Next Article in Journal
Investigation on Mechanical and Thermal Properties of 3D-Printed Polyamide 6, Graphene Oxide and Glass-Fibre-Reinforced Composites under Dry, Wet and High Temperature Conditions
Next Article in Special Issue
Physico-Chemical Study of the Possibility of Utilization of Coal Ash by Processing as Secondary Raw Materials to Obtain a Composite Cement Clinker
Previous Article in Journal
Structure and Properties of Polystyrene-Co-Acrylonitrile/Graphene Oxide Nanocomposites
Previous Article in Special Issue
Innovative Use of Single-Use Face Mask Fibers for the Production of a Sustainable Cement Mortar
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Studies of Utilization of Technogenic Raw Materials in the Synthesis of Cement Clinker from It and Further Production of Portland Cement

by
Nurgali Zhanikulov
1,
Bayan Sapargaliyeva
2,
Aktolkyn Agabekova
3,
Yana Alfereva
4,
Aidin Baidibekova
5,
Samal Syrlybekkyzy
6,*,
Lazzat Nurshakhanova
7,
Farida Nurbayeva
6,
Gulzhan Sabyrbaeva
6,
Yergazy Zhatkanbayev
8,
Pavel Kozlov
9,*,
Aizhan Izbassar
10 and
Olga Kolesnikova
11,*
1
Academician E.A. Buketov Karaganda University, Karaganda 100028, Kazakhstan
2
Abai Kazakh National Pedagogical University, Almaty 050010, Kazakhstan
3
Department of Electrical Engineering, H. A. Yassavi International Kazakh-Turkish University, Turkestan 161200, Kazakhstan
4
M.V. Lomonosov Moscow State University, 119991 Moscow, Russia
5
Department of Higher Mathematics and Physics for Technical Specialties, M. Auezov South Kazakhstan University, Shymkent 160012, Kazakhstan
6
Department of Ecology and Geology, Sh. Yesenov Caspian University of Technology and Engineering, Aktau 130002, Kazakhstan
7
Department of Petrochemical Engineering, Sh. Yesenov Caspian University of Technology and Engineering, Aktau 130002, Kazakhstan
8
Department of Chemistry, M. Auezov South Kazakhstan University, Shymkent 160012, Kazakhstan
9
Polytechnic Institute, Far Eastern Federal University, 690922 Vladivostok, Russia
10
Department of Construction Engineering, Sh. Yesenov Caspian University of Technology and Engineering, Aktau 130002, Kazakhstan
11
M. Auezov South Kazakhstan University, Shymkent 160012, Kazakhstan
*
Authors to whom correspondence should be addressed.
J. Compos. Sci. 2023, 7(6), 226; https://doi.org/10.3390/jcs7060226
Submission received: 20 April 2023 / Revised: 21 May 2023 / Accepted: 25 May 2023 / Published: 1 June 2023
(This article belongs to the Special Issue Composites for Construction Industry)

Abstract

:
Four series of experiments were carried out to study the possibility of replacing clay and an iron-containing component with tefritobasalt and lead slag as part of the initial charge for Portland cement. The experiments were carried out at atmospheric pressure and a temperature of 1350 °C. It was shown that the replacement of clay and an iron-containing component with tefritobasalt and lead slag as part of the initial charge in the cement industry will lead to a decrease in temperature by 100 °C in the technological scheme of production and a reduction in energy consumption, since the theoretical specific consumption of raw materials is 1.481 t/t of clinker, which is approximately 70 kg lower than in traditional mixtures. The content of non-traditional components in total was 24.69%. In addition, tefritobasalts improved clinker formation processes, contributed to a decrease in the firing temperature, and intensified the clinker firing process. A small amount of lead slag (5.06%) introduced into the mixture changed the structure of the clinker and improved the process of mineral formation while also improving roasting and reducing the anthropogenic impact on the environment through the disposal of man-made waste. The strength of the experimental composite cements was tested after 7 and 28 days on small samples measuring 2 × 2 × 2 cm. The physicomechanical characteristics and structure of composite cements were studied.

1. Introduction

The main raw materials in the production of cement are carbonate and clay rocks, as well as some types of industrial and technogenic waste [1,2,3,4,5,6,7,8,9]. Due to the depletion of stocks of high-quality traditional raw materials, in the future, the maximum use of industrial man-made waste and the search for new types of raw materials that can be used in the production of Portland cement are required [10,11,12,13,14,15,16,17,18,19].
Igneous rocks such as basalts, tefritobasalts, etc. are non-traditional materials and industrial wastes that allow for a reduction in energy costs during the firing of cement clinker. It is also possible to completely replace the traditional clay component with coal mining waste from coal mines.
Studies on the use of basalts, tefritobasalts, lead slags, coal mining waste, etc. in the production of Portland cement as raw materials have been carried out [2,20,21,22,23,24,25,26,27]. In the technological cycle for the production of Portland cement clinker, the possibility of replacing clay with basalt as the main source of aluminosilicate substance has been shown. According to the results of physical and mechanical tests, the strength of cement obtained using basalt raw materials is not lower than conventional cement. The resulting clinker consists mainly of alite, belite, tricalcium aluminate, and ferrite. The effectiveness of using basalts to reduce the melting temperature of the mixture by 60–100 °C and accelerate the process of clinker formation has also been shown [28,29,30,31,32,33,34].
The possibility of using basalts as an additional cementing material in the composition of cement for wells has been shown [35,36,37,38,39,40,41].
The use of technogenic raw materials leads to significant changes in the content of heavy metals in clinker. This applies mainly to zinc, lead, and nickel. Studies have focused on the effect of zinc on the sintering process. It has been established that even the first percent of Zn content causes a decrease in the content of tricalcium aluminate in the clinker. Alloyed cements were at least as reactive as the reference cement [42,43,44,45].
The purpose of our research was to study the possibility of using tefritobasalt and lead slag as substitutes for clay and iron-containing components during clinker firing in the production of Portland cement. If implemented, this would be able to improve the state of the natural environment due to the utilization of man-made waste and conserve resources due to the involvement of waste instead of natural raw materials that need to be explored, mined, and maintained to maintain quarries [5,17,19,21,28]. Thus, there would also be an economic advantage to the proposed study [13,23].

