Next Article in Journal
Carbon Nitride-Based Catalysts for High Pressure CO2 Photoreduction
Next Article in Special Issue
Mechanistic Studies into the Selective Production of 2,5-furandicarboxylic Acid from 2,5-bis(hydroxymethyl)furan Using Au-Pd Bimetallic Catalyst Supported on Nitrated Carbon Material
Previous Article in Journal
Special Issue “10th Anniversary of Catalysts: Biocatalysis in Analysis and Synthesis—Past, Present and Future”
Previous Article in Special Issue
Effects of Ca-Compounds on the Gases Formation Behavior during Molten Salts Thermal Treatment of Bio-Waste
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhancing the Catalytic Activity of Eggshell-Derived CaO Catalyst and Its Application in Biodiesel Production from Waste Chicken Fat

1
Department of Chemical Engineering, COMSATS University Islamabad, Lahore Campus, Lahore 54000, Pakistan
2
Department of Chemistry, The University of Lahore, 1-km Defence Road, Lahore 54000, Pakistan
3
Department of Petroleum and Chemical Engineering, College of Engineering, Sultan Qaboos University, Muscat 123, Oman
4
Department of Sustainable and Renewable Energy Engineering, University of Sharjah, Sharjah 27272, United Arab Emirates
5
Department of Environmental Sciences, COMSATS University Islamabad, Vehari Campus, Vehari 61100, Pakistan
*
Authors to whom correspondence should be addressed.
Catalysts 2022, 12(12), 1627; https://doi.org/10.3390/catal12121627
Submission received: 28 October 2022 / Revised: 8 December 2022 / Accepted: 9 December 2022 / Published: 12 December 2022

Abstract

:
The comparatively greater cost of producing biodiesel in comparison to petroleum diesel is one of the key drawbacks. Eggshells and leftover chicken fat are examples of poultry wastes that can be used to produce biodiesel at a low cost as catalysts and oil, respectively. In this study, eggshell-derived CaO and its doping with sodium methoxide catalyst for enhancing catalytic activity was synthesized for the transesterification of waste chicken fat and characterized by FT-IR and XRD analyses. XRD studies confirmed the crystalline structure of the developed catalyst and doping of sodium with eggshell-derived CaO. The transesterification reaction was performed at different reaction parameters such as the catalyst loading, the methanol to oil ratio, the reaction temperature, and the reaction time. The biodiesel produced at the maximum yield was evaluated by gas chromatography mass spectrometry analysis. A maximum yield of 96% biodiesel was obtained with catalyst loading of 2 wt% of oil, as well as a methanol to oil ratio of 13:1 at 60 °C in 1 h. The output demonstrates that eggshell waste is a potentially accessible source of biomass-derived nano catalyst for the synthesis of biodiesel using chicken fat as a feedstock.

1. Introduction

From the last decade, there has been a continuous growing trend in producing and consuming biofuels in transportation and energy sectors [1,2,3]. Because of the problems and limitations associated with the use of fossil fuels, there will be a significant rise in the demand for biofuels in near future [4,5,6]. Biodiesel is one of the important biofuels and an alternative energy source because it possesses similar properties, such as a high flash point, a high cetane number, good lubricity, non-toxicity, low sulfur contamination, etc., compared to conventional fossil fuels [7,8]. Biodiesel is a mono alkyl ester of long-chain fatty acids produced from a range of feedstocks, e.g., edible and non-edible oils. Moreover, non-edible oils are in forefront of this field as these overcome the food vs. fuel debate caused by edible oils [9,10]. In the last few years, biodiesel has been produced and is being consumed as an alternative to fossil fuel on a commercial scale, and it is available in fuel markets [11,12]. The type of feedstock used is the main economic contributor, which can have a remarkable effect on biodiesel production costs [13,14]. Therefore, using non-edible oils can reduce biodiesel production costs and biomass-derived feedstock costs. Hence, the choice of feedstock is very important in biodiesel production [15,16].
Animal fat wastes (AFWs) are easily available in slaughterhouses, farms, and restaurants in almost every region of the world. AFWs are not consumed by human beings directly and are disposed of in an open environment, which causes several health and environmental problems [17]. That is why the valorization of such wastes to derive different valuable products is more desirable to address these concerns. Numerous reports in the literature indicate that AFWs are composed of free fatty acids and triglycerides as major constituents [18]. Hence, AFWs such as tallow, lard, and chicken fats are considered as more promising candidates for producing biodiesel [19]. The biodiesel produced from using AFWs has number of advantages such as lower emissions and a higher cetane number due to the high ratio of free fatty acids in AFWs. Moreover, the biodiesel obtained from AFWs exhibits oxidative stable characteristics [20]. The transesterification of the oil extracted from AFWs is an obvious and affordable way to produce biodiesel from AFWs feedstock [21,22]. Transesterification is the reaction of triglycerides with methanol in presence of suitable catalyst and operating conditions. The reaction rate of transesterification can be faster when using different catalysts [23,24]. Different homogeneous and heterogeneous catalysts have been developed for optimum biodiesel production. Among these, heterogeneous catalysts are superior due to their ease of separation from the final products [25]. However, the synthesis of chemical and biological heterogeneous catalysts is a challenging task from an industrial point of view [26]. To cope with this problem, biomass-derived nano catalysts are intriguing due to their cost effective synthesis [27].
Waste eggshells are easily accessible as chicken eggs are commonly consumed on a regular basis by humans. Waste eggshells are recognized as one of the suitable precursors of CaO, an active catalyst in biodiesel production [28]. However, the lone CaO is not stable and active as it requires a long period of time to achieve equilibrium conversion [29]. Therefore, it has been attempted here to develop a CaO-based catalyst to overcome the mentioned problem. Various reports are available in the literature, which focus on enhancing the properties of CaO by using different chemical and physical ways [30,31,32,33]. It can also be concluded that the activity and stability of CaO can be improved by doping with a suitable dopant [34,35,36]. It has been reported that the employment of CaO as a catalyst resulted in an 80% increase in the amount of biodiesel produced. The yield with tungsten- and molybdenum-doped CaO, on the other hand, was roughly 96% [37,38]. The generation of biodiesel from Nahor oil (Mesua ferrea Linn) with lithium-doped CaO achieves an efficiency of approximately 94% [39]. In the process of producing methyl esters, the methoxide ions function as an active catalyst. This chemical unit is responsible for attacking the triglyceride molecules and resulting in the production of methyl esters. It is regenerated at the end of each reaction step when a hydrogen ion is stripped from a nearby methanol molecule. To the best of the authors’ knowledge, there has been no report in the literature indicating a CaO-based catalyst with sodium methoxide as an active species. Hence, herein, a CaO-based catalyst with enhanced activity due to incorporation with sodium methoxide (CH3NaO) as an active species is reported. Further, the designed catalyst CaO-CH3NaO was used in the transesterification reaction of oil obtained from chicken fat. Additionally, this work also emphasizes the optimization of catalyst loading and reacting conditions for transesterification to produce biodiesel from chicken fat oil.

