Open access peer-reviewed chapter

Simulation and Optimization of an Integrated Process Flow Sheet for Cement Production

Written By

Oluwafemi M. Fadayini, Adekunle A. Obisanya, Gloria O. Ajiboye, Clement Madu, Tajudeen O. Ipaye, Taiwo O. Rabiu, Shola J. Ajayi and Joseph T. Akintola

Submitted: 17 September 2020 Reviewed: 27 November 2020 Published: 09 March 2021

DOI: 10.5772/intechopen.95269

From the Edited Volume

Cement Industry - Optimization, Characterization and Sustainable Application

Edited by Hosam El-Din Mostafa Saleh

Chapter metrics overview

798 Chapter Downloads

View Full Metrics

Abstract

In this study the process flow diagram for the cement production was simulated using Aspen HYSYS 8.8 software to achieve high energy optimization and optimum cement flow rate by varying the flow rate of calcium oxide and silica in the clinker feed. Central composite Design (C.C.D) of Response Surface Methodology was used to design the ten experiments for the simulation using Design Expert 10.0.3. Energy efficiency optimization is also carried out using Aspen Energy Analyser. The optimum cement flow rate is found from the contour plot and 3D surface plot to be 47.239 tonnes/day at CaO flow rate of 152.346 tonnes/day and the SiO2 flow rate of 56.8241 tonnes/day. The R2 value of 0.9356 determined from the statistical analysis shows a good significance of the model. The overall utilities in terms of energy are found to be optimised by 81.4% from 6.511 x 107 kcal/h actual value of 1.211 x 107 kcal/h with 297.4 tonnes/day the carbon emission savings.

Keywords

  • central composite design
  • optimisation
  • response surface methodology
  • cement production
  • design expert

1. Introduction

Cement is a fine greyish or whitish inorganic, non-metallic powder commonly used as a binding agent in construction materials. It consists of pyroprocessed chemically combined hydraulic cement materials such as calcareous, siliceous, argillaceous and ferriferous [1]. Cement forms paste when mixed with water, which later becomes hard due to cement, mineral hydrate formation when solidified [2]. The various types of cement and their applications such as Portland, Siliceous fly ash, calcareous, slag and Fume silica cement differ by the amount of SiO2, Al2O3, Fe2O3, CaO, MgO, SO3 and other materials such as Na2O and K2O composition [3]. Economic growth and urbanisation have made cement one of the most consumed commodity in world with global annual production increase from 3.3 Gt in 2010 to current 4.1 Gt which is still expected to grow moderately in the next decade due to expected infrastructure development in India and other developing Asian and African countries [4, 5]. Cement production consists of three sections: fuel and raw material processing, production of clinker via pyroprocessing and grinding and blending of cement clinker nodules with additive materials such as gypsum and anhydrite for different types of cement types [6]. Natural occurring limestone is ground and mixed in required proportion with silicon and aluminium source such as clay and sand and iron-containing compounds to form a homogenous raw mix called raw meal. The raw meal is then pyroproccessed at a high temperature of about 1450 °C in rotary kiln system where it is dried, preheated, calcined and sintered into cement clinker. The pyroprocessing can be dry, wet, semi-dry or semi-wet and their selection depends on the moisture content of the raw meal, rotary kiln configuration and energy cost. The wet process is cheaper with a high-quality product but very high energy intensity because of the high moisture content of about 36% in raw meal. The dry process is usually more compact with low operational cost and energy consumption compared with the wet process but with lesser product homogeneity [2]. The clinker produced is further grinded with about 5% gypsum which prevents pre-set and controls the hydration rate of the cement. Other types of cement are produced by blending with hydraulic, pozzolanic or inert materials [7]. Cement production processes are energy-intensive and generate huge greenhouse emissions with the clinker energy intensity of about 3.4 GJ/t in 2018 [4] generating about 4% of the global CO2 emission [8]. Strategies identified to reduce the emissions in cement production include improving heat recovery and energy efficiency [9, 10, 11], switching to low carbon source of energy [12], feedstock and material substitute [13, 14, 15], reducing the clinker-to-cement ratio [16] and advancing technology innovations such as carbon capture and storage [17, 18]. Cement and concrete technology modelling and simulation have also been used to improve energy efficiency and usage [19, 20, 21, 22].