2. Materials and Research Methods

To achieve the set goals, we studied the chemical and mineral compositions of raw materials and the clinker obtained from them. Scanning (raster) electron microscopic analysis of raw materials and baked clinkers was performed in the laboratory of local methods for studying substances at the Department of Petrology and Volcanology, Faculty of Geology, M.V. Lomonosov Moscow State University using an energy dispersive spectrometer (EDS) (INCA-Energy 350) based on a JEOL JSM-6480LV (Tokyo, Japan) microscope.
The content of the micro components in the initial material and products of the experiment was measured using the ICP-MS (Horiba, France) method at the Department of Geochemistry, M.V. Lomonosov Moscow State University. Samples for analysis were prepared by sintering with sodium carbonate. A 0.1 g sample was mixed with 0.3 g anhydrous sodium carbonate (Merck, Rahway, NJ, USA, Suprapur®) in an agate mortar and transferred to a corundum crucible. Sintering was carried out at 800 °C for one hour in a muffle furnace. The resulting sintered pellets were treated with 5 mL of HNO3:HCl:HF (10:2:1), and after dissolution, diluted to 50 mL with deionized water (EasyPure®). For measurement, the solution was diluted with 3% nitric acid. The sintering method was tested and compared with multi-acid microwave decomposition of several standard samples [21].
The measurements were carried out on a high-resolution mass spectrometer with ionization in an inductively coupled plasma Thermo Element 2. Separation of ions was carried out using an analyzer with double focusing—magnetic and electrostatic modes. The ions were detected using an electron multiplier, which remained linear in the range from 1 to 1 × 1010 ions per second. Calibration of the sensitivity of the instrument over the entire mass scale was carried out using reference 68-element solutions (ICP-MS-68A, HPS, solutions A and B), which included all analyzed elements in the samples. To control the quality of measurements and take into account the drift in the sensitivity of the instrument, analyses of the samples were alternated with analyses of the reference sample with a frequency of 1:10. Indium at a concentration of 10 ppb was introduced into the samples as an internal standard. The analysis error ranged from 0.5 to 2 rel.%.
The X-ray diffraction data of the samples were obtained with a STOE-STADIMP (Darmstadt, Germany) powder diffract meter, with a curved Ge (111) monochromatic providing strictly monochromatic Co Kα1-radiation. Data collection was carried out in the mode of stepwise overlapping of scanning areas from 6° to 90° by 2θ using a position-sensitive linear detector, the capture angle of which was 5° by 2θ, with a channel width of 0.02°. The phase composition was determined using the software package WinXPowSoftware // STOE&CIEGmbH, 2000 and Match! Software // CrystalImpactGbR, 2016, with their associated PDF-2 Powder Database (ICDD-2013).

2.1. Limestone of the Sastobe Deposit

The deposit is located in the Tyulkubas district of the Turkestan region (Kazakhstan). Limestone reserves are approximately 70 million tons. The productive stratum is composed of limestones of the Tournaisian Stage of the Carboniferous Age.
The bedlike limestone deposit has a northwest strike and a steep dip to the southwest. Its length is 1200 m, with a width of 320 m and a thickness of 670 m. The physical and mechanical properties of the limestones are bulk density—2.68–2.75 g/cm3, water absorption, approximately—0.1–0.54%, loosening coefficient—1.24, and dry compression density—475–940 kg/cm2. Limestones of the Sastobe deposit have special molecular strength and frost resistance [22]. The X-ray of limestone is shown in Figure 1.
According to X-ray phase analysis, limestone consists mainly of calcite and quartz; the diffraction maxima of calcite were noted on the diffraction pattern: d = 3.86, 3.03, 2.49, 2.28, 2.09, 1.91, 1.87, and 1.60 Å. Quartz is also present: d = 3.34, 3.14, 2.84, 1.62, 1.52, and 1.47 Å. The content of oxides (SiO2 and Al2O3) is low—2.67 and 0.26 wt.%, respectively, and the content of Fe2O3 is very low—0.57 wt.%. The content of MgO is insignificant—0.88 wt.%. The limestone is pure and highly basic, and the content of CaO is more than 52 wt.%. The content of alkalis is within the normal range (less than 0.54 wt.%).

2.2. Coal Mining Waste

Coal mines are located in the city of Lenger, Turkestan region (Kazakhstan). More than 6 million tons of waste have been accumulated in the storage facilities. The repository occupies approximately 25 hectares of land [23,24]. The X-ray of coal mining waste is shown in Figure 2.
According to the results of X-ray phase analysis, the coal mining wastes consist of kaolinite, gypsum, quartz, and a small amount of illite. Diffraction maxima were recorded: quartz d = 4.29, 3.35, 2.46, 2.28, 1.82, and 1.54 Å. There are also minerals of gypsum dihydrate d = 3.07, 2.24, 2.13, and 1.85 Å; calcite d = 2.28, 1.86, and 1.60 Å; dolomite d = 1.80, 1.54, and 1.11 Å; kaolinite clay minerals d = 7.207, 3.584, and 1.673 Å; illite (hydromica) d = 3.798, 3.257, and 1.983 Å; and carbon d = 2.132, 2.568, and 4.484 Å. The content of coal in the composition of coal mining waste was 15.23%. The content of silicon oxide was more than 55 wt.%, Al2O3—10.6 wt.%, and the carbon content in the coal mining waste was more than 15 wt.%. They can replace the aluminosilicate component in the raw mix.