2. Result and Discussions

2.1. FTIR Analysis

Figure 1 shows the FTIR analysis CaO-CH3NaO catalysts with different loadings and the uncalcined CaO-CH3NaO catalyst. The band of CaO appeared at around 649 cm−1, which indicates the presence of Ca-O vibration [40]. The area looked different in uncalcined and calcined CaO at 649 cm−1. The single bond region includes 1500–500 cm−1, and CaO peaks observed at 649 cm−1, which are a strong indication of its presence. NaO stretching was observed at 719 cm−1 in calcined and un-calcined groups. As concentration increases, the broader and strong peak observed shows stretching and an absorbance band of NaO. At 1090 cm−1, the C-OH stretching is observed in that region. At 1390cm−1, the broader peak is shown due to the presence of CH3. As concentration increases and after calcination at 700 °C, the CH3 has a strong broader peak, indicating its confirmation. The presence of a peak for the produced catalyst at 3680 cm−1 is due to the formation of OH in Ca(OH)2 during H2O adsorption by calcium oxide [40].

2.2. XRD Characterization

XRD studies confirmed the crystalline structure of CH3NaO-doped CaO catalyst using different doping ratios of dopant, as is shown in Figure 2. For the 5% CaO-CH3NaO catalyst, the peaks at 31.66°, 36.72°, and 52.92° correspond to CaO (COD entry no. 96-900-6706) [41], while the peaks at 29.84 °, 42.70°, and 52.96° were found regarding sodium (COD entry no. 96-901-1003) [42]. The peaks at 31.78, 36.86, and 53.12 are CaO peaks (COD entry no. 96-900-6703) [41], and the peaks at 29.84°, 42.70°, and 52.96° are sodium peaks (COD entry no. 96-901-1003) [42] in 10% CaO-CH3NaO catalyst. As far 15% CaO-CH3NaO catalyst is concerned, the angles 32.24°, 37.38°, and 53.90° correspond to CaO (COD entry no. 96-720-0687) [43], and 29.90°, 42.80°, and 53.08° correspond to sodium (COD entry no. 96-900-8545) [44]. In 20% CaO-CH3NaO catalyst, CaO peaks (COD entry no. 96-900-6746) were found at 30.82°, 35.74°, and 51.44° [41], while sodium peaks were found at 29.84°, 42.70°, and 52.96° (COD entry no. 96-901-1003) [42]. In summary, it is clear from XRD studies that the calcined CaO XRD spectra of calcined Cao without sodium doping contain peaks corresponding to CaO only. No peaks relevant to sodium were found. Meanwhile, the XRD spectra of calcined CaO doped with sodium clearly showed additional peaks, indicating the successful doping of sodium responsible for enhancing the catalytic activity of calcined CaO.

