Next Article in Journal
Assessing Construction Constraint Factors on Project Performance in the Construction Industry
Next Article in Special Issue
Post-High-Temperature Exposure Repeated Impact Response of Steel-Fiber-Reinforced Concrete
Previous Article in Journal
Experiment Study on the Hysteretic Performance of a Novel Replaceable Beam-to-Column Joint with Energy-Dissipating Steel Hinge
Previous Article in Special Issue
Fresh and Hardened Properties of Self-Compacting Concrete Comprising a Copper Slag
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Improving the Performances of a Mortar for 3D Printing by Mineral Modifiers

1
Department of Building Materials Science, Products and Structures, Belgorod State Technological University Named after V.G. Shoukhov, 308012 Belgorod, Russia
2
Central Research and Design Institute of the Ministry of Construction, Housing and Utilities of the Russian Federation, 119331 Moscow, Russia
3
Polytechnic Institute, Far Eastern Federal University, 690922 Vladivostok, Russia
4
Peter the Great St. Petersburg Polytechnic University, 195251 St. Petersburg, Russia
5
Department of Civil Engineering, College of Engineering, Prince Sattam Bin Abdulaziz University, Alkharj 16273, Saudi Arabia
6
Department of Civil Engineering, Faculty of Engineering and IT, Amran University, Amran 9677, Yemen
7
Department of Civil and Environmental Engineering, Universiti Teknologi Petronas, Perak Seri Iskandar 32610, Malaysia
8
LECIV—Civil Engineering Laboratory, UENF—State University of the Northern Rio de Janeiro, Av. Alberto Lamego, 2000, Campos dos Goytacazes, Rio de Janeiro 28013-602, Brazil
*
Authors to whom correspondence should be addressed.
Buildings 2022, 12(8), 1181; https://doi.org/10.3390/buildings12081181
Submission received: 5 July 2022 / Revised: 27 July 2022 / Accepted: 4 August 2022 / Published: 7 August 2022

Abstract

:
Erection of buildings using 3D printing has great potential. However, its mass use for high-rise buildings is hampered by the lack of cement mortars with the required technical characteristics, the most important of which is high plastic strength (in the first minutes after pouring). The significance of the work (novelty) lies in the creation of a composite binder using a mineral modifier obtained by joint grinding up to 500 m2/kg of bentonite clay, chalk, and sand. A comprehensive study of the developed mortars was carried out from the standpoint of the necessary characteristics for volumetric concreting of high-rise thin-walled buildings. A composite binder for high-strength composites (compressive strength up to 70 MPa) has been obtained, which can provide effective mortars for 3D-additive high-rise construction technologies. The influence of the genetic characteristics of the modifier components on the properties of the composite binder has been established. The hydration process in this system of hardening concrete of the optimal composition proceeds more intensively due to the significantly larger specific surface of the mineral modifier components, which act as an active additive and activators of the crystallization of new growths. It has been proven that the features of mortars of high-strength fine-grained composites for 3D-additive technologies of high-rise buildings must meet special properties, such the rheotechnological index and the bearing capacity of the freshly formed layer (plastic strength or dimensional stability). Compared with a conventional mortar, the plastic strength of the developed one increases much faster (in 15 min, it is 762.2 kPa, in contrast to 133.0 kPa for the control composition). Thus, the strength remains sufficient for 3D printing of high-rise buildings and structures.

1. Introduction

In recent years, the construction of buildings using 3D-printing technology has received significant development [1,2,3]. Walls produced by a 3D printer have various shapes with well-defined tolerances, which is an important quality of this technology [4,5,6]. First, the printer casts the inner, outer, and middle layers of the wall to the specified level and then holds them together with a zigzag profile “in the form of stiffeners” (Figure 1). As a result, the hollow wall is filled with a lightweight concrete mixture; it has increased thermal insulation capacity and lightness [7,8,9]. One of the advantages of 3D-additive technologies is the speed of construction [10]. With 3D printing, it can construct a building of 250 m2 in 18–20 h [11] and with the use of high-strength concrete, which reduces the thickness of the structural elements [12,13]. Thus, for example, the thickness of a wall with B60 class concrete can be 8 cm; the same thickness can have a ceiling. This technology helps to save material and energy resources and reduces material consumption. In addition, according to the experience of Chinese builders, the possibility of using recyclable materials in this technology, such as crushed construction waste, has become obvious [14].
There is a wide range of types of mixes for 3D concreting, which is increasingly expanding every year. In particular, Markin et al. [15] developed a foam concrete obtained by digital technology. Yakovlev et al. [16] engaged in color printing issues using nanocomposites. Similar results are shown in the article of Muller et al. [17], describing the algorithms for the automatic printing of three-dimensional models of different facades. Despite a certain lack of knowledge, a search is underway for new materials for 3D printing. At the same time, some search is aimed at energy-efficient materials and technologies, such as geopolymers [18,19]. Klyuev et al. [20] developed 3D concretes using fiber, which can significantly improve their performances. Based on a historical approach to the science of concrete, Van Damme [10] calls for the study of radical changes in three key aspects of the use of concrete: reinforcement, binder content, and methods of implementation. More precisely, it is envisaged that, in parallel with the introduction of robotic manufacturing methods, digital technologies may become key to the implementation of several innovations, such as reinforcement without rebar using nonconvex granular materials; compression-optimized concrete structures using topology optimization, architectural geometry, and 3D-printed formwork; and truly digital concrete by combining massive data collection and deep learning. Khalil et al. [21] used calcium sulfoaluminate cements for setting the control of 3D-printing mortars. Yuan et al. [22] developed a feasible method for measuring the buildability of a fresh 3D-printing mortar. Kruger [23] studied in detail the thixotropy characteristics of concrete (saturated with nanoparticles) for 3D printing. Rahul et al. [24] studied some mechanical characterization of 3D-printable concrete.
The 3D-additive construction technologies can provide a significant shorter construction time for buildings and structures compared with traditional methods [25]. Experiments carried out on laser 3D printers show the possibility of obtaining the following positive aspects:
  • Reduction of time and labor costs [26,27];
  • Solution of complex design projects, regardless of shape and size [28,29];
  • High-performance characteristics of the resulting structures (the ability to operate in seismically and climatically difficult zones) [30].
Work is underway to address the following 3D-printing problems [31]:
  • Creation of a regulatory technological base [32];
  • Development of raw material supply for production [33];
  • Reducing the cost of designing and assembling 3D printer equipment [34].
The main issue that hinders the widespread use of 3D-additive technologies in the construction industry is the lack of a theoretical basis for creating a new generation of composites based on effective composite binders and the optimization of compositions with new types of energy-saving raw materials [29,35]. With the help of this technology, the construction of critical structures has not yet been carried out, and the existing traditional types of cement harden for a long time, complicating the production process [36,37]. It is also difficult to obtain an effective binder with clearly pronounced thixotropic properties when the viscosity decreases upon mechanical action and increases upon the termination of mechanical action [13,38].
Therefore, to solve these problems, it is necessary to develop theoretical foundations for creating a new generation of composites for high-rise construction using 3D-additive printing based on effective raw materials and high-quality composite binders [39]. The composite binder is obtained most often using Portland cement and a complex of various organomineral additives, with the use of mechanical activation, which ensures high strength and durability [40,41,42,43,44,45,46,47,48,49,50]. The effect of additives used in the composite binder that reduce water demand (retarders) and increase the rheology of the mixture (accelerators) on the hydration of cement minerals is especially significant when they are combined [51]. Heat release in the early stages depends on the content of the retarding or accelerating component in the additive [52,53].
Recently, organomineral additives have become widespread, in which chemical additives and dispersed mineral additives–fillers are combined into a single system [54]. In this case, a higher effect on the structure formation of concrete is achieved, and the best results of modifying the structure and properties of concrete are obtained [55,56]. Of great importance is the use of low-water demand binders based on Portland cement, for which some mineral additives and a superplasticizer can be added with grinding (regrinding) of cement, which increases the positive effect of its use [57]. The use of mechanoactivation and organomineral additives, such as the microsilica and superplasticizer system, makes it possible to increase the active ability of the composite binder and use it for 3D-printing purposes [58].
In the specific conditions of the construction of buildings using 3D-additive technologies, considering the purpose, special requirements arise for the binder and its structure formation [59]. This technology requires the use of special modifiers containing both mineral and organic components in their composition, ensuring the binder’s high efficiency [60,61,62,63,64,65]. The basis of 3D-additive technology in the construction of frame multistorey buildings is a specially developed material, such as fast-setting high-strength fine-grained concrete [24,66]. For its manufacture, a special composite binder is used, based on Portland cement [23].
Special requirements are imposed on the concrete mix and cement paste of a high-strength mortar for 3D-additive technology, forming a strong bond of fine aggregate, modifier, and other particles to a single conglomerate [22]. The rheological properties of the concrete (mortar) mix largely determine the future quality of concrete and products [21]. The most important rheological characteristic of a concrete mixture is workability or formability [67,68], that is, the ability of the mixture to spread and take a given shape while maintaining solidity and homogeneity [69,70]. Workability depends on the mortar mixture’s fluidity when filling the mold (i.e., the ability to deform without breaking the continuity) [71,72]. Based on the analysis of modern literature, gaps in the modern knowledge of 3D printing are indicated. It is necessary to continue improving the fresh and physical and mechanical properties, which will allow the construction of high-rise and thin-walled buildings. To do this, it is necessary to involve local substandard materials as much as possible, including natural and man-made waste.
This research article aims to determine the features of molding sand on composite binders for 3D printing of high-rise buildings. The composite binder was obtained by mixing Portland cement, a hyperplasticizer, and a mineral modifier. The mineral modifier, in turn, was obtained by grinding bentonite clay, chalk, and sand to a specific surface area of 500 m2/kg. The tasks for achieving the aim include the study of rheotechnological (fresh) behavior, the indicator of the bearing capacity of the freshly laid layer (plastic strength), and the physical and mechanical properties of fine-grained concrete composites (mortars). The significance of the paper (novelty) lies in the creation of a composite binder using a mineral modifier obtained by joint grinding up to 500 m2/kg mix of bentonite clay, chalk, and sand. A comprehensive study of the developed mortars was carried out from the standpoint of the necessary characteristics for 3D concreting of high-rise thin-walled buildings.

