Next Article in Journal
Review on Test Benches Studying Sliding Electrical Contact and Synthesis of Experimental Results
Next Article in Special Issue
Biodiesel Additives Synthesis Using Solid Heteropolyacid Catalysts
Previous Article in Journal
UAV Photogrammetry Application for Determining the Influence of Shading on Solar Photovoltaic Array Energy Efficiency
Previous Article in Special Issue
The Initial Boiling Point of Lubricating Oil as an Indicator for the Assessment of the Possible Contamination of Lubricating Oil with Diesel Oil
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Transesterification of Algae Oil and Little Amount of Waste Cooking Oil Blend at Low Temperature in the Presence of NaOH

1
Department of Mechanical Engineering, School of Engineering, University of Petroleum and Energy Studies, Energy Acres, Bidholi, Dehradun 248007, Uttarakhand, India
2
Department of Mechanical Engineering, Phonics Group of Institutions, Roorkee 247667, India
3
Department of Electrical Engineering, School of Electronics and Electrical Engineering, Lovely Professional University, Phagwara 144402, India
4
Electrical Engineering Department, I. K. G. Punjab Technical University, Jalandhar 144601, India
5
Department of Electrical Engineering, Graphic Era (Deemed to be University), Dehradun 248002, India
6
Graphic Era Hill University, Dehradun 248002, India
7
Applied Science Research Center, Applied Science Private University, Amman 11931, Jordan
8
Department of Mechanical Engineering, College of Engineering, King Khalid University, Abha 61421, Saudi Arabia
*
Authors to whom correspondence should be addressed.
Energies 2023, 16(3), 1293; https://doi.org/10.3390/en16031293
Submission received: 5 December 2022 / Revised: 14 January 2023 / Accepted: 20 January 2023 / Published: 25 January 2023
(This article belongs to the Special Issue Recent Progress in Biodiesel and IC Engines)

Abstract

:
The present study describes the single-step transesterification method of biodiesel production from high free fatty acid (FFA) waste cooking oil blended with algae oil using a homogeneous base catalyst. Due to high FFA contents, two step transesterification is needed to convert oil into biodiesel and therefore the high FFA content of waste cooking oil is decreased by blending it with low FFA content algae oil, which would further lead only to single step transesterification of low FFA oil. The design and optimization studies were conducted using Response Surface Methodology (RSM). The box-Behnken design technique is applied to optimize the three process parameters, i.e., catalyst concentration (0–2 wt%), methanol concentration (v/v) (20–60%) and reaction time (60–180 min) at a uniform reaction temperature of 50 °C. The result of the current study indicates that an effective biodiesel yield of 92% can be obtained at the optimized condition of catalyst concentration of 1.5% (w/w), methanol/oil ratio of 21:1 and reaction time of 110 min at a constant reaction temperature of 50 °C. This analysis clearly shows that this study can resolve the storage problem of high FFA oils from different feedstock and RSM can be successfully used to model the reaction to maximize the biodiesel yield.