2.3. GC-MS Analysis

The GC-MS analysis was performed to determine the composition of biodiesel obtained from chicken fat oil. Figure 3 shows the chromatogram of biodiesel obtained using 5% catalyst loading (sample B1), while Figure 4 and Figure 5 show the chromatograms of biodiesel generated using 10% (sample B2) and 15% (sample B3) catalyst loading, respectively. Moreover, the detailed summary of the GC-MS analysis of biodiesel obtained by using CaO catalysts with 5, 10, and 15% loadings of sodium methoxide is presented in Table 1. The biodiesel produced using 5% CaO-CH3NaO was composed of methyl esters at retention peaks of 18.46, 19.76, 33.49, and 35.57, with octenyl ester at 44.52. Similarly, the biodiesel produced using 10% CaO-CH3NaO contained butyl esters at peaks of 8.24 and 11.41 and methyl esters at peaks of 14.71, 17.01, 20.12, 26.62, and 36.85. Meanwhile, the biodiesel produced using 15% CaO-CH3NaO comprised methyl esters at peaks of 8.22, 29.86, 33.64, 37.04, 40.35, and 44.5.

2.4. Effect of Reaction Parameters

2.4.1. Effect of Oil to Methanol Ratio

Methanol to oil molar ratio is one of the most important elements affecting biodiesel formation. Even though the transesterification reaction requires a 3:1 M ratio to produce three moles of ester and one mole of glycerol, most experiments revealed that more alcohol was needed to complete the reaction [45]. The methanol to oil ratio ranged from 6:1 to 14:1 at 55–70 °C in the investigation. The biodiesel yield was increased when the methanol to oil ratio was increased, as illustrated in Figure 6. When the molar ratio is greater than 13:1, the biodiesel output decreases. The yield of biodiesel decreases if the molar ratio is increased over 13:1 because too much methanol and glycerol dissolve in methanol, which stops the reaction. Additionally, the reaction inhibition would cause a reaction direction to reverse and reduce the yield. On the other hand, the reactions would not continue, and no fatty acid methyl ester (FAME) would be produced if the molar ratio were too low.

2.4.2. Effect of Catalyst Loading

The choice of the catalyst and its concentration are crucial in the transesterification reaction since they both influence how well methanol and oil combine. The catalyst’s loadings wt.% of oil varied from 1 to 5 wt.% to examine the biodiesel yield (Figure 7). The yield increases with time with an oil catalyst loading of 2wt.% and subsequently falls after reaching the maximum point that kept the others factor constant, as given in Figure 8. The reason was that as more catalyst was added, the viscosity of the reaction mixture was changed as a result of soap formation and the difficulties in separating biodiesel from glycerol and catalyst. Therefore, a catalyst concentration that is either too low or too high would reduce the output of biodiesel.

2.4.3. Effect of Temperature

Figure 8 shows the effect of the reaction temperature on biodiesel production from catalyzed transesterification. The biodiesel yield was improved when the temperature was increased. When the temperature is increased from 45 to 60 °C, the biodiesel yield increased. This was due to the fact that as the temperature rises, so does the miscibility of alcohol in the waste chicken fat oil, as does the average speed (kinetic energy) of the molecules, resulting in more successful collisions and higher reaction rates. The viscosity of the reaction mixture was reduced as the reaction temperature was increased, lowering the diffusion barrier between the distinct phases (oil, methanol, and catalyst). As the temperature rises to 65 °C, methanol evaporation causes the cavitation bubble to become supersaturated with methanol vapor, cushioning the collapse. Bubbles collapse with less force as a result, reducing mass transfer and yield. The effect of temperature on biodiesel yield was investigated in this study at temperatures ranging from 55 to 70 °C. By keeping the remaining parameters constant, the reaction at 65 °C yielded the highest yield. At 55 °C, the reaction was partial, resulting in reduced conversion. Further increasing the temperature to 65 °C yielded the best results; however, as the temperature was increased, the yield dropped due to the evaporation of alcohol, resulting in a loss in mass transfer.

2.4.4. Effect of Reaction Time

At reaction temperature of 60 °C and 1 h reaction time, the highest biodiesel yield of 96% was produced, as is shown in Figure 9. The yield was found to be lower during the shortest reaction time due to incorrect diffusion between the three phases of the reaction mixture. Increasing the reaction time beyond 1 h, on the other hand, resulted in a reduced yield. The longer reaction time leads to a reduction in biodiesel yield due to the reversible reaction of transesterification resulting in the loss of esters, as well as soap formation.

2.4.5. Comparative Study of Biodiesel Yield from Waste Chicken Fat

Table 2 shows the comparative study of biodiesel yield from waste chicken fat as feedstock using CaO and different catalysts. The catalysts used for biodiesel production from waste chicken fat other than CaO showed lower yields of biodiesel, as shown in Table 2. However, using the CaO catalyst, the biodiesel yield was obtained to 90.2%. Similarly, a biodiesel yield up to 94.5% was observed using CaO/CuFe2O4 but with catalyst loading higher than 2 wt%. In this study, the developed catalyst produced biodiesel up to a maximum of 96% with 2 wt% catalyst loading due to enhanced activity.

3. Materials and Methods

3.1. Materials

Chicken fat and waste eggshells were collected from the nearby slaughterhouse and bakery shop, respectively, and washed with distilled water several times to eliminate impurities such as dust particles and other odorous wastes.