2. Materials and Methods

2.1. Characteristics of the Raw Materials and Experimental Program Design

In the construction industry, 3D-printing has not yet become widespread, so the selection of raw materials for this purpose is an important and urgent task at the stage of scientific development. Figure 2 shows a flowchart of the mix preparation.

2.1.1. Mineral Modifier

A number of studies have been carried out to study the fractional ratios of the base, such as Portland cement and the mineral modifier (MM), accelerating the setting time of the cement paste to obtain composite binders. A highly active mineral modifier for a composite binder was developed in this work. The effective ratio of the MM components was determined: bentonite clay powder, chalk, and sand (Table 1). Various degrees of grinding of the components were produced, amounting from 300 to 600 m2/kg. Grinding was carried out in a ball mill for 1 h. Then the powdered materials were mixed with Portland cement and a hyperplasticizer. The chemical composition of used raw materials is listed in Table 1.
It is known that in a hardening cement-containing system, an excess of calcium ions can cause type I corrosion (leaching). Therefore, the minimum amount of calcium carbonate was limited in the composition of the mineral modifier. Various proportions of the mineral modifier components were investigated (Table 2). It was established that the optimal ratio of the components for reducing the amount of calcium carbonate is the following: sand/bentonite clay flour/chalk = 2:2:1 (by wt.).
The components were joint milled to varying degrees of fineness. A hyperplasticizer and the mineral modifier with a specific surface area of 500 m2/kg at amounts of 1%, 3%, 5%, 7%, and 10% (by weight of cement) were added to Portland cement to obtain the composite binder. Mix ID MM1-500 was used in this case.

2.1.2. Composite Binder

Portland cement CEM I 42.5N (Belgorodsky cement, Belgorod, Russia), a highly active mineral modifier (MM), and a Melflux 5581 hyperplasticizer (HP) (BASF Construction Additives, Ludwigshafen, Germany) were used for the manufacture of the composite binder (CB) (Table 3).
In this part of the experimental program, various factors were investigated, that is, differences between the concrete mixtures from Table 3: (i) the degree of grinding of Portland cement (specific surface area of 300 and 500 m2/kg), (ii) the presence or absence of a hyperplasticizer, and (iii) the presence or absence of mineral modifiers.
The equal flowability of all mixtures was achieved by varying the content of water (162–193 L) and the hyperplasticizer (0–1.35 kg) with an equal content of ground and unground cement (450 kg) and sand (1500 kg).

2.1.3. Mortar Mix

A screening of quartzite sandstone (QS) from the Lebedinsky mining and processing plant (Russia) was chosen as a fine aggregate for high-strength concrete using 3D-printing technology (Table 4). Its interaction with the surface of the modifier particles and Portland cement clinker ensures the creation of a high homogeneity of the structure, a decrease in stresses in the interfacial transition zone, and an increase in adhesion to the cement paste [73,74,75,76,77]. Quartz sand used in the mortar mixes is shown for comparison.
BISEAL CEL (Drizoro, Madrid, Spain) was used as an air-entraining additive (AE). For comparison, in the control composition mix B, quartz sand (S) with a size modulus of 2.2 (Razdolnenskoe deposit, Razdolnoe, Russia) was used. Table 5 lists the proportions of concrete mixes.

2.2. Laboratory Equipment and Research Methods

Figure 3 shows a flowchart of the laboratory equipment. A MicroSizer 201 laser particle analyzer (Nauchnye pribory, Moscow, Russia) was used, which allows one to determine particles with sizes from 0.2 to 600 µm. The specific surface area of raw materials and binders was determined by gas permeability using a PSH-11 device.
An ARL X’TRA device (Thermo Fisher Scientific, Waltham, MA, USA) was used for X-ray diffraction analysis. The samples were ground to powders, which should have a grain size of less than 300 μm in order for the effects associated with radiation scattering to be reproducible. Samples were precompressed into tablets without any filler. The supply voltage from the AC mains was 220 V. The power consumption (with attachments) was no more than 10 kVA. The range of angular displacement of the detection unit was from −8 to +160 degrees. The readout resolution of the angle sensor was 0.00025 degrees.
The microstructure was studied using TESCAN MIRA 3 LMU (TESCAN, Brno, Czech) and LEO SUPRA 50VP (SUPRA, Ludwigshafen, Germany) high-resolution scanning electron microscopes. The methodology used to acquire the SEM image involved polishing the surface of the samples. Intact specimens cast specifically for microstructure assessment were used. Mortar samples after stripping and curing were sawn with a cutter into pieces 25 × 60 × 60 mm in size. The essence of the method for determining the setting time is that the pestles of various sections penetrate into the concrete mixture and fix the time required for their penetration to a depth of 25 ± 2 mm at two values of penetration resistance (4.0 and 24.0 MPa), corresponding start and end time.
The flowability (slump flow) of the cement composite (mortar) was determined by its lower diameter, which was formed as a result of the flow of the mortar. The diameter was determined by measuring the slump flow in two mutually perpendicular directions with an accuracy of 0.5 cm.
The slump flow was determined, rounded to the nearest centimeter, as the arithmetic mean of the results of two tests. The density of cement pastes and mortars was determined by dividing mass by volume. The extrudability of the mixes, characterized by the rheotechnological index (RTI), was determined by Formula (1) [23]:
RTI = 80 − h + 11
where 80 is the distance of the maximum possible immersion of the working body of the viscometer in mm, h the immersion value in mm, and 11 the idle in mm.
The indicator of the bearing capacity of the freshly formed layer (plastic strength, PS) was determined by Formula (2) in kPa [26]:
P S = P S
where P is the pressure exerted on the mix at which it begins to be squeezed out of the mold in kg, and S is the area of the stamp used for compression in m2.
The average density of mortar specimens was determined as follows: three cube specimens with an edge of 70 mm, identical in hardening conditions, were weighed with an accuracy of 0.5%, and the dimensions of the faces were measured with a caliper (the average of three measurements, such two parallel faces on one side and a height in middle on this side). The average density was calculated by dividing a mass of a specimen by its volume. In compressive and flexural strength tests, a specimen (cubes with an edge of 100 mm and prisms with a size of 40 × 40 × 160 mm, respectively) was loaded continuously at a constant rate of load increase of 0.6 ± 0.2 MPa/s until its destruction. A generalized list of all compositions and tests is given in Table 6.

3. Results and Discussion

3.1. Obtaining the Mineral Modifiers

Table 7 lists the effect of the ratio of the components of the mineral modifier on the strength characteristics of cement composite.
For all compositions, an increase in compressive strength was noted in comparison with the control composition. At the same time, the maximum effect (almost 50%) was achieved when using the mineral modifier with the ratio of the components, such sand/bentonite clay powder: chalk as 2:2:1 (by mass) at a specific surface area of 500 m2/kg.