1. Introduction

The planet is dealing with the issue of the depletion of conventional energy sources and their hazardous environmental pollution. Due to this, researchers have decided to develop new sources of energy that can play a promising role as alternative fuels [1,2]. Among the multiple renewable sources of energy, biomass is the most easily accessible resource, having a low cost and being abundant worldwide [3]. Biofuels like biodiesel, biooil, syngas and methane obtained from biomass have huge potential for commercial usage [4,5]. Since there are multiple challenges for their large scale production, it is essential to develop novel technologies to fulfil the energy needs of tomorrow [6,7].
Biodiesel, in particular, has been identified as a beneficial substitute, as biodiesel and its blends used with a Compression Ignition (CI) engine create a significant reduction in the number of emissions such as CO and unburnt hydrocarbons, with no sulfur compounds, as it is sulfur-free [8,9]. Biodiesel can be made through the transesterification of vegetable oil, animal fats, dry biomass and other non-edible oils with alcohol and a catalyst [10]. Producing biodiesel from edible oils disturbs food security worldwide and also raises the price of biodiesel [11] Thus, more focus is being given to the non-edible sources for biodiesel synthesis so that the production cost of biodiesel can be lowered [12,13].
The application of waste cooking oil (WCO) has increased over the years and gained the attention of various countries for use as a feedstock for biodiesel production [14]. Using WCO, the cost of biodiesel can be lowered and waste disposal reduced. WCO is the oil being used for food preparation and is no longer usable. The main concern in using WCO for high-quality biodiesel production is the high free-fatty acid (FFA) concentration. Various pretreatment processes are required to reduce the FFA of WCO to make it suitable for transesterification [15]. Various researchers have studied WCO for producing biodiesel, using different heterogeneous catalysts to prevent the need for pretreatment of high FFA oils [16,17].
Mohadesi et al. [18] used KOH or clinoptilolite as a heterogeneous catalyst for biodiesel synthesis from WCO. The calcination time of the catalyst was 300 min. at 400 °C. A considerable yield of 97.45% was recorded for optimum production at an 8.1 wt% catalyst concentration and a 2.25:01 methanol and oil ratio for a reaction time (Rt) of 13.4 min at a reaction temperature (RT) of 65 °C. Hsiao et al. [19] studied Ca(OH)2 as a heterogeneous catalyst for microwave-assisted transesterification of WCO. The calcination time for the catalyst was 120 min at 65 °C, with a considerable yield of 98.2% at 65 °C obtained for the conditions of 75-min reaction time, a 4.0 wt% catalyst concentration and 8:01 methanol/oil ratio. Aghel et al. [20] evaluated kettle limescale as a catalyst for biodiesel synthesis from WCO in a pilot-scale microreactor. Limescale was calcined at 900 °C for 120 min for catalyst preparation. The maximum production of 93.41% was achieved for the reaction conditions of 8.87 wt% catalyst concentration, 1.7:03 methanol/oil ratio with Rt of 15 min at a T of 61.7 °C. Degfie et al. [21] used the CaO heterogeneous catalyst to produce biodiesel from WCO. The catalyst was calcined at 500 °C. The optimum combination of 1% catalyst and an 8:1 methanol to oil ratio resulted in a 96% yield in 90 min at 50 °C. Komintarachat and Chuepeng [22] developed a CaO-KCl catalyst from green mussel shells by calcination at 500 °C for 300 min to produce biodiesel from WCO. An optimum biodiesel production rate of 97% was achieved, but at the cost of high energy consumption in catalyst preparation.
The literature review showed that heterogeneous catalysts are difficult to prepare and require a high amount of energy for the calcination process. Some base heterogeneous catalysts are sensitive to FFA content, thus single-step transesterification is not possible when using these catalysts [23]. Homogeneous catalysts are easily available and have become more important with the commercialization of biodiesel production worldwide [24]. Mohadesi et al. [25] manufactured biodiesel from WCO in a pilot prototype reactor. The esterification process was carried out at 1% H2SO4 to keep the FFA of WCO below 1% at 06:01 methanol and oil concentration at a RT of 65 °C. The oil obtained was further trans-esterified with a 1.16 wt% KOH catalyst concentration. Biodiesel production of 98.26% was achieved using a 9.4:01 methanol and oil ratio at a RT of 62.4 °C.
Algae is supported by many researchers as a renewable source of energy and at the same time it has a smaller amount of FFA, which is good for the trans-esterification process [26].
From the literature review, it was found that homogeneous catalysts are more popular and easier to use. These catalysts require the pretreatment of high-FFA oils. As a result, they require a two-step transesterification process: an acid catalyzed transesterification process to lower the FFA content of the oil, followed by a base catalyzed transesterification process to produce biodiesel. This whole process, consisting of two steps of transesterification, is both a time and energy consuming process. Algae is receiving attention due to its high productivity and oil yield which is of low FFA, and due to this reason it can be converted to biodiesel using single step transesterification. On the other hand, WCO requires a two-step transesterification process to convert it into biodiesel due to high FFA. For this reason, to use WCO for biodiesel production with single step transesterification, the FFA of the oil can be reduced by blending the oil with low-FFA algae oil; therefore, in the present study, the production of biodiesel from high-FFA WCO blended with low-FFA algal oil using a homogeneous catalyst in a single-step transesterification process is proposed.

2. Materials and Methods

2.1. Materials

For this experiment, WCO was collected from city restaurants and algae oil was brought in from the Soley Biotechnology Institute. Analytical reactants included NaOH and Methanol, which were obtained from Naveen Traders, Roorkee. The properties of the fuels, WCO and algae oil were calculated using the usual techniques described in IS:15607 and then obtained after refinement. The FFA content of WCO was found to be 20%, which was very high for the base-catalyzed transesterification process. Only 0.50% of algal oil was discovered to contain FFA. Using a base catalyst causes the saponification reaction, which results in the creation of soap when FFA levels are more than 1.0%. Thus, there is a need to reduce the FFA of WCO below or equal to 1%. The two oils (algae oil: WCO) mentioned above were blended in a volumetric proportion to reduce the FFA content of WCO to 1% to make it suitable for the homogeneous-base catalyzed transesterification-process. The titration process was performed to obtain the optimum blend ratio of algae oil: WCO. The final blend ratio of 1:40 was found to be appropriate (algal oil is 40 times the volume of WCO) and the FFA was reduced to 1.0%. Table 1 shows the various fuel characteristics of WCO, algae oil and their blend used in this experimental study.

