Science and Technology Studies, Leiden University, The Netherlands) from Scopus database, where a co-occurrence network was performed to analyze the selected keywords (refer to Figure 2a). This technique offers graphic (visual) representation of co-occurrence networks to be visualized and to infer regarding relationships between entities. In this case, the size of each node represents the appearance co-occurrence, and its color represents links that connect different entities related to the topic.
To measure the influence of the organic residue on the potential use as building material, a variance analysis using the two-way ANOVA test was performed to a group of data ( n = 557 ) to determine the influence the specific keyword has on the potential use. Factors (type of organic residue and constructive use) and variable (compressive strength) were considered. Normality (Shapiro-Wilks test) and variance homogeneity (Levene test) were previously verified, considering database parametricity. Therefore, parametric (Tukey test) or non-parametric (Kruskal Wallis test) variance analysis were used. Infostat 2020e software for ANOVA analysis and a significance level of 0.05 were used. This software is used for both descriptive statistical and advanced methods of statistical modeling and multivariate analysis. Database is only limited by available memory, admitting ordinal, and nominal variables. Relationship between properties (physicochemical, mechanic, and thermal) were based on multiple correlation analysis. Physicochemical properties, such as density (kg/cm3), ash content (%, weight/weight or w/w), extractive content (%, w/w), lignin (%, w/w), cellulose (%, w/w), hemicellulose (%, w/w), moisture content (%, w/w), and pH were evaluated. Meanwhile, mechanic properties, such as compressive strength (kgf/cm2), flexural strength (kgf/cm2) and tensile strength (kgf/cm2) were considered. Finally, thermal conductivity was established as thermal property. These properties were evaluated for organic residues and/or compounds (organic residue + traditional building material). Normality (Shapiro-Wilks test) was previously verified, considering database parametricity. Parametric (Pearson test) or non-parametric (Spearman test) multiple correlation were considered as appropriate. Correlation coefficient ( r < 0.5 ) and significance ( p < 0.05 ) were considered to establish regression trends ( r 2 > 0.5 ). OriginPro 2020b (Version 2021, Northampton, MA 01060, United States) software for correlation analysis was used. This software allows performing multivariate correlation analysis with parametric and non-parametric tests, offering summary tables that show correlation and significance coefficients. In addition, it has applications for peak fitting, surface fitting, statistics, and signal processing. It also offers excellent graphing and representation of results.

4. Results

4.1. Bibliometric Analysis

Bibliometric analyses conducted in VOSviewer software were employed to synthesize patterns of knowledge production in the organic residues literature. Specific keyword was used in order to analyze the research trend, that have evolved in this knowledge base. The analysis calculates the number of times that an specific keyword has been declared in the review database. Because this analysis scanned the use of organic residues rather than any other co-citation, its results reflect patterns of using organic residues as construction material in the ‘broader literature’. Thus, to some extent, this analysis overcomes a limitation of ‘traditional citation analysis’ limited to a particular document repository (e.g., Scopus, Web of Science).
Co-citation analyses have also been used as the basis for the ‘visualization of similarities’ (VOS), a powerful approach to network mapping [43]. Co-citation analysis assumes that when two scholars are frequently ‘cited together’ by other authors, they tend to share a similarity in theoretical perspective [44]. Author co-citation analysis in VOSviewer transforms patterns of author co-citation into a social network map that visualizes similarities among the authors in a particular literature [43]. In this research, bibliometric network to describe the co-occurrence map of the selected database (organic residues as building material) was established. Figure 2a describes the co-occurrence map generated using VOSviewer. The most prevalent keywords, due to their occurrence and represented by their larger size were: building material, waste management, recycling, and construction materials. These keywords are closely related to the topic addressed in this study, and represent 54% of the scientific publications. The keywords that are not directly related to the topic of this study were: incineration, heavy metals, leaching, landfill, adsorption, among others. These keywords correspond to 46% of the total amount of scientific publications. Therefore, there is a significant amount of publications that address the issues raised in this study within the Scopus database, focusing on building materials and the use of residues in construction applications.
Figure 2b shows the temporal evolution of studies (scientific publications) between 1992 and 2020. The solid black line represents the total number of scientific publications ( n = 1416 ). The selected database (solid red line) corresponds to articles that fulfilled the previously detailed selection phases, corresponding to 11% of the total ( n = 158 ) (see Appendix A which contains all database for the criteria defined in Table 2). From 2007, the number of studies show an annual average growth rate of 28.7%. The year 2020 is the year with the largest number of studies selected with a total of n = 31 scientific publications that represents the 12%. Most of the scientific publications were published between 2012 and 2020, representing the 92% of the total database.
The results on organic residues have been categorized and shown in Figure 3. This group of scientific publications were divided depending on the construction implementation or use, the origin of the organic residues, and the type of organic residue. Figure 3a describes the constructive use given to the organic residue in the scientific publication analyzed. The results showed that the organic residues have been used as a replacement of cement powder in mortar and concrete (∼60%), and also as a replacement/reinforcement in blocks and bricks. To a lesser extent, organic residues have been implemented as partial replacement of the constitutive component in insulating panels (12.7%). Figure 3b illustrates the origin of the organic residue studied in this research. In this regard, 81.7% of the scientific publications declare to study residues coming from the agricultural industry, whereas 8.9% comes from the fishing industry, 6.3% from sludge coming from wastewater treatment plants, and 3.2% from the livestock industry. In this paper, the effect of different sources of organic residues on the mechanical/thermal properties was taken into account to establish a statistical relationship with the physicochemical property. These sources correspond to agricultural, farming, fishing, sludge and livestock residues. In order to measure the influence of the type of the organic residue as a potential construction material, we use the average compressive strength of compounds using these organic residues. On the other hand, thermal properties were related to the physicochemical properties of the organic residues from all sources. Since the agricultural industry is the primary source of organic residues, a more detailed analysis of it is made. In Figure 3c, the different types of organic residues used as potential building material from the agricultural industry are shown. The first three major type of agricultural industry are: rice industry that represents the 14.6%, wood industry with the 9.5%, and the paper industry with the 9.5% of the total agricultural residues. These studies constitute 33.5% of all the scientific publications analyzed. Less frequently, there are organic residues from the production of sugar cane (7.6%), oil palm (5.1%), biomass (3.8%), hemp (3.8%), and coconut (3.8%), which have been studied as building material.
In summary, the bibliometric analysis showed that 54% of the scientific publications identified with the specific keyword were related to the potential use of organic residues as building material. In addition, the amount of scientific publications related to the revaluation of residues in construction presented an annual growth rates of ∼30%. Specifically, 32% of the scientific publications evaluated as potential constructive use of organic residues have an specific application in mortar compounds, followed by concrete, blocks and bricks, and panels. Finally, 82% of the scientific publications analyzed used organic residues from the agricultural industry (rice, wood and paper), followed by those from the fishing industry and the wastewater sanitation sector (sludge).

4.2. Influence of Type of Organic Residue on Its Potential Constructive Application

Figure 4 shows the behavior of compressive strength measured for compounds made up of mixed between organic residues and conventional materials (i.e., cement, aggregates) at ranges from 5 to 30% (v/v). Figure 4a displays how the compressive strength varies, according to the different constructive uses ( n = 557 ). The average compressive strength of the mortar is 309.4 Kgf/cm2; while that, concrete, block/brick, and insulating panels report average compressive strengths of 263.8, 142.1, and 40.7 Kgf/cm2, respectively. Therefore, in terms of constructive application, insulating panels tend to show significantly the least compressive strength ( p < 0.05 ) with respect to the other applications.
Figure 4b shows behavior of compressive strengths respect to the origin of the organic residue ( n = 557 ). The average compressive strength of compounds using organic residues from livestock industry corresponded to 359.6 Kgf/cm2, followed by agricultural ( 215.6 Kgf/cm2), fishing ( 211.2 Kgf/cm2) and finally wastewater sanitation ( 202.2 Kgf/cm2). The livestock industry presented significantly the most compressive strengths ( p < 0.05 ) with respect to the agricultural, fishing and wastewater sanitation (sludge) industries. Figure 4c represents the values obtained for compressive strength, according to organic residues from agricultural industry ( n = 430 ). The most reported agricultural residues were rice, wood, paper, sugar cane, oil palm, hemp, wheat with average compressive strenght of 220, 200, 120, 300, 410, 90, and 205 Kgf/cm2, respectively. Therefore, there is differences between the compressive strengths depend on the type of agricultural residue. Indeed, oil palm is the only residue that presents the most significant differences ( p < 0.05 ) with respect to the other agricultural residues.
In general, a key parameter that characterize any material utilized in the construction industry is the compressive strength. Therefore, the comparison of this universal parameter for classify building materials, allows to establish differences between different materialities. In this research, the influence of type of organic residue on compressive strength (constructive application) was observed. In fact, an average of 287 Kgf/cm2 was obtained for compounds used as mortar or concrete. Typical values of compresive strengths for mortar and concrete vary between 50.8 and 611.8 Kgf/cm2 [5]. Therefore, organic residues generate similar compressive strengths to typical. However, compressive strength ranges (0.8–629 Kgf/cm 2 ) for compounds used as concrete were slightly smaller than mortar. It was observed mainly when organic residues were used as aggregates replacement. This could be improved, if they replace to the cement, thanks to the fact that some organic residues (e.g., sludge ash, 76.2% [3CaO · SiO2 + 2CaO · SiO2 + 3CaO · Al2O3 + 3CaO · Al2O3 · Fe2O3]) have close chemical characteristics to the Portland cement (93–95% [3CaO · SiO2 + 2CaO · SiO2 + 3CaO · Al2O3 + 3CaO · Al2O3 · Fe2O3]) [15,45]. On the other hand, lower compressive strengths can be found in block and brick, as well as in insulating panels, as expected. Typical compressive strengths for blocks, bricks, and insulating panel vary between 37 and 205 Kgf/cm2, which are similar to reported by compounds in this research [32,46]. Finally, the agricultural residues from the livestock industry presented the most compressive strenght (60–650 Kgf/cm 2 ), thanks to its similar inorganic chemical composition when the organic material is volatilized ( T > 500 °C) (e.g., manure ash, 78.4% [3CaO · SiO2 + 2CaO · SiO2 + 3CaO · Al2O3 + 3CaO · Al2O3 · Fe2O3]) to other conventional material as the cement [47]. Therefore, a key factor in determining their constructive use is in its composition.