Optimization is a mathematical technique used to find the best solution to objective function (s) by maximising the desired variable and minimising the undesired variables under some set of constraints with the sole aim of improving performance and cost [23]. The optimisation technique in cement and concrete studies can be broadly classified as a meta-heuristic approach and statistical experimental design methods [24]. The meta-heuristic approach is an iterative method that intelligently exploits search space at learning strategies. It includes Genetic Algorithm (GAs), Particle swarm optimization (PSO), Harmony Search (HS), Ant Colony Optimization (ACO), Charged System Search (CSS), Big Bang-Big Crunch (BB-BC), Artificial Bee Colony algorithm (ABC), spherical interpolation of the objective function, Colliding bodies optimization (CBO), Vibrating Particles System (VPS), simulated annealing, krill herd (KH), Whale Optimization Algorithm (WOA), hybrid Harmony Search, force method and genetic algorithm, mine and improved mine blast algorithms [24, 25, 26] which are modelled from natural and social behaviours as well as physics laws.

Statistical experimental design methods are widely used to obtain desired optimise solution for a set of constraints [27]. Response Surface Methodology (RSM) is a statistical optimisation technique used to model and analyse a process to determine the effect of independent multivariable on the process response and to evaluate the relations between these variables [28]. RSM is based on understanding the topography of the response surface to determine the most appropriate response region [29]. RSM experimental design can be categorised into Box–Behnken Design (BBD), Central Composite Design (CCD), Dohlert design, Mixture response and three-level factorial design [30, 31, 32]. The BBD is created from 3 level factorial design [32] and gives quadratic response model with three minimum number of factors requiring three levels of factors (upper, centre, lower) for each factor and specific positioning of design points [33, 34]. The CCD is developed from the 2 factorial design and gives the quadratic response model with five levels for each factor. Hence it is more robust and insensitive to missing data or experimental runs [34].

In recent years, Response Surface Methodology has been applied to optimise several chemical processes such as extraction [29], adsorption [23], pharmaceutical wastewater treatment [28], leaching [35]. Studies on cement production optimisation have been carried out on clinker simulation using AspenTech [36], cement raw materials blending using a general nonlinear time-varying model [37], cement grinding using population balance model [6], clinker chemistry and kiln energy efficiency using metaheuristic optimization techniques [38], numerical and computational fluid dynamics study of cement calciner [16]. RSM has been efficient and accurate in studies on cement and concrete technology [39, 40, 41, 42, 43]. This study focused on the simulation of an integrated wet cement process flow sheet using Aspen HYSYS and optimisation of the cement production rate at minimum raw material feed using CCD of response surface methodology.

Advertisement

2. Methodology

2.1 Cement production simulation

Aspen HYSYS 8.8 was used for the steady-state simulation of the integrated process flow sheet for the cement production [44]. Within the simulation environment, topological optimization (proper arrangement of equipment) was done to enable very high energy savings or optimization. A pure component such as water, CO2 and air are added as conventional components, while non-conventional components are added as hypothetical components to the HYSYS environment based on their physical properties (molecular weight & density). Figure 1 shows the block diagram of the production of cement while the HYSYS process flow diagram for the cement production simulation is shown in Figure A1 in the appendix. Limestone is decomposed in the first reactor to give off CO2 as gas, while the produced CaO is the feed to the Section A reactor to react with silicate to form Calcium disilicate. The produced Calcium disilicate reacts further with the unreacted CaO in reactor B to produce Calcium trisilicate. Calcium oxide (CaO) further reacts with Aluminium oxide to produce tricalcium aluminate, another constituent of cement, while the final product component is produced in section C, where CaO reacts with Aluminate and Ferric oxide to produce tetracalciumaluminoferrite. These separate components produced at a different section of the simulated Kiln are mixed to achieve a matrix compound of the cement product, having over 70% of CaO.

Figure 1.

Process flow diagram for the production of cement.