3.2. Methods

3.2.1. Oil-Extraction Method

This step involved oil extraction from chicken fat, which was achieved by placing chicken fat in microwave oven until the oil was produced. To remove the water content, the collected oil was heated over 100 °C before being purified. The waste chicken fat oil was filtered for further purification and stored.

3.2.2. Determination of Acid Value

High free fatty acids (FFA) reduce biodiesel yield by generating soap during the transesterification cycle. The acid value was determined using the following equation:
Acid value (AV) = V × C × 56.11/m
where ‘V’ represents the volume of the KOH solution (mL), ‘C’ represents the KOH concentration (mol L−1), and ‘m’ represents the mass of oil extracted (gram).
The acid value (AV) was 0.86 mg KOH/g of oil, which equates to 0.43% free fatty acids (FFA).

3.2.3. Catalyst Synthesis and Characterization

The sodium-methoxide-doped CaO was prepared by the simple impregnation method [34]. The complete synthesis procedure is as follows. The washed waste eggshells were crushed into tiny pieces and heated to 105 °C for 24 h in a hot air oven to eliminate the moisture. The dried shells were then ground with a pestle and mortar followed by calcination at 700 °C for 5 h in a muffle furnace. The obtained powder was labeled as an active CaO catalyst and was stored in tightly closed vial for further use. Further, the aqueous solution of sodium methoxide with a suitable concentration was prepared and then drop-wise added to CaO powder carefully. The resulting mixture was placed in an oven for drying for 10 h at 110 °C. After drying, the resulting solid was named the CaO-CH3NaO catalyst and was used for the transesterification process and characterized. The X-ray diffraction (XRD) patterns were acquired by using the PRO analytical X-ray diffractometer. Fourier transform infrared spectroscopy (FTIR) spectra were recorded by infrared spectroscopy with KBr pellets in the range 400–4000 cm−1. The signal resolution of 4 cm−1 with 40 scans was applied. The same procedure was followed to develop different catalysts with varying loadings of dopant. Figure 10 shows the complete synthetic procedure of the CaO-CH3NaO catalyst.

3.2.4. Transesterification of Chicken Fat Oil

The transesterification reaction was carried in a two-neck round-bottom flask (250 mL) connected with a refluxing condenser on a magnetic stirrer. An appropriate volume of oil and methanol was poured into a flask followed by the addition of the synthesized catalyst. The reaction was conducted under continuous stirring with different catalyst loadings and a methanol to oil ratio at suitable temperatures. The catalyst was removed by centrifugation after the reaction was completed. The resulting mixture after centrifugation was transferred into a separating funnel. The biodiesel and glycerol layers were separated in a separating funnel after 34 h. The biodiesel layer was carefully collected and heated up to 100 °C to remove any unreacted methanol. The biodiesel sample was sent for GC-MS analysis. The FAME of biodiesel was characterized by a Perkin Elmer (Waltham, MA, USA) Clarus 600 gas chromatograph mass spectrometry (GC-MS) using a DB-Wax column. Helium was used as a carrier gas with constant flow of 1.0 mL/min, with a 1 μL sample injected with a mass spectrometer set to scan a frequency range of 40–550 amu. The initial oven temperature was set to 80 °C and held for 5 min before being increased to 240 °C at a heating rate of 4 °C/min. Finally, the biodiesel yield was calculated as given below. In addition, multiple reactions were performed to study the effect of different parameters such as the catalyst loading (1–5 wt%), the methanol to oil ratio (6–14), the temperature (55–70 °C), and the reaction time on biodiesel yield (30–180 min).
Y i e l d = M a s s   o f   b i o d i e s e l   i n   g r a m s M a s s   o f   o i l   t a k e n   i n   g r a m s × 100

4. Conclusions

Waste chicken fat and eggshell are examples of poultry wastes that could serve as inexpensive sources of raw materials for the production of biodiesel. In this work, used chicken eggshells were calcined at 700 °C and effectively turned into CaO and doped with sodium methoxide for an effective catalyst for biodiesel production. The waste chicken fat that was recovered using a straightforward non-solvent approach was trans-esterified (valorized) using a heterogenous catalyst made from chicken eggshell. The maximum yield of 96% biodiesel was obtained with catalyst loading of 2 wt% of oil, and a methanol to oil ratio of 13:1 at 60 °C in 1 h. The analysis of the resulting biodiesel provides additional evidence for the viability of the raw materials for the synthesis of biodiesel. This research has established the transesterification of oil from chicken slaughter waste using a low-cost, non-solvent extraction method and a low-cost catalyst source. A future study for a suitable and feasible commercialization of the process is proposed, including optimization studies, eggshell catalyst treatment with metal oxide reagents, and pilot-scale investigations of the procedure.