4.3. Influence of Organic Residue Composition on Its Constructive Application

Table 3 resumes the correlation analysis (Spearman multiple correlation) established between physicochemical (organic residue) and mechanical/thermal (compounds) properties of n = 595 scientific publications. Of these relationships, those that meet the database robustness requirements ( p < 0.05 ) and correlation ( r > 0.5 ) correspond to 18.7%. Of the latter, the correlations between organic residue (physicochemical properties) with their compounds (mechanical/thermal properties) reached 53% and relationships, being the following: mechanical—physicochemical, mechanical—thermal, physicochemical—physicochemical, physicochemical—mechanical, and physicochemical—thermal. These last two correlations are those of main interest for this research.
The density and lignin content were well correlated with compressive, flexural and tensile strengths (physicochemical—mechanical properties). Meanwhile, the density and extractives content were well correlated with the thermal conductivity (physicochemical—thermal properties).

4.3.1. Physicochemical Properties (Organic Residue) with Respect to Mechanical Properties (Compounds)

Figure 5 shows the linear correlations obtained between the physicochemical properties of the organic residue and the mechanical properties of the compounds. Figure 5a shows the relationship between the residue density (0.08–2.52 g/cm 3 ) and the compressive strength of the compound (0.09–611.83 kg/cm 2 ), reaching adequate levels of correlation ( r = 0.669 ; p = 0.001 ).

4.3.2. Physicochemical Properties (Organic Residue) with Respect to Thermal Properties (Compound)

Figure 6 shows the non linear correlations between the physicochemical properties of the residue and the thermal properties of the compound. Figure 6a indicates the relationship between the residue density (0.08–1.4 g/cm 3 ) and the thermal conductivity of the compound (0.05–1.13 W/mK), obtaining adequate correlation indicators ( r = 0.537 ; p = 0.001 ). Figure 6b shows the relationship between the extractives content from organic residue (1.5–15.58%), w/w) and the thermal conductivity of the compound (0.09–0.33 W/mK), which showed adequate correlations ( r = 0.176 ; p = 0.006 ). Therefore, materials made up of organic residues with a density between the ranges 0.08 and 0.25 g/cm3 and 1.18 and 1.45 g/cm3 have a lower thermal conductivity (0.05–0.34 W/mK). Whereas, in materials with incorporation of organic residues with high extractives contents (12–15.6%, w/w), lower values are also obtained in the thermal conductivity (0.09–0.14 W/mK).
Thermal conductivity is the physical property that measures the ability of materials to conduct heat through its mass. Therefore, the less the thermal conductivity a material generates, the better insulating properties in construction applications it has. Indeed, conventional building materials, such as cement and aggregates forming concrete reach thermal conductivities up to 1.7 W/mK, being increased up to 2.2 W/mK when water saturation ratio has reached 100%, v/v [48]. In this bibliometric analysis, compounds using organic residues (≥0.5%, v/v) reached thermal conductivities from 0.1 to 1.9 W/mK. Moreover, the moisture content of these residues (≥70%, v/v) increases their density and therefore decreases their insulating properties. Effectively, the obtained polynomial regression shown in Figure 6a displays that this behavior only occurs for densities close to the water density (approx. 1.0 g/cm3). On the other hand, low thermal conductivity are obtained when densities are greater than 1.0 g/cm3. This can be explained due to the a greater amount of extractives within the organic residues composition.
The extractives from organic residues are natural polymers, that is, macro molecules formed by monomers joined by covalent bonds that prevents the free flow of electrons. Therefore, a higher extractives content (≥2%, w/w) could improve thermal properties, reaching values between 0.35 and 0.09 W/mK (refer to Figure 6b). In specific, non wood biomass offers better thermal conditions than wood biomass, due to its extractives content (5–15%, w/w) [49,50].

5. Conclusions

The bibliometric analysis in this engineering sciences field establishes that 54% of the scientific publications that study organic residues and building materials focus on the reuse and revaluation of organic residues, incorporating them as additives or as replacement, total or partial, to the conventional materials (i.e., cement, aggregates, and water). The number of studies that address these issues show an annual growth rate of 30%. Mortar is presented as the most recurrent construction use (32%), followed by concrete, blocks and bricks, and insulating panels. 82% of the organic residues studied come from the agricultural industry. Regarding the results obtained on the compressive strength of the compounds, mortar (309 Kgf/cm2) is projected as a potential structural constructive use of organic residue with compounds percentages between 5 and 30%), v/v. The tensile strength of cement based materials was statistically correlated with materials incorporating organic residues. To measure this mechanical parameter properly, density of the residue was used as a dependent parameter. To this end, the proportion of the variance for the dependent variable in our regression model was interpreted as the R 2 value of 0.73 . The residue density is directly related with the extractives content within the vegetable biomass composition. As a consequence, as the extractive content increases, the density of the organic residue increases, and the tensile strength also increases. The opposite outcome is found for organic residues with high moisture content, where density usually decreases. The relationship between the lignin content (dependant variable) of the residue and the flexural strength of the compound (independent variable) was established using a regression analysis as shown in Table 3 and Figure 5c. The proportion of the variance for the dependent variable in the regression model was interpreted as the R 2 value of 0.84 , which demonstrate good correlation. From this results, organic residues with lower lignin content (7.4–9.6%, w/w), have higher flexural strength (83–99 kg/cm 2 ). In specific, lignin is an amorphous molecule that contains an internal structure with aliphatic chains (polyphenolic polymer), which gives rigidity to the plant cell wall. Consequently, organic residues from herbaceous biomass (lignin content from hard/soft woods range between 21 and 32%) have better chances of improving flexural properties in compounds. On the other hand, the use in panels (41 Kgf/cm2) would be the most relevant application for non-structural purposes. The identification of the physicochemical properties of the organic residues allow us to predict the mechanical and thermal behavior that the material composed of this residue will have. The statistical analysis carried out showed adequate levels of correlation ( r > 0.5 , p < 0.05 ) between some physicochemical properties (density and lignin content) of organic residues and mechanical/thermal properties (compressive strength, tensile strength, flexural strength, and thermal conductivity) of the compounds.

Author Contributions

Conceptualization, S.Y. and C.A.V.-A.; methodology, S.Y. and C.A.V.-A.; experimental data acquisition and processing, S.Y., C.M., B.V., and C.A.V.-A.; formal analysis, S.Y., C.M., B.V., and C.A.V.-A.; investigation, S.Y. and C.A.V.-A.; resources, S.Y. and C.A.V.-A.; writing—original draft preparation, S.Y. and C.A.V.-A.; writing—review and editing, S.Y. and C.A.V.-A.; visualization, S.Y. and C.A.V.-A.; project administration, S.Y. and C.A.V.-A.; funding acquisition, S.Y. and C.A.V.-A. All authors have read and agreed to the published version of the manuscript.