2.2 Aspen Hysys simulation

Aspen Hysys was used for the steady-state simulation of the integrated process flow sheet for the cement production. Within the simulation environment, topological optimization (proper arrangement of equipment) was done to enable very high energy savings or optimization. A pure component such as water, CO2 and air are added as conventional components, while non-conventional components are added as hypothetical components to the HYSYS environment based on their physical properties (molecular weight and density). Based on the process description, the different reactions taking place in each simulated reactor, as presented in the flowchart are:

CaCO3CaO+CO2Limestone decompositionE1
2CaO+SiO2Ca2SiO4SectionAE2
CaO+Ca2SiO4Ca3SiO5SectionBE3
3CaO+Al2O3Ca3Al2O6SectionCE4
4CaO+Al2O3+Fe2O3Ca4Al2Fe2O10SectionDE5

The various products in the various sections of the process reactors are; Tricalcium silicate (Ca2SiO4) which is responsible for early strength and the initial set of the cement; Dicalcium silicate (Ca3SiO5) which increases the strength as it age; Tricalcium aluminate (Ca3Al2O6) which contributes to the concrete strength development in the first few days but least desirable due to its reactiveness with sulphate containing soils and water; Tetracaliumaluminoferrite (Ca4Al2Fe2O10) which reduces clinkering temperature. The equipment design parameters employed in this work are provided in Table 1.

Simple separatorDelta PStream fractions
0.000Solids in vapour0.0100
Solids in liquid0.0100
Liquid in bottoms0.0100
ExchangerDelta P (bar)Delta T (°C)Duty (kcal/h)
Cooler0.0006552.691 x0 107
Component SplitterTop Temp. (°C)Bottom Temp. (°C)Top Pressure (bar)Bottom Pressure (bar)Duty
Splitter 13030112.054 x 105
Splitter 212521252111.825 x 104
Splitter 312521252111.923 x 104
ReactorsDelta PVessel volume (m3)Liquid level (%)Liquid volume (m3)Duty (kcal/h)
CaCO3 decomp reactor0.00000.000050.0025.002.6764 x 107
Section A reactor0.000050.0050.0025.007.6209 x 106
Section B reactor0.000050.0050.0025.002.6131 x 106
Section C reactor0.000050.0050.0025.004.5634 x 105
Section D reactor0.000050.0050.0025.007.4721 x 105

Table 1.

Equipment design parameter.

The flow rate of the major raw materials for the production of cement in the clinkering reactor as depicted by Eqns. (6–12) are carefully chosen based on the standard provided by Winter N. B. [45]. The Chemical parameters based on the oxide composition are very useful in describing clinker characteristics. The following parameters are widely used.

  1. Lime Saturation Factor (LSF): is the measure of the ratio of alite to belite in the clinker. It is estimated by the ratio of CaO to the sum of other three main oxides SiO2, Fe2O3 and Al2O3. The equation is given by:

LSF=CaO2.8SiO2+1.2Al2O3+0.65Fe2O3E6
LSF=1902.860+1.215+0.6510=0.98E7

LSF values in clinkers range between 0.92–0.98. The LSF value of 0.98 falls within an acceptable range.

  1. Silica Ratio (SR): This is also known as Silica Modulus. The expression of SR is given as:

SR=SiO2Al2O3+Fe2O3E8

Based on the experimental design for the simulated cement production process. The flow rate of Al2O3 and Fe2O3 are 15 tonnes/day and 10 tonnes/day respectively. The low level and high levels of SiO2 are found to be 50 tonnes/day and 60 tonnes/day respectively. Hence, the SR values are the high and low value of the SiO2 flow rate are calculated as follows:

SR=5015+10=2.0E9
SR=6015+10=2.4E10

A high silicate ratio means that more calcium silicates are present in the clinker and less aluminate and ferrite. SR is typical, between 2.0 and 3.0. The SR values of 2.0 and 2.4 fall within an acceptable range of 2.0 and 3.0.

  1. Aluminate Ratio (AR): This is the ratio of aluminate and ferrite phases in the clinker. AR value ranges between 1–4 in Portland clinkers. The flow rate of Al2O3 and Fe2O3 used in the process simulation are 15 tonnes/day and 10 tonnes/day respectively. The equation governing the AR of the oxide is given by

AR=Al2O3Fe2O3E11
AR=1510=1.5E12

The mass flow and corresponding clinker quality parameters are presented in Table 2.

Mass flow (tonnes/day)Clinker quality parameter
CaOSiO2Al2O3Fe2O3LSFSRAR
High level1906015100.982.41.5
Low level1355015100.822.01.5

Table 2.

Raw material mass flow and clinker quality parameter.