Author Contributions

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

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

Murid Hussain would like to thank the Higher Education Commission (HEC), Pakistan, for its partial support under TDF Project (TDF02- 011).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sajid, M.; Raheem, A.; Ullah, N.; Asim, M.; Ur Rehman, M.S.; Ali, N. Gasification of municipal solid waste: Progress, challenges, and prospects. Renew. Sustain. Energy Rev. 2022, 168, 112815. [Google Scholar] [CrossRef]
  2. Osman, A.I.; Qasim, U.; Jamil, F.; Al-Muhtaseb, A.H.; Jrai, A.A.; Al-Riyami, M.; Al-Maawali, S.; Al-Haj, L.; Al-Hinai, A.; Al-Abri, M.; et al. Bioethanol and biodiesel: Bibliometric mapping, policies and future needs. Renew. Sustain. Energy Rev. 2021, 152, 111677. [Google Scholar] [CrossRef]
  3. Inayat, A.; Jamil, F.; Ghenai, C.; Kamil, M.; Bokhari, A.; Waris, A.; Rafiq, S.; Khurram, S. Biodiesel synthesis from neem oil using neem seeds residue as sustainable catalyst support. Biomass Convers. Biorefinery 2021, 1–6. [Google Scholar] [CrossRef]
  4. Bashir, M.A.; Wu, S.; Zhu, J.; Krosuri, A.; Khan, M.U.; Ndeddy Aka, R.J. Recent development of advanced processing technologies for biodiesel production: A critical review. Fuel Processing Technol. 2022, 227, 107120. [Google Scholar] [CrossRef]
  5. Zhang, Y.; Duan, L.; Esmaeili, H. A review on biodiesel production using various heterogeneous nanocatalysts: Operation mechanisms and performances. Biomass Bioenergy 2022, 158, 106356. [Google Scholar] [CrossRef]
  6. Ali Qamar, O.; Jamil, F.; Hussain, M.; Al-Muhtaseb, A.H.; Inayat, A.; Waris, A.; Akhter, P.; Park, Y.-K. Feasibility-to-applications of value-added products from biomass: Current trends, challenges, and prospects. Chem. Eng. J. 2022, 454, 140240. [Google Scholar] [CrossRef]
  7. Basumatary, S.F.; Patir, K.; Das, B.; Saikia, P.; Brahma, S.; Basumatary, B.; Nath, B.; Basumatary, B.; Basumatary, S. Production of renewable biodiesel using metal organic frameworks based materials as efficient heterogeneous catalysts. J. Clean. Prod. 2022, 358, 131955. [Google Scholar] [CrossRef]
  8. Zulkepli, S.; Lee, H.V.; Rahman, N.A.; Chuan, L.T.; Show, P.L.; Chen, W.-H.; Juan, J.C. Highly active iron-promoted hexagonal mesoporous silica (HMS) for deoxygenation of triglycerides to green hydrocarbon-like biofuel. Fuel 2022, 308, 121860. [Google Scholar] [CrossRef]
  9. Shaah, M.A.H.; Hossain, M.S.; Salem Allafi, F.A.; Alsaedi, A.; Ismail, N.; Ab Kadir, M.O.; Ahmad, M.I. A review on non-edible oil as a potential feedstock for biodiesel: Physicochemical properties and production technologies. RSC Adv. 2021, 11, 25018–25037. [Google Scholar] [CrossRef]
  10. Singh, D.; Sharma, D.; Soni, S.L.; Inda, C.S.; Sharma, S.; Sharma, P.K.; Jhalani, A. A comprehensive review of biodiesel production from waste cooking oil and its use as fuel in compression ignition engines: 3rd generation cleaner feedstock. J. Clean. Prod. 2021, 307, 127299. [Google Scholar] [CrossRef]
  11. Abu-Jrai, A.M.; Jamil, F.; Al-Muhtaseb, A.H.; Baawain, M.; Al-Haj, L.; Al-Hinai, M.; Al-Abri, M.; Rafiq, S. Valorization of waste Date pits biomass for biodiesel production in presence of green carbon catalyst. Energy Convers. Manag. 2017, 135, 236–243. [Google Scholar] [CrossRef]
  12. Sitepu, E.K.; Heimann, K.; Raston, C.L.; Zhang, W. Critical evaluation of process parameters for direct biodiesel production from diverse feedstock. Renew. Sustain. Energy Rev. 2020, 123, 109762. [Google Scholar] [CrossRef]
  13. Ningaraju, C.; Yatish, K.V.; Sakar, M. Simultaneous refining of biodiesel-derived crude glycerol and synthesis of value-added powdered catalysts for biodiesel production: A green chemistry approach for sustainable biodiesel industries. J. Clean. Prod. 2022, 363, 132448. [Google Scholar] [CrossRef]
  14. Chong, J.W.R.; Tan, X.; Khoo, K.S.; Ng, H.S.; Jonglertjunya, W.; Yew, G.Y.; Show, P.L. Microalgae-based bioplastics: Future solution towards mitigation of plastic wastes. Environ. Res. 2022, 206, 112620. [Google Scholar] [CrossRef] [PubMed]