Funding

S.Y. acknowledges the financial support from Universidad de Santiago de Chile, Usach, through project DICYT N° 052018YC, Dirección de Investigación Científica y Tecnológica, Dicyt. C.A.V.-A. and S.Y. acknowledge financial support from Universidad de Santiago de Chile, Usach, though project grant DGT-TR2007. C.A.V.-A., C.M. and B.V. acknowledge the financial support from Fondecyt (project grant 11190352).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We want to pay a fair tribute with our work, to all the people who have suffered and died from COVID-19, in addition to the medical staff, for their work and dedication in these difficult times.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A. List of References Used in This Review Article

1.
Abbas, S., Kazmi, S. M., Munir, M. J. (2017). Potential of rice husk ash for mitigating the alkali-silica reaction in mortar bars incorporating reactive aggregates. Construction and Building Materials, 132, 61–70.
2.
Acosta, Y. L., Abreu, M. C. O. (2005). La digestión anaerobia. Aspectos teóricos. Parte I. ICIDCA. Sobre los Derivados de la Caña de Azúcar. 39, 35–48.
3.
Agoua, E., Allognon-Houessou, E., Adjovi, E., Togbedji, B. (2013). Thermal conductivity of composites made of wastes of wood and expanded polystyrene. Construction and Building Materials, 41, 557–562.
4.
Aksoğan, O., Binici, H., Ortlek, E. (2016). Durability of concrete made by partial replacement of fine aggregate by colemanite and barite and cement by ashes of corn stalk, wheat straw and sunflower stalk ashes. Construction and Building Materials, 106, 253–263.
5.
Alabdulkarem, A., Ali, M., Iannace, G., Sadek, S., Almuzaiqer, R. (2018). Thermal analysis, microstructure and acoustic characteristics of some hybrid natural insulating materials. Construction and Building Materials, 187, 185–196.
6.
Alavez-Ramirez, R., Chiñas-Castillo, F., Morales-Dominguez, V. J., Ortiz-Guzman, M. (2012). Thermal conductivity of coconut fibre filled ferrocement sandwich panels. Construction and Building Materials, 37, 425–431.
7.
Alavez-Ramirez, R., Montes-Garcia, P., Martínez-Reyes, J., Altamirano-Juárez, D. C., Gochi-Ponce, Y. (2012). The use of sugarcane bagasse ash and lime to improve the durability and mechanical properties of compacted soil blocks. Construction and building materials, 34, 296–305.
8.
Ali, M. E., Alabdulkarem, A. (2017). On thermal characteristics and microstructure of a new insulation material extracted from date palm trees surface fibers. Construction and Building Materials, 138, 276–284.
9.
Algin, H. M., Turgut, P. (2008). Cotton and limestone powder wastes as brick material. Construction and Building Materials, 22(6), 1074–1080.
10.
Alkaya, E., Demirer, G. N. (2013). Greening of production in metal processing industry through process modifications and improved management practices. Resour. Conserv. Recycl. 77, 89–96.
11.
Aluko, O. G., Yatim, J. M., Kadir, M. A. A., Yahya, K. (2020). A review of properties of bio-fibrous concrete exposed to elevated temperatures. Construction and Building Materials, 260, 119671.
12.
Aquim, P. M., Hansen, É., Gutterres, M. (2019). Water reuse: An alternative to minimize the environmental impact on the leather industry. J. Environ. Manage. 230, 456–463.
13.
Aouba, L., Bories, C., Coutand, M., Perrin, B., Lemercier, H. (2016). Properties of fired clay bricks with incorporated biomasses: cases of olive stone flour and wheat straw residues. Construction and Building Materials, 102, 7–13.
14.
Arif, E., Clark, M. W., Lake, N. (2016). Sugar cane bagasse ash from a high efficiency co-generation boiler: Applications in cement and mortar production. Construction and Building Materials, 128, 287–297.
15.
Ataie, F. F., Riding, K. A. (2016). Influence of agricultural residue ash on early cement hydration and chemical admixtures adsorption. Construction and Building Materials, 106, 274–281.
16.
Awwad, E., Mabsout, M., Hamad, B., Farran, M. T., Khatib, H. (2012). Studies on fiber-reinforced concrete using industrial hemp fibers. Construction and Building Materials, 35, 710–717.
17.
Aziz, F. N. A. A., Bida, S. M., Nasir, N. A. M., Jaafar, M. S. (2014). Mechanical properties of lightweight mortar modified with oil palm fruit fibre and tire crumb. Construction and Building Materials, 73, 544–550.
18.
Baenla, J., Mbah, J. B., Ndjock, I. D. L., Elimbi, A. (2019). Partial replacement of low reactive volcanic ash by cassava peel ash in the synthesis of volcanic ash based geopolymer. Construction and Building Materials, 227, 116689.
19.
Barbieri, V., Gualtieri, M. L., Siligardi, C. (2020). Wheat husk: A renewable resource for bio-based building materials. Construction and Building Materials, 251, 118909.
20.
Barbosa, R., Lapa, N., Dias, D., Mendes, B. (2013). Concretes containing biomass ashes: Mechanical, chemical, and ecotoxic performances. Construction and Building Materials, 48, 457–463.
21.
Beltrán, M. G., Barbudo, A., Agrela, F., Jiménez, J. R., de Brito, J. (2016). Mechanical performance of bedding mortars made with olive biomass bottom ash. Construction and Building Materials, 112, 699—707.
22.
Berger, F., Gauvin, F., Brouwers, H. J. H. (2020). The recycling potential of wood waste into wood-wool/cement composite. Construction and Building Materials, 260, 119786.
23.
Berra, M., Mangialardi, T., Paolini, A. E. (2015). Reuse of woody biomass fly ash in cement-based materials. Construction and Building Materials, 76, 286–296.
24.
Binici, H., Aksogan, O., Sevinc, A. H., Cinpolat, E. (2015). Mechanical and radioactivity shielding performances of mortars made with cement, sand and egg shells. Construction and Building Materials, 93, 1145–1150.
25.
Binici, H., Eken, M., Kara, M., Dolaz, M. (2014). An environment-friendly thermal insulation material from sunflower stalk, textile waste and stubble fibers. Construction and Building Materials, 51, 24–33.
26.
Bołtryk, M., Pawluczuk, E. (2014). Properties of a lightweight cement composite with an ecological organic filler. Construction and Building Materials, 51, 97–105.
27.
Buratti, C., Belloni, E., Lascaro, E., Merli, F., Ricciardi, P. (2018). Rice husk panels for building applications: Thermal, acoustic and environmental characterization and comparison with other innovative recycled waste materials. Construction and Building Materials, 171, 338–349.
28.
Chabannes, M., Bénézet, J. C., Clerc, L., Garcia-Diaz, E. (2014). Use of raw rice husk as natural aggregate in a lightweight insulating concrete: An innovative application. Construction and Building Materials, 70, 428–438.
29.
Chabi, E., Lecomte, A., Adjovi, E. C., Dieye, A., Merlin, A. (2018). Mix design method for plant aggregates concrete: Example of the rice husk. Construction and Building Materials, 174, 233–243.
30.
Cheah, C. B., Ramli, M. (2012). Mechanical strength, durability and drying shrinkage of structural mortar containing HCWA as partial replacement of cement. Construction and Building Materials, 30, 320–329.
31.
Chen, Y., Wang, T., Zhou, M., Hou, H., Xue, Y., Wang, H. (2018). Rice husk and sewage sludge co-combustion ash: Leaching behavior analysis and cementitious property. Construction and Building Materials, 163, 63–72.
32.
Corrêa, A. A. R., Mendes, L. M., Barbosa, N. P., de Paula Protásio, T., de Aguiar Campos, N., Tonoli, G. H. D. (2015). Incorporation of bamboo particles and “synthetic termite saliva” in adobes. Construction and Building Materials, 98, 250–256.
33.
de Moraes Pinheiro, S. M., Font, A., Soriano, L., Tashima, M. M., Monzó, J., Borrachero, M. V., Payá, J. (2018). Olive-stone biomass ash (OBA): An alternative alkaline source for the blast furnace slag activation. Construction and Building Materials, 178, 327–338.
34.
El-Hawary, M. M., Hamoush, S. A. (1994). Strength of concrete reinforced with treated feathers. Construction and Building Materials, 8(3), 173–176.
35.
El Wardi, F. Z., Khabbazi, A., Cherki, A. B., Khaldoun, A. (2020). Thermomechanical study of a sandwich material with ecological additives. Construction and Building Materials, 252, 119093.
36.
Eliche-Quesada, D., Iglesias-Godino, F. J., Pérez-Villarejo, L., Corpas-Iglesias, F. A. (2014). Replacement of the mixing fresh water by wastewater olive oil extraction in the extrusion of ceramic bricks. Construction and Building Materials, 68, 659–666.
37.
Eliche-Quesada, D., Corpas-Iglesias, F. A., Pérez-Villarejo, L., Iglesias-Godino, F. J. (2012). Recycling of sawdust, spent earth from oil filtration, compost and marble residues for brick manufacturing. Construction and Building Materials, 34, 275–284.