2.3. Tefritobasalt

The Daubaba deposit is located in the Tyulkubas district of the Turkestan region (Kazakhstan). The deposit of Daubaba tefritobasalts has an inclined northeast direction. The length of the deposit is 2200 m, the width is almost 1200 m, and the thickness is from 13 to 70 m. The reserves of tefritobasalts are approximately 20 million tons. Tefritobasalts have a porphyritic structure (Figure 3).
Porphyritic phenocrysts are euhedral short-columnar grains of clinopyroxene. The composition of clinopyroxene corresponds to the isomorphic series diopside–hedenbergite. The grain size reaches 0.5 mm in length. It makes up approximately 30% of the total volume of the breed. Phenocrysts also include relics of plagioclase grains replaced by an aggregate of secondary minerals. The predominant secondary mineral is zeolite (analcime). The relics have elongated prismatic outlines. Their size was 0.5–1 mm in length and approximately 0.1 mm in width. The content in the breed was approximately 15%. Plagioclase was almost completely replaced. The exact composition of the original mineral cannot be determined.
The sample contained isometric grains of magnetite ranging in size from hundredths of a millimeter to ≈0.2 mm. It contained impurities of Al, Mn, V, Si, Cr, and Zn. The amount of magnetite in the rock did not exceed 15%. Small (up to 0.1 mm in size) grains of chlorite, calcite, alkali feldspar, and apatite were also found. Their total content did not exceed 15%. The X-ray of tefritobasalt is shown in Figure 4.
The diffraction peaks of the following minerals were noted on the X-ray pattern of tefritobasalt: pyroxene d = 3.62, 3.26, 2.78, 2.30, 2.19, 1.97, and 1.94 Å; plagioclase d = 4.02, 3.43, 2.68, 2.59, and 2.14 Å; biotite d = 4.78, 4.60, 3.52, 1.66, 1.60, and 1.54 Å; anorthite d = 4.17, 3.13, 2.24, 1.90, 1.84, and 1.51 Å; and olivine d = 2.97, 2.48, 2.34, 2.03, 1.75, 1.62, 1.57, and 1.49 Å. According to the results of the X-ray phase analysis, olivine and clinochlore were also found in the rock in small amounts up to 5%. The groundmass was composed of a submicron aggregate, the mineral composition of which, apparently, was close to that of phenocrysts.
The temperature at the beginning of the melting of tefritobasalts, determined by the dilatometric method, was 1280 °C. At 1450 °C, the rock powder (sifted through a 02 sieve) completely melted, clarified, and homogenized within 45 min [25]. The physical and mechanical properties of the tefritobasalts were as follows: density = 2.0 g/cm3 and compressive strength = 147.6–195.8 MPa. According to chemical analysis, tefritic basalt contains SiO2—45.54 wt.%, Al2O3—more than 10 wt.%, and Fe2O3—approximately 8.5 wt.%. The magnesium content was 6.95 wt.%. It contained a significant amount of alkalis (K2O + Na2O), which was 6.54 wt.%.

2.4. Lead Slag

The slag plant used was JSC “Yuzhpolymetal”, which is located in the city of Shymkent (Kazakhstan). The balance reserves of lead slags are approximately 2 million tons [26,27]. Lead slags contain up to 37.25% Fe2O3 and can replace the corrective additive. In addition, lead slags contain up to 14% CaO and partially replace the carbonate component. The X-ray of lead slag is shown in Figure 5.
According to the results of the X-ray phase analysis of lead slag, the diffraction maxima were recorded: fayalite d = 3.77, 2.81, 2.31, 1.77, and 1.58 Å; wustite d = 4.27, 2.14, 1.52, and 1.50 Å; hematite d = 3.66, 2.69, 2.42, 2.22, 1.83, 1.66, and 1.48 Å; melilite d = 2.51, 2.04, 1.87, 1.74, 1.71, and 1.54 Å; and zinc spinel d = 1.90 and 1.62 Å.
According to the microprobe study, lead production slag has an incompletely crystalline porphyry structure (Figure 6). The bulk was composed mainly of glass. The amount of glass was approximately 60% of the total volume of the sample. The structure of the ground mass was heterogeneous. It alternated areas with vitrophyric and hyalopilitic structures. In the most crystallized areas, an increase in the content of zinc and sulfur and a decrease in the amount of calcium were noted.
Phenocrysts are represented by various sulfides that form rounded isometric aggregates. Their size varied from a few microns to ≈0.5 mm. The mineral composition was different. There were aggregates composed of approximately 95% chalcocite (Cu2S) and 5% galena (PbS), and aggregates composed mainly of sphalerite (ZnS) with a small amount (up to 5%) of chalcopyrite (CuFeS2) and galena. Subhedral grains of hematite (Fe2O3) were present as phenocrysts. Their average size was 10 µm. The total amount in the sample was approximately 20%.
Based on the data on the chemical composition of the raw materials in the ROCS program [28], a four-component calculation of the composition of the raw batch was made, consisting of “Limestone, tefritobasalt, coal mining waste, and lead slag”. For the optimal modular characteristics of the resulting clinkers, the quantitative ratio of the initial components, as well as the calculated chemical and mineral composition of the clinker, were determined.
The raw materials were crushed and sieved through a No. 008 sieve. As a result, the surface area of particles per 1 g of material (specific surface area) was 3118–3326 cm2/g, and the average particle size was 5.67–5.92 µm. In accordance with the obtained calculated composition, a mixture was made from them. This mixture, on a hydraulic press under a pressure of 20 MPa, was molded into tablets with a diameter of 15 mm and a height of 10 mm.
The firing of the tablets was carried out in an electric laboratory furnace, SX-2–18TP (NanYang, China), at M. Auezov South Kazakhstan University (Kazakhstan). The temperature regime of firing fully corresponded to the temperature regime of a rotary industrial kiln. The rise in temperature to 1350 °C occurred within 2–2.5 h. Exposure at the maximum temperature was carried out for 30 min. A general view of the furnace, unfired pellets, and clinkers obtained at a temperature of 1350 °C is shown in Figure 7.
The duration of clinker firing was controlled by the degree of absorption of CaO into the clinker minerals. The burned tablets were crushed. The qualitative content of free CaO was determined microscopically. The quantitative content of CaO in the clinkers was determined using the ethylene–glycerate method [29].

