Next Article in Journal
A Transmission Price Design Considering the Marginal Benefits of the Transmission and Spatiotemporal Information of Electricity Demand
Previous Article in Journal
Energy Resilience in Telecommunication Networks: A Comprehensive Review of Strategies and Challenges
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Thermal Characterization, Kinetic Analysis and Co-Combustion of Sewage Sludge Coupled with High Ash Ekibastuz Coal

1
Chemical and Materials Engineering, School of Engineering and Digital Sciences, Nazarbayev University, Astana 010000, Kazakhstan
2
Mechanical and Aerospace Engineering, School of Engineering and Digital Sciences, Nazarbayev University, Astana 010000, Kazakhstan
*
Authors to whom correspondence should be addressed.
Energies 2023, 16(18), 6634; https://doi.org/10.3390/en16186634
Submission received: 8 August 2023 / Revised: 24 August 2023 / Accepted: 11 September 2023 / Published: 15 September 2023
(This article belongs to the Section J: Thermal Management)

Abstract

:
Efficient utilization of natural resources and possible valorization of solid waste materials such as sewage sludge into secondary materials via thermal conversion and simultaneously recovering energy is vital for sustainable development. The continuous increase in metropolises leads to an enormous production of wet sewage sludge, which creates major environmental and technical issues. In this paper, the samples of sewage sludge from Astana’s waste water treatment plant are analyzed for their thermochemical properties, followed by thermogravimetric and kinetic analysis using the Flynn–Wall–Ozawa and Kissinger–Akahira–Sunose methods. Overall, the calorific value of sewage sludge sample was 18.87 MJ/kg and was comparable to that of the bituminous coal samples. The activation energy varied from 140 to 410 kJ/mol with changing conversion from 0.1 to 0.7. Further, mono-combustion and co-combustion experiments of the sewage sludge with high ash bituminous coal were conducted using the laboratory scale bubbling fluidized bed rig, respectively. The difference in NOx emissions between mono-combustion of sewage sludge and co-combustion with coal were at around 150 ppm, while this value for SO2 was similar in average, but fluctuates between 150 and 350 ppm. Overall, the findings of this study will be useful in developing a co-combustion technology for a sustainable disposal of municipal sewage sludge.

1. Introduction

Renewable energy is currently undergoing active development through various approaches, including georesources’ development [1], electrification of transport, advanced energy storage technologies [2], and the development of solar panels [3]. Alternative source to recover energy is the municipal sewage sludge (SS), via various thermochemical methods [4]. SS is the residual solid waste produced during the treatment of wastewater [4]. SS typically consists of organic and inorganic compounds, along with toxic and hazardous substances [5]. SS is derived from wastewater treatment plants (WWTP) and thus has a high moisture content, which leads to the accumulation of large volumes and subsequently high operational costs during treatment [6]. It is reported that the handling and disposal of SS roughly takes 50% of the operational cost of secondary wastewater treatment plants in Europe [7]. The growing population and the associated rate of urbanization have led to an increase in sewage sludge volumes. For example, considering local demography, the population of Kazakhstan is expected to reach around 23.5 million by 2050 [8], and thus high volumes of SS are produced. The high amount of SS can cause serious environmental issues if proper waste management strategies are not planned, and hence developing sustainable treatment solutions is imperative.
An Increase in population also leads to a higher demand for energy resources, which brings additional sources of GHG emissions. The energy requirements of Kazakhstan are mainly covered by coal-fired power plants (~70% of the total energy demand) [9,10], followed by gas, hydro, and renewable energy sources. Meanwhile, traditional coal-fired power plants may pose serious risks and be subject to accidents, including those involved in handling coal [5,11,12]. In the context of current work, valorization of available sewage sludge with minimum GHG emissions can help the country in mitigating GHG emissions and reduce coal usage [13]. In addition, Kazakhstan has initiated a new decarburization strategy to reach carbon neutrality by 2060 [14]. Hereby, a sustainable use of the waste for energy recovery is going to be a part of the ongoing strategy to reduce GHG emissions.
At present, there are 180 WWTPs under operation in Kazakhstan [15]. Most of the WWTPs in the country use mechanical and biological treatment coupled with tertiary treatment with sand filtration and ultraviolet (UV) disinfection [15]. In 2016, 583 million tons of wastewater was treated with roughly 52 million tons of sewage sludge produced in Kazakhstan [15,16]. The WWTP at the capital city produces 300–350 tons of sewage sludge per day, with an initial moisture content of around 98%. The generated sewage sludge passes through centrifugal dryers after which the moisture is reduced to an approximately 70%, mixed with a 0.1% aqueous solution of polyacrylamide to promote the flocculation, and stored in open drying beds, followed by disposal in the sanitary landfills [17]. Alternatively, the disposal of SS via thermal methods, particularly incineration, are widely practiced in European countries like the Netherlands, Switzerland, Germany, Belgium, Austria, and Slovenia [18].
Among the various thermal treatment methods, combustion has received a greater amount of attention mainly due to better conversion efficiency for solid wastes. Combustion reduces high volumes, and disposes of organic matter, while simultaneously recovering the energy from the waste [19]. In addition, the combustion of SS alleviates bad odor, and can thermally wipe out some of the toxic components [20]. However, SS has low calorific value and inefficient combustion performance. Hence, blending with higher calorific solid fuels is proposed [21,22]. Furthermore, fluidized bed combustion of sewage sludge has advantages due to the high fuel flexibility, lower net CO2, and control of SO2, NOx emissions [23,24,25]. On the contrary, one of the main disadvantages of SS combustion is its high content of nitrogen, which leads to the higher formation of NOx, including NO + NO2 and N2O [25].
Thermal characterization of the blended SS has been widely analyzed in the literature via thermogravimetric and kinetics studies. Liu et al. [26], for example, performed kinetic analysis for a pure SS using two isoconversional methods, Flynn–Wall–Ozawa (FWO) and Kissinger–Akahira–Sunose (KAS), under oxy-fuel and air environments [26]. According to the results of Liu et al. [26], SS demonstrated more complex combustion behavior than the coal/hydrochar samples, yielding a maximum activation energy of 184.14 kJ/mol at a conversion rate of 0.2. Likewise, the activation energy of high ash sewage sludge combustion was analyzed by Naqvi et al. [27], who observed an increase in activation energy with the conversion, although there were some fluctuations [27]. The authors reported that the initial increase in the activation energy was due to the presence of cellulosic compounds, while the intermediate porous nature of SS served as a reason for a drop in the activation energy at high conversion [27]. More recently, the co-combustion of coffee residues and SS was investigated using thermogravimetric Fourier transform infrared spectroscopy (TG-FTIR) and the results showed a better combustion performance for the blends; however, high amounts of toxic gases were released [28]. The authors reported that during the initial combustion rates (<0.5), the apparent activation energy of blends was greater than that of SS due to the increase in organic compounds and release of volatiles. However, at a high conversion rate, the activation energy decreases because of the gradual expansion of the pores, making the reaction fast. While in most of the studies, the kinetic values were determined by using the model-free (iso-conversional) method, the projected thermal decomposition may not align with the real decomposition in SS due to complexity of structure [29]. Therefore, the disagreement can be adjusted by the modeling of the projected results at three different heating rates and with two isoconversional methods, KAS and FWO.
Furthermore, several experiments are available in the literature on the co-combustion of SS with different types of coals in various reactors as summarized in Table 1. The results of the literature review indicate that the experiments were performed in similar experimental conditions, and the temperature used in the experiment in this work is within the range presented in the literature. In addition, the literature studies demonstrate that the proportions of SS and coal in the feed affect the emissions, and this information was applied in the analysis of the experimental results.
To the best of our knowledge, there is a lack of data in the literature on SS and its co-combustion behavior with high ash bituminous coal similar to those produced in Kazakhstan. To this end, this work presents the mono and co-combustion of sewage sludge. In addition, thermogravimetric (TGA) analysis was used to investigate thermal degradation and the combustion behavior. The weight loss of material, kinetics, and combustion characteristics were precisely revealed via TGA. This analysis facilitated the comprehension of the synergistic effects of co-combustion of coal and of sewage sludge. Furthermore, flue gas measurements in terms of SO2, NOx and CO2 from mono-combustion of SS and co-combustion of SS and blended with coal with proportion of 75%/25% in a laboratory scale bubbling fluidized bed (BFB) unit were presented.