3.2. Development of the Composite Binder

During the development of the composite binder, the influence of the genesis of the MM components and the fineness of the grinding of Portland cement on the characteristics of the cement paste was studied (Table 8).
The results obtained allowed us to establish the following: Specimens made on cement of factory grinding fineness (300 m2/kg), which included the mineral modifier, were characterized by an increase in compressive strength by 39% and flexural strength by 38%. This was due to the most complete degree of hydration in comparison with the control specimen (specimen CB-1), as can be seen in Figure 4.
As a result, the effective content of the mineral modifier components in the composite binder was determined, as well as their effect on its hardening and the properties of the mortar. The fullness of hydration of the samples of the obtained composite binder was investigated using X-ray diffraction analysis according to the change in the intensity of reflections of alite and belite in an XRD pattern (d = 2.76 and 2.78 Å), which is most indicative for such systems.
The composition on Portland cement subjected to final grinding to a specific surface area of 500 m2/kg (specimen CB-7) with the mineral modifier had strength indicators about 40%–42% higher than those for control specimens with a large number of peaks related to new growths. This fact is explained by the higher specific surface area of the cement and, accordingly, the higher activity in comparison with the factory one.
Cement pastes based on a composite binder with the addition of the mineral modifier and the hyperplasticizer, both in the case of using cement of factory fineness of grinding and after extra grinding, showed a significant increase in strength indicators (specimens CB-4 and CB-8). The overall increase in compressive and flexural strength was up to 61%. The completeness of hydration was maximal. It should be noted that the mineral modifier acts as a catalyst for the formation of secondary calcium silicate hydrates, simultaneously binding Ca(OH)2. This is clearly seen from the comparison of the control specimen CB-1 and the specimens CB-3, CB-7, and CB-8.
The carbonate phase, which is the center of crystallization, provides tight contacts with the cement matrix due to epitoxic bonds between crystalline hydrates and the mineral modifier. Based on these prerequisites, the CaCO3 particles in the composition of the composite binder serve as a chemical activator of the growth of the interaction between CB and the aggregates. At the same time, no obvious peaks of ettringite were found in the developed compositions, which indicates a targeted control of the structure formation of cement composites with a predominance of low-basic calcium silicate hydrates.
Therefore, using the developed composite binder with the mineral modifier makes it possible to obtain a cement paste with a high-density structure, characterized by an intensive process of phase formation, hardening, and strength gain. As a result, the products have high construction and technical properties, low creep, and high durability. In a composite binder with the mineral modifier at an amount of 3% by weight of cement, the established ratio of components is characterized by high strength indicators compared with the unmodified binder. The increase in strength is due to the reduced water demand of the concrete mixture and the completeness of filling the pore space with hydrated new growths in the resulting cement paste (Table 9). Hence, due to the intensification of the interaction processes between clinker minerals and the mineral modifier and water, a rapid formation of crystalline hydrates occurs, which ensures the production of a material with high construction and technical properties. Cement paste with the addition of 3% MM is characterized by the greatest increase in compressive strength (46.2%).
The fineness and method of grinding the MM components of the composite binders were studied by preliminary grinding to various fineness of grinding, such as 400, 500, and 600 m2/kg. The flowability of the mix with the MM content in the range of 1% to 10% was constant and amounted to 160 mm of slump flow. Therefore, for further research, the previously established composition of the composite binder was adopted, containing MM at an amount of 3%. From this mix, test cubes were prepared and tested within the 28 days age (Table 10).
The growth of the strength indicators of the composite binder occurs when the specific surface area of the components is 500 m2/kg. An increase in the fineness of grinding above this value does not lead to a significant increase in strength. Based on this, for further studies, the specific surface of the MM was taken at the level of 500 m2/kg. The water demand of the composite binder was studied on a mix with different contents of the mineral modifier (Figure 5). The results showed that an MM content of 3%–5% makes it possible to reduce the water/cement ratio from 0.4 to 0.3 without reducing the flowability of the mix.
Figure 5 shows that the water demand of the developed composite binder rises with an increase in the fineness of grinding and with an increase in the amount of the mineral modifier in the mix. A unique decrease in water demand was noted at the content of the mineral modifier in the amount of 5% and its specific surface area of 500 m2/kg. This is due to the ideal granulometry with the required amount of rounded particles.
Tests of the control specimens for compressive strength confirmed the optimal ratio of the components of the mix (Table 11). A significant increase in the strength indicators of samples at the age of 14 days (up to 85%) was obtained.
The highest compressive strength value, 67.5 MPa, is observed in specimens of the composite binder with 3% addition of the mineral modifier. The acceleration of the structure formation process is noticeable in samples with Portland cement, a hyperplasticizer, and a mineral modifier, the components of which are crushed to 500 m2/kg. This is due to the combined action of these components in the phase formation processes.
The setting time of the composite binder with the mineral modifier differs from conventional cement paste. The beginning of the setting of the composite binder is reduced, as well as the end of this process, but the time period between them is somewhat extended. This is due to the activating effect of Portland cement particles, due to their modification when the components are ground with a mineral modifier, which accelerates the hydration reaction of clinker minerals (Figure 6).
For 3D printing, this will have a positive effect, since it will help to eliminate cracking during hydration. The results show that the start of setting is 1 h and 56 min and the end of setting is 2 h and 45 min. The end of the setting of the composite binder is associated with the addition of the mineral modifier, which reduces the setting time of the binder and improves its performances. SEM images of the composite binder confirm the consolidation of cement paste (Figure 7). This fact, of course, affects the strength indicators in the direction of their increase.

4. Performances of the Mortar Mix

Studies of the rheological properties of the fine-grained high-strength mortar have shown its full suitability for the process of forming a supporting structure using 3D-additive technology and high strength after hardening. Thus, the behavior of the high-strength fine-grained concrete ensures its efficient casting during 3D printing and obtaining of a stable structure of a wall structure (Figure 8).
It has been established (Figure 9) that the plastic strength (dimensional stability) of fine-grained concrete in mix B grows much faster than that of control fine-grained concrete: after 15 min, it is 762.2 KPa for the developed composition, and 133.0 kPa for the control composition.
Special requirements are also imposed on the cement paste of a high-strength fine-grained mortar for 3D-additive technology, forming a strong bond of fine aggregate, a modifier, and other particles into a single conglomerate. The strength properties of structural new growths in cement paste, the strength bonds of the interfacial transition zone of fine particles, and the size and nature of their surface determine the kinetics of phase formation processes in the hardening mortar system (Figure 10).
According to Figure 10, there are a smaller number of visible pores for the developed mix A than the control mix B. This is important because the increased moisture in the pores causes an increase in volume in the hardening cement composite, and in air-dry conditions, it shrinks. These phenomena largely depend on the composition of the concrete mixture: with an increase in the amount of the binder and modifiers, shrinkage and swelling of the cement stone are minimized. Therefore, when designing a mortar for 3D-additive technologies, it is necessary to consider the shrinkage deformations of the cement paste since they are undesirable due to its importance, especially for the formless casting method.
Testing the mortar specimens confirmed the features of the strength characteristics of the experimental structure erected using 3D printing and the optimal ratio of the composition components and the relationship between the rheological and strength characteristics of a mortar (Table 12). It has been established that the use of the composite binder and aggregates with a defective structure from the screening out of crushing of quartzite sandstone of the greenschist degree of metamorphism intensifies the process of hardening of concrete and mortar mixes, densifies the structure and strengthens products based on cement compositions, reduces creep and deformation characteristics, and increases strength. The increased surface area of the filler also determines its high structure-forming role during the interaction. The implementation of these conditions in the developed mortar leads to an intensification of the structure formation process and an acceleration of the strength gain of the mortar.
According to Table 12, the highest value of the compressive strength at the age of 28 days (67.5 MPa) is observed in the CB specimens with 3% addition of the mineral modifier. The acceleration of the structure formation process is noticeable in samples with Portland cement, the superplasticizer, and the modifier, the components of which are crushed to 500 m2/kg. This acceleration is due to the combined action of the components of the mineral modifier in the phase formation processes.
The 3D printing of the mix can be effectively used due to the high adhesion between the particles and the layers of the laid mixture; the time of hardening and concrete curing is maximally reduced so that the previous concrete layer can withstand the weight of subsequent layers without deforming. With 3D printing, the concrete layers, starting from the first, immediately begin to be subjected to stress, so the binder must be fast-hardening. In order to ensure the dimensional stability of the structure, the yield strength of concrete must be minimal. However, at the same time, the moldable mixt must be flowable enough to carry out extrusion. The 3D-printing augmentation of a building structure can be effective when using a mixture of sufficient fluidity and fast stabilization of properties after application. It was found that the change in the yield point in the first 40 min is linear.
Tests of the strength characteristics of fine-grained high-strength mortars for 3D-additive technologies using a special method showed that cube specimens with an edge of 100 mm cut from the printed wall of the experimental structure have a lower value of the compressive strength than specimens formed in 10 × 10 × 10 cm molds amounting to 15%–20%. At the same time, differences in strength indicators are also manifested depending on the direction of cutting specimens from the wall (Table 13).
The decrease in the strength indicators of the wall section at the contact points of the mixture layers compared with the cube specimen strength is established. At the same time, there is a difference in the strength characteristics of the layers of the printed wall themselves due to the direction of the sample cutting. The test direction (breaking load) along the forming direction of the 3D printer shows a slightly higher strength than that perpendicular to the forming direction. The shape of the layer can explain this phenomenon after solidification of the mixture, which is an ellipse in the cross section. The load applied to the ellipse’s minor axis will be greater than that applied to the major axis.