38.
Erkmen, L., Yavuz, H. I., Kavci, E., Sari, M. (2020). A new environmentally friendly insulating material designed from natural materials. Construction and Building Materials, 255, 119357.
39.
Escalera, E., Garcia, G., Terán, R., Tegman, R., Antti, M. L., Odén, M. (2015). The production of porous brick material from diatomaceous earth and Brazil nut shell ash. Construction and Building Materials, 98, 257–264.
40.
Esmeray, E., Atıs, M. (2019). Utilization of sewage sludge, oven slag and fly ash in clay brick production. Construction and Building Materials, 194, 110–121.
41.
Esteves, T. C., Rajamma, R., Soares, D., Silva, A. S., Ferreira, V. M., & Labrincha, J. A. (2012). Use of biomass fly ash for mitigation of alkali-silica reaction of cement mortars. Construction and Building Materials, 26(1), 687–693.
42.
Farooqi, M. U., Ali, M. (2019). Effect of pre-treatment and content of wheat straw on energy absorption capability of concrete. Construction and Building Materials, 224, 572–583.
43.
Ferrándiz-Mas, V., Bond, T., García-Alcocel, E., Cheeseman, C. R. (2014). Lightweight mortars containing expanded polystyrene and paper sludge ash. Construction and Building Materials, 61, 285–292.
44.
Ferraro, R. M., Nanni, A. (2012). Effect of off-white rice husk ash on strength, porosity, conductivity and corrosion resistance of white concrete. Construction and Building Materials, 31, 220–225.
45.
Foti, D., Cavallo, D. (2018). Mechanical behavior of concretes made with non-conventional organic origin calcareous aggregates. Construction and Building Materials, 179, 100–106.
46.
Frías, M., Rodríguez, O., de Rojas, M. I. S., Villar-Cociña, E., Rodrigues, M. S., Junior, H. S. (2017). Advances on the development of ternary cements elaborated with biomass ashes coming from different activation process. Construction and Building Materials, 136, 73–80.
47.
Frías, M., Rodríguez, O., De Rojas, M. S. (2015). Paper sludge, an environmentally sound alternative source of MK-based cementitious materials. A review. Construction and Building Materials, 74, 37–48.
48.
Fu, Q., Yan, L., Ning, T., Wang, B., Kasal, B. (2020). Behavior of adhesively bonded engineered wood–Wood chip concrete composite decks: Experimental and analytical studies. Construction and Building Materials, 247, 118578.
49.
Fusade, L., Viles, H., Wood, C., Burns, C. (2019). The effect of wood ash on the properties and durability of lime mortar for repointing damp historic buildings. Construction and Building Materials, 212, 500–513.
50.
Gao, X., Yang, Y., Deng, H. (2011). Utilization of beet molasses as a grinding aid in blended cements. Construction and Building Materials, 25(9), 3782–3789.
51.
Goel, G., & Kalamdhad, A. S. (2017). An investigation on use of paper mill sludge in brick manufacturing. Construction and Building Materials, 148, 334–343.
52.
Görhan, G., Şimşek, O. (2013). Porous clay bricks manufactured with rice husks. Construction and Building Materials, 40, 390–396.
53.
Gunasekaran, K., Annadurai, R., Kumar, P. S. (2012). Long term study on compressive and bond strength of coconut shell aggregate concrete. Construction and Building Materials, 28(1), 208–215.
54.
Guo, A., Aamiri, O. B., Satyavolu, J., Sun, Z. (2019). Impact of thermally modified wood on mechanical properties of mortar. Construction and Building Materials, 208, 413–420.
55.
Gupta, S., Kua, H. W. (2020). Application of rice husk biochar as filler in cenosphere modified mortar: Preparation, characterization and performance under elevated temperature. Construction and Building Materials, 253, 119083. G
56.
upta, S., Kua, H. W., Dai Pang, S. (2018). Biochar-mortar composite: Manufacturing, evaluation of physical properties and economic viability. Construction and Building Materials, 167, 874–889.
57.
Haik, R., Bar-Nes, G., Peled, A., Meir, I. A. (2020). Alternative unfired binders as lime replacement in hemp concrete. Construction and Building Materials, 241, 117981.
58.
Hamada, H., Tayeh, B., Yahaya, F., Muthusamy, K., Al-Attar, A. (2020). Effects of nano-palm oil fuel ash and nano-eggshell powder on concrete. Construction and Building Materials, 261, 119790.
59.
Hamdaoui, O., Ibos, L., Mazioud, A., Safi, M., Limam, O. (2018). Thermophysical characterization of Posidonia Oceanica marine fibers intended to be used as an insulation material in Mediterranean buildings. Construction and Building Materials, 180, 68–76.
60.
Hernández-Olivares, F., Bollati, M. R., Del Rio, M., Parga-Landa, B. (1999). Development of cork–gypsum composites for building applications. Construction and Building Materials, 13(4), 179–186.
61.
Hernández-Olivares, F., Medina-Alvarado, R. E., Burneo-Valdivieso, X. E., Zúñiga-Suárez, A. R. (2020). Short sugarcane bagasse fibers cementitious composites for building construction. Construction and Building Materials, 247, 118451.
62.
Husni, H., Nazari, M. R., Yee, H. M., Rohim, R., Yusuff, A., Ariff, M. A. M., Junaidi, M. U. M. (2017). Superhydrophobic rice husk ash coating on concrete. Construction and Building Materials, 144, 385–391.
63.
Jauberthie, R., Rendell, F., Tamba, S., Cisse, I. (2000). Origin of the pozzolanic effect of rice husks. Construction and Building Materials, 14(8), 419–423.
64.
Joshaghani, A., Moeini, M. A. (2017). Evaluating the effects of sugar cane bagasse ash (SCBA) and nanosilica on the mechanical and durability properties of mortar. Construction and building materials, 152, 818–831.
65.
Jumaat, M. Z., Alengaram, U. J., Ahmmad, R., Bahri, S., Islam, A. S. (2015). Characteristics of palm oil clinker as replacement for oil palm shell in lightweight concrete subjected to elevated temperature. Construction and Building Materials, 101, 942-951.
66.
Kairytė, A., Kremensas, A., Balčiūnas, G., Matulaitienė, I., Członka, S., Sienkiewicz, N. (2020). Evaluation of self-thermally treated wood plastic composites from wood bark and rapeseed oil-based binder. Construction and Building Materials, 250, 118842.
67.
Kairytė, A., Vaitkus, S., Kremensas, A., Pundienė, I., Członka, S., Strzelec, K. (2019). Moisture-mechanical performance improvement of thermal insulating polyurethane using paper production waste particles grafted with different coupling agents. Construction and Building Materials, 208, 525–534.
68.
Karim, M. R., Hashim, H., Razak, H. A. (2016). Thermal activation effect on palm oil clinker properties and their influence on strength development in cement mortar. Construction and Building Materials, 125, 670–678.
69.
Katare, V. D., Madurwar, M. V. (2017). Experimental characterization of sugarcane biomass ash–A review. Construction and Building Materials, 152, 1–15.
70.
Kaya, G. G., Yilmaz, E., Deveci, H. (2018). Sustainable bean pod/calcined kaolin reinforced epoxy hybrid composites with enhanced mechanical, water sorption and corrosion resistance properties. Construction and Building Materials, 162, 272–279.
71.
Kazmi, S. M. S., Munir, M. J., Patnaikuni, I., Wu, Y. F. (2017). Pozzolanic reaction of sugarcane bagasse ash and its role in controlling alkali silica reaction. Construction and Building Materials, 148, 231–240.
72.
Kazmi, S. M., Abbas, S., Saleem, M. A., Munir, M. J., Khitab, A. (2016). Manufacturing of sustainable clay bricks: Utilization of waste sugarcane bagasse and rice husk ashes. Construction and building materials, 120, 29–41.
73.
Khan, R., Jabbar, A., Ahmad, I., Khan, W., Khan, A. N., Mirza, J. (2012). Reduction in environmental problems using rice-husk ash in concrete. Construction and Building Materials, 30, 360–365.
74.
Khan, K., Ullah, M. F., Shahzada, K., Amin, M. N., Bibi, T., Wahab, N., Aljaafari, A. (2020). Effective use of micro-silica extracted from rice husk ash for the production of high-performance and sustainable cement mortar. Construction and Building Materials, 258, 119589.
75.
Khan, M., Ali, M. (2018). Effect of super plasticizer on the properties of medium strength concrete prepared with coconut fiber. Construction and Building Materials, 182, 703–715.
76.
Khan, W., Shehzada, K., Bibi, T., Islam, S. U., Khan, S. W. (2018). Performance evaluation of Khyber Pakhtunkhwa Rice Husk Ash (RHA) in improving mechanical behavior of cement. Construction and Building Materials, 176, 89–102.
77.
Khoshnava, S. M., Rostami, R., Ismail, M., Rahmat, A. R., Ogunbode, B. E. (2017). Woven hybrid Biocomposite: Mechanical properties of woven kenaf bast fibre/oil palm empty fruit bunches hybrid reinforced poly hydroxybutyrate biocomposite as non-structural building materials. Construction and Building Materials, 154, 155–166.
78.