  15. Jamil, F.; Saleem, M.; Qamar, O.A.; Khurram, M.S.; Al-Muhtaseb, A.H.; Inayat, A.; Akhter, P.; Hussain, M.; Rafiq, S.; Yim, H.; et al. State-of-the-art catalysts for clean fuel (methyl esters) production—A comprehensive review. J. Phys. Energy 2022, 5, 014005. [Google Scholar] [CrossRef]
  16. Chanthon, N.; Ngaosuwan, K.; Kiatkittipong, W.; Wongsawaeng, D.; Appamana, W.; Quitain, A.T.; Assabumrungrat, S. High-efficiency biodiesel production using rotating tube reactor: New insight of operating parameters on hydrodynamic regime and biodiesel yield. Renew. Sustain. Energy Rev. 2021, 151, 111430. [Google Scholar] [CrossRef]
  17. Liu, N.; Jiang, J. Valorisation of food waste using salt to alleviate inhibition by animal fats and vegetable oils during anaerobic digestion. Biomass Bioenergy 2020, 143, 105826. [Google Scholar] [CrossRef]
  18. Ndiaye, M.; Arhaliass, A.; Legrand, J.; Roelens, G.; Kerihuel, A. Reuse of waste animal fat in biodiesel: Biorefining heavily-degraded contaminant-rich waste animal fat and formulation as diesel fuel additive. Renew. Energy 2020, 145, 1073–1079. [Google Scholar] [CrossRef]
  19. Sai Akhil, U.; Alagumalai, A. A Short Review on Valorization of Slaughterhouse Wastes for Biodiesel Production. ChemistrySelect 2019, 4, 13356–13362. [Google Scholar] [CrossRef]
  20. Kirubakaran, M.; Selvan, V.A.M. A comprehensive review of low cost biodiesel production from waste chicken fat. Renew. Sustain. Energy Rev. 2018, 82, 390–401. [Google Scholar] [CrossRef]
  21. Riaz, I.; Shafiq, I.; Jamil, F.; Al-Muhtaseb, A.H.; Akhter, P.; Shafique, S.; Park, Y.-K.; Hussain, M. A review on catalysts of biodiesel (methyl esters) production. Catal. Rev. 2022, 1–53. [Google Scholar] [CrossRef]
  22. Taufiq-Yap, Y.H.; Teo, S.H.; Rashid, U.; Islam, A.; Hussien, M.Z.; Lee, K.T. Transesterification of Jatropha curcas crude oil to biodiesel on calcium lanthanum mixed oxide catalyst: Effect of stoichiometric composition. Energy Convers. Manag. 2014, 88, 1290–1296. [Google Scholar] [CrossRef]
  23. Munyentwali, A.; Li, H.; Yang, Q. Review of advances in bifunctional solid acid/base catalysts for sustainable biodiesel production. Appl. Catal. A Gen. 2022, 633, 118525. [Google Scholar] [CrossRef]
  24. Thangaraj, B.; Solomon, P.R.; Muniyandi, B.; Ranganathan, S.; Lin, L. Catalysis in biodiesel production—A review. Clean Energy 2019, 3, 2–23. [Google Scholar] [CrossRef] [Green Version]
  25. Zhang, H.; Li, H.; Hu, Y.; Rao, K.T.V.; Xu, C.; Yang, S. Advances in production of bio-based ester fuels with heterogeneous bifunctional catalysts. Renew. Sustain. Energy Rev. 2019, 114, 109296. [Google Scholar] [CrossRef]
  26. Zhang, H.; Zhang, L.-L.; Tan, X.; Li, H.; Yang, S. Catalytic high-yield biodiesel production from fatty acids and non-food oils over a magnetically separable acid nanosphere. Ind. Crops Prod. 2021, 173, 114126. [Google Scholar] [CrossRef]
  27. Tan, X.; Zhang, H.; Li, H.; Yang, S. Electrovalent bifunctional acid enables heterogeneously catalytic production of biodiesel by (trans)esterification of non-edible oils. Fuel 2022, 310, 122273. [Google Scholar] [CrossRef]
  28. Mansir, N.; Teo, S.H.; Rashid, U.; Saiman, M.I.; Tan, Y.P.; Alsultan, G.A.; Taufiq-Yap, Y.H. Modified waste egg shell derived bifunctional catalyst for biodiesel production from high FFA waste cooking oil. A review. Renew. Sustain. Energy Rev. 2018, 82, 3645–3655. [Google Scholar] [CrossRef]
  29. Sronsri, C.; Sittipol, W.; U-yen, K. Performance of CaO catalyst prepared from magnetic-derived CaCO3 for biodiesel production. Fuel 2021, 304, 121419. [Google Scholar] [CrossRef]
  30. Kesserwan, F.; Ahmad, M.N.; Khalil, M.; El-Rassy, H. Hybrid CaO/Al2O3 aerogel as heterogeneous catalyst for biodiesel production. Chem. Eng. J. 2020, 385, 123834. [Google Scholar] [CrossRef]
  31. Pandit, P.R.; Fulekar, M.H. Biodiesel production from microalgal biomass using CaO catalyst synthesized from natural waste material. Renew. Energy 2019, 136, 837–845. [Google Scholar] [CrossRef]