5. Conclusions

The 3D-additive technologies for high-rise concrete structures are recently becoming very imperative research topics in many construction industries worldwide, the promising direction in modern building technology. Additionally, a high plastic strength is deemed to be the most important property that should be taken into account by researchers and developers during the design of a durable mortar for 3D printing.
Based on the analysis of recent literature, gaps in the knowledge on 3D printing are indicated. It is necessary to continue improving the fresh and physical and mechanical properties, which will allow the construction of high-rise and thin-walled buildings. To do this, it is necessary to involve local substandard materials as much as possible, including natural and man-made waste.
For this reason, this study aimed to develop a high-strength mortar for 3D-printing building with increased plastic strength using a highly active mineral modifier, as well as to identify the features of the molding mixture on composite binders for 3D printing of high-rise buildings, including rheotechnological behavior, an indicator of the bearing capacity of a freshly formed layer (plastic strength) and the physical and mechanical properties of the mortar. Based on the results, the following conclusions were made:
  • A highly active mineral modifier has been developed, consisting of bentonite clay, chalk, and sand with a specific surface area of 500 m2/kg.
  • A multicomponent composite binder for high-strength mortars has been obtained, which, according to its characteristics, can provide effective molding sands for 3D-additive technologies of high-rise construction.
  • The effect of the genetic characteristics of the modifier components on the properties of the composite binder has been established. A composite binder with a compressive strength of up to 70 MPa has been obtained. At the same time, the process of hydration in this hardening system of concrete of optimal composition is more intensive due to the significantly larger specific surface area of the components of the mineral modifier, which act as an active mineral additive and activators of crystallization of new growths.
  • The features of molding sands of high-strength fine-grained composites for 3D-additive technologies of high-rise structures, which must meet special properties, such as the rheotechnological index (RTI) and the indicator of the bearing capacity of the freshly formed layer (plastic strength or dimensional stability), have been established. In comparison with ordinary fine-grained concrete, the plastic strength (dimensional stability) of the developed fine-grained concrete grows much faster (within 15 min, it is 762.2 KPa and 133.0 KPa for the control composition).
  • A high-strength mortar, obtained by 3D-additive technology, shows a decrease in strength compared with specimens formed in standard forms (up to 15%). At the same time, the strength remains sufficient for 3D-printing of high-rise buildings and structures.
The significance of the paper (novelty) lies in the creation of a composite binder using a mineral modifier obtained by joint grinding up to 500 m2/kg of bentonite clay, chalk, and sand. A comprehensive study of the developed mortars was carried out from the standpoint of the necessary characteristics for 3D concreting of high-rise thin-walled buildings.
The results obtained agree well with the papers cited in the first section. The values of fresh and physical–mechanical characteristics are acceptable for the construction of high-rise and thin-walled buildings.
The prospects for further research is considered in the direction of the possibility of obtaining modified binders to create various cementitious products, expanding the range of local raw materials used to obtain modified binders, and to continue research in the direction of studying the features of the course of structure formation processes using various modifying additives.

Author Contributions

Conceptualization, V.L., A.T., R.F., M.A. (Mugahed Amran), M.A. (Mujahid Ali) and A.R.G.d.A.; methodology, V.L., A.T., R.F., M.A. (Mugahed Amran), M.A. (Mujahid Ali) and A.R.G.d.A.; software, V.L., A.T., R.F., M.A. (Mugahed Amran), M.A. (Mujahid Ali) and A.R.G.d.A.; validation, V.L., A.T., R.F., M.A. (Mugahed Amran), M.A. (Mujahid Ali) and A.R.G.d.A.; formal analysis, V.L., A.T., R.F., M.A. (Mugahed Amran), M.A. (Mujahid Ali) and A.R.G.d.A.; investigation, V.L., A.T., R.F., M.A. (Mugahed Amran), M.A. (Mujahid Ali) and A.R.G.d.A.; resources, V.L., A.T., R.F., M.A. (Mugahed Amran), M.A. (Mujahid Ali) and A.R.G.d.A.; data curation, V.L., A.T., R.F., M.A. (Mugahed Amran), M.A. (Mujahid Ali) and A.R.G.d.A.; writing—original draft preparation, V.L., A.T., R.F., M.A. (Mugahed Amran), M.A. (Mujahid Ali) and A.R.G.d.A.; writing—review and editing, V.L., A.T., R.F., M.A. (Mugahed Amran), M.A. (Mujahid Ali) and A.R.G.d.A.; visualization, V.L., A.T., R.F., M.A. (Mugahed Amran), M.A. (Mujahid Ali) and A.R.G.d.A.; supervision, V.L., A.T., R.F., M.A. (Mugahed Amran), M.A. (Mujahid Ali) and A.R.G.d.A.; project administration, V.L., A.T., R.F., M.A. (Mugahed Amran), M.A. (Mujahid Ali) and A.R.G.d.A.; funding acquisition, V.L., A.T., R.F., M.A. (Mugahed Amran), M.A. (Mujahid Ali) and A.R.G.d.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by RSF Grant No. 22-19-20115, https://rscf.ru/project/22-19-20115/ (accessed on 23 April 2022) and the Government of the Belgorod Region, Agreement No. 3 of 24 March 2022.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors gratefully acknowledge the financial support given by the Deanship of Scientific Research at Prince Sattam bin Abdulaziz University, Al-Kharj, Saudi Arabia, and the cooperation of the Department of Civil Engineering, Faculty of Engineering and IT, Amran University, Yemen, for this research.

Conflicts of Interest

The authors declare no conflict of interest. The research does not involve human participants and/or animals.