2. Materials and Methods

2.1. Sample Preparation

Samples of SS were collected from the WWTP located in Astana, the capital city of Kazakhstan. The samples were preheated in a muffle furnace (Carbolite, Neuhausen, Germany) at 105 °C for 24 h to remove the excess moisture and deactivate potential bacteria and microbes, as has been suggested in the literature [36,37]. Despite pre-drying, which consequently resulted in a significant weight loss due to the moisture evaporation, the remaining SS fraction was unsuitable for the batch feeding. As seen in Figure 1A, it had a slurry-like consistency with non-uniform particle sizes. Therefore, a hand press tool was used to make uniform pellets. The final shape of the SS pellet is demonstrated in Figure 1B. The mass of the individual pellets varied in the range of 0.55 ± 0.05 g. The bituminous high ash coal from the North of Kazakhstan (Ekibastuz) was used as a supplementary fuel during co-combustion experiments and was pelletized along with SS. Details thermal and chemical properties of the Ekibastuz coal can be found elsewhere [38].

2.2. Thermochemical Properties of the SS and Coal

The thermal properties of SS, in general, are non-uniform [23,39] and, hence, characterization is essential for the local sludge. Standards including ISO 18134-3:2023, ISO 18122:2022, ISO 18123:2023were used to determine proximate analysis and were performed in the muffle furnace [40,41,42]. Elemental analysis was performed using a UNICUBE® (Paris, France) micro elemental analyzer. Calibrated bomb calorimeter (Etalon®, Almaty, Kazakhstan) was used to determine the calorific values. These analyses were repeated three times and the average values were reported and the standard deviation was less than 1%. Table 2 shows the thermal properties of SS in comparison with the high ash bituminous coal used in the experiments [43]. As seen the SS samples have low fixed carbon but ~three times higher volatile matter than the coal. In addition, the ash content of SS was around 32%. Ultimate analysis demonstrated that oxygen, sulfur, and nitrogen presence in SS, considering its organic origin, were also approximately three times higher than the bituminous coal, whereas the carbon content was similar. The HHV value for the SS samples was 18.87 MJ/kg and was comparable to coal samples.
Prior to the batch combustion of SS pellets in the bubbling fluidized bed rig, a series of thermogravimetric analyses were conducted to explore the weight loss behavior of the samples. Dry and powdered SS and its blends with coal particles were used in Simultaneous Thermal Analyzer (STA) 6000 between 30 and 900 °C, with a heating rate of 15 °C/min and constant air flow rate. Furthermore, the kinetic analysis of TGA data for SS samples were performed using the free-model iso-conversional methods based on the 10, 15, 20 °C/min heating rate in the nitrogen environment. The proximate and ultimate analyses performed demonstrate similar results of Kazakhstan’s SS compared to those of China’s SS (range of C = 30.9–40.5%; H = 5.06–5.8%; N = 5.36–6.84%; S = 0.61–1.28%; O = 12.3–19.7%) [44] and those of Thailand’s SS (range of C = 9.0–31.1%; H = 2.0–4.6%; N = 1.5–4.3%; S = 0.7–1.8%; O = 15.7–27.5%) [45]. The variations can be explained by heterogeneous compositions of sewage sludge samples, atmospheric conditions, and different moisture content in case of the proximate analysis.