2.3 Multivariate design of experiment

The central composite design of response surface methodology was used to analyse the effect of CaO and SiO2 on cement production rate. The total number of experimental runs (N) required for n independent variables and nc number of replica centre points is given by Eq. 13

N=2n+2n+ncE13

Design Expert 10.0.3 software was used to generate the experimental design from the ten experimental runs to study the combined effect of two variables on the response. For two variables factor in the experiments; four factorial points (2n), four axial points (2n) and two replicates at the central points (nc) at distance α = 1.414 from the centre were used for the CCD design. A polynomial empirical model was developed from the ten experimental runs to correlate the response with the independent variables. The mathematical expression can be expressed as:

Y=βo+i=1nβiXi+i=1nβiiXi2+1ijnβijXiXj+εE14

Where Y = Predicted response, βo = constant coefficient, βi = Linear coefficient, βii = Quadratic, βij = Interaction coefficients and ε = model random error, n the number of variable factors, Xi and Xj are the coded values of the variable parameters [35].

The response generated function distance from the centre α=2n/4. The codes are calculated as a function of the range the factors as shown in Table 3.

CodeMathematical relationship
Lower axial point- αXmin
Lower level- 1Xmax+Xmin/2αXmaxXmin/2
Centre point0Xmax+Xmin/2
Upper level1Xmax+Xmin/2+αXmaxXmin/2
Upper axial pointαXmax

Table 3.

Relationship between the variable values and their assigned codes.

The central composite experimental design for the synthesis of cement via simulation is depicted in Table 4. The mass flow rate of CaO and SiO2 measured in tonnes/day are the independent variables or predictors which are studied for their effect on the response variable (cement flow rate) at a constant Al2O3 and Fe2O3 flow rates.

FactorNameUnitsLowHigh
AFlow rate of CaOtonnes/day135190123.609201.391
BFlow rate of SiO2tonnes/day506047.928962.0711

Table 4.

Design of Experiment using central composite (C.C.D) Design of Response Surface Methodology (R.S.M).

2.4 Model fitting and statistical analysis

The interaction between the variables and the response data as well as the statistical parameters were analysed graphically by analysis of variance (ANOVA) in the Design-Expert software. Regression analysis, significance, F-test, surface and contour plots of the response were also generated from the software. A probability value of 95% confidence level was used to evaluate the significance of the model terms and coefficients.

Advertisement

3. Results and discussion

3.1 Simulation and optimisation of cement flow

All simulations were done in duplicate and the experimental design were generated by the Central Composite Design (C.C.D) of the Design-Expert Software, which resulted in a total of 10 experimental (simulation) runs and the results of the experiments (simulations) are shown in Table 5.

Run No.FactorsResponse
Coded levelsActual valuesThe flow rate of cement
ABABHYSYS SimulationC.C.D Model
11.0001.0001906046.645.6674
2−1.000−1.0001355030.7431.9851
30.0000.000162.55546.0446.04
40.000−1.414162.547.928930.5731.271
51.000−1.0001905044.342.1551
6−1.4140.000123.6095546.0443.4867
70.0000.000162.55546.0446.04
8−1.0001.0001356046.649.0574
91.4140.000201.3915546.0448.2808
100.0001.414162.562.071146.8445.8265

Table 5.

Simulation and predicted results from central composite design (C.C.D).

The change in mean response in cement flow per unit increase in variable occurs when other predictors area kept constant and is estimated by the coefficient of estimation and is presented Table 6.

CoefficientStandard95% CI95% CI
FactorEstimatedfErrorLowHighVIF
Intercept46.0410.8644.0048.08
A-Flowrate of CaO1.7010.680.0803.311.00
B-Flow rate of Silica5.1510.683.536.761.00
AB−3.3910.97−5.67−1.111.00
A2−0.07810.73−1.811.651.02
B2−3.7510.73−5.48−2.011.02

Table 6.

Coefficient estimation for cement flow rate in terms of coded factors.

The empirical quadratic equation for the optimal cement product rate as a function of CaO and SiO2 mass flow in coded form as derived from Table 7 was obtained according to the CCD and is given in Eq. 15

SourceSum of squaresdfMeanF valuep-value (Prob > F)
Model379.61575.9220.350.0005
A22.98122.986.160.0421
B211.861211.8656.780.0001
AB45.97145.9712.320.0099
A20.04210.0420.0110.9180
B297.60197.6026.160.0014
Residual26.1273.73
Lack of Fit26.1238.71
Pure Error0.00040.000
Cor. Total405.7312

Table 7.