References

  1. Amran, M.; Murali, G.; Fediuk, R.; Vatin, N.; Vasilev, Y.; Abdelgader, H. Palm oil fuel ash-based eco-efficient concrete: A critical review of the short-term properties. Materials 2021, 14, 332. [Google Scholar] [CrossRef] [PubMed]
  2. Muthukrishnan, S.; Ramakrishnan, S.; Sanjayan, J. Effect of alkali reactions on the rheology of one-part 3D printable geopolymer concrete. Cem. Concr. Compos. 2021, 116, 103899. [Google Scholar] [CrossRef]
  3. Muthukrishnan, S.; Ramakrishnan, S.; Sanjayan, J. Effect of microwave heating on interlayer bonding and buildability of geopolymer 3D concrete printing. Constr. Build. Mater. 2020, 265, 120786. [Google Scholar] [CrossRef]
  4. Lesovik, V.; Fediuk, R.; Amran, M.; Alaskhanov, A.; Volodchenko, A.; Murali, G.; Uvarov, V.; Elistratkin, M. 3d-printed mortars with combined steel and polypropylene fibers. Fibers 2021, 9, 79. [Google Scholar] [CrossRef]
  5. Klyuev, S.; Klyuev, A.; Fediuk, R.; Ageeva, M.; Fomina, E.; Amran, M.; Murali, G. Fresh and mechanical properties of low-cement mortars for 3D printing. Constr. Build. Mater. 2022, 338, 127644. [Google Scholar] [CrossRef]
  6. Amran, M.; Abdelgader, H.S.; Onaizi, A.M.; Fediuk, R.; Ozbakkaloglu, T.; Rashid, R.S.; Murali, G. 3D-printable alkali-activated concretes for building applications: A critical review. Constr. Build. Mater. 2022, 319, 126126. [Google Scholar] [CrossRef]
  7. Lukuttsova, N.; Pykin, A.; Kleymenicheva, Y.; Suglobov, A.; Efremochkin, R. Nano-additives for composite building materials and their environmental safety. Int. J. Appl. Eng. Res. 2016, 11, 7561–7565. [Google Scholar]
  8. Yoo, D.Y.; Banthia, N. Mechanical properties of ultra-high-performance fiber-reinforced concrete: A review. Cem. Concr. Compos. 2016, 73, 267–280. [Google Scholar] [CrossRef]
  9. Li, Z.; Ding, Z.; Zhang, Y. Development of sustainable cementitious materials. In Proceedings of the International Workshop on Sustainable Development and Concrete Technology, Beijing, China, 20–21 May 2004. [Google Scholar]
  10. Van Damme, H. Concrete material science: Past, present, and future innovations. Cem. Concr. Res. 2018, 112, 5–24. [Google Scholar] [CrossRef]
  11. Inozemtcev, A.; Korolev, E.; Qui, D.T. Study of mineral additives for cement materials for 3D-printing in construction. In Proceedings of the IOP Conference Series: Materials Science and Engineering, Moscow, Russia, 25–27 April 2018. [Google Scholar]
  12. Suntharalingam, T.; Gatheeshgar, P.; Upasiri, I.; Poologanathan, K.; Nagaratnam, B.; Corradi, M.; Nuwanthika, D. Fire performance of innovative 3D printed concrete composite wall panels—A Numerical Study. Case Stud. Constr. Mater. 2021, 15, e00586. [Google Scholar] [CrossRef]
  13. Fediuk, R.S.; Smoliakov, A.K.; Timokhin, R.A.; Batarshin, V.O.; Yevdokimova, Y.G. Using thermal power plants waste for building materials. In Proceedings of the IOP Conference Series: Earth and Environmental Science, Saint-Petersburg, Russia, 23–24 March 2017; Volume 87. [Google Scholar]
  14. Xiao, J.; Han, N.; Zhang, L.; Zou, S. Mechanical and microstructural evolution of 3D printed concrete with polyethylene fiber and recycled sand at elevated temperatures. Constr. Build. Mater. 2021, 293, 123524. [Google Scholar] [CrossRef]
  15. Markin, V.; Nerella, V.N.; Schröfl, C.; Guseynova, G.; Mechtcherine, V. Material design and performance evaluation of foam concrete for digital fabrication. Materials 2019, 12, 2433. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Yakovlev, A.V.; Milichko, V.A.; Vinogradov, V.V.; Vinogradov, A.V. Inkjet Color Printing by Interference Nanostructures. ACS Nano 2016, 10, 3078–3086. [Google Scholar] [CrossRef] [PubMed]
  17. Müller, P.; Zeng, G.; Wonka, P.; Van Gool, L. Image-based procedural modeling of facades. ACM Trans. Graph. 2007, 26, 85. [Google Scholar] [CrossRef]
  18. Zhao, J.; Tong, L.; Li, B.; Chen, T.; Wang, C.; Yang, G.; Zheng, Y. Eco-friendly geopolymer materials: A review of performance improvement, potential application and sustainability assessment. J. Clean. Prod. 2021, 307, 127085. [Google Scholar] [CrossRef]
  19. Singh, N.B. Fly ash-based geopolymer binder: A future construction material. Minerals 2018, 8, 299. [Google Scholar] [CrossRef] [Green Version]
  20. Klyuev, S.V.; Klyuev, A.V.; Shorstova, E.S. Fiber concrete for 3-D additive technologies. Mater. Sci. Forum 2019, 974, 367–372. [Google Scholar] [CrossRef]
  21. Khalil, N.; Aouad, G.; El Cheikh, K.; Rémond, S. Use of calcium sulfoaluminate cements for setting control of 3D-printing mortars. Constr. Build. Mater. 2017, 157, 382–391. [Google Scholar] [CrossRef]
  22. Yuan, Q.; Li, Z.; Zhou, D.; Huang, T.; Huang, H.; Jiao, D.; Shi, C. A feasible method for measuring the buildability of fresh 3D printing mortar. Constr. Build. Mater. 2019, 27, 116600. [Google Scholar] [CrossRef]
  23. Kruger, J.; Zeranka, S.; Zijl, G. An ab initio approach for thixotropy characterisation of (nanoparticle-infused) 3D printable concrete. Constr. Build. Mater. 2019, 224, 372–386. [Google Scholar] [CrossRef]
  24. Rahul, A.V.; Santhanam, M.; Meena, H.; Ghani, Z. Mechanical characterization of 3D printable concrete. Constr. Build. Mater. 2019, 227, 116710. [Google Scholar] [CrossRef]
  25. Muthukrishnan, S.; Kua, H.W.; Yu, L.N.; Chung, J.K.H. Fresh Properties of Cementitious Materials Containing Rice Husk Ash for Construction 3D Printing. J. Mater. Civ. Eng. 2020, 32, 4020195. [Google Scholar] [CrossRef]
  26. Berde, N.N.; Sanap, S.B.; Thorat, S.G. Study of impact and fatigue on 3D printed composites. Mater. Today Proc. 2021, 47, 2376–2378. [Google Scholar] [CrossRef]
  27. Jabir, H.A.; Abid, S.R.; Murali, G.; Ali, S.H.; Klyuev, S.; Fediuk, R.; Vatin, N.; Promakhov, V.; Vasilev, Y. Experimental Tests and Reliability Analysis of the Cracking Impact Resistance of UHPFRC. Fibers 2020, 8, 74. [Google Scholar] [CrossRef]
  28. Chen, K.; Yu, L.; Cui, Y.; Jia, M.; Pan, K. Optimization of printing parameters of 3D-printed continuous glass fiber reinforced polylactic acid composites. Thin-Walled Struct. 2021, 164, 107717. [Google Scholar] [CrossRef]
  29. Fediuk, R.; Yushin, A. Composite binders for concrete with reduced permeability. In Proceedings of the IOP Conference Series: Materials Science and Engineering, Tomsk, Russia, 9–11 November 2015; Volume 116. [Google Scholar]
  30. Schuldt, S.J.; Jagoda, J.A.; Hoisington, J.A.; Delorit, J.D. A systematic review and analysis of the viability of 3D-printed construction in remote environments. Autom. Constr. 2021, 125, 103642. [Google Scholar] [CrossRef]
  31. Kruger, J.; du Plessis, A.; van Zijl, G. An investigation into the porosity of extrusion-based 3D printed concrete. Addit. Manuf. 2021, 37, 101740. [Google Scholar] [CrossRef]
  32. Mohan, M.K.; Rahul, A.V.; Van Tittelboom, K.; De Schutter, G. Rheological and pumping behaviour of 3D printable cementitious materials with varying aggregate content. Cem. Concr. Res. 2021, 139, 106258. [Google Scholar] [CrossRef]