2.3. Experimental Apparatus and Test Conditions for Combustion

A laboratory scale BFB rig was used for the combustion experiments. The schematic illustration of the rig was shown in Figure 2. The rig was heated electrically with a three-zone heater (Nabertherm) to maintain the furnace temperature. The BFB rig consists of two vertically aligned stainless steel tubes with diameters of outer tube being 89 mm and inner tube being 48 mm. Other elements of the rig include a fuel feeding point at the top, pressure drop manometers, gas supply cylinders/flowmeters, filters to collect fly ash, and a flue gas analyzer. Inert sand was placed as the bed material, which was held by the two-layer stainless steel meshes. The gas supply system was split into primary and secondary streams, connected to the bottom of the rig and to the feeding point, respectively. The gas inlet was controlled by mass flow meters (Bronckhorst). Mechanical batch feeding was organized from the top of the rig and was supported by pneumatic transport, if necessary, based on the fuel density. Combustion products at the exit line of the furnace were filtered from fly ash particles and measured in real-time mode using the calibrated Gasmet Dx 4000 FTIR (Vantaa, Finland).
Operating conditions of the lab-scale BFB rig were summarized in Table 3. Combustion tests were conducted at the atmospheric pressure and the bed temperature was kept at 850 °C for all tests. Primary air supply was kept constant at 3.0 L/m (at STP) to ensure the superficial gas velocity being above the minimum fluidization velocity. The pressure drop between the bottom and the top of the reactor was measured using a U-tube manometer and was around 2.05 kPa.
Experimental results on flue gas analysis presented in this work reported the maximum peak values among each batch for NOx, SO2 (in ppm), CO2 and O2 (in vol%). The completion of the batch pellet combustion was monitored using stabilization of the measured gases and recovery of oxygen in the flue gas, which took approximately 5 min per batch.

3. Results and Discussion

3.1. Thermogravimetric Analysis of the SS and the Blends with Coal

Figure 3A,B shows the thermogravimetric analysis (TGA)/derivative gravimetry (DTG) curves of SS and blends with bituminous coal samples at different weight ratio: (a) pure SS, (b) 25%/75% (SS/coal wt%), (c) 50%/50% (SS/coal wt%) and (d) 75%/25% (SS/coal wt%). The final weights of the samples tested in the TGA varied between 30% and 34%, which was in good agreement with the ash content obtained during the proximate analysis. As seen from Figure 3A, no moisture release was seen until 100 °C for all samples in the thermograms, as the samples were pre-dried and stored in the air tight packs. A further increase in the temperature from 200 to 600 °C leads to a sharp weight loss curve for SS samples, followed by the blended ratio of 25 wt% SS, 50 wt% SS and 75 wt% SS. Chen et al. [46] has recently highlighted three main stages of SS combustion during TGA. The first stage of decomposition occurs in the temperature range of 100–250 °C that corresponds to the release of volatiles [46]. In our experiments, the first peak of decomposition for SS was observed at 190 °C, while for blended tests, the peak shifted to higher temperature due to the presence of coal and its low volatile content. The peak was observed at a temperature of 220 °C and 230 °C for blend ratio 75 wt% and 50 wt% of SS, respectively. There was no sharp peak observed for the blend ratio of 25 wt% SS in which coal was predominant, because of the low volatile content of the coal.
The second stage corresponds to the decomposition of hemicellulose and cellulose, which contributes to a notable weight loss [47] as also seen in Figure 3B. In addition, the high rate of weight loss can be explained by the reaction of volatiles and other reactive structures in the presence of air [30]. The second stage follows immediately after the main oxidation and has a small peak at ~420 °C. Moreover, due to the continuous weight loss, it can be noted that the occurrence of other reactions with more complex compounds. Some of these reactions can involve interactions with heavy metals, as most of the organic volatiles were decomposed during the main oxidation process. Overall, oxidation is the longest stage with more than 60% of weight loss between 200 and 600 °C. Figure 3A further showed a relatively small weight loss at temperatures higher than 800 °C. According to Chen et al. [48], this last stage corresponds to the burnout stage. Magdziarz and Wilk [49] suggested that the burnout happens due to the small part of the carbon left in the sample [49]. However, Chen et al. [48] argued that the burnout stage corresponds to the decomposition of the remaining CaO, as reported to be 18.9 wt%, within the sludge as has been proposed by Folgueras et al. [30]. Regression was performed for the data presented in Figure 3 and is available in the Supplementary Material (Figure S1, Equations (S1)–(S4)).

3.2. Kinetic Analysis of SS Samples

Next, the kinetic analysis of the sewage sludge was carried out using three different heating rates and in nitrogen environment. The rate of solid phase reactions for heterogeneous sewage sludge samples can be expressed as follows:
d α d t = k ( t ) f ( a ) = A e x p ( E R T ) ( 1 α )  
where /dt is the rate of conversion, k(t) is the rate of constant, f(α) depends on the reaction model, α is the fractional conversion, A is the pre-exponential Arrhenius factor (1/s), E is the apparent activation energy, R is the universal gas constant, T is the reaction temperature (K). Herein, the iso-conversional method with the FWO and KAS models was used to calculate the activation energy from experimental data on TGA analysis of sewage sludge at 10, 15, and 20 °C/min [50]. The result for TGA in the nitrogen environment is presented in the Supplementary Material (Figure S1). The KAS model obeys the following expression:
l n ( β T 2 ) = l n [ A E R g ( α ) ] E R T  
where β is the heating rate (K/min), A is the pre-exponential factor (1/min), α is the conversion factor, and g(α) is a decomposition mechanism function. The FWO model [51] employs a correlation between the sample heating rate, activation energy, and inverse temperature, which was first approximated using Doyle’s method [52] for the temperature integral. The equation is as follows:
l o g ( β ) = l o g [ A E R g ( α ) ] 2.315 0.457 E R T
Using Equations (2) and (3), the activation energy of pure SS at conversion factors from 0.1 to 0.7 was evaluated, as shown in Figure 4a (FWO) and Figure 4b (KAS). The average apparent activation energies determined via FWO and KAS are 236.9 kJ/mol and 239.4 kJ/mol, respectively, with R2 above 0.9. The deviation of 1.04% in mean activation energies showed a good agreement between the KAS and FWO methods.
A relationship between the activation energy and a conversion factor was depicted in Figure 5. As observed from Figure 5, an activation energy shows an increase with a conversion factor, particularly at a conversion of 0.6~0.7. According to Liu et al. [26], the maximum activation energy was determined to be 184.14 kJ/mol at the conversion value of 0.2, and the reason for this difference could be different environments: 30% O2 with 70% CO2, and 100% N2 in this study. The activation energy depends on several parameters including activated molecule concentration, diffusion limitation, and organic contaminants [53]. Syed-Hassan et al. [54] reported that SS is a complex mixture of organic and inorganic compounds, as well as diverse live and dead microbes [54]. Proteins, carbohydrates, and lipids are the significant components of sludge and correspond to 76–97% of the sludge’s volatiles [55]. Up to 60% of weight loss accounts for the volatile release and organic fraction combustion, while the sharp increase in activation energy from 0.6 to 0.7 corresponds to the degradation of organic compounds in sludge composition. The correlation coefficients were shown in Figure S3, accounting for almost the same R2 values for both models.