  32. Lani, N.S.; Ngadi, N.; Inuwa, I.M. New route for the synthesis of silica-supported calcium oxide catalyst in biodiesel production. Renew. Energy 2020, 156, 1266–1277. [Google Scholar] [CrossRef]
  33. Badnore, A.U.; Jadhav, N.L.; Pinjari, D.V.; Pandit, A.B. Efficacy of newly developed nano-crystalline calcium oxide catalyst for biodiesel production. Chem. Eng. Process.-Process Intensif. 2018, 133, 312–319. [Google Scholar] [CrossRef]
  34. Das, V.; Tripathi, A.M.; Borah, M.J.; Dunford, N.T.; Deka, D. Cobalt-doped CaO catalyst synthesized and applied for algal biodiesel production. Renew. Energy 2020, 161, 1110–1119. [Google Scholar] [CrossRef]
  35. Zul, N.A.; Ganesan, S.; Hamidon, T.S.; Oh, W.-D.; Hussin, M.H. A review on the utilization of calcium oxide as a base catalyst in biodiesel production. J. Environ. Chem. Eng. 2021, 9, 105741. [Google Scholar] [CrossRef]
  36. Poosumas, J.; Ngaosuwan, K.; Quitain, A.T.; Assabumrungrat, S. Role of ultrasonic irradiation on transesterification of palm oil using calcium oxide as a solid base catalyst. Energy Convers. Manag. 2016, 120, 62–70. [Google Scholar] [CrossRef]
  37. Shin, H.-Y.; Ryu, J.-H.; Bae, S.-Y.; Crofcheck, C.; Crocker, M. Lipid extraction from Scenedesmus sp. microalgae for biodiesel production using hot compressed hexane. Fuel 2014, 130, 66–69. [Google Scholar] [CrossRef] [Green Version]
  38. Meher, L.C.; Kulkarni, M.G.; Dalai, A.K.; Naik, S.N. Transesterification of karanja (Pongamia pinnata) oil by solid basic catalysts. Eur. J. Lipid Sci. Technol. 2006, 108, 389–397. [Google Scholar] [CrossRef]
  39. Cho, Y.B.; Seo, G. High activity of acid-treated quail eggshell catalysts in the transesterification of palm oil with methanol. Bioresour. Technol. 2010, 101, 8515–8519. [Google Scholar] [CrossRef]
  40. Chukanov, N.V.; Chervonnyi, A.D. Infrared Spectroscopy of Minerals and Related Compounds, 1st ed.; Springer: Cham, Switzerland, 2016; p. 1109. [Google Scholar]
  41. Fiquet, G.; Richet, P.; Montagnac, G. High-temperature thermal expansion of lime, periclase, corundum and spinel. Phys. Chem. Miner. 1999, 27, 103–111. [Google Scholar] [CrossRef]
  42. Barrett, C.S. X-ray study of the alkali metals at low temperatures. Acta Crystallogr. 1956, 9, 671–677. [Google Scholar] [CrossRef]
  43. Verbraeken, M.C.; Suard, E.; Irvine, J.T.S. Structural and electrical properties of calcium and strontium hydrides. J. Mater. Chem. 2009, 19, 2766–2770. [Google Scholar] [CrossRef]
  44. Wyckoff, R.W.G. Crystal Structures 1, 2nd ed.; Interscience Publishers: New York, NY, USA, 1963; pp. 7–83. [Google Scholar]
  45. Eboibi, B.E.; Eboibi, O.; Okputu, J.; Okpohwo, K.A. Production and analysis of biodiesel from Jatropha curcas seed. J. Appl. Sci. Environ. Manag. 2018, 22, 26–33. [Google Scholar] [CrossRef] [Green Version]
  46. Mata, T.M.; Cardoso, N.; Ornelas, M.; Neves, S.; Caetano, N.S. Evaluation of Two Purification Methods of Biodiesel from Beef Tallow, Pork Lard, and Chicken Fat. Energy Fuels 2011, 25, 4756–4762. [Google Scholar] [CrossRef]
  47. Alptekin, E.; Canakci, M.; Sanli, H. Methyl ester production from chicken fat with high FFA. In Proceedings of the World Renewable Energy Congress-Sweden, Linköping, Sweden, 8–13 May 2011; pp. 319–326. [Google Scholar]
  48. Odetoye, T.E.; Agu, J.O.; Ajala, E.O. Biodiesel production from poultry wastes: Waste chicken fat and eggshell. J. Environ. Chem. Eng. 2021, 9, 105654. [Google Scholar] [CrossRef]
  49. Seffati, K.; Honarvar, B.; Esmaeili, H.; Esfandiari, N. Enhanced biodiesel production from chicken fat using CaO/CuFe2O4 nanocatalyst and its combination with diesel to improve fuel properties. Fuel 2019, 235, 1238–1244. [Google Scholar] [CrossRef]
Figure 1. FTIR analysis of uncalcined and calcined CH3NaO-doped CaO catalyst with different doping ratios.
Figure 1. FTIR analysis of uncalcined and calcined CH3NaO-doped CaO catalyst with different doping ratios.
Catalysts 12 01627 g001
Figure 2. XRD studies of calcined CaO and CH3NaO-doped CaO catalyst with different doping ratios.