3. Results and Discussion

In accordance with the results of calculating the composition of the raw mixture (Table 1), the optimal characteristics of the clinker for these initial conditions were as follows: SC = 0.94; silicate module n = 2.02; and alumina modulus p = 0.95. The saturation factor (SC) of the clinker is the ratio of CaO remaining after saturation of Al2O3 and Fe2O3 to 3CaO·Al2O3 and 4CaO·Al2O3·Fe2O3 to the amount of CaO that is necessary to completely saturate the silica to 3CaO·SiO2. The SC value of the factory clinkers varied from 0.85 to 0.95, depending on the composition and properties of the raw materials, the firing conditions, etc. The silicate module is the ratio of the SiO2 content to the sum of the oxides A12O3 and Fe2O3. The n module value ranged from 1.7 to 4.0. The alumina modulus is the ratio of the Al2O3 content to the Fe2O3 content and characterizes the composition of the aluminoferrite phase in the clinker. The values of the p modulus ranged from 0.9 to 3.0. The theoretical specific consumption of raw materials is 1.481 t/t of clinker, which is approximately 70 kg lower than in traditional mixtures. The content of non-traditional components in the mixture was 24.69% (Table 2).
Figure 8 shows the dependence of the content of alite (3CaO·SiO2) in the clinker on the value of SC and the silicate and alumina modules of the raw batch. With an increase in the SC from 0.8 to 0.95 and an n modulus from 1.5 to 3.5, the alite content increased from 40% to 70%. With the silicate modulus n = 2.02, the value of the alumina modulus was p = 0.95. An increase in the silicate modulus contributed to an increase in the alumina modulus as well as an increase in the content of tricalcium aluminate. According to the results of calculating the composition of the raw charge consisting of “Limestone, tefritobasalt, coal mining waste, and lead slag” is suitable for producing sulfate-resistant Portland slag cement clinkers of the CEMIII/AC grade, according to GOST 22266-2013 [30]. The calculated chemical and mineral composition of the clinker contained no more than 7% of 3CaO·Al2O3, and the amount of Al2O3 and MgO did not exceed 5%. The total content of 3CaO·Al2O3 + 4CaO·Al2O3·Fe2O3 was 18.07% and did not exceed the maximum allowable values (3CaO·Al2O3 + 4CaO·Al2O3·Fe2O3 < 22%). Alite in the composition of the clinker was 57.88%, and in belite (2CaO·SiO2) it was 18.82%.
Clinker firing is characterized by a complex physical and chemical process. The reactions occurring in the process of clinker formation determine the quality of the finished product and its phase composition.
Clinker fired at a temperature of 1350 °C reached almost complete assimilation of calcium oxide within 30 min, and the content of free CaO in the resulting clinker was 0.2% (Table 3). The quality of the clinker was high, and the chemical and mineralogical composition met the requirements. It has been established that in the studied raw mixtures, the process of clinker formation ends at 1350 °C (i.e., 100 °C lower than in traditional raw mixes) [31,32,33,34,35,36,37,38,39,40].
On the radiograph of the clinker obtained by burning at a temperature of 1350 °C of energy and resource-saving raw mixtures, lines characteristic of free CaO are found (Figure 9).
When burning the energy resource-saving raw material mixture at 1350 °C, the following clinker minerals were formed: alite (C3S) d = 2.19, 2.62, 2.78, and 3.05 Å; belite (C2S) d = 1.73, 1.93, 2.74, and 2.88 Å; tricalcium aluminate (C3A) d = 1.87, 1.90, 2.95, and 4.04 Å; four-calcium aluminoferrite (C4AF) d = 2.03, 2.13, and 2.66 Å The peaks characteristic of free CaO are absent.
Figure 10 shows micrographs of the polished section of the resulting clinker. The synthesized clinker had a full-crystalline structure and a massive texture. The main crystalline phases were evenly distributed over the sample [32,33,34,35,36,37,38,39,40,41,42,43]. Alite was represented by isometric subhedral crystals 20–50 µm in size. Belite inclusions were observed in large rhombohedral alite crystals. The content of alite in the composition of the clinker was 57.88%. Belite was represented by small grains up to 10 µm in size. The crystals had a round, near-isometric shape. Belite grains were in direct contact with alite grains, which indicates their formation by reactions in the solid state. The belite content in the clinker composition was 18.82%.
Along the grain boundaries of alite and belite, there was an aluminate phase—a dark intermediate substance—and an aluminoferrite phase—a light intermediate substance. The intermediate substance was present in sufficient quantity (3CaO·Al2O3 + 4CaO·Al2O3·Fe2O3 = 18.07%). In traditional clinker, the aluminoferrite phase is 14%. The aluminoferrite phase in all parameters (hydraulic activity, hardening rate, and strength characteristics) occupied an intermediate position between 3CaO·SiO2 and 2CaO·SiO2. Hardening products have increased water and corrosion resistance. In clinker sinters based on tefritobasalt, well-formed, regular-shaped alite crystals are formed. Tefritobasalt improved and accelerated the processes of clinker formation, helped reduce the firing temperature (the beginning of melting was 1280 °C), and intensified the process of firing the clinker. The traditional raw mix was heated to clinker sintering at 1450–1500 °C.
A small amount of lead slag (5.06%) introduced into the mixture changed the structure of the clinker and improved the process of mineral formation. The addition of lead slag to the raw mix improved fire ability and calcium silicate formation at a lower temperature. Zinc oxide contained in the composition of the slag contributed as a mineralizer to the dissolution of the residue of free lime at 1350 °C. Its presence affected the course of solid reactions and the formation of silicates and calcium aluminates. During the firing process, the material lost a significant amount of these metals. In the finished product, their content was low. According to the ICPMS, the content of lead in the clinker was 41 ppm, and zinc was 197 ppm.
The microstructure of traditional cement clinker is uniformly granular and finely crystalline. The optimal size of alite crystals is 20–40 microns. The content of tricalcium aluminate was approximately 8%. The saturation coefficient of the raw mixture was in the range of 0.90–0.92.
The conducted research shows the possibility of replacing traditional raw materials with man-made industrial waste such that economically beneficial phenomena, including resource conservation and utilization of man-made waste, are possible, while reducing the anthropogenic impact on the region.
In a laboratory ball mill, cement was mixed with 5% gypsum. The residue on sieves No. 02 and No. 008 was determined every 10 min, and the total grinding time was 30 min. A study of the kinetics of grinding showed that the burning temperature and the amount of slag introduced into the raw material mixture had a noticeable effect on the grinding process of cements. After 30 min of grinding, the cement residue on sieve No. 008 was 13%. The specific surface of the cements was determined using a PSX-K. According to the results of the analysis, the specific surface of the cements was 3245 cm2/g, and the average particle size was 5.97 microns.
In the laboratory, the physicomechanical properties of the experimental Portland cement were tested in small samples of 2 × 2 × 2 cm after 7 and 28 days [41,42,43,44,45,46]. As can be seen from the data in Table 4, the tensile strength of the obtained cements increased with increasing hardening time. After 28 days, the compressive strength of the cements was 41.8 MPa. These indicators correspond to the cement grade for strength M400 according to GOST 10178-85 or strength class 42.5 according to GOST 31108-2016.
These studies confirm and supplement previously conducted similar studies by both international scientific centers and scientific communities [31,32,33,36,38,43,44,45,46].