3.3. Exit Gas Analysis

Lastly, to examine the combustion characteristics of the SS blend fuel, the combustion of SS-bituminous briquettes was performed. A total of 40 batch combustion tests carried out done with pellets of pure SS and 20 with a blend ratio of SS/coal 75%/25% in the BFB. Figure 6a shows the concentrations (maximum concentration for CO2 and minimum for O2 during the batch run) of O2 and CO2 in the exit gas line during the series of batch feeding. As can be observed, the lowest concentrations of O2 were observed between 1.6 and 4.0 vol%, which ensured a complete combustion of the dropped pellets in the BFB rig. It can be seen from Figure 6a that the general trend of O2 peaks demonstrated a slight fluctuation with an average concentration being around 2.5%. These fluctuations in the consumption of O2 peaks can be explained by the heterogeneous nature of sewage sludge.
Meanwhile, the CO2 concentration was obtained within the range of 5 to 8%, with an average value of 6.5 vol.% between the 40 batch experiments that were carried out. Figure 6b further illustrates the highest detected concentrations of SO2 and NOx during the mono-combustion of SS samples. Peak value for SO2 concentration was determined to be 347 ppm, while for NOx, the value was 442 ppm. Among different batches, the trend showed a slight increase for NOx, while SO2 emissions did not demonstrate any significant deviations. The average values and standard deviations were found for the data presented in Figure 6 and are available in the Supplementary Material (Figures S4 and S5).
The combustion of blends with 75 wt% SS was further carried out for another 20 batches after the mono-combustion of SS. The results for emissions in terms of NOx and SO2 gases are presented in Figure 7a,b. As can be seen in Figure 7a, NOx emissions highlighted a sharp decrease in concentrations under co-combustion mode. The trend was in line with the amount of nitrogen available in coal as compared to the pure SS. Likewise, Figure 7b shows the SO2 concentrations in the blends, which fluctuated between 150 and 350 ppm, with an average value remaining similar to those in the case of mono-combustion of SS. A possible explanation for the fluctuation could be the interconversion of SO2 to H2SO3 in the presence of coal, within the stipulated time of batch feeding; however, further analysis is required to explore that fluctuation. Average values and standard deviations were found for the data presented in Figure 7 and is available in the Supplementary Material (Figures S6 and S7). Table 4 summarizes the results based on mean and standard deviation values based on 60 experimental data points for each gas mentioned above.

4. Conclusions

This work presents a systematic study on the evaluation of the thermochemical properties of local sewage sludge and investigates the mono and co-combustion of SS and with high ash Ekibastuz coal in the TGA and on a laboratory-scale BFB rig. The thermo-chemical results revealed a high potential of the pre-dried SS for the combustion and in blends with the coal. The measured calorific value of SS (18.87 MJ/kg) was similar to those for the bituminous high ash coal (19.40 MJ/kg), while the fixed carbon content was significantly lower (2.69 wt% to 38.90 wt%). TGA and DTG analyses highlight a sensible weight loss of SS during the devolatilization process. The kinetic analysis of Astana’s SS samples showed comparable results available in the literature. Lastly, co-combustion in the BFB and the analysis of the flue gas revealed low CO2 emissions and comparable SO2 and NOx emissions. The combustion of the pure SS resulted in 6.81 ± 0.6 vol% max CO2, 356 ± 26 ppm max NOx, 247 ± 39 ppm max SO2, while that of the blend (75% SS/25% coal) in 2.2 ± 0.34 vol% max CO2, 204 ± 47 ppm max NOx, 249 ± 71 ppm max SO2. Further experiments at varying operating conditions are required to optimize the SO2 and NOx emissions. Moreover, future research is necessary to broaden the scope of co-combustion studies covering different types of coal beyond Ekibastuz coal. An exploration of the following operational factors is needed: temperature, fuel mixture ratios, and residence time, might further help improve combustion efficiency and emissions profiles. The results can be used in consideration of the implementation of SS as fuel at local incineration sites.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/en16186634/s1, Figure S1: Regression analysis for the TGA curves for SS and bituminous coal at different blend ratios at heating of 15 °C/min in air; Figure S2: The thermographs of sewage sludge as obtained from the TGA experiments in nitrogen environment at (A) 10 (B) 15 (C) 20 °C/min; Figure S3: The exponential fit for both FWO and KAS models; Figure S4: Analysis for O2 and CO2 concentrations emissions during the mono-combustion of SS in BFB rig; Figure S5: Analysis for SO2 and NOx concentrations emissions during the mono-combustion of SS in BFB rig; Figure S6: Analysis for SO2 emissions; Figure S7: Analysis for NOx emissions.