Kizinievič, O., Kizinievič, V., Malaiškienė, J. (2018). Analysis of the effect of paper sludge on the properties, microstructure and frost resistance of clay bricks. Construction and Building Materials, 169, 689–696.
79.
Kochova, K., Gauvin, F., Schollbach, K., Brouwers, H. J. H. (2020). Using alternative waste coir fibres as a reinforcement in cement-fibre composites. Construction and Building Materials, 231, 117121.
80.
Koohestani, B., Koubaa, A., Belem, T., Bussière, B., Bouzahzah, H. (2016). Experimental investigation of mechanical and microstructural properties of cemented paste backfill containing maple-wood filler. Construction and Building Materials, 121, 222–228.
81.
Kuo, W. T., Wang, H. Y., Shu, C. Y., Su, D. S. (2013). Engineering properties of controlled low-strength materials containing waste oyster shells. Construction and Building Materials, 46, 128–133.
82.
Kupaei, R. H., Alengaram, U. J., Jumaat, M. Z. B., Nikraz, H. (2013). Mix design for fly ash based oil palm shell geopolymer lightweight concrete. Construction and Building Materials, 43, 490–496.
83.
Lertsatitthanakorn, C., Atthajariyakul, S., Soponronnarit, S. (2009). Techno-economical evaluation of a rice husk ash (RHA) based sand–cement block for reducing solar conduction heat gain to a building. Construction and Building Materials, 23(1), 364–369.
84.
Liu, H. Y., Wu, H. S., Chou, C. P. (2020). Study on engineering and thermal properties of environment-friendly lightweight brick made from Kinmen oyster shells & sorghum waste. Construction and Building Materials, 246, 118367.
85.
Liu, M., Zhao, Y., Xiao, Y., Yu, Z. (2019). Performance of cement pastes containing sewage sludge ash at elevated temperatures. Construction and Building Materials, 211, 785–795.
86.
Liu, Z., Chen, Q., Xie, X., Xue, G., Du, F., Ning, Q., Huang, L. (2011). Utilization of the sludge derived from dyestuff-making wastewater coagulation for unfired bricks. Construction and Building Materials, 25(4), 1699–1706.
87.
Luo, L., Li, K., Fu, W., Liu, C., Yang, S. (2020). Preparation, characteristics and mechanisms of the composite sintered bricks produced from shale, sewage sludge, coal gangue powder and iron ore tailings. Construction and Building Materials, 232, 117250.
88.
Madrid, M., Orbe, A., Carré, H., García, Y. (2018). Thermal performance of sawdust and lime-mud concrete masonry units. Construction and Building Materials, 169, 113–123.
89.
Madrid, M., Orbe, A., Rojí, E., Cuadrado, J. (2017). The effects of by-products incorporated in low-strength concrete for concrete masonry units. Construction and Building Materials, 153, 117–128.
90.
Marques, B., Tadeu, A., António, J., Almeida, J., de Brito, J. (2020). Mechanical, thermal and acoustic behaviour of polymer-based composite materials produced with rice husk and expanded cork by-products. Construction and Building Materials, 239, 117851.
91.
Martínez-García, C., González-Fonteboa, B., Carro-López, D., Martínez-Abella, F. (2019). Design and properties of cement coating with mussel shell fine aggregate. Construction and Building Materials, 215, 494–507.
92.
Martínez-García, C., González-Fonteboa, B., Martínez-Abella, F., Carro-López, D. (2017). Performance of mussel shell as aggregate in plain concrete. Construction and building materials, 139, 570–583.
93.
Memon, S. A., Javed, U., Khushnood, R. A. (2019). Eco-friendly utilization of corncob ash as partial replacement of sand in concrete. Construction and Building Materials, 195, 165–177.
94.
Mohseni, E., Kazemi, M. J., Koushkbaghi, M., Zehtab, B., Behforouz, B. (2019). Evaluation of mechanical and durability properties of fiber-reinforced lightweight geopolymer composites based on rice husk ash and nano-alumina. Construction and Building Materials, 209, 532–540.
95.
Moraes, M. J. B., Moraes, J. C. B., Tashima, M. M., Akasaki, J. L., Soriano, L., Borrachero, M. V., Payá, J. (2019). Production of bamboo leaf ash by auto-combustion for pozzolanic and sustainable use in cementitious matrices. Construction and Building Materials, 208, 369–380.
96.
Moraes, J. C. B., Akasaki, J. L., Melges, J. L. P., Monzó, J., Borrachero, M. V., Soriano, L., Tashima, M. M. (2015). Assessment of sugar cane straw ash (SCSA) as pozzolanic material in blended Portland cement: Microstructural characterization of pastes and mechanical strength of mortars. Construction and Building Materials, 94, 670–677.
97.
Moretti, J. P., Nunes, S., Sales, A. (2018). Self-compacting concrete incorporating sugarcane bagasse ash. Construction and building materials, 172, 635–649.
98.
Muntohar, A. S., Rahman, M. E. (2014). Lightweight masonry block from oil palm kernel shell. Construction and Building Materials, 54, 477–484.
99.
Muntohar, A. S. (2011). Engineering characteristics of the compressed-stabilized earth brick. Construction and Building Materials, 25(11), 4215–4220.
100.
Muñoz, P., Mendívil, M. A., Letelier, V., Morales, M. P. (2019). Thermal and mechanical properties of fired clay bricks made by using grapevine shoots as pore forming agent. Influence of particle size and percentage of replacement. Construction and Building Materials, 224, 639–658.
101.
Muñoz, P., Morales, M. P., Mendívil, M. A., Juárez, M. C., Muñoz, L. (2014). Using of waste pomace from winery industry to improve thermal insulation of fired clay bricks. Eco-friendly way of building construction. Construction and Building Materials, 71, 181–187.
102.
Mymrin, V., Meyer, S. A. S., Alekseev, K. P., Pawlowsky, U., Fernandes, L. H., Scremim, C. B., Catai, R. E. (2014). Manufacture of a construction material using the physicochemical properties of ash and sludge wastes from MDF board production. Construction and Building Materials, 50, 184–189.
103.
Naqi, A., Siddique, S., Kim, H. K., Jang, J. G. (2020). Examining the potential of calcined oyster shell waste as additive in high volume slag cement. Construction and Building Materials, 230, 116973.
104.
Nguyen, D. H., Boutouil, M., Sebaibi, N., Baraud, F., Leleyter, L. (2017). Durability of pervious concrete using crushed seashells. Construction and Building Materials, 135, 137–150.
105.
Nguyen, D. H., Boutouil, M., Sebaibi, N., Leleyter, L., Baraud, F. (2013). Valorization of seashell by-products in pervious concrete pavers. Construction and Building Materials, 49, 151–160.
106.
Novais, R. M., Carvalheiras, J., Senff, L., Labrincha, J. A. (2018). Upcycling unexplored dregs and biomass fly ash from the paper and pulp industry in the production of eco-friendly geopolymer mortars: A preliminary assessment. Construction and Building Materials, 184, 464–472.
107.
Olacia, E., Pisello, A. L., Chiodo, V., Maisano, S., Frazzica, A., Cabeza, L. F. (2020). Sustainable adobe bricks with seagrass fibres. Mechanical and thermal properties characterization. Construction and Building Materials, 239, 117669.
108.
Olutoge, F. A., Adesina, P. A. (2019). Effects of rice husk ash prepared from charcoal-powered incinerator on the strength and durability properties of concrete. Construction and Building Materials, 196, 386–394.
109.
Ozturk, S., Sutcu, M., Erdogmus, E., Gencel, O. (2019). Influence of tea waste concentration in the physical, mechanical and thermal properties of brick clay mixtures. Construction and Building Materials, 217, 592–599.
110.
Pacheco-Torgal, F., Jalali, S. (2011). Cementitious building materials reinforced with vegetable fibres: A review. Construction and Building Materials, 25(2), 575–581.
111.
Pinto, J., Vieira, B., Pereira, H., Jacinto, C., Vilela, P., Paiva, A., Varum, H. (2012). Corn cob lightweight concrete for non-structural applications. Construction and Building Materials, 34, 346–351.
112.
Pliya, P., Cree, D. (2015). Limestone derived eggshell powder as a replacement in Portland cement mortar. Construction and Building Materials, 95, 1–9.
113.
Rajput, D., Bhagade, S. S., Raut, S. P., Ralegaonkar, R. V., Mandavgane, S. A. (2012). Reuse of cotton and recycle paper mill waste as building material. Construction and Building Materials, 34, 470–475.
114.
Rashid, K., Haq, E. U., Kamran, M. S., Munir, N., Shahid, A., Hanif, I. (2019). Experimental and finite element analysis on thermal conductivity of burnt clay bricks reinforced with fibers. Construction and Building Materials, 221, 190–199.
115.
Raut, S. P., Sedmake, R., Dhunde, S., Ralegaonkar, R. V., Mandavgane, S. A. (2012). Reuse of recycle paper mill waste in energy absorbing light weight bricks. Construction and Building Materials, 27(1), 247–251.
116.