4. Conclusions

  • The conducted research shows the possibility of replacing traditional raw materials with man-made industrial waste such that economically beneficial phenomena, including resource conservation and utilization of man-made waste, are possible, while reducing the anthropogenic impact on the region.
  • In the studied initial experimental compositions, the replacement of the clay- and iron-containing component with tefritobasalt and lead slag led to a decrease in the melting temperature of the mixture by 100 °C relative to traditional compositions and is environmentally and economically feasible.
  • In the samples obtained from the four-component raw material mixture “Limestone—tefritobasalt—coal mining waste—lead slag”, it is clear that the synthesized clinker has a full-crystalline structure with a content of 57.88% alite and 18.82% belite.
  • According to the calculation results, SC = 0.94, silicate modulus n = 2.02, and alumina modulus p = 0.95. The theoretical specific consumption of raw materials is 1481 t/t of clinker, which is approximately 70 kg lower than in traditional mixtures. The content of non-traditional components in total was 24.69%.
  • According to the results of chemical analysis, the content of free CaO in the clinker was 0.2%, which meets the requirements of GOST. The quality of the synthesized clinker was high, the minerals were formed correctly, and the size of the alite crystals was 20–50 microns.
  • The strength of the obtained Portland cement after 28 days under compression was 41.8 MPa, which corresponds to the cement grade of strength M400 according to GOST 10178-85 or strength class 42.5 according to GOST 31108-2016.

Author Contributions

Conceptualization, N.Z. and O.K.; methodology, B.S.; software, A.A.; validation, Y.A., A.B. and S.S.; formal analysis, L.N.; investigation, F.N.; resources, G.S.; data curation, Y.Z.; writing—original draft preparation, P.K.; writing—review and editing, A.I.; visualization, O.K.; supervision, N.Z.; project administration, P.K.; funding acquisition, N.Z. All authors have read and agreed to the published version of the manuscript.

Funding

The study is funded by the authors.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing is not applicable to this article.