ANOVA results for the statistical model for cement flow rate.

C=46.04+1.70A+5.15B3.39AB0.078A23.75B2E15

The test of the significance and the adequacy of the model and its coefficients lack fitness which was based on F-value or P-value at 95% confidence level was tested from analysis of variance (ANOVA) and the result is presented in Table 7. The result shows that the model at an F value of 20.35 and a very low P-value of 0.0005 indicates that the statistical regression model was significant. The result also shows that A, B, ABand B2 are significant terms.

The regression statistical analysis is summarised in Table 8. The R squared value of 0.9356 is in good agreement with the adjusted R-square value of 0.8896, showing a good fit of the model, as the closer the R squared value to 1.00, the more significant the model. The adequate precision of 13.55 indicates the low noise level and a strong signal for optimisation. Hence, this indicates the two predictors (flow rate of CaO and flow rate of SiO2) could predict the flow rate of cement, thus the model equation, contour plot and 3D surface plot could be used to predict the response (flow rate of the cement).

ParameterValues
R-Squared0.9356
Adj R-Squared0.8896
Pred R-Squared0.5422
Mean43.69
Std. Dev.1.93
C.V. %4.42
Adeq Precision13.554

Table 8.

Statistical information for the statistical model for cement flow rate.

3.2 Surface plots

The contour plot which shows the possible relationship between the CaO, SiO2 and cement product mass flow is presented in Figure 2. The darker red regions indicate higher C (response) values. Here, the optimum flow rate of cement is found from the isolines to be 47.748 tonnes/day at a flow rate of 152.346 tonnes/day of CaO and 56.8241 tonnes/day SiO2.

Figure 2.

Contour surface plot showing the effects of the flow rate of CaO and flow rate of SiO2 on the flow rate of cement.

The three-dimensional (3D) response surface plots obtained from the model equation using Design Expert 10.03 is depicted in Figure 3. This depicts the effect of the flow rate of CaO, the flow rate of SiO2 on the flow rate of cement. The flow rate of cement was observed to increase with an increase in the flow rate of CaO. Conversely, increasing the flow rate of SiO2 did not increase the flow rate of cement. Hence, the major predictor in cement production in the clinkering section is the flow rate of CaO.

Figure 3.

Response surface plot of the effects of CaO and SiO2 mass flow on cement production.

The optimization plot of the cement output is shown in Figure 4. The optimum cement flow rate of 47.239 tonnes/day is found to be at CaO flow rate of 152.346 tonnes/day and SiO2 flow rate of 56.8241 tonnes/day.

Figure 4.

Optimization plot showing the effects of the flow rate of CaO and flow rate of SiO2 on the flow rate of cement.

3.3 Energy optimization

Aspen Energy Analyser was used to determine the percentage of energy savings based on converged steady-state simulation of the process flow sheet in Figure 1. The total energy savings as a function of process utilities and carbon emissions are present in Figure 5.

Figure 5.

Energy savings at optimum feed (CaO and SiO2) rate and product (cement) flow rate.

The overall utilities in terms of energy are found to be optimised from the actual value of 6.511 x 107 kcal/h to 1.211 x 107 kcal/h and indicating available energy savings of 5.3 x 107 kcal/h, with overall energy savings of 81.40% which also correspond to 297.4 tonnes/day carbon emission reduction.

Advertisement

4. Conclusion

Process flow diagram for the cement production was simulated to achieve high energy optimization and optimum cement flow rate by minimising the flow rate of the feed (CaO and SiO2). Central composite Design (C.C.D) of Response Surface Methodology used to design the experiment for the simulation using Design Expert 10.0.3. The optimum cement flow rate is found from surface and contour plots to be 47.239 tonnes/day at CaO flow rate of 152.346 tonnes/day and SiO2 flow rate of 56.8241 tonnes/day. The R squared value of 0.9356 determined from the statistical analysis shows a very high significance of the model. Energy efficiency optimization is also carried out using Aspen Energy Analyser. The overall utilities in terms of energy are found to be optimised by 81.4 % from 6.511 x 107 kcal/hactual value to 1.211 x 107 kcal/h with 297.4 tonnes/day the carbon emission savings.

Further work could be performed on fault identification and diagnosis of the process plant. Incorporated with an automated plant to guarantee the safety of workers, reduce environmental problems and increase yield to sustain production improvement.