Author Contributions

Conceptualization, D.S. and Y.S.; Methodology, D.Z.; Validation, D.Z. and K.Z.; Formal analysis, A.T.; Investigation, M.A. and A.T.; Writing—original draft, M.A., D.Z. and A.T.; Writing—review & editing, D.S. and Y.S.; Visualization, K.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Ministry of Education and Science of the Republic of Kazakhstan, Project number IRN: AP14872091 (Project name: Comparative assessment of municipal sewage sludge pyrolysis, gasification and incineration in the pilot-scale fluidized bed rig) and the APC was funded by Nazarbayev University research grant: 11022021FD2905 (Project name: Efficient thermal valorization of municipal sewage sludge in fluidized bed systems: Advanced experiments with process modeling).

Data Availability Statement

The data presented in this study are available per request from the corresponding author.

Acknowledgments

The authors gratefully acknowledge the support provided by the duty technologists at Wastewater Treatment Plant of Astana city (Astana Sy Arnasy) for preparation of samples.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kongar-Syuryun, C.B.; Aleksakhin, A.V.; Eliseeva, E.N.; Zhaglovskaya, A.V.; Klyuev, R.V.; Petrusevich, D.A. Modern Technologies Providing a Full Cycle of Geo-Resources Development. Resources 2023, 12, 50. [Google Scholar] [CrossRef]
  2. Shchurov, N.I.; Dedov, S.I.; Malozyomov, B.V.; Shtang, A.A.; Martyushev, N.V.; Klyuev, R.V.; Andriashin, S.N. Degradation of Lithium-Ion Batteries in an Electric Transport Complex. Energies 2021, 14, 8072. [Google Scholar] [CrossRef]
  3. Malozyomov, B.V.; Martyushev, N.V.; Voitovich, E.V.; Kononenko, R.V.; Konyukhov, V.Y.; Tynchenko, V.; Kukartsev, V.A.; Tynchenko, Y.A. Designing the Optimal Configuration of a Small Power System for Autonomous Power Supply of Weather Station Equipment. Energies 2023, 16, 5046. [Google Scholar] [CrossRef]
  4. Di Giacomo, G.; Romano, P. Evolution and Prospects in Managing Sewage Sludge Resulting from Municipal Wastewater Purification. Energies 2022, 15, 5633. [Google Scholar] [CrossRef]
  5. Zhang, X.; Wang, X.; Wang, D. Immobilization of Heavy Metals in Sewage Sludge during Land Application Process in China: A Review. Sustainability 2017, 9, 2020. [Google Scholar] [CrossRef]
  6. Quan, L.M.; Kamyab, H.; Yuzir, A.; Ashokkumar, V.; Hosseini, S.E.; Balasubramanian, B.; Kirpichnikova, I. Review of the application of gasification and combustion technology and waste-to-energy technologies in sewage sludge treatment. Fuel 2022, 316, 123199. [Google Scholar] [CrossRef]
  7. Kacprzak, M.; Neczaj, E.; Fijałkowski, K.; Grobelak, A.; Grosser, A.; Worwag, M.; Rorat, A.; Brattebo, H.; Ålmas, A.; Singh, B.R. Sewage sludge disposal strategies for sustainable development. Environ. Res. 2017, 156, 39–46. [Google Scholar] [CrossRef]
  8. UNFPA, We, Kazakhstan. Population Situation Analysis of the Republic of Kazakhstan. 2019, Ministry of National Economy of the Republic of Kazakhstan Committee on Statistics. p. 65. Available online: https://kazakhstan.unfpa.org/sites/default/files/pub-pdf/07_FEB_UNFPA_Report_20pager_ENG_PREVIEW%20%281%29_0.pdf (accessed on 10 September 2023).
  9. KAZENERGY, The National Energy Report 2019. Available online: https://www.kazenergy.com/upload/document/energy-report/NationalReport19_en.pdf (accessed on 10 September 2023).
  10. Sarbassov, Y.; Kerimbay, A.; Tokmurzin, D.; Tosato, G.; De Miglio, R. Electricity and heating system in Kazakhstan: Exploring energy efficiency improvement paths. Energy Policy 2013, 60, 431–444. [Google Scholar] [CrossRef]
  11. Bosikov, I.I.; Martyushev, N.V.; Klyuev, R.V.; Savchenko, I.A.; Kukartsev, V.V.; Kukartsev, V.V.; Tynchenko, Y.A. Modeling and Complex Analysis of the Topology Parameters of Ventilation Networks When Ensuring Fire Safety While Developing Coal and Gas Deposits. Fire 2023, 6, 95. [Google Scholar] [CrossRef]
  12. Malozyomov, B.V.; Golik, V.I.; Brigida, V.; Kukartsev, V.V.; Tynchenko, Y.A.; Boyko, A.A.; Tynchenko, S.V. Substantiation of Drilling Parameters for Undermined Drainage Boreholes for Increasing Methane Production from Unconventional Coal-Gas Collectors. Energies 2023, 16, 4276. [Google Scholar] [CrossRef]
  13. UNFCCC, Fourth Biennial Report of the Republic of Kazakhstan to the UN Framework Convention on Climate Change. 2019. Available online: https://unfccc.int/sites/default/files/resource/BR4_en.pdf (accessed on 10 September 2023).
  14. Partnership for Action on Green Economy, Kazakhstan Finalizes the Carbon Neutrality Strategy until 2060, UN PAGE—Partnership for Action on Green Economy. 2022. Available online: https://www.un-page.org/news/kazakhstan-finalizes-the-carbon-neutrality-strategy-until-2060/ (accessed on 10 September 2023).