Acknowledgments

This research was supported by the JSC Center for International Programs “Bolashak” of the Ministry of Science and Higher Education of the Republic.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhanikulov, N.; Khudyakova, T.; Taimassov, B.; Sarsenbayev, B.; Dauletiarov, M.; Kolesnikov, A.; Karshygayev, R. Receiving Portland Cement from Technogenic Raw Materials of South Kazakhstan. Eurasian Chem.-Technol. J. 2019, 21, 333. [Google Scholar] [CrossRef]
  2. Kolesnikova, O.; Vasilyeva, N.; Kolesnikov, A.; Zolkin, A. Optimization of raw mix using technogenic waste to produce cement clinker. Min. Inf. Anal. Bull 2022, 60, 103–115. [Google Scholar] [CrossRef]
  3. Zhantasov, M.K.; Bimbetova, G.Z.; Kolesnikov, A.S.; Sadyrbayeva, A.S.; Orynbasarov, A.K.; Kutzhanova, A.N.; Turemuratov, R.S.; Botabaev, N.E.; Zhantasova, D. Examination of optimal parameters of oxy-ethylation of fatty acids with a view to obtaining demulsifiers for deliquefaction in the system of skimming and treatment of oil: A method to obtain demulsifier from fatty acids. Chem. Today 2016, 34, 72–77. [Google Scholar]
  4. Zhanikulov, N.N.; Kolesnikov, A.S.; Taimasov, B.T.; Zhakipbayev, B.Y.; Shal, A.L. Influence of industrial waste on the structure of environmentally friendly cement clinker. Complex Use Miner. Resour. 2022, 4, 84–91. [Google Scholar] [CrossRef]
  5. Zhakipbaev, B.Y.; Zhanikulov, N.N.; Kolesnikova, O.G.; Akhmetova, E.K.; Kuraev, R.M.; Shal, A.L. Review of technogenic waste and methods of its processing for the purpose of complex utilization of tailings from the enrichment of non-ferrous metal ores as a component of the raw materials mixture in the production of cement clinker. Rasayan J. Chem. 2021, 14, 997–1005. [Google Scholar]
  6. Taimasov, B.T.; Borisov, I.N.; Dzhanmuldaeva, Z.K.; Dauletiarov, M.S. Research on obtaining low energy cements from technogenic raw materials. J. Chem. Technol. Metall. 2020, 55, 814–823. [Google Scholar]
  7. Hassaan, M. Basalt rock as an alternative raw material in Portland cement manufacture. Mater. Lett. 2001, 50, 172–178. [Google Scholar] [CrossRef]
  8. Siziakova, E.; Ivanov, P.; Boikov, A. Hydrocarboaliminate calcium application in aluminum processing for production of special cement brands. J. Phys. Conf. Ser. 2018, 1118, 012018. [Google Scholar] [CrossRef]
  9. Abd El-Hafiz, N.A.; Abd El-Moghny, M.W.; El-Desoky, H.M.; Afifi, A.A. Characterization and technological behavior of basalt raw materials for Portland cement clinker production. IJISET-Int. J. Innov. Sci. Eng. Technol. 2015, 2, 1–22. [Google Scholar]
  10. Taimasov, B.T. Chemical Technology of Binding Materials; Evero: Almaty, Kazakhstan, 2015. [Google Scholar]
  11. Gorshkov, V.S.; Aleksandrov, S.E.; Ivashchenko, S.I.; Gorshkova, I.V. Complex Processing and Use of Metallurgical Slags in Construction; Stroyizdat: Moscow, Russia, 1985. [Google Scholar]
  12. Klassen, V.K.; Borisov, I.N.; Manuilov, V.E. Technogenic Materials in Cement Production, Monograph; Publishing House of BSTU: Belgorod, Russia, 2008. [Google Scholar]
  13. Taimasov, B.T.; Khudyakova, T.M.; Zhanikulov, N.N. Integrated Use of Natural and Technogenic Raw Materials in the Production of Low-Energy Cements; Auezov SKSU: Shymkent, Kazakhstan, 2017. [Google Scholar]
  14. Reed, S.D. b. Electron Probe Microanalysis and Scanning Electron Microscopy in Geology; Technosphere: Moscow, Russia, 2008. [Google Scholar]
  15. Bishimbaev, V.K.; Rusnak, V.V.; Egorov, Y.V. Mineral and Raw Material and Technological Base of the South-Kazakhstan Cluster of Building and Silicate Materials; Rarity: Almaty, Kazakhstan, 2009. [Google Scholar]
  16. Kulikova, E.Y.; Balovtsev, S.V.; Skopintseva, O.V. Complex estimation of geotechnical risks in mine and underground construction. Sustain. Develop. Mount.Territ. 2023, 15, 7–16. [Google Scholar] [CrossRef]
  17. Taimasov, B.T.; Sarsenbayev, B.K.; Khudyakova, T.M.; Kolesnikov, A.S.; Zhanikulov, N.N. Development and testing of low-energy-intensive technology of receiving sulphate-resistant and road portlandcement. Eurasian Chem.-Technol. J. 2017, 19, 347–355. [Google Scholar] [CrossRef]
  18. Kolesnikov, A.S.; Serikbaev, B.E.; Zolkin, A.L.; Kenzhibaeva, G.S.; Isaev, G.I.; Botabaev, N.E.; Shapalov, S.K.; Kolesnikova, O.G.; Iztleuov, G.M.; Suigenbayeva, A.Z.; et al. Processing of Non-Ferrous Metallurgy Waste Slag for its Complex Recovery as a Secondary Mineral Raw Material. Refract. Ind. Ceram. 2021, 62, 375–380. [Google Scholar] [CrossRef]
  19. Kolesnikova, O.; Syrlybekkyzy, S.; Fediuk, R.; Yerzhanov, A.; Nadirov, R.; Utelbayeva, A.; Agabekova, A.; Latypova, M.; Chepelyan, L.; Volokitina, I.; et al. Thermodynamic Simulation of Environmental and Population Protection by Utilization of Technogenic Tailings of Enrichment. Materials 2022, 15, 6980. [Google Scholar] [CrossRef] [PubMed]
  20. Müller, D.; Groves, D.I. Mineral Resource Reviews. In Potassic Igneous Rocks and Associated Gold-Copper Mineralization; Springer International Publishing: Cham, Switzerland, 2019; ISBN 978-3-319-92978-1. [Google Scholar]
  21. Donayev, A.; Kolesnikov, A.; Shapalov, S.; Sapargaliyeva, B.; Ivakhniyuk, G. Studies of waste from the mining and metallurgical industry, with the determination of its impact on the life of the population. News Natl. Acad. Sci. Repub. Kazakhstan Ser. Geol. Tech. Sci. 2022, 4, 55–68. [Google Scholar] [CrossRef]
  22. Utelbayeva, A.; Kolesnikov, A.; Aldiyarov, Z.; Dossybekov, S.; Esimov, E.; Duissenbekov, B.; Fediuk, R.; Vatin, N.I.; Yermakhanov, M.; Mussayeva, S. Experimental Analysis of the Stress State of a Prestressed Cylindrical Shell with Various Structural Parameters. Materials 2022, 15, 4996. [Google Scholar] [CrossRef]
  23. Kolesnikov, A.; Sapargaliyeva, B.; Bychkov, A.Y.; Alferyeva, Y.O.; Syrlybekkyzy, S.; Altybaeva, Z.; Nurshakhanova, L.; Seidaliyeva, L.; Suleimenova, B.; Zhidebayeva, A.; et al. Thermodynamic modeling of the formation of the main minerals of cement clinker and zinc fumes in the processing of toxic technogenic waste of the metallurgical industry. Rasayan J. Chem. 2022, 15, 2181–2187. [Google Scholar] [CrossRef]