Advertisement

5. Recommendations

This research work sought to recommend the following concerns in which cement production could be improved:

  1. Research and development (R&D) in the cement production, processing and utilisation should be encouraged. This will play a vital role in the construction industry, operation and maintenance of efficient road network and effective transportation system.

  2. Automation of cement and kiln sections of the cement production is recommended

  3. Optimization of the cement production can be tailored into the fabrication of high tech cement equipment and parts.

  4. Optimization of the limestone crusher to quantify the amount of crushed limestone is needed.

  5. Looking into future the results obtained in this research will open up several important possibilities in the cement production at optimum conditions. This will have a multiplier effect on infrastructural amenities development.

Advertisement

Conflict of interest

There is no conflict of interest associated with this work.

Advertisement

Figure A1.

HYSYS process flow simulation diagram for the production of cement.

References

  1. 1. EPA, “Mineral Products Industry,” in Compilation of air pollutant emission factors. Volume I: stationary point and area sources, AP-42 5th., Research Triangle Park, NC: U.S. Environmental protection agency, 1995
  2. 2. N. Martin, E. Worrell, and L. Price, “Energy Efficiency and Carbon Dioxide Emissions Reduction Opportunities in the U.S. Cement Industry,” Berkeley, 1999. doi: 10.2172/751775
  3. 3. S. P. Dunuweera and R. M. G. Rajapakse, “Cement Types, Composition, Uses and Advantages of Nanocement, Environmental Impact on Cement Production, and Possible Solutions,” Adv. Mater. Sci. Eng., vol. 2018, 2018, doi: 10.1155/2018/4158682
  4. 4. IEA, “Cement,” Paris, 2020. [Online]. Available: https://www.iea.org/reports/cement
  5. 5. TERI, “CEMENT INDUSTRY; Trends Report,” New Delhi, 2017. [Online]. Available: http://www.teriin.org/library/files/Cement-Industry-Trends-Report2017.pdf
  6. 6. A. Jankovic, W. Valery, and E. Davis, “Cement grinding optimisation,” Miner. Eng., vol. 17, no. 11–12, pp. 1075–1081, 2004, doi: 10.1016/j.mineng.2004.06.031
  7. 7. D. Olsen, S. Goli, D. Faulkner, and A. Mckane, “Opportunities for Energy Efficiency and Demand Response in the California Cement Industry,” no. December. Lawrence Berkeley National Laboratory, 2010
  8. 8. J. G. J. Olivier and J. A. H. W. Peters, “Trends in Global CO2 and Total Greenhouse Gas Emissions: Report 2019,” The Hague, 2020. [Online]. Available: www.pbl.nl/en
  9. 9. H. Mikulčić, M. Vujanović, N. Markovska, R. V. Filkoski, M. Ban, and N. Duić, “CO2 emission reduction in the cement industry,” Chem. Eng. Trans., vol. 35, pp. 703–708, 2013, doi: 10.3303/CET1335117
  10. 10. S. Zhang, H. Ren, W. Zhou, Y. Yu, and C. Chen, “Assessing air pollution abatement co-benefits of energy efficiency improvement in cement industry: A city-level analysis,” J. Clean. Prod., vol. 185, pp. 761–771, 2018, doi: 10.1016/j.jclepro.2018.02.293
  11. 11. M. J. S. Zuberi and M. K. Patel, “Bottom-up analysis of energy efficiency improvement and CO2 emission reduction potentials in the Swiss cement industry,” J. Clean. Prod., vol. 142, pp. 4294–4309, 2017, doi: 10.1016/j.jclepro.2016.11.178
  12. 12. N. Chatziaras, C. S. Psomopoulos, and N. J. Themelis, “Use of waste-derived fuels in cement industry: a review,” Manag. Environ. Qual. An Int. J., vol. 27, no. 2, pp. 178–193, 2016, doi: 10.1108/MEQ-01-2015-0012
  13. 13. E. Marchetti, “Use of Agricultural Wastes as Supplementary Cementitious Materials,” KTH ROYAL INSTITUTE OF TECHNOLOGY, 2020