  33. Chen, Y.; Çopuroğlu, O.; Rodriguez, C.R.; de Mendonca Filho, F.F.; Schlangen, E. Characterization of air-void systems in 3D printed cementitious materials using optical image scanning and X-ray computed tomography. Mater. Charact. 2021, 173, 110848. [Google Scholar] [CrossRef]
  34. Sun, J.; Xi, J.; Li, Z.; Feng, X. Experimental study on the thermal performance of a 3D printed concrete prototype building. Energy Build. 2021, 241, 110965. [Google Scholar] [CrossRef]
  35. Wang, L.; Liu, Y.; Yang, Y.; Li, Y.; Bai, M. Bonding performance of 3D printing concrete with self-locking interfaces exposed to compression–shear and compression–splitting stresses. Addit. Manuf. 2021, 42, 101992. [Google Scholar] [CrossRef]
  36. Wu, Y.-C.; Yang, Q.; Kong, X.; Zhi, P.; Xiao, J. Uncertainty quantification for the representative volume element of geometrically monoclinic 3D printed concrete. Int. J. Solids Struct. 2011, 226–227, 111102. [Google Scholar] [CrossRef]
  37. Volodchenko, A.A.; Lesovik, V.S.; Cherepanova, I.A.; Volodchenko, A.N.; Zagorodnjuk, L.H.; Elistratkin, M.Y. Peculiarities of non-autoclaved lime wall materials production using clays. In Proceedings of the IOP Conference Series: Materials Science and Engineering, Tomsk, Russia, 4–6 December 2017. [Google Scholar]
  38. Ghourchian, S.; Butler, M.; Krüger, M.; Mechtcherine, V. Modelling the development of capillary pressure in freshly 3D-printed concrete elements. Cem. Concr. Res. 2021, 145, 106457. [Google Scholar] [CrossRef]
  39. Murcia, D.H.; Genedy, M.; Taha, M.M.R. Examining the significance of infill printing pattern on the anisotropy of 3D printed concrete. Constr. Build. Mater. 2020, 262, 120559. [Google Scholar] [CrossRef]
  40. Comminal, R.; da Silva, V.R.L.; Andersen, T.J.; Stang, H.; Spangenberg, J. Modelling of 3D concrete printing based on computational fluid dynamics. Cem. Concr. Res. 2020, 138, 106256. [Google Scholar] [CrossRef]
  41. De Azevedo, A.R.G.; Cruz, A.S.A.; Marvila, M.T.; de Oliveira, L.B.; Monteiro, S.N.; Vieira, C.M.F.; Fediuk, R.; Timokhin, R.; Vatin, N.; Daironas, M. Natural fibers as an alternative to synthetic fibers in reinforcement of geopolymer matrices: A comparative review. Polymers 2021, 13, 2493. [Google Scholar] [CrossRef]
  42. Amran, M.; Fediuk, R.; Murali, G.; Avudaiappan, S.; Ozbakkaloglu, T.; Vatin, N.; Karelina, M.; Klyuev, S.; Gholampour, A. Fly ash-based eco-efficient concretes: A comprehensive review of the short-term properties. Materials 2021, 14, 4264. [Google Scholar] [CrossRef]
  43. Onaizi, A.M.; Huseien, G.F.; Lim, N.H.A.S.; Amran, M.; Samadi, M. Effect of nanomaterials inclusion on sustainability of cement-based concretes: A comprehensive review. Constr. Build. Mater. 2021, 306, 124850. [Google Scholar] [CrossRef]
  44. Amran, M.; Al-Fakih, A.; Chu, S.H.; Fediuk, R.; Haruna, S.; Azevedo, A.; Vatin, N. Long-term durability properties of geopolymer concrete: An in-depth review. Case Stud. Constr. Mater. 2021, 15, e00661. [Google Scholar] [CrossRef]
  45. Amran, M.; Fediuk, R.; Abdelgader, H.S.; Murali, G.; Ozbakkaloglu, T.; Lee, Y.H.; Lee, Y.Y. Fiber-reinforced alkali-activated concrete: A review. J. Build. Eng. 2021, 45, 103638. [Google Scholar] [CrossRef]
  46. Lesovik, V.; Volodchenko, A.; Fediuk, R.; Mugahed Amran, Y.H.; Timokhin, R. Enhancing performances of clay masonry materials based on nanosize mine waste. Constr. Build. Mater. 2021, 269, 121333. [Google Scholar] [CrossRef]
  47. Haruna, S.; Mohammed, B.S.; Wahab, M.M.A.; Kankia, M.U.; Amran, M.; Gora, A.M. Long-Term Strength Development of Fly Ash-Based One-Part Alkali-Activated Binders. Materials 2021, 14, 4160. [Google Scholar] [CrossRef] [PubMed]
  48. Lesovik, V.; Chernysheva, N.; Fediuk, R.; Amran, M.; Murali, G.; de Azevedo, A.R.G. Optimization of fresh properties and durability of the green gypsum-cement paste. Constr. Build. Mater. 2021, 287, 123035. [Google Scholar] [CrossRef]
  49. Lesovik, V.; Volodchenko, A.; Fediuk, R.; Mugahed Amran, Y.H. Improving the Hardened Properties of Nonautoclaved Silicate Materials Using Nanodispersed Mine Waste. J. Mater. Civ. Eng. 2021, 33, 4021214. [Google Scholar] [CrossRef]
  50. Amran, M.; Huang, S.S.; Debbarma, S.; Rashid, R.S.M. Fire resistance of geopolymer concrete: A critical review. Constr. Build. Mater. 2022, 324, 126722. [Google Scholar] [CrossRef]
  51. Geng, Z.; She, W.; Zuo, W.; Lyu, K.; Pan, H.; Zhang, Y.; Miao, C. Layer-interface properties in 3D printed concrete: Dual hierarchical structure and micromechanical characterization. Cem. Concr. Res. 2020, 138, 106220. [Google Scholar] [CrossRef]
  52. Li, Z.; Li, W.; Ma, J.; Sanjayan, J.; Feng, D. Strength and ductility enhancement of 3D printing structure reinforced by embedding continuous micro-cables. Constr. Build. Mater. 2020, 264, 120296. [Google Scholar] [CrossRef]
  53. Abid, S.R.; Murali, G.; Amran, M.; Vatin, N.; Fediuk, R.; Karelina, M. Evaluation of mode II fracture toughness of hybrid fibrous geopolymer composites. Materials 2021, 14, 349. [Google Scholar] [CrossRef]
  54. Baduge, S.K.; Navaratnam, S.; Abu-Zidan, Y.; McCormack, T.; Nguyen, K.; Mendis, R.; Zhang, G.; Aye, B. Improving performance of additive manufactured (3D printed) concrete: A review on material mix design, processing, interlayer bonding, and reinforcing methods. Structures 2021, 29, 1597–1609. [Google Scholar] [CrossRef]
  55. Lao, W.; Li, W.; Tjahjowidodo, T. Variable-geometry nozzle for surface quality enhancement in 3D concrete printing. Addit. Manuf. 2021, 37, 101638. [Google Scholar] [CrossRef]
  56. Siddika, A.; Amin, M.R.; Rayhan, M.A.; Islam, M.S.; Al Mamun, M.A.; Alyousef, R.; Mugahed Amran, Y.H. Performance of sustainable green concrete incorporated with fly ash, rice husk ash, and stone dust. Acta Polytech. 2021, 61, 279–291. [Google Scholar] [CrossRef]
  57. Santana, H.A.; Amorim, N.S., Jr.; Ribeiro, D.V.; Cilla, M.S.; Dias, C.M.R. 3D printed mesh reinforced geopolymer: Notched prism bending. Cem. Concr. Compos. 2021, 116, 103892. [Google Scholar] [CrossRef]
  58. Archez, J.; Texier-Mandoki, N.; Bourbon, X.; Caron, J.F.; Rossignol, S. Shaping of geopolymer composites by 3D printing. J. Build. Eng. 2021, 34, 101894. [Google Scholar] [CrossRef]
  59. Ding, T.; Xiao, J.; Zou, S.; Wang, Y. Hardened properties of layered 3D printed concrete with recycled sand. Cem. Concr. Compos. 2020, 113, 103724. [Google Scholar] [CrossRef]
  60. Zhang, C.; Hou, Z.; Chen, C.; Zhang, Y.; Mechtcherine, V.; Sun, Z. Design of 3D printable concrete based on the relationship between flowability of cement paste and optimum aggregate content. Cem. Concr. Compos. 2019, 104, 103406. [Google Scholar] [CrossRef]