Rubia-García, M. D., Yebra-Rodríguez, Á., Eliche-Quesada, D., Corpas-Iglesias, F. A., López-Galindo, A. (2012). Assessment of olive mill solid residue (pomace) as an additive in lightweight brick production. Construction and Building Materials, 36, 495–500.
117.
Russ, W., Mörtel, H., Meyer-Pittroff, R. (2005). Application of spent grains to increase porosity in bricks. Construction and Building Materials, 19(2), 117-126.
118.
Saeli, M., Senff, L., Tobaldi, D. M., Carvalheiras, J., Seabra, M. P., Labrincha, J. A. (2020). Unexplored alternative use of calcareous sludge from the paper-pulp industry in green geopolymer construction materials. Construction and Building Materials, 246, 118457.
119.
Saeli, M., Senff, L., Tobaldi, D. M., Seabra, M. P., Labrincha, J. A. (2019). Novel biomass fly ash-based geopolymeric mortars using lime slaker grits as aggregate for applications in construction: Influence of granulometry and binder/aggregate ratio. Construction and Building Materials, 227, 116643.
120.
Safi, B., Saidi, M., Daoui, A., Bellal, A., Mechekak, A., Toumi, K. (2015). The use of seashells as a fine aggregate (by sand substitution) in self-compacting mortar (SCM). Construction and Building Materials, 78, 430–438.
121.
Sales, A., De Souza, F. R., Dos Santos, W. N., Zimer, A. M., Almeida, F. D. C. R. (2010). Lightweight composite concrete produced with water treatment sludge and sawdust: Thermal properties and potential application. Construction and building materials, 24(12), 2446–2453.
122.
Sathiparan, N., Rupasinghe, M. N., Pavithra, B. H. (2017). Performance of coconut coir reinforced hydraulic cement mortar for surface plastering application. Construction and Building Materials, 142, 23–30.
123.
Schackow, A., Effting, C., Barros, V. G., Gomes, I. R., da Costa Neto, V. S., Delandréa, M. S. (2020). Permeable concrete plates with wastes from the paper industry: Reduction of surface flow and possible applications. Construction and Building Materials, 250, 118896.
124.
Sekar, S. K. (2016). Mechanical and fracture characteristics of eco-friendly concrete produced using coconut shell, ground granulated blast furnace slag and manufactured sand. Construction and Building materials, 103, 1–7.
125.
Shafigh, P., Mahmud, H. B., Jumaat, M. Z., Zargar, M. (2014). Agricultural wastes as aggregate in concrete mixtures–A review. Construction and Building Materials, 53, 110–117.
126.
Shakouri, M., Exstrom, C. L., Ramanathan, S., Suraneni, P. (2020). Hydration, strength, and durability of cementitious materials incorporating untreated corn cob ash. Construction and Building Materials, 243, 118171.
127.
Sheridan, J., Sonebi, M., Taylor, S., Amziane, S. (2020). The effect of a polyacrylic acid viscosity modifying agent on the mechanical, thermal and transport properties of hemp and rapeseed straw concrete. Construction and Building Materials, 235, 117536.
128.
Sirico, A., Bernardi, P., Belletti, B., Malcevschi, A., Dalcanale, E., Domenichelli, I., Moretti, E. (2020). Mechanical characterization of cement-based materials containing biochar from gasification. Construction and Building Materials, 246, 118490.
129.
Taallah, B., Guettala, A. (2016). The mechanical and physical properties of compressed earth block stabilized with lime and filled with untreated and alkali-treated date palm fibers. Construction and Building Materials, 104, 52–62.
130.
Tioua, T., Kriker, A., Barluenga, G., Palomar, I. (2017). Influence of date palm fiber and shrinkage reducing admixture on self-compacting concrete performance at early age in hot-dry environment. Construction and Building Materials, 154, 721–733.
131.
Ukwatta, A., Mohajerani, A. (2017). Characterisation of fired-clay bricks incorporating biosolids and the effect of heating rate on properties of bricks. Construction and Building Materials, 142, 11–22.
132.
Ukwatta, A., Mohajerani, A., Setunge, S., Eshtiaghi, N. (2015). Possible use of biosolids in fired-clay bricks. Construction and Building Materials, 91, 86–93.
133.
Varhen, C., Carrillo, S., Ruiz, G. (2017). Experimental investigation of Peruvian scallop used as fine aggregate in concrete. Construction and Building Materials, 136, 533–540.
134.
Vegas, I., Urreta, J., Frías, M., García, R. (2009). Freeze–thaw resistance of blended cements containing calcined paper sludge. Construction and Building Materials, 23(8), 2862–2868.
135.
Viel, M., Collet, F., Lanos, C. (2019). Development and characterization of thermal insulation materials from renewable resources. Construction and Building Materials, 214, 685–697.
136.
Villamar, C. A., Rodríguez, D. C., López, D., Peñuela, G., Vidal, G. 2013. Effect of the generation and physical-chemical characterization of swine and dairy cattle slurries on treatment technologies. Waste Manag. Res. 31, 820–828.
137.
Villamizar, M. C. N., Araque, V. S., Reyes, C. A. R., Silva, R. S. (2012). Effect of the addition of coal-ash and cassava peels on the engineering properties of compressed earth blocks. Construction and Building Materials, 36, 276–286.
138.
Walker, R., Pavia, S., Mitchell, R. (2014). Mechanical properties and durability of hemp-lime concretes. Construction and Building Materials, 61, 340–348.
139.
Wang, T., Xue, Y., Zhou, M., Lv, Y., Chen, Y., Wu, S., Hou, H. (2017). Hydration kinetics, freeze-thaw resistance, leaching behavior of blended cement containing co-combustion ash of sewage sludge and rice husk. Construction and Building Materials, 131, 361–370.
140.
Wu, F., Yu, Q., Liu, C., Brouwers, H. J. H., Wang, L. (2019). Effect of surface treatment of apricot shell on the performance of lightweight bio-concrete. Construction and Building Materials, 229, 116859.
141.
Wu, F., Liu, C., Zhang, L., Lu, Y., Ma, Y. (2018). Comparative study of carbonized peach shell and carbonized apricot shell to improve the performance of lightweight concrete. Construction and Building Materials, 188, 758–771.
142.
Xie, X., Gou, G., Wei, X., Zhou, Z., Jiang, M., Xu, X., Hui, D. (2016). Influence of pretreatment of rice straw on hydration of straw fiber filled cement based composites. Construction and Building Materials, 113, 449–455.
143.
Xu, R., He, T., Da, Y., Liu, Y., Li, J., Chen, C. (2019). Utilizing wood fiber produced with wood waste to reinforce autoclaved aerated concrete. Construction and Building Materials, 208, 242–249.
144.
Xu, W., Lo, T. Y., Memon, S. A. (2012). Microstructure and reactivity of rich husk ash. Construction and Building Materials, 29, 541–547.
145.
Yadav, A. L., Sairam, V., Srinivasan, K., Muruganandam, L. (2020). Synthesis and characterization of geopolymer from metakaolin and sugarcane bagasse ash. Construction and Building Materials, 258, 119231.
146.
Yan, S., Sagoe-Crentsil, K., Shapiro, G. (2012). Properties of cement mortar incorporating de-inking waste-water from waste paper recycling. Construction and Building Materials, 29, 51–55.
147.
Yaras, A. (2020). Combined effects of paper mill sludge and carbonation sludge on characteristics of fired clay bricks. Construction and Building Materials, 249, 118722.
148.
Yuzer, N., Cinar, Z., Akoz, F., Biricik, H., Gurkan, Y. Y., Kabay, N., Kizilkanat, A. B. (2013). Influence of raw rice husk addition on structure and properties of concrete. Construction and Building Materials, 44, 54–62.
149.
Zak, P., Ashour, T., Korjenic, A., Korjenic, S., Wu, W. (2016). The influence of natural reinforcement fibers, gypsum and cement on compressive strength of earth bricks materials. Construction and Building Materials, 106, 179–188.
150.
Zhang, T., Dieckmann, E., Song, S., Xie, J., Yu, Z., Cheeseman, C. (2018). Properties of magnesium silicate hydrate (MSH) cement mortars containing chicken feather fibres. Construction and Building Materials, 180, 692–697.
151.
Zhang, Z. J., Liu, J. B., Li, B., Yu, G. X., Li, L. (2019). Experimental study on factors affecting the physical and mechanical properties of shell lime mortar. Construction and Building Materials, 228, 116726.
152.
Zhou, W., Yang, F., Zhu, R., Dai, G., Wang, W., Wang, W., Wang, Z. (2020). Mechanism analysis of pore structure and crystalline phase of thermal insulation bricks with high municipal sewage sludge content. Construction and Building Materials, 263, 120021.
153.
Zunino, F., Lopez, M. (2017). A methodology for assessing the chemical and physical potential of industrially sourced rice husk ash on strength development and early-age hydration of cement paste. Construction and Building Materials, 149, 869–881.
154.
Zwicky, D. (2020). Mechanical properties of organic-based lightweight concretes and their impact on economic and ecological performances. Construction and Building Materials, 245, 118413.