  14. 14. A. Naqi and J. G. Jang, “Recent progress in green cement technology utilizing low-carbon emission fuels and raw materials: A review,” Sustain., vol. 11, no. 2, 2019, doi: 10.3390/su11020537
  15. 15. R. Maddalena, J. J. Roberts, and A. Hamilton, “Can Portland cement be replaced by low-carbon alternative materials? A study on the thermal properties and carbon emissions of innovative cement,” J. Clean. Prod., vol. 186, no. April, pp. 933–942, 2018, doi: 10.1016/j.jclepro.2018.02.138
  16. 16. H. Mikulcic, M. Vujanovic, and N. Duic, “Improving the Sustainability of Cement Production by Using Numerical Simulation of Limestone Thermal Degradation and Pulverized Coal Combustion in a Cement Calciner,” J. Clean. Prod., vol. 88, pp. 262–271, 2015
  17. 17. P. Markewitz et al., “Carbon capture for CO2 emission reduction in the cement industry in Germany,” Energies, vol. 12, no. 12, pp. 1–27, 2019, doi: 10.3390/en12122432
  18. 18. WSP Parson Brinkerhoff and DNV GL, “Industrial Decarbonisation & Energy Efficiency Roadmaps to 2050: Cement,” 2015
  19. 19. C. Csernyei and A. G. Straatman, “Numerical modelling of a rotary cement kiln with improvements to shell cooling,” Int. J. Heat Mass Transf., vol. 102, pp. 610–621, 2016, doi: 10.1016/j.ijheatmasstransfer.2016.06.058
  20. 20. H. R. Goshayeshi and F. K. Poor, “Modeling of Rotary Cement Kiln (In Persian),” Energy Power Eng., vol. 8, pp. 23–33, 2016
  21. 21. S. Sadighi, M. Shirvani, and A. Ahmad, “Rotary cement kiln coating estimator: Integrated modelling of the kiln with shell temperature measurement,” Can. J. Chem. Eng., vol. 89, no. 1, pp. 116–125, Feb. 2011, doi: 10.1002/cjce.20365
  22. 22. K. V. Sabarish, M. Akish Remo, and P. Paul, “Optimizing the Concrete Materials by Taguchi Optimization Method,” IOP Conf. Ser. Mater. Sci. Eng., vol. 574, no. 1, 2019, doi: 10.1088/1757-899X/574/1/012002
  23. 23. S. Saini, J. Chawla, R. Kumar, and I. Kaur, “Response surface methodology (RSM) for optimization of cadmium ions adsorption using - C 16–6 - 16 incorporated mesoporous MCM - 41,” SN Appl. Sci., vol. 1, no. 8, pp. 1–10, 2019, doi: 10.1007/s42452-019-0922-5
  24. 24. C. TAO, “Optimization of Cement Production and Hydration for Improved,” UNIVERSITY OF FLORIDA, 2017
  25. 25. F. Omidinasab and V. Goodarzimehr, “A hybrid particle swarm optimization and genetic algorithm for truss structures with discrete variables,” J. Appl. Comput. Mech., vol. 6, no. 3, pp. 593–604, 2020, doi: 10.22055/JACM.2019.28992.1531
  26. 26. L. J. Li, Z. B. Huang, and F. Liu, “A heuristic particle swarm optimization method for truss structures with discrete variables,” Comput. Struct., vol. 87, no. 7–8, pp. 435–443, Apr. 2009, doi: 10.1016/j.compstruc.2009.01.004
  27. 27. M. J. Simon, “Concrete Mixture Optimization Using Statistical Methods: Final Report,” 2003. [Online]. Available: http://www.fhwa.dot.gov/publications/research/infrastructure/pavements/03060/03060.pdf
  28. 28. S. Barisci and O. Turkay, “Optimization and modelling using the response surface methodology (RSM) for ciprofloxacin removal by electrocoagulation,” Water Sci. Technol., vol. 73, no. 7, pp. 1673–1679, 2016, doi: 10.2166/wst.2015.649
  29. 29. A. Y. Aydar, “Utilization of Response Surface Methodology in Optimization of Extraction of Plant Materials,” in Statistical Approaches With Emphasis on Design of Experiments Applied to Chemical Processes, V. Silva, Ed. InTech, 2018, pp. 157–168
  30. 30. S. A. Adeyeye, “Banana Drying Kinetics,” in Banana Nutrition - Function and Processing Kinetics, IntechOpen, 2020, pp. 0–20