  61. Hong, D.L.H.; Mohammed, B.S.; Al-Fakih, A.; Wahab, M.M.A.; Liew, M.S.; Mugahed Amran, Y.H. Deformation properties of rubberized ecc incorporating nano graphene using response surface methodology. Materials 2020, 13, 2831. [Google Scholar] [CrossRef] [PubMed]
  62. Amran, M.; Lee, Y.H.; Fediuk, R.; Murali, G.; Mosaberpanah, M.A.; Ozbakkaloglu, T.; Lee, Y.Y.; Vatin, N.; Klyuev, S.; Karelia, M. Palm Oil Fuel Ash-Based Eco-Friendly Concrete Composite: A Critical Review of the Long-Term Properties. Materials 2021, 14, 7074. [Google Scholar] [CrossRef] [PubMed]
  63. Chakrawarthi, V.; Avudaiappan, S.; Amran, M.; Dharmar, B.; Jesuarulraj, L.R.; Fediuk, R.; Aepuru, R.; Vatin, N.; Flores, E.S. Impact Resistance of Polypropylene Fibre-Reinforced Alkali–Activated Copper Slag Concrete. Materials 2021, 14, 7735. [Google Scholar] [CrossRef]
  64. Avudaiappan, S.; Prakatanoju, S.; Amran, M.; Aepuru, R.; Flores, E.I.S.; Das, R.; Gupta, R.; Fediuk, R.; Vatin, N. Experimental Investigation and Image Processing to Predict the Properties of Concrete with the Addition of Nano Silica and Rice Husk Ash. Crystals 2021, 11, 1230. [Google Scholar] [CrossRef]
  65. Amran, M.; Huang, S.S.; Onaizi, A.M.; Murali, G.; Abdelgader, H.S. Fire spalling behavior of high-strength concrete: A critical review. Constr. Build. Mater. 2022, 341, 127902. [Google Scholar] [CrossRef]
  66. Muthukrishnan, S.; Gupta, S.; Kua, H.W. Application of rice husk biochar and thermally treated low silica rice husk ash to improve physical properties of cement mortar. Theor. Appl. Fract. Mech. 2019, 104, 102376. [Google Scholar] [CrossRef]
  67. Arularasi, V.; Thamilselvi, P.; Avudaiappan, S.; Flores, E.I.S.; Amran, M.; Fediuk, R.; Vatin, N.; Karelina, M. Rheological behavior and strength characteristics of cement paste and mortar with fly ash and GGBS admixtures. Sustainability 2021, 13, 9600. [Google Scholar] [CrossRef]
  68. Subash, N.; Avudaiappan, S.; Adish Kumar, S.; Amran, M.; Vatin, N.; Fediuk, R.; Aepuru, R. Experimental Investigation on Geopolymer Concrete with Various Sustainable Mineral Ashes. Materials 2021, 14, 7596. [Google Scholar] [CrossRef] [PubMed]
  69. Ma, G.; Salman, N.S.; Wang, L.; Wang, F. A novel additive mortar leveraging internal curing for enhancing interlayer bonding of cementitious composite for 3D printing. Constr. Build. Mater. 2020, 244, 118305. [Google Scholar] [CrossRef]
  70. Raju, S.; Rathinam, J.; Dharmar, B.; Rekha, S.; Avudaiappan, S.; Amran, M.; Usanova, K.I.; Fediuk, R.; Guindos, P.; Ramamoorthy, R.V. Cyclically Loaded Copper Slag Admixed Reinforced Concrete Beams with Cement Partially Replaced with Fly Ash. Materials 2022, 15, 3101. [Google Scholar] [CrossRef]
  71. Nair, S.A.O.; Panda, S.; Santhanam, M.; Sant, G.; Neithalath, N. A critical examination of the influence of material characteristics and extruder geometry on 3D printing of cementitious binders. Cem. Concr. Compos. 2020, 112, 103671. [Google Scholar] [CrossRef]
  72. Zeyad, A.M.; Magbool, H.M.; Amran, M.; Mijarsh, M.J.A.; Almalki, A. Performance of high-strength green concrete under the influence of curing methods, volcanic pumice dust, and hot weather. Arch. Civ. Mech. Eng. 2022, 22, 134. [Google Scholar] [CrossRef]
  73. Makul, N.; Fediuk, R.; Amran, M.; Zeyad, A.M.; Murali, G.; Vatin, N.; Klyuev, S.; Ozbakkaloglu, T.; Vasilev, Y. Use of recycled concrete aggregates in production of green cement-based concrete composites: A review. Crystals 2021, 11, 232. [Google Scholar] [CrossRef]
  74. Fediuk, R.; Timokhin, R.; Mochalov, A.; Otsokov, K.; Lashina, I. Performance properties of high-density impermeable cementitious paste. J. Mater. Civ. Eng. 2019, 31, 04019013. [Google Scholar] [CrossRef]
  75. Fediuk, R.S.; Lesovik, V.S.; Mochalov, A.V.; Otsokov, K.A.; Lashina, I.V.; Timokhin, R.A. Composite binders for concrete of protective structures. Mag. Civ. Eng. 2018, 6, 208–218. [Google Scholar] [CrossRef]
  76. Fediuk, R.; Smoliakov, A.; Stoyushko, N. Increase in composite binder activity. In Proceedings of the IOP Conference Series: Materials Science and Engineering, Tomsk, Russia, 9–11 June 2016; Volume 156. [Google Scholar]
  77. Bessmertnyi, V.S.; Lesovik, V.S.; Krokhin, V.P. The reducing effect of argon in the plasma treatment of high-melting nonmetallic materials (a review). Glas. Ceram. 2001, 58, 362–364. [Google Scholar] [CrossRef]
Figure 1. Construction of buildings using 3D-concrete printing.
Figure 1. Construction of buildings using 3D-concrete printing.
Buildings 12 01181 g001
Figure 2. Flowchart of the mix preparation.
Figure 2. Flowchart of the mix preparation.
Buildings 12 01181 g002
Figure 3. Flowchart of the laboratory equipment.
Figure 3. Flowchart of the laboratory equipment.
Buildings 12 01181 g003
Figure 4. XRD pattern of the specimens CB (1, 3, 7, and 8).
Figure 4. XRD pattern of the specimens CB (1, 3, 7, and 8).
Buildings 12 01181 g004
Figure 5. Dependence of the water demand of the mix on the MM1 amount and its specific surface.
Figure 5. Dependence of the water demand of the mix on the MM1 amount and its specific surface.
Buildings 12 01181 g005
Figure 6. Dependence of the setting time of the CB-8 on the content of the MM1-500.
Figure 6. Dependence of the setting time of the CB-8 on the content of the MM1-500.
Buildings 12 01181 g006
Figure 7. SEM images of the specimen CB-8 at magnifications: (a) ×10,000, (b) ×30,000, (c) ×50,000, (d) ×1500, (e) ×3000, (f) ×5000.
Figure 7. SEM images of the specimen CB-8 at magnifications: (a) ×10,000, (b) ×30,000, (c) ×50,000, (d) ×1500, (e) ×3000, (f) ×5000.
Buildings 12 01181 g007
Figure 8. Mix A made with a 3D printer.
Figure 8. Mix A made with a 3D printer.
Buildings 12 01181 g008
Figure 9. Plastic strength of the cement composites.
Figure 9. Plastic strength of the cement composites.
Buildings 12 01181 g009
Figure 10. SEM images: (a) mix A; (b) mix B.
Figure 10. SEM images: (a) mix A; (b) mix B.
Buildings 12 01181 g010
Table 1. Chemical content and fractions of the MM components.
Table 1. Chemical content and fractions of the MM components.
DepositOxides (%)Content of the Fractions (wt. %)
SiO2Al2O3Fe2O3CaOOtherSandySiltyClayey
Bentonite clay powder75.4111.046.61-6.9445–5132–4014–19
Sand98.230.450.81 0.51100--
Chalk0.540.44-54.744.32-100-
Table 2. Mix design of MM.
Table 2. Mix design of MM.
Mix IDProportionsSpecific Surface Area (m2/kg)
Bentonite Clay PowderChalkSand
MM1-500212500
MM1-300212300
MM1-400212400
MM1-600212600
MM2-500112500
MM3-500211500
MM4-500121500
Table 3. Mix proportions of CB.
Table 3. Mix proportions of CB.
Mix IDContent of CB (kg)Water (L)Sand (kg)Water-to-Cement Ratio
CEM I
(300 m2/kg)
CEM I
(500 m2/kg)
HPMM
CB-1450---18915000.42
CB-2450-1.35-17115000.38
CB-3450--13.518015000.40
CB-4450-1.3513.516215000.36
CB-5-450--19815000.44
CB-6-4501.35-18915000.42
CB-7-450-13.519315000.43
CB-8-4501.3513.518915000.42
Table 4. Characteristics of quartzite sandstone and sand.
Table 4. Characteristics of quartzite sandstone and sand.