References

  1. Zuo, J.; Zhao, Z.Y. Green building research–current status and future agenda: A review. Renew. Sustain. Energy Rev. 2014, 30, 271–281. [Google Scholar] [CrossRef]
  2. Betts, M.; Robinson, G.; Burton, C.; Leonard, J.; Sharda, A.; Whittington, T. Global Construction 2030: A Global Forecast for the Construction Industry to 2030; Global Construction Perspectives and Oxford Economics: Bracknell, UK, 2015. [Google Scholar]
  3. Edenhofer, O. (Ed.) Climate Change 2014: Mitigation of Climate Change: Working Group III Contribution to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change; Cambridge University Press: New York, NY, USA, 2014. [Google Scholar]
  4. Dunning, R.J.; Lord, A. Preparing for the climate crisis: What role should land value capture play? Land Use Policy 2020, 99, 104867. [Google Scholar] [CrossRef]
  5. Naik, T.R. Sustainability of Concrete Construction. Pract. Period. Struct. Des. Constr. 2008, 13, 98–103. [Google Scholar] [CrossRef] [Green Version]
  6. Arıoğlu Akan, M.Ö.A.; Dhavale, D.G.; Sarkis, J. Greenhouse gas emissions in the construction industry: An analysis and evaluation of a concrete supply chain. J. Clean. Prod. 2017, 167, 1195–1207. [Google Scholar] [CrossRef]
  7. Zareei, S.A.; Ameri, F.; Bahrami, N.; Shoaei, P.; Moosaei, H.R.; Salemi, N. Performance of sustainable high strength concrete with basic oxygen steel-making (BOS) slag and nano-silica. J. Build. Eng. 2019, 25, 100791. [Google Scholar] [CrossRef]
  8. Habert, G.; Miller, S.A.; John, V.M.; Provis, J.L.; Favier, A.; Horvath, A.; Scrivener, K.L. Environmental impacts and decarbonization strategies in the cement and concrete industries. Nat. Rev. Earth Environ. 2020, 1, 559–573. [Google Scholar] [CrossRef]
  9. Jiménez, L.F.; Dominguez, J.A.; Vega-Azamar, R.E. Carbon footprint of recycled aggregate concrete. Adv. Civ. Eng. 2018, 2018. [Google Scholar] [CrossRef]
  10. ChileCubica. Dosificaciones de Hormigones y Morteros. 2006. Available online: www.chilecubica.com (accessed on 17 May 2021).
  11. Alengaram, U.J.; Muhit, B.A.A.; Jumaat, M.Z.b. Utilization of oil palm kernel shell as lightweight aggregate in concrete—A review. Constr. Build. Mater. 2013, 38, 161–172. [Google Scholar] [CrossRef]
  12. Miller, S.A.; Horvath, A.; Monteiro, P.J.M. Impacts of booming concrete production on water resources worldwide. Nat. Sustain. 2018, 1, 69–76. [Google Scholar] [CrossRef]
  13. Rashad, A.M. A preliminary study on the effect of fine aggregate replacement with metakaolin on strength and abrasion resistance of concrete. Constr. Build. Mater. 2013, 44, 487–495. [Google Scholar] [CrossRef]
  14. Tuan, B.L.A.; Hwang, C.L.; Lin, K.L.; Chen, Y.Y.; Young, M.P. Development of lightweight aggregate from sewage sludge and waste glass powder for concrete. Constr. Build. Mater. 2013, 47, 334–339. [Google Scholar] [CrossRef]
  15. Khan, K.; Ullah, M.F.; Shahzada, K.; Amin, M.N.; Bibi, T.; Wahab, N.; Aljaafari, A. Effective use of micro-silica extracted from rice husk ash for the production of high-performance and sustainable cement mortar. Constr. Build. Mater. 2020, 258, 119589. [Google Scholar] [CrossRef]
  16. Karbassi, A.; Jafari, H.; Yavari, A.; Hoveidi, H.; Kalal, H.S. Reduction of environmental pollution through optimization of energy use in cement industries. Int. J. Environ. Sci. Technol. 2010, 7, 127–134. [Google Scholar] [CrossRef] [Green Version]
  17. Zhang, S.; Ren, H.; Zhou, W.; Yu, Y.; Chen, C. Assessing air pollution abatement co-benefits of energy efficiency improvement in cement industry: A city level analysis. J. Clean. Prod. 2018, 185, 761–771. [Google Scholar] [CrossRef] [Green Version]
  18. Huntzinger, D.N.; Eatmon, T.D. A life-cycle assessment of Portland cement manufacturing: Comparing the traditional process with alternative technologies. J. Clean. Prod. 2009, 17, 668–675. [Google Scholar] [CrossRef]
  19. Huang, Y.; Xu, C.; Li, H.; Jiang, Z.; Gong, Z.; Yang, X.; Chen, Q. Utilization of the black tea powder as multifunctional admixture for the hemihydrate gypsum. J. Clean. Prod. 2019, 210, 231–237. [Google Scholar] [CrossRef]
  20. Kaza, S.; Yao, L.; Bhada-Tata, P.; Van Woerden, F. What a Waste 2.0: A Global Snapshot of Solid Waste Management to 2050; World Bank Publications: Washington, DC, USA, 2018. [Google Scholar]
  21. ONU para la Alimentación y la Agricultura. FAOSTAT Residuos Agrícolas. 2019. Available online: www.fao.org (accessed on 17 May 2021).
  22. Felix, M.; Elsamahy, E. The Efficiency of Using Different Outer Wall Construction Materials to Achieve Thermal Comfort in Various Climatic Zones. Energy Procedia 2017, 115, 321–331. [Google Scholar] [CrossRef]
  23. Abuelnuor, A.A.A.; Omara, A.A.M.; Saqr, K.M.; Elhag, I.H.I. Improving indoor thermal comfort by using phase change materials: A review. Int. J. Energy Res. 2018, 42, 2084–2103. [Google Scholar] [CrossRef]
  24. Falasca, S.; Ciancio, V.; Salata, F.; Golasi, I.; Rosso, F.; Curci, G. High albedo materials to counteract heat waves in cities: An assessment of meteorology, buildings energy needs and pedestrian thermal comfort. Build. Environ. 2019, 163, 106242. [Google Scholar] [CrossRef]
  25. Chen, K.; Wang, J.; Yu, B.; Wu, H.; Zhang, J. Critical evaluation of construction and demolition waste and associated environmental impacts: A scientometric analysis. J. Clean. Prod. 2021, 287, 125071. [Google Scholar] [CrossRef]
  26. Erkmen, J.; Yavuz, H.I.; Kavci, E.; Sari, M. A new environmentally friendly insulating material designed from natural materials. Constr. Build. Mater. 2020, 255, 119357. [Google Scholar] [CrossRef]
  27. Fu, Q.; Yan, L.; Ning, T.; Wang, B.; Kasal, B. Behavior of adhesively bonded engineered wood—Wood chip concrete composite decks: Experimental and analytical studies. Constr. Build. Mater. 2020, 247, 118578. [Google Scholar] [CrossRef]
  28. Gupta, S.; Kua, H.W. Application of rice husk biochar as filler in cenosphere modified mortar: Preparation, characterization and performance under elevated temperature. Constr. Build. Mater. 2020, 253, 119083. [Google Scholar] [CrossRef]
  29. Haik, R.; Bar-Nes, G.; Peled, A.; Meir, I.A. Alternative unfired binders as lime replacement in hemp concrete. Constr. Build. Mater. 2020, 241, 117981. [Google Scholar] [CrossRef]
  30. Kochova, K.; Gauvin, F.; Schollbach, K.; Brouwers, H.J.H. Using alternative waste coir fibres as a reinforcement in cement-fibre composites. Constr. Build. Mater. 2020, 231, 117121. [Google Scholar] [CrossRef]
  31. Martínez-García, C.; González-Fonteboa, B.; Carro-López, D.; Martínez-Abella, F. Carbonation evolution of lime putty coatings with mussel shell aggregate. Constr. Build. Mater. 2020, 264, 120165. [Google Scholar] [CrossRef]
  32. Muñoz, P.; Letelier, V.; Muñoz, L.; Bustamante, M.A. Adobe bricks reinforced with paper & pulp wastes improving thermal and mechanical properties. Constr. Build. Mater. 2020, 254, 119314. [Google Scholar] [CrossRef]
  33. Pérez-Villarejo, L.; Eliche-Quesada, D.; Martín-Pascual, J.; Martín-Morales, M.; Zamorano, M. Comparative study of the use of different biomass from olive grove in the manufacture of sustainable ceramic lightweight bricks. Constr. Build. Mater. 2020, 231, 117103. [Google Scholar] [CrossRef]
  34. Zwicky, D. Mechanical properties of organic-based lightweight concretes and their impact on economic and ecological performances. Constr. Build. Mater. 2020, 245, 118413. [Google Scholar] [CrossRef]
  35. Hamada, H.; Tayeh, B.; Yahaya, F.; Muthusamy, K.; Al-Attar, A. Effects of nano-palm oil fuel ash and nano-eggshell powder on concrete. Constr. Build. Mater. 2020, 261, 119790. [Google Scholar] [CrossRef]
  36. Hamood, A.; Khatib, J.M.; Williams, C. The effectiveness of using Raw Sewage Sludge (RSS) as a water replacement in cement mortar mixes containing Unprocessed Fly Ash (u-FA). Constr. Build. Mater. 2017, 147, 27–34. [Google Scholar] [CrossRef]