  15. Druz, N. First Step: Opportunities to Reduce Methane Emissions in the Wastewater Treatment Sector in Kazakhstan [Russian]. 2018. Available online: https://nur.nu.edu.kz/bitstream/handle/123456789/6123/Book%20%28ru%29.pdf?sequence=2&isAllowed=y (accessed on 10 September 2023).
  16. Zhumabayev, D.; Bakdolotov, A.; De Miglio, R.; Litvak, V.; Baibakisheva, A.; Sarbassov, Y.; Baigarin, K. Kazakhstan’s Road to Net Zero GHG Emissions, ed. G. Tosato. 2022. Available online: https://nur.nu.edu.kz/bitstream/handle/123456789/6123/Book%20%28eng%29.pdf?sequence=1&isAllowed=y (accessed on 10 September 2023).
  17. Ospanov, K.; Myrzahmetov, M.; Zapparov, M. Study of the Products of Pyrolysis Recycling. Sewage Sludge in the Aeration Station Almaty, Kazakhstan. Procedia Eng. 2015, 117, 288–295. [Google Scholar] [CrossRef]
  18. Eurostat, Sewage Sludge Production and Disposal from Urbanwastewater (in Dry Substance (d.s)). 2020. Available online: http://ec.europa.eu/eurostat/tgm/refreshTableAction.do?tab=table&plugin=1&pcode=ten00030&language=en (accessed on 10 September 2023).
  19. lincoln.ne.gov, Solid Waste Plan 2040. 2012. Available online: https://www.lincoln.ne.gov/City/Departments/LTU/Utilities/Solid-Waste-Management/SWMP (accessed on 10 September 2023).
  20. Magdziarz, A.; Kosowska-Golachowska, M.; Kijo-Kleczkowska, A.; Środa, K.; Wolski, K.; Richter, D.; Musiał, T. Analysis of sewage sludge ashes from air and oxy-fuel combustion in a circulating fluidized-bed. In Proceedings of the 1st International Conference on the Sustainable Energy and Environment Development (SEED 2016), Kraków, Poland, 17–19 May 2016; Kosowska-Golachowska, M., Ed.; [Google Scholar] [CrossRef]
  21. Otero, M.; Calvo, L.F.; Gil, M.V.; García, A.I.; Morán, A. Co-combustion of different sewage sludge and coal: A non-isothermal thermogravimetric kinetic analysis. Bioresour. Technol. 2008, 99, 6311–6319. [Google Scholar] [CrossRef]
  22. Nadziakiewicz, J.; Kozioł, M. Co-combustion of sludge with coal. Appl. Energy 2003, 75, 239–248. [Google Scholar] [CrossRef]
  23. Fonts, I.; Azuara, M.; Gea, G.; Murillo, M.B. Study of the pyrolysis liquids obtained from different sewage sludge. J. Anal. Appl. Pyrolysis 2009, 85, 184–191. [Google Scholar] [CrossRef]
  24. Žnidarčič, A.; Katrašnik, T.; Zsély, I.G.; Nagy, T.; Seljak, T. Sewage sludge combustion model with reduced chemical kinetics mechanisms. Energy Convers. Manag. 2021, 236, 114073. [Google Scholar] [CrossRef]
  25. Moško, J.; Pohořelý, M.; Zach, B.; Svoboda, K.; Durda, T.; Jeremiáš, M.; Šyc, M.; Václavková, S.; Skoblia, S.; Beňo, Z.; et al. Fluidized Bed Incineration of Sewage Sludge in O2/N2 and O2/CO2 Atmospheres. Energy Fuels 2018, 32, 2355–2365. [Google Scholar] [CrossRef]
  26. Liu, T.; Lang, Q.; Xia, Y.; Chen, Z.; Li, D.; Ma, J.; Gai, C.; Liu, Z. Combination of hydrothermal carbonization and oxy-fuel combustion process for sewage sludge treatment: Combustion characteristics and kinetics analysis. Fuel 2019, 242, 265–276. [Google Scholar] [CrossRef]
  27. Naqvi, S.R.; Tariq, R.; Hameed, Z.; Ali, I.; Taqvim, S.A.; Naqvi, M.; Niazi, M.B.K.; Noor, T.; Farooq, W. Pyrolysis of high-ash sewage sludge: Thermo-kinetic study using TGA and artificial neural networks. Fuel 2018, 233, 529–538. [Google Scholar] [CrossRef]
  28. Ni, Z.; Bi, H.; Jiang, C.; Sun, H.; Zhou, W.; Tian, J.; Lin, Q. Investigation of co-combustion of sewage sludge and coffee industry residue by TG-FTIR and machine learning methods. Fuel 2022, 309, 122082. [Google Scholar] [CrossRef]
  29. Sunphorka, S.; Chalermsinsuwan, B.; Piumsomboon, P. Artificial neural network model for the prediction of kinetic parameters of biomass pyrolysis from its constituents. Fuel 2017, 193, 142–158. [Google Scholar] [CrossRef]
  30. Folgueras, M.; Díaz, R.; Xiberta, J.; Prieto, I. Thermogravimetric analysis of the co-combustion of coal and sewage sludge. Fuel 2003, 82, 2051–2055. [Google Scholar] [CrossRef]
  31. Shimizu, T.; Toyono, M.; Ohsawa, H. Emissions of NOx and N2O during co-combustion of dried sewage sludge with coal in a bubbling fluidized bed combustor. Fuel 2007, 86, 957–964. [Google Scholar] [CrossRef]
  32. Kowarska, B.; Baron, J.; Kandefer, S.; Zukowski, W. Incineration of Municipal Sewage Sludge in a Fluidized Bed Reactor. Engineering 2013, 5, 125–134. [Google Scholar] [CrossRef]
  33. Han, X.; Niu, M.; Jiang, X.; Liu, J. Combustion Characteristics of Sewage Sludge in a Fluidized Bed. Ind. Eng. Chem. Res. 2012, 51, 10565–10570. [Google Scholar] [CrossRef]
  34. Leckner, B.; Åmand, L.; Lücke, K.; Werther, J. Gaseous emissions from co-combustion of sewage sludge and coal/wood in a fluidized bed. Fuel 2004, 83, 477–486. [Google Scholar] [CrossRef]
  35. Åmand, L.; Leckner, B. Metal emissions from co-combustion of sewage sludge and coal/wood in fluidized bed. Fuel 2004, 83, 1803–1821. [Google Scholar] [CrossRef]