Fine AggregateSieve Fractions (mm)Average Grain Size (mm)Packing DensityWater Absorption
(% wt.)
Elastic Modulus (MPa)
Quartzite sandstone5–2.53.070.590.49100.2
2.5–1.251.750.6930.5196.2
0.63–0.3140.430.7840.5390.7
0.14–0.10.110.8580.5687.0
Sand5–2.52.970.600.5997.1
2.5–1.251.650.6980.6192.1
0.63–0.3140.390.7940.6387.0
0.14–0.10.090.8630.6681.0
Table 5. Proportions of concrete mixes (mortars).
Table 5. Proportions of concrete mixes (mortars).
Content (kg/m3)Theoretical Density (kg/m3)
CEM IFine AggregateHPAEMMWaterW/C
QS (mm)S
2.5–1.251.25–0.630.63–0.315
Mix A3735804804901.00.16121540.421935
Mix B45015001940.401950
Table 6. Generalized list of all compositions and tests.
Table 6. Generalized list of all compositions and tests.
CompositionsTests
Mineral modifiers
MM1-500Specific surface area
MM1-300Specific surface area
MM1-400Specific surface area
MM1-600Specific surface area
MM2-500Specific surface area
MM3-500Specific surface area
MM4-500Specific surface area
Composite binders
CB-1Average density, compressive and flexural strength at 7 and 28 days, XRD
CB-2Average density, compressive and flexural strength at 7 and 28 days
CB-3Average density, compressive and flexural strength at 7 and 28 days
CB-4Average density, compressive and flexural strength at 7 and 28 days, XRD
CB-5Average density, compressive and flexural strength at 7 and 28 days
CB-6Average density, compressive and flexural strength at 7 and 28 days
CB-7Average density, compressive and flexural strength at 7 and 28 days, XRD
CB-8Average density, compressive strength at 3, 7, 14, and 28 days, flexural strength at 7 and 28 days, XRD, SEM, setting time
CB-8 (at 1% MM1-500)Compressive strength at 3, 7, 14, and 28 days
CB-8 (at 5% MM1-500)Compressive strength at 3, 7, 14, and 28 days
CB-8 (at 7% MM1-500)Compressive strength at 3, 7, 14, and 28 days
CB-8 (at 10% MM1-500)Compressive strength at 3, 7, 14, and 28 days
CEMI + 0.5%HP + 3%MM1-500Compressive strength at 28 days
CEMI + 0.5%HP + 3%MM2-500Compressive strength at 28 days
CEMI + 0.5%HP + 3%MM3-500Compressive strength at 28 days
CEMI + 0.5%HP + 3%MM4-500Compressive strength at 28 days
CEMI + 0.5%HP + 3%MM1-300Compressive and flexural strength at 28 days
CEMI + 0.5%HP + 3%MM1-400Compressive and flexural strength at 28 days
CEMI + 0.5%HP + 3%MM1-600Compressive and flexural strength at 28 days
Concrete
Mix APlastic strength, SEM, RTI, compressive strength at 7 and 28 days
Mix BPlastic strength, SEM, RTI, compressive strength at 7 and 28 days
Table 7. Compressive strength of the composite binder.
Table 7. Compressive strength of the composite binder.
CEM I + 0.5% HP + 3.0% MM with a Ratio of Sand: Bentonite Clay Powder: Chalk (by wt.)Compressive Strength at 28 Days, MPa)Strength Gain (%)Compressive Strength at 28 Days (MPa)Strength Gain (%)
500 m2/kg300 m2/kg
46.3 ± 0.146.3 ± 0.1
2:2:167.7 ± 0.146.258.7 ± 0.126.8
2:1:162.7 ± 0.135.451.7 ± 0.111.7
1:2:163.3 ± 0.136.352.3 ± 0.113.0
1:1:260.4 ± 0.129.851.4 ± 0.111.0
Table 8. Physical and mechanical properties of the CB.
Table 8. Physical and mechanical properties of the CB.
Mix IDAverage Density (kg/m3)Compressive Strength (MPa) at the AgeFlexural Strength (MPa) at the Age
7 Days28 Days7 Days28 Days
CB-1195014.5 ± 0.0543.2 ± 0.11.6 ± 0.054.7 ± 0.05
CB-2194616.3 ± 0.0548.8 ± 0.11.8 ± 0.055.3 ± 0.05
CB-3195116.2 ± 0.0548.4 ± 0.11.7 ± 0.055.2 ± 0.05
CB-4194918.1 ± 0.0560.1 ± 0.12.2 ± 0.056.0 ± 0.05
CB-5195118.7 ± 0.0555.9 ± 0.12.1 ± 0.056.1 ± 0.05
CB-6195021.1 ± 0.0563.1 ± 0.12.4 ± 0.056.8 ± 0.05
CB-7195220.9 ± 0.0562.2 ± 0.12.3 ± 0.056.7 ± 0.05
CB-8195123.4 ± 0.0569.9 ± 0.12.6 ± 0.057.4 ± 0.05
Table 9. Dependence of the compressive strength of the composite binder on the dosage of the mineral modifier.
Table 9. Dependence of the compressive strength of the composite binder on the dosage of the mineral modifier.
BinderCompressive Strength
at 28 Days (MPa)
Strength Gain (%)
CEM I46.3 ± 0.1
CEM I + 0.5% HP + 1.0% MM1-50049.3 ± 0.16.5
CEM I + 0.5% HP + 2.0% MM1-50058.5 ± 0.126.3
CEM I + 0.5% HP + 3.0% MM1-50067.7 ± 0.146.2
CEM I + 0.5% HP + 5.0% MM1-50059.6 ± 0.128.7
CEM I + 0.5% HP + 7.0% MM1-50058.3 ± 0.125.9
CEM I + 0.5% HP + 10.0% MM1-50055.2 ± 0.119.2
Table 10. Influence of the fineness of grinding of MM on the strength of CB.
Table 10. Influence of the fineness of grinding of MM on the strength of CB.
Water-to-Cement RatioSpecific Surface Area of the MM (m2/kg)Strength at 28 Days (MPa)
CEM IMMHPCompressiveFlexural
99.5-0.50.2830048.2 ± 0.18.3 ± 0.05
96.530.50.2940060.2 ± 0.18.9 ± 0.05
96.530.50.3250067.5 ± 0.19.5 ± 0.05
96.530.50.3460065.8 ± 0.19.4 ± 0.05
Table 11. Compressive strength gain kinetics for composite binder specimens.
Table 11. Compressive strength gain kinetics for composite binder specimens.
Age (Days)Compressive Strength of the Composite Binder CB-8 (MPa) with the MM1-500 Content (%)
135710
322.4 ± 0.0523.6 ± 0.0523.4 ± 0.0523.2 ± 0.0522.9 ± 0.05
732.1 ± 0.0533.8 ± 0.0533.4 ± 0.0533.2 ± 0.0532.8 ± 0.05
1454.5 ± 0.157.4 ± 0.156.8 ± 0.156.4 ± 0.155.7 ± 0.1
2864.1 ± 0.167.5 ± 0.166.8 ± 0.166.3 ± 0.165.5 ± 0.1
Table 12. Relationship between rheological and strength characteristics of a mortar.
Table 12. Relationship between rheological and strength characteristics of a mortar.
MixRTIPlastic Strength (Dimensional Stability) (kPa) at Test Time (Min)Compressive Strength (MPa) at the Age (Days)
1153550728
Mix A2560.0762.21744.41855.635.0 ± 0.0567.5 ± 0.1
Mix B7222.2133.0627.01444.522.5 ± 0.0542.8 ± 0.1
Table 13. Physical and mechanical properties of specimens depending on the method of casting and testing.
Table 13. Physical and mechanical properties of specimens depending on the method of casting and testing.
Type of SpecimensPhysical and Mechanical Properties
Average Density (kg/m3)Compressive Strength (MPa), Age (Days)
3728
Molded in standard shapes191012.3 ± 0.0528.5 ± 0.0567.1 ± 0.05
Cut and tested along the 3D-printing direction18509.4 ± 0.0520.7 ± 0.0560.8 ± 0.05
Cut and tested across the 3D-printing direction18708.2 ± 0.0518.5 ± 0.0555.9 ± 0.05
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Lesovik, V.; Tolstoy, A.; Fediuk, R.; Amran, M.; Ali, M.; de Azevedo, A.R.G. Improving the Performances of a Mortar for 3D Printing by Mineral Modifiers. Buildings 2022, 12, 1181. https://doi.org/10.3390/buildings12081181

AMA Style

Lesovik V, Tolstoy A, Fediuk R, Amran M, Ali M, de Azevedo ARG. Improving the Performances of a Mortar for 3D Printing by Mineral Modifiers. Buildings. 2022; 12(8):1181. https://doi.org/10.3390/buildings12081181

Chicago/Turabian Style

Lesovik, Valery, Aleksandr Tolstoy, Roman Fediuk, Mugahed Amran, Mujahid Ali, and Afonso R. G. de Azevedo. 2022. "Improving the Performances of a Mortar for 3D Printing by Mineral Modifiers" Buildings 12, no. 8: 1181. https://doi.org/10.3390/buildings12081181

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