  37. Malaiskiene, J.; Vaiciene, M.; Giosuè, C.; Tittarelli, F. The impact of bitumen roofing production waste (BTw) on cement mortar properties. Constr. Build. Mater. 2020, 234, 117350. [Google Scholar] [CrossRef]
  38. Şeyma Seyrek, E.; Yalçin, E.; Yilmaz, M.; Vural Kök, B.; Arslanoğlu, H. Effect of activated carbon obtained from vinasse and marc on the rheological and mechanical characteristics of the bitumen binders and hot mix asphalts. Constr. Build. Mater. 2020, 240, 117921. [Google Scholar] [CrossRef]
  39. Alabdulkarem, A.; Ali, M.; Iannace, G.; Sadek, S.; Almuzaiqer, R. Thermal analysis, microstructure and acoustic characteristics of some hybrid natural insulating materials. Constr. Build. Mater. 2018, 187, 185–196. [Google Scholar] [CrossRef]
  40. Farooqi, M.U.; Ali, M. Effect of pre-treatment and content of wheat straw on energy absorption capability of concrete. Constr. Build. Mater. 2019, 224, 572–583. [Google Scholar] [CrossRef]
  41. Hallinger, P.; Kovačević, J. A Bibliometric Review of Research on Educational Administration: Science Mapping the Literature, 1960 to 2018. Rev. Educ. Res. 2019, 89, 335–369. [Google Scholar] [CrossRef]
  42. Minhas, M.R.; Potdar, V. Decision Support Systems in Construction: A Bibliometric Analysis. Buildings 2020, 10, 108. [Google Scholar] [CrossRef]
  43. van Eck, N.J.; Waltman, L. Visualizing Bibliometric Networks. In Measuring Scholarly Impact: Methods and Practice; Ding, Y., Rousseau, R., Wolfram, D., Eds.; Springer International Publishing: Cham, Switzerland, 2014; pp. 285–320. [Google Scholar] [CrossRef]
  44. White, H.D.; McCain, K.W. Visualizing a discipline: An author co-citation analysis of information science, 1972–1995. J. Am. Soc. Inf. Sci. 1998, 49, 327–355. [Google Scholar] [CrossRef]
  45. Wang, K.S.; Chiou, I.J.; Chen, C.H.; Wang, D. Lightweight properties and pore structure of foamed material made from sewage sludge ash. Constr. Build. Mater. 2005, 19, 627–633. [Google Scholar] [CrossRef]
  46. Yang, L.; Jing, M.; Lu, L.; Zhu, X.; Zhao, P.; Chen, M.; Li, L.; Liu, J. Effects of modified materials prepared from wastes on the performance of flue gas desulfurization gypsum-based composite wall materials. Constr. Build. Mater. 2020, 257, 119519. [Google Scholar] [CrossRef]
  47. Zhou, S.; Zhang, X.; Chen, X. Pozzolanic activity of feedlot biomass (cattle manure) ash. Constr. Build. Mater. 2012, 28, 493–498. [Google Scholar] [CrossRef]
  48. Zhang, W.; Min, H.; Gu, X.; Xi, Y.; Xing, Y. Mesoscale model for thermal conductivity of concrete. Constr. Build. Mater. 2015, 98, 8–16. [Google Scholar] [CrossRef]
  49. Stokke, D.D.; Wu, Q.; Han, G. Introduction to Wood and Natural Fiber Composites; John Wiley & Sons: Hoboken, NJ, USA, 2013. [Google Scholar]
  50. Tarasov, D.; Leitch, M.; Fatehi, P. Lignin–carbohydrate complexes: Properties, applications, analyses, and methods of extraction: A review. Biotechnol. Biofuels 2018, 11, 269. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Generation of agricultural residues (1980–2019): (a) worldwide; and (b) by continent (Source: Figures created by the authors based on [21]).
Figure 1. Generation of agricultural residues (1980–2019): (a) worldwide; and (b) by continent (Source: Figures created by the authors based on [21]).
Buildings 12 00597 g001
Figure 2. Bibliometric analysis based on the specific keyword “organic residues as building materials”: (a) VOSviewer map displaying the most prevalent keywords for n = 654 ; and (b) temporal evolution of studies related to organic residues in the construction industry between 1992 and 2020 for n = 1416 .
Figure 2. Bibliometric analysis based on the specific keyword “organic residues as building materials”: (a) VOSviewer map displaying the most prevalent keywords for n = 654 ; and (b) temporal evolution of studies related to organic residues in the construction industry between 1992 and 2020 for n = 1416 .
Buildings 12 00597 g002
Figure 3. Categorization of organic residues ( n = 158 ) by: (a) constructive use; (b) origin; and (c) type of agricultural residue.
Figure 3. Categorization of organic residues ( n = 158 ) by: (a) constructive use; (b) origin; and (c) type of agricultural residue.
Buildings 12 00597 g003
Figure 4. Compressive strength behaviour by (a) constructive use; (b) industry; and (c) agricultural industry. (n = 557). (*) there are significant differences ( p < 0.05 ).
Figure 4. Compressive strength behaviour by (a) constructive use; (b) industry; and (c) agricultural industry. (n = 557). (*) there are significant differences ( p < 0.05 ).
Buildings 12 00597 g004
Figure 5. Physicochemical and mechanical properties correlations. (a) Residue density-compressive strength; (b) Residue density-tensile strength; and, (c) Lignin content-flexural strength.
Figure 5. Physicochemical and mechanical properties correlations. (a) Residue density-compressive strength; (b) Residue density-tensile strength; and, (c) Lignin content-flexural strength.
Buildings 12 00597 g005
Figure 6. Physicochemical and thermal properties correlations. (a) Residue density-thermal conductivity; and (b) Extractives content-thermal conductivity.
Figure 6. Physicochemical and thermal properties correlations. (a) Residue density-thermal conductivity; and (b) Extractives content-thermal conductivity.
Buildings 12 00597 g006
Table 1. Constraining criteria of bibliometric analysis (confirmation phase).
Table 1. Constraining criteria of bibliometric analysis (confirmation phase).
Criteria TypeInclusionExclusion
Temporality1992–2020<1992
Residue typeOrganicInorganic
ApplicationBuilding materialEnergy or agriculture
Specific informationPhysicochemical, mechanical
and/or thermal properties
No information
Table 2. Processing information phases.
Table 2. Processing information phases.
ObjectiveFactorVariableAnalysis
Bibliometric analysisYear of publicationNumber of scientific publicationCo-ocurrence and state-of-the-art
Residue vs. useOrganic residue and constructive useCompressive strengthtwo-way ANOVA
Constructive applicationPhysicochemical properties (organic residue or compound)Mechanical properties Thermal propertiesMultiple correlation
Table 3. Analysis of correlation and regression between physicochemical, mechanical and thermal properties.
Table 3. Analysis of correlation and regression between physicochemical, mechanical and thermal properties.
VariableCorrelation
Analysis
Regression
Analysis
DependentIndependentp-ValuerCoefficientsEquation R 2
Residue density—
R D (g/cm 3 )
Compressive strength—
C S (kg/cm 2 )
<0.010.67 a = 29.440 ± 21.212
b = 186.185 ± 23.246
C S = a + b ( R D ) 0.40
Residue density—
R D (g/cm 3 )
Tensile strength—
T S (kg/cm 2 )
<0.010.72 a = 6.764 ± 2.315
b = 16.124 ± 1.725
T S = a + b ( R D ) 0.73
Lignin content—
L C (%/, w/w)
Flexural strength—
F S (kg/cm 2 )
<0.01−0.95 a = 120.130 ± 14.214
b = 3.295 ± 0.639
F S = a + b ( L C ) 0.84
Residue density—
R D (g/cm 3 )
Thermal conductivity—
T C (W/Km)
<0.010.54 a = 0.143 ± 0.072
b = 2.563 ± 0.261
c = 1.691 ± 0.199
T C = a + b ( R D ) + c ( R D ) 2 0.68
Extractives content—
E C (%, w/w)
Thermal conductivity—
T C (W/Km)
0.01 −0.72 a = 0.100 ± 0.030
b = 0.525 ± 0.227
c = 0.578 ± 0.146
T C = a + b c E C 0.89
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yanez, S.; Márquez, C.; Valenzuela, B.; Villamar-Ayala, C.A. A Bibliometric-Statistical Review of Organic Residues as Cementitious Building Materials. Buildings 2022, 12, 597. https://doi.org/10.3390/buildings12050597

AMA Style

Yanez S, Márquez C, Valenzuela B, Villamar-Ayala CA. A Bibliometric-Statistical Review of Organic Residues as Cementitious Building Materials. Buildings. 2022; 12(5):597. https://doi.org/10.3390/buildings12050597

Chicago/Turabian Style

Yanez, Sergio, Constanza Márquez, Benjamín Valenzuela, and Cristina Alejandra Villamar-Ayala. 2022. "A Bibliometric-Statistical Review of Organic Residues as Cementitious Building Materials" Buildings 12, no. 5: 597. https://doi.org/10.3390/buildings12050597

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