  36. Soria-Verdugo, A.; Kauppinen, J.; Soini, T.; García-Gutiérrez, L.M.; Pikkarainen, T. Pollutant emissions released during sewage sludge combustion in a bubbling fluidized bed reactor. J. Waste Manag. 2020, 105, 27–38. [Google Scholar] [CrossRef]
  37. Gao, N.; Kamran, K.; Quan, C.; Williams, P.T. Thermochemical conversion of sewage sludge: A critical review. Prog. Energy Combust. Sci. 2020, 79, 100843. [Google Scholar] [CrossRef]
  38. Suleimenova, B.; Aimbetov, B.; Shah, D.; Anthony, E.J.; Sarbassov, Y. Attrition of high ash Ekibastuz coal in a bench scale fluidized bed rig under O2/N2 and O2/CO2 environments. Fuel Process. Technol. 2021, 216, 106775. [Google Scholar] [CrossRef]
  39. Fedorov, A.; Dubinin, Y.V.; Yeletsky, P.; Fedorov, I.A.; Shelest, S.N.; Yakovlev, V. Combustion of sewage sludge in a fluidized bed of catalyst: ASPEN PLUS model. J. Hazard. Mater. 2021, 405, 124196. [Google Scholar] [CrossRef] [PubMed]
  40. ISO 18134-3:2023; Solid Biofuels—Determination of Moisture Content—Part 3: Moisture in General Analysis Sample. 2023. Available online: https://www.iso.org/standard/83193.html (accessed on 10 September 2023).
  41. ISO 18122:2022; Solid Biofuels—Determination of Ash Content. 2022. Available online: https://www.iso.org/standard/83190.html (accessed on 10 September 2023).
  42. ISO 18123:2023; Solid Biofuels—Determination of Volatile Matter. 2023. Available online: https://www.iso.org/standard/83192.html (accessed on 10 September 2023).
  43. Saparov, A.; Kulmukanova, L.; Mostafavi, E.; Sarbassov, Y.; Inglezakis, V.; Anthony, E.J.; Shah, D. Development and validation of a novel process model for fluidized bed combustion: Application for efficient combustion of low-grade coal. Can. J. Chem. Eng. 2021, 99, 1510–1519. [Google Scholar] [CrossRef]
  44. Wang, Z.; Chen, D.; Song, X.; Zhao, L. Study on the combined sewage sludge pyrolysis and gasification process: Mass and energy balance. Environ. Technol. 2012, 33, 2481–2488. [Google Scholar] [CrossRef]
  45. Thipkhunthod, P.; Meeyoo, V.; Rangsunvigit, P.; Kitiyanan, B.; Siemanond, K.; Rirksomboon, T. Predicting the heating value of sewage sludges in Thailand from proximate and ultimate analyses. Fuel 2005, 84, 849–857. [Google Scholar] [CrossRef]
  46. Chen, J.; Sun, Y.; Zhang, Z. Evolution of trace elements and polluting gases toward clean co-combustion of coal and sewage sludge. Fuel 2020, 280, 118685. [Google Scholar] [CrossRef]
  47. Vamvuka, D.; Salpigidou, N.; Kastanaki, E.; Sfakiotakis, S. Possibility of using paper sludge in co-firing applications. Fuel 2009, 88, 637–643. [Google Scholar] [CrossRef]
  48. Chen, W.; Wang, F.; Kanhar, A.H. Sludge Acts as a Catalyst for Coal during the Co-Combustion Process Investigated by Thermogravimetric Analysis. Energies 2017, 10, 1993. [Google Scholar] [CrossRef]
  49. Magdziarz, A.; Wilk, M. Thermogravimetric study of biomass, sewage sludge and coal combustion. Energy Convers. Manag. 2013, 75, 425–430. [Google Scholar] [CrossRef]
  50. Akahira, T.; Sunuse, T.T. Joint Convention of Four Electrical Institutes; Chiba Institute of Technology: Chiba, Japan, 1971; pp. 22–31. Available online: https://www.researchgate.net/publication/284662595_Joint_convention_of_four_electrical_institutes (accessed on 10 September 2023).
  51. Ozawa, T. A new method of analyzing thermogravimetric data. Bull. Chem. Soc. Jpn. 1965, 38, 1881–1886. [Google Scholar] [CrossRef]
  52. Doyle, C.D. Series approximations to the equation of thermogravimetric data. Nature 1965, 207, 290–291. [Google Scholar] [CrossRef]
  53. Fang, M.X.; Shen, D.; Li, Y.X.; Yu, C.J.; Luo, Z.Y.; Cen, K.F. Kinetic study on pyrolysis and combustion of wood under different oxygen concentrations by using TG-FTIR analysis. J. Anal. Appl. Pyrolysis 2006, 77, 22–27. [Google Scholar] [CrossRef]
  54. Syed-Hassan, S.S.A.; Wang, Y.; Hu, S.; Su, S.; Xiang, J. Thermochemical processing of sewage sludge to energy and fuel: Fundamentals, challenges and considerations. Renew. Sustain. Energy Rev. 2017, 80, 888–913. [Google Scholar] [CrossRef]
  55. Li, M.; Xiao, B.; Wang, X.; Liu, J. Consequences of sludge composition on combustion performance derived from thermogravimetry analysis. J. Waste Manag. 2015, 35, 141–147. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (A) Pre-dried SS samples, (B) SS fuel pellet of 0.55 ± 0.05 g.
Figure 1. (A) Pre-dried SS samples, (B) SS fuel pellet of 0.55 ± 0.05 g.
Energies 16 06634 g001
Figure 2. The bubbling fluidized bed apparatus with supplementary equipment.
Figure 2. The bubbling fluidized bed apparatus with supplementary equipment.
Energies 16 06634 g002
Figure 3. (A) TGA, (B) DTG curves for SS and bituminous coal at different blend ratios at heating of 15 °C/min in air.