  24. Baibatsha, A.B. Geology of Mineral Deposits; KazNTU: Almaty, Kazakhstan, 2008. [Google Scholar]
  25. Seignez, N.; Bulteel, D.; Damidot, D.; Gauthier, A.; Potdevin, J.L. Weathering of metallurgical slag heaps: Multi-experimental approach of the chemical behaviours of lead and zinc. WIT Trans. Ecol. Environ. 2006, 92, 31–40. [Google Scholar]
  26. Trubaev, P. Methodological Guide to the Application of the Rocs Program; V.G. Shukhov BSTU: Belgorod, Russia, 2006. [Google Scholar]
  27. Butt, Y.M.; Timashev, V.V. Workshop on Chemical Technology of Binders; Higher School: Moscow, Russia, 1973. [Google Scholar]
  28. Naizabekov, A.; Volokitina, I.; Kolesnikov, A. Changes In Microstructure And Properties Of Austenitic Steel AISi 316 during High-Pressure Torsion. J. Chem. Technol. Metallur. 2022, 57, 809–815. [Google Scholar]
  29. Kuznetsova, T.V.; Samchenko, S.V. Microscopy of Cement Production Materials; MIKHiS: Moscow, Russia, 2007. [Google Scholar]
  30. Fediuk, R. High-strength fibrous concrete of Russian Far East natural materials. IOP Conf. Ser. Mater. Sci. Eng. 2016, 116, 012020. [Google Scholar] [CrossRef]
  31. Ponzi, G.G.D.; dos Santos, V.H.J.M.; Martel, R.B.; Pontin, D.; Stepanha, A.S.D.G.E.; Schütz, M.K.; Menezes, S.C.; Einloft, S.M.; Vecchia, F.D. Basalt powder as a supplementary cementitious material in cement paste for CCS wells: Chemical and mechanical resistance of cement formulations for CO2 geological storage sites. Int. J. Greenh. Gas Control. 2021, 109, 103337. [Google Scholar] [CrossRef]
  32. Bolio-Arcero, H.; Glasser, F.P. Zinc oxide in cement clinkering: Part 1 systems CaO–ZnO–Al2O3 and CaO–ZnO–Fe2O3. Adv. Cem. Res. 1998, 25, 10. [Google Scholar]
  33. Gineys, N. Influence de la Teneuren Elements Métalliques du Clinker sur les Proprieties Techniques et Environnementales du Ciment Portland. Ph.D. Thesis, Université Lille Nord de France, Lille, France, 2011. [Google Scholar]
  34. Matusiewicz, A.; Bochenek, A.; Szelag, H.; Kurdowski, W. Pewne zagadnienia zwiazane z podwyzszona zawartoscia cynku w klinkierze i w produkowanym z niego cemencie. Cem. Wapno Beton 2011, 78, 332–341. [Google Scholar]
  35. Kolesnikov, A.; Fediuk, R.; Amran, M.; Klyuev, S.; Klyuev, A.; Volokitina, I.; Naukenova, A.; Shapalov, S.; Utelbayeva, A.; Kolesnikova, O.; et al. Modeling of Non-Ferrous Metallurgy Waste Disposal with the Production of Iron Silicides and Zinc Distillation. Materials 2022, 15, 2542. [Google Scholar] [CrossRef] [PubMed]
  36. Auyesbek, S.T.; Sarsenbayev, N.B.; Sarsenbayev, B.K.; Khudyakova, T.M.; Aimenov, Z.T.; Abdiramanova, K.S.; Aubakirova, T.S.; Sauganova, G.R.; Kolesnikova, O.G.; Karshyga, G.O.; et al. Thermal Insulating Materials Based on Magnesium-Containing Technogenic Raw Materials. Rasayan J. Chem. 2023, 16, 413–421. [Google Scholar] [CrossRef]
  37. Sergeeva, I.V.; Botabaev, N.E.; Al’Zhanova, A.Z.; Ashirbaev, K.A. Thermodynamic simulation of chemical and phase transformations in the system of oxidized manganese ore—Carbon. Izv. Ferr. Metall. 2017, 60, 759–765. [Google Scholar]
  38. Auyesbek, S.; Sarsenbayev, N.; Abduova, A.; Sarsenbayev, B.; Uderbayev, S.; Aimenov, Z.; Kenzhaliyeva, G.; Akishev, U.; Aubakirova, T.; Sauganova, G.; et al. Man-Made Raw Materials for the Production of Composite Silicate Materials Using Energy-Saving Technology. J. Compos. Sci. 2023, 7, 124. [Google Scholar] [CrossRef]
  39. Kolesnikov, A.; Fediuk, R.; Kolesnikova, O.; Zhanikulov, N.; Zhakipbayev, B.; Kuraev, R.; Akhmetova, E.; Shal, A. Processing of Waste from Enrichment with the Production of Cement Clinker and the Extraction of Zinc. Materials 2022, 15, 324. [Google Scholar] [CrossRef]
  40. Zhangabay, N.; Suleimenov, U.; Utelbayeva, A.; Kolesnikov, A.; Baibolov, K.; Imanaliyev, K.; Moldagaliyev, A.; Karshyga, G.; Duissenbekov, B.; Fediuk, R.; et al. Analysis of a Stress-Strain State of a Cylindrical Tank Wall Vertical Field Joint Zone. Buildings 2022, 12, 1445. [Google Scholar] [CrossRef]
  41. Naizabekov, A.B.; Kolesnikov, A.S.; Latypova, M.A.; Fedorova, T.D.; Mamitova, A.D. Current Trends to Obtain Metals and Alloys with Ultrafine-Grained Structure. Progr. Phys. Met. 2022, 23, 629–657. [Google Scholar]
  42. Kolesnikov, A.S.; Zhanikulov, N.N.; Syrlybekkyzy, S.; Kolesnikova, O.G.; Shal, A.L. Utilization of Waste from the Enrichment of Non-Ferrous Metal Ores as Secondary Mineral Raw Materials in the Production of Cement Clinker. Ecol. Ind. Russ. 2023, 27, 19–23. [Google Scholar]
  43. Efremova, S.V. Scientific and technical solutions to the problem of utilization of waste from plant- and mineral-based industries. Russ. J. Gen. Chem. 2012, 82, 963–968. [Google Scholar] [CrossRef]
  44. Volokitina, I.E. Evolution of the Microstructure and Mechanical Properties of Copper under ECAP with Intense Cooling. Met. Sci. Heat Treat. 2020, 62, 253–258. [Google Scholar] [CrossRef]
  45. Volokitina, I.; Vasilyeva, N.; Fediuk, R.; Kolesnikov, A. Hardening of Bimetallic Wires from Secondary Materials Used in the Construction of Power Lines. Materials 2022, 15, 3975. [Google Scholar] [CrossRef]
  46. Kolesnikov, A.; Zhanikulov, N.; Zhakipbayev, B.; Kolesnikova, O.; Kuraev, R. Thermodynamic modeling of the synthesis of the main minerals of cement clinker from technogenic raw materials. Complex Use Miner. Resour. 2021, 318, 24–34. [Google Scholar] [CrossRef]
Figure 1. X-ray of limestone.
Figure 1. X-ray of limestone.
Jcs 07 00226 g001
Figure 2. X-ray of coal mining waste.
Figure 2. X-ray of coal mining waste.
Jcs 07 00226 g002
Figure 3. Micrograph of tefritobasalts. Legend: CPx—clinopyroxene, Chl—chlorite, Cal—calcite, Mt—magnetite, Pl—plagioclase relic.
Figure 3. Micrograph of tefritobasalts. Legend: CPx—clinopyroxene, Chl—chlorite, Cal—calcite, Mt—magnetite, Pl—plagioclase relic.