Figure 2. XRD studies of calcined CaO and CH3NaO-doped CaO catalyst with different doping ratios.
Catalysts 12 01627 g002
Figure 3. GC-MS analysis of sample B1.
Figure 3. GC-MS analysis of sample B1.
Catalysts 12 01627 g003
Figure 4. GC-MS analysis of sample B2.
Figure 4. GC-MS analysis of sample B2.
Catalysts 12 01627 g004
Figure 5. GC-MS analysis of sample B3.
Figure 5. GC-MS analysis of sample B3.
Catalysts 12 01627 g005
Figure 6. Effect of methanol to oil ratio on biodiesel production.
Figure 6. Effect of methanol to oil ratio on biodiesel production.
Catalysts 12 01627 g006
Figure 7. Effect of catalyst loadings on biodiesel production.
Figure 7. Effect of catalyst loadings on biodiesel production.
Catalysts 12 01627 g007
Figure 8. Effect of reaction temperature on biodiesel production.
Figure 8. Effect of reaction temperature on biodiesel production.
Catalysts 12 01627 g008
Figure 9. Effect of reaction time on biodiesel production.
Figure 9. Effect of reaction time on biodiesel production.
Catalysts 12 01627 g009
Figure 10. Steps involved in complete synthesis of CH3NaO-doped CaO catalyst.
Figure 10. Steps involved in complete synthesis of CH3NaO-doped CaO catalyst.
Catalysts 12 01627 g010
Table 1. Summary of GC-MS analysis of biodiesel produced using CaO-CH3NaO catalysts with different doping ratios.
Table 1. Summary of GC-MS analysis of biodiesel produced using CaO-CH3NaO catalysts with different doping ratios.
5% CaO-CH3NaO10% CaO-CH3NaO15% CaO-CH3NaO
Ret. TimeCompositionMwRet. TimeCompositionMwRet. TimeCompositionMw
8.23Cyclohexene,1-methyl-52708.23Cyclohexene,1-methyl-51288.22Methyl ethyl ester186
13.905-dimethyl-benzene methanol3988.24Isobutyl ester12811.09Encalyptol284
18.466-heptanoic acid, methyl ester43811.413-methyl butyl ester28414.84Cyclo hexene296
19.762,4-hexadenoic acid, methyl ester26414.712-propenoic acid, methyl ester29616.84Citronellol208
22.273-hexadiene-1-ol10017.017-methylene-2,4,4-timethyl, methyl ester32022.242-cyclo hexe-1one296
32.426- Octenoic acid21417.61Mono-(2,2-dimethyl cyclo hexyl acid) ester25026.34Napthalene156
33.49Hexadecanoic acid, methyl ester43820.12Hexa decanoic acid, methyl ester13829.86Docosaptenoic acid, methyl ester214
35.577Pentadecanoic acid, methyl ester46626.626-octenoic acid, methyl ester28433.64Hexa decenoic acid, methyl ester310
44.527-dimethyl-6-octenyl ester38934.65Tetra methyl thioctic acid16037.04Hepta deconic acid, methyl ester348
36.859-octenoic acid, methyl ester20340.3511-eicosenoic acid, methyl ester408
44.49Mono ethyl hexyl ester30744.5Docasenoic acid, methyl ester450
Table 2. Comparative study of biodiesel from waste chicken fat using different catalysts.
Table 2. Comparative study of biodiesel from waste chicken fat using different catalysts.
Sr. NoCatalystsCatalyst Loading, wt%Biodiesel Yield, %Reference
1.KOH0.567[46]
2.NaOH171.3[47]
3.KOMe0.588.5[47]
4.KOH0.876.8[46]
5.CaO290.2[48]
6.CaO/CuFe2O4394.5[49]
7.CaO-CH3NaO296This study
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Saleem, M.; Jamil, F.; Qamar, O.A.; Akhter, P.; Hussain, M.; Khurram, M.S.; Al-Muhtaseb, A.H.; Inayat, A.; Shah, N.S. Enhancing the Catalytic Activity of Eggshell-Derived CaO Catalyst and Its Application in Biodiesel Production from Waste Chicken Fat. Catalysts 2022, 12, 1627. https://doi.org/10.3390/catal12121627

AMA Style

Saleem M, Jamil F, Qamar OA, Akhter P, Hussain M, Khurram MS, Al-Muhtaseb AH, Inayat A, Shah NS. Enhancing the Catalytic Activity of Eggshell-Derived CaO Catalyst and Its Application in Biodiesel Production from Waste Chicken Fat. Catalysts. 2022; 12(12):1627. https://doi.org/10.3390/catal12121627

Chicago/Turabian Style

Saleem, Muhammad, Farrukh Jamil, Obaid Ali Qamar, Parveen Akhter, Murid Hussain, Muhammad Shahzad Khurram, Ala’a H. Al-Muhtaseb, Abrar Inayat, and Noor Samad Shah. 2022. "Enhancing the Catalytic Activity of Eggshell-Derived CaO Catalyst and Its Application in Biodiesel Production from Waste Chicken Fat" Catalysts 12, no. 12: 1627. https://doi.org/10.3390/catal12121627

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