Figure 3. (A) TGA, (B) DTG curves for SS and bituminous coal at different blend ratios at heating of 15 °C/min in air.
Energies 16 06634 g003
Figure 4. Kinetic plots for pure sewage sludge in the N2 environment based on heating rate of 10, 15, and 20 °C/min for (a) FWO and (b) KAS.
Figure 4. Kinetic plots for pure sewage sludge in the N2 environment based on heating rate of 10, 15, and 20 °C/min for (a) FWO and (b) KAS.
Energies 16 06634 g004
Figure 5. The relationship between activation energy vs. conversion for both FWO and KAS models.
Figure 5. The relationship between activation energy vs. conversion for both FWO and KAS models.
Energies 16 06634 g005
Figure 6. (a) O2 and CO2 concentrations (b) SO2 and NOx emissions during the mono-combustion of SS in BFB rig.
Figure 6. (a) O2 and CO2 concentrations (b) SO2 and NOx emissions during the mono-combustion of SS in BFB rig.
Energies 16 06634 g006
Figure 7. Effect of blending with coal on (a) NOx emissions (b) SO2 emissions.
Figure 7. Effect of blending with coal on (a) NOx emissions (b) SO2 emissions.
Energies 16 06634 g007
Table 1. Literature review on co-combustion of SS samples with different coal samples.
Table 1. Literature review on co-combustion of SS samples with different coal samples.
Type of ReactorBlend Solid FuelTemperatureKey Observations
TGA [30]10% SS/coal, 50% SS/coal800 °C50 wt% blends had two regions of reactivity. At T < 350 °C reactivity were similar to sludge, while for T > 350 °C had reactivity close to coal
Bubbling fluidized bed [31]SS with
High, medium, and low volatile bituminous coal
850 °CNOx increased during SS feeding, while only slight change was noted N2O emissions
Fluidized bed combustor [25]100% SS720–930 °CNOx and SO2 increased with an increase in bed temperature; N2O slightly reduced.
Fluidized bed reactor [32]100% SS,
30% SS/62% coal/8% CaCO3
800 °CDecline of SO2 emissions during the co-combustion of SS and coal with CaCO3
Fluidized bed reactor [33]100% SS700 °CMoisture content of the sludge affects the conversion of SO2 to H2SO3, but has no effect on NOx emissions
Fluidized bed [34]100% coal, blends with 70%, 55%, 30%, and pure SS800–900 °CEmissions from combustion and optimal blends based on emissions were investigated
Fluidized bed [35]100% coal
13% coal/87% SS
46% coal/54% SS
840 °CHeavy metals content during co-combustion were analyzed
Table 2. Proximate and ultimate analysis of Kazakhstan’s SS and Ekibastuz bituminous coal.
Table 2. Proximate and ultimate analysis of Kazakhstan’s SS and Ekibastuz bituminous coal.
Proximate (wt% as Received)Ultimate Analysis (wt%, Dry Ash Free)
Sewage sludgeBituminous coal Sewage sludgeBituminous coal
FC2.6938.90C48.2183.45
VM62.8221.55H5.695.35
Ash32.2238.24N5.491.40
Moisture2.271.31S1.620.60
HHV (MJ/kg)18.8719.40O *38.999.2
* By difference.
Table 3. Operating conditions of the laboratory-scale BFB rig during the SS pellets combustion.
Table 3. Operating conditions of the laboratory-scale BFB rig during the SS pellets combustion.
ParametersValue
Bed temperature, °C850 *
Operating pressure, kPa101.3
Weight of bed material, g390
Superficial gas velocity, m/s0.112
Pressure drop, kPa2.05
Primary air flow rate, lpm3.0
Secondary air flow rate, lpm0 **
Average pellet surface area, mm2226
Pellet weight (one batch), g0.55 ± 0.05
* Can be operated up to 950 °C; ** Used only for pneumatic transport.
Table 4. Mean and standard deviation values for flue gas compositions during batch combustion of SS and blend.
Table 4. Mean and standard deviation values for flue gas compositions during batch combustion of SS and blend.
ParameterMean
Pure SSBlend (75% SS/25% Coal)
Pellet mass (g)0.55 ± 0.5
O2 min (vol%)2.3 ± 0.64.6 ± 0.5
max CO2 (vol%)6.81 ± 0.62.2 ± 0.34
max NOx (ppm)356 ± 26204 ± 47
max SO2 (ppm)247 ± 39249 ± 71
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

Aidabulov, M.; Zhakupov, D.; Zhunussova, K.; Temireyeva, A.; Shah, D.; Sarbassov, Y. Thermal Characterization, Kinetic Analysis and Co-Combustion of Sewage Sludge Coupled with High Ash Ekibastuz Coal. Energies 2023, 16, 6634. https://doi.org/10.3390/en16186634

AMA Style

Aidabulov M, Zhakupov D, Zhunussova K, Temireyeva A, Shah D, Sarbassov Y. Thermal Characterization, Kinetic Analysis and Co-Combustion of Sewage Sludge Coupled with High Ash Ekibastuz Coal. Energies. 2023; 16(18):6634. https://doi.org/10.3390/en16186634

Chicago/Turabian Style

Aidabulov, Madiyar, Daulet Zhakupov, Khabiba Zhunussova, Aknur Temireyeva, Dhawal Shah, and Yerbol Sarbassov. 2023. "Thermal Characterization, Kinetic Analysis and Co-Combustion of Sewage Sludge Coupled with High Ash Ekibastuz Coal" Energies 16, no. 18: 6634. https://doi.org/10.3390/en16186634

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