Jcs 07 00226 g003
Figure 4. X-ray diffraction pattern of tefritobasalt from the Daubaba deposit.
Figure 4. X-ray diffraction pattern of tefritobasalt from the Daubaba deposit.
Jcs 07 00226 g004
Figure 5. X-ray of lead slag.
Figure 5. X-ray of lead slag.
Jcs 07 00226 g005
Figure 6. Micrographs of lead slag with various types of phenocrysts. Symbols: (a) Gal—galena, Chs—chalcocite, Mt—magnetite, (b) Chp—chalcopyrite, and Sph—sphalerite.
Figure 6. Micrographs of lead slag with various types of phenocrysts. Symbols: (a) Gal—galena, Chs—chalcocite, Mt—magnetite, (b) Chp—chalcopyrite, and Sph—sphalerite.
Jcs 07 00226 g006
Figure 7. General view of the furnace (a), unfired pellets (b), and clinkers (c), obtained at a temperature of 1350 °C.
Figure 7. General view of the furnace (a), unfired pellets (b), and clinkers (c), obtained at a temperature of 1350 °C.
Jcs 07 00226 g007
Figure 8. Dependence of the content of alite (C3S) in the clinkers on the value of modules (n and p) and the saturation coefficient in the raw mixtures.
Figure 8. Dependence of the content of alite (C3S) in the clinkers on the value of modules (n and p) and the saturation coefficient in the raw mixtures.
Jcs 07 00226 g008
Figure 9. Radiographs of the clinkers.
Figure 9. Radiographs of the clinkers.
Jcs 07 00226 g009
Figure 10. Micrographs of the clinker (a)—there was an aluminate phase a dark intermediate substance, (b)—and an aluminoferrite phase—a light intermediate substance.
Figure 10. Micrographs of the clinker (a)—there was an aluminate phase a dark intermediate substance, (b)—and an aluminoferrite phase—a light intermediate substance.
Jcs 07 00226 g010
Table 1. Chemical composition of raw materials, waste, and igneous rocks.
Table 1. Chemical composition of raw materials, waste, and igneous rocks.
NameThe Chemical Composition, mas. %
SiO2Al2O3Fe2O3CaOMgOSO3Na2OK2OTiO2ClCr2O3ZnOPbOCuOLoss of IgnitionOtherTotal
Limestone3.871.040.5752.830.880.10-0.120.0180.020.012---40.200.37100.0
Tefrito
basalt
45.5410.708.5310.666.950.204.042.500.910.0170.007---5.374.57100.0
Coal mining waste55.5010.602.013.210.700.79-2.350.38-----24.080.38100.0
Lead slag25.946.4437.2514.716.150.041.241.360.350.0060.0634.340.520.940.65-100.0
Table 2. Results of calculating the composition of the raw batch and the resulting clinker.
Table 2. Results of calculating the composition of the raw batch and the resulting clinker.
MaterialsSiO2Al2O3Fe2O3CaOMgOSO3L.W.COther
Limestone Sastobe3.871.040.5752.830.880.1040.71-
Tefritobasalt Daubaba45.5410.708.5310.666.950.207.929.50
Coal mining waste from the Lenger mines55.5010.602.013.210.700.7924.083.11
Lead slag of Yuzhpolimetall JSC plant25.946.4437.2514.716.150.040.109.37
By component chemical composition of the raw batchComponent content
kg/kg clinker%
Limestone2.910.780.4339.780.660.07530.65-1.116075.31%
Tefritobasalt9.202.161.722.151.410.0401.611.920.157510.63%
Coal mining waste0.580.110.020.040.010.0080.250.0330.157510.63%
Lead slag0.880.221.270.510.210.0010.0030.3210.05083.43%
Raw mix13.593.283.4542.482.290.1332.522.271.4818100.00%
By component chemical composition of the clinkerComponent content
Limestone4.321.160.6458.950.980.112--66.17%
Tefritobasalt13.633.202.553.192.080.060-2.84514.38%
Coal mining waste0.860.170.030.050.010.012-0.04914.39%
Lead slag1.310.331.890.750.310.002-0.4765.06%
clinker20.144.865.1162.953.390.19-3.37100.00%
Chemical composition of raw mix and clinkerSCnpTEC. (kcal/kg)Gfuel. kg ref.fuel/t cl)
Raw mix13.593.283.4542.482.290.1332.522.270.942.022.02--
clinker20.144.865.1162.953.390.19-3.370.940.950.95364.1196
Mineralogical composition of the clinker
Minerals3CaO·SiO22CaO·SiO23CaO·Al2O34CaO·Al2O3·Fe2O3CaSO4MgO
wt.%57.8818.826.4611.610.323.39
Symbols: C3S—alite (3CaO·SiO2), C2S—belite (2CaO·SiO2), C3A—tricalcium aluminate (3CaO·Al2O3), C4AF—tetracalcium aluminoferrite (4CaO·Al2O3·Fe2O3), LOI—loss on ignition.
Table 3. Specific consumption of raw materials and the amount of free CaO in the clinker.
Table 3. Specific consumption of raw materials and the amount of free CaO in the clinker.
MixComposition of the Raw Batch, wt.%Specific Consumption of the Raw Materials, t/t of ClinkerSCModulesAmount of Free CaO at 1350 °C, %
Lime StoneTefrito BasaltCoal Mining WasteLead SlagLime StoneTefrito BasaltCoal Mining WasteLead Slagnp
166.1714.3814.395.061.11600.15750.15750.05080.942.020.950.2
Table 4. Physicomechanical properties of experimental Portland cement.
Table 4. Physicomechanical properties of experimental Portland cement.
CementSieve Residue,%Specific Surface Area, cm2/gStrength of Small Samples 2 × 2 × 2 cm, (MPa)
No. 02No. 0087 Day28 Day
Cement3.313324527.441.8
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhanikulov, N.; Sapargaliyeva, B.; Agabekova, A.; Alfereva, Y.; Baidibekova, A.; Syrlybekkyzy, S.; Nurshakhanova, L.; Nurbayeva, F.; Sabyrbaeva, G.; Zhatkanbayev, Y.; et al. Studies of Utilization of Technogenic Raw Materials in the Synthesis of Cement Clinker from It and Further Production of Portland Cement. J. Compos. Sci. 2023, 7, 226. https://doi.org/10.3390/jcs7060226

AMA Style

Zhanikulov N, Sapargaliyeva B, Agabekova A, Alfereva Y, Baidibekova A, Syrlybekkyzy S, Nurshakhanova L, Nurbayeva F, Sabyrbaeva G, Zhatkanbayev Y, et al. Studies of Utilization of Technogenic Raw Materials in the Synthesis of Cement Clinker from It and Further Production of Portland Cement. Journal of Composites Science. 2023; 7(6):226. https://doi.org/10.3390/jcs7060226

Chicago/Turabian Style

Zhanikulov, Nurgali, Bayan Sapargaliyeva, Aktolkyn Agabekova, Yana Alfereva, Aidin Baidibekova, Samal Syrlybekkyzy, Lazzat Nurshakhanova, Farida Nurbayeva, Gulzhan Sabyrbaeva, Yergazy Zhatkanbayev, and et al. 2023. "Studies of Utilization of Technogenic Raw Materials in the Synthesis of Cement Clinker from It and Further Production of Portland Cement" Journal of Composites Science 7, no. 6: 226. https://doi.org/10.3390/jcs7060226

Article Metrics

Back to TopTop