  31. 31. M. N. Chollom, S. Rathilal, F. M. Swalaha, B. F. Bakare, and E. K. Tetteh, “Comparison of response surface methods for the optimization of an up-flow anaerobic sludge blanket for the treatment of slaughterhouse wastewater,” Environ. Eng. Res., vol. 25, no. 1, pp. 114–122, 2020, doi: 10.4491/eer.2018.366
  32. 32. A. I. Khuri, “Response Surface Methodology and Its Applications In Agricultural and Food Sciences,” Biometrics Biostat. Int. J., vol. 5, no. 5, pp. 155–163, 2017, doi: 10.15406/bbij.2017.05.00141
  33. 33. B. Olawoye, A comprehensive handout on central composite design (CCD), no. July. 2016
  34. 34. Shari, “Choosing the Best Design for Process Optimization,” 2017. https://www.statease.com/blog/choosing-best-design-process-optimization/ (accessed Nov. 27, 2020)
  35. 35. S. Kumar, H. Meena, S. Chakraborty, and B. C. Meikap, “International Journal of Mining Science and Technology Application of response surface methodology ( RSM ) for optimization of leaching parameters for ash reduction from low-grade coal,” Int. J. Min. Sci. Technol., vol. 28, no. 4, pp. 621–629, 2018, doi: 10.1016/j.ijmst.2018.04.014
  36. 36. B. Hokfors, M. Eriksson, and E. Viggh, “Modelling the cement process and cement clinker quality,” Adv. Cem. Res., vol. 26, no. 6, pp. 311–318, 2014, doi: 10.1680/adcr.13.00050
  37. 37. X. Li, H. Yu, and M. Yuan, “Modeling and Optimization of Cement Raw Materials Blending Process,” Math. Probl. Eng., vol. 2012, 2012, doi: 10.1155/2012/392197
  38. 38. C. Tao, “Optimization of cement production and hydration for improved performance, energy conservation, and cost,” UNIVERSITY OF FLORIDA, 2017
  39. 39. G. Cibilakshmi and J. Jegan, “ A DOE approach to optimize the strength properties of concrete incorporated with different ratios of PVA fibre and nano-Fe 2 O 3,” Adv. Compos. Lett., vol. 29, p. 2633366X2091388, 2020, doi: 10.1177/2633366x20913882
  40. 40. T. F. Awolusi, O. L. Oke, O. O. Akinkurolere, and O. D. Atoyebi, “Comparison of response surface methodology and hybrid-training approach of artificial neural network in modelling the properties of concrete containing steel fibre extracted from waste tyres,” Cogent Eng., vol. 6, no. 1, 2019, doi: 10.1080/23311916.2019.1649852
  41. 41. M. Sonebi and M. T. Bassuoni, “Investigating the effect of mixture design parameters on pervious concrete by statistical modelling,” Constr. Build. Mater., vol. 38, pp. 147–154, Jan. 2013, doi: 10.1016/j.conbuildmat.2012.07.044
  42. 42. K. E. Alyamac, E. Ghafari, and R. Ince, “Development of eco-efficient self-compacting concrete with waste marble powder using the response surface method,” J. Clean. Prod., vol. 144, pp. 192–202, Feb. 2017, doi: 10.1016/j.jclepro.2016.12.156
  43. 43. L. Soto-Pérez, V. López, and S. S. Hwang, “Response Surface Methodology to optimize the cement paste mix design: Time-dependent contribution of fly ash and nano-iron oxide as admixtures,” Mater. Des., vol. 86, pp. 22–29, Dec. 2015, doi: 10.1016/j.matdes.2015.07.049
  44. 44. R. C. Carpio, L. D. S. Coelho, R. J. Silva, and A. B. Jorge, “Case Study in Cement Kilns Alternative Secondary Fuels Mixing Using Sequential Quadratic Programming, Genetic Algorithms, and Differential Evolution,” 2005, doi: 10.1590/S1678-58782008000400010
  45. 45. N. Winter, “Understanding Cement,” 2005. https://www.understanding-cement.com/clinker.html (accessed Nov. 02, 2020)

Written By

Oluwafemi M. Fadayini, Adekunle A. Obisanya, Gloria O. Ajiboye, Clement Madu, Tajudeen O. Ipaye, Taiwo O. Rabiu, Shola J. Ajayi and Joseph T. Akintola

Submitted: 17 September 2020 Reviewed: 27 November 2020 Published: 09 March 2021