2.2. Method

This research intends to increase the transesterification of a mixture or blend of WCO and algae oil with respect to three process variables. Based on the literature review, the ranges of these parameters were decided [27]. These factors are: (a) reaction duration (60.00 to 180.00 min); (b) catalyst concentration (NaOH) (0.0% to 2.0% w/w); and (c) volumetric ratio of the methanol-oil blend (20.0% to 60.0% v/v). Due to the high reactivity of homogeneous catalysts, NaOH was used in the study [28]. Jain et al. [29] looked into the kinematics of transesterification of WCO and found 50 °C to be the optimum reaction temperature for acquiring a higher FAME yield. Thus, the transesterification reaction was performed at a steady RT of 50 °C. The experiments were created with RSM and the Box-Behnken Design (BBD) technique [30]. The statistical analysis and optimizations were carried out using Design of experiments

2.3. Experimental Procedure

Design of Experiments

There were a total of 17 trials designed with RSM based on BBD to obtain the optimal combinations of three process variables. The three varying parameters stated earlier were evaluated to determine their ideal blend using this experimental design. Table 2 displays the three parameters, A (reaction time (Rt) in minutes), B (methanol to oil ratio in volume percent) and C (catalyst amount in w/w%), at three different stages of the reaction mechanism within the predetermined range. A list of 17 experiments designed to gain experimental results for biodiesel yield is shown in Table 3.

2.4. Transesterification

The blend of WCO and algae oil was trans-esterified by adding methanol (CH3OH), a reactant with a base catalyst, i.e., sodium hydroxide (NaOH), for the production of methyl ester. With the aid of a heating plate stirrer and a batching reactor, all of the planned tests were carried out at a maintained temperature of 50 °C as illustrated in Figure 1. To produce pure biodiesel, pure FAME and glycerol layers were segregated by applying a separator funnel.

2.5. Statistical Analysis

An analysis of variance (ANOVA), a regression analysis and the development of the response-surface graphs were the three analytical stages that were used to analyze the relationship among the varying parameters and the response yield. With a confidence level of 95.0% (p = 0.05), Design Expert-11 has been used to assess the statistically significant results of the entire quadratic polynomial model. As the response values for the tests indicated, the highest experimental yield percentage rates were used. Then the developed model was employed for analyzing the interactive impact of individual parameters. Surface regression analysis was also conducted using the polynomial equation (Equation (1)):
R = β 0 + j = 1 k β j   P j + j = 1 k β j j   P j 2   + i = 1 j 1 i = 2 k   β i j P i P j + ε
where Pi and Pj are the uncoded independent parameters, ε is the random disturbance term, k is the number of independent parameters maximized for testing (k = 3 in this work) and β0 is the regression coefficient. R is the result (% Fatty-Acid Methyl-Ester Yield).

3. Results and Discussions

3.1. Transesterification Process

During the transesterification of the blend, two distinct response yields were obtained: the actual and expected value. The 17 trials yielded the empirical yield and RSM’s analysis of the experimental outputs led to the prediction of the yield. The three independent process parameters, namely A: reaction time (Rt), B: methanol/oil proportion and C: catalyst concentration (CC), are listed in Table 4. These two response rates are expressed as percentages. The link between both the outcome and the input parameters was characterized using second-order polynomial equations, including interaction effects, by using multiple regression analysis on the experimental response. Equation (2) displays the finalised techniques developed by coded factors.
Yield % = + 28.6475 + 0.1810 A 0.6176 B + 98.3164 C + 0.01186 A   B 0.0512 A C 0.3175 B C 2.4063 10 3 A 2 0.0107 B 2 27.8689 C 2
where A is Rt, B is Methanol/Oil proportion and C is CC.
The graph was plotted for the anticipated value and actual value of response yield (%) as presented in Figure 2. The graph shows that the anticipated values around the zero-error line are scattered at lower values of yield, with some of the points of higher yield values being near the experimental values. Therefore, it confirms the validity of the model created to set up a significant relationship between the response yield of the oil blend and the three independent process variables.

3.2. ANOVA

The most appropriate method for determining how accurate the tests and modelling are is analysis of variance (ANOVA). Variance analysis for performed experiments is shown in Table 5. The quadratic model was formed and analysed using RSM. The model was found to be significant and the lack of fit was insignificant. The study of the coefficient of determination (R2) provided more insight into the model’s suitability. A model can be deemed appropriate specifically if the R2 value is higher than 0.95, showing that the model can account for up to 95% of the variability of the response data. The observed R2 value of 0.9904 demonstrates that the parameter estimates and experimentally obtained results correspond well. Additionally, there was a 0.2 discrepancy between the anticipated R2 and modified R2 values, which is a respectable level of agreement.

3.3. Effects of Process Variables on Yield

Figure 3 depicts how changes in reaction time (Rt), methanol/oil ratio and catalyst concentration (CC) affect the blend’s responsive yields. As Rt of the method increases (curve A), the yield shows an increasing trend, but after reaching the point of optimum yield, a small decrease in yield % was observed. Due to excess Rt, the biodiesel yield decreases because of the occurrence of the reverse reaction. At the initial level, the methanol/oil ratio (curve B) results in a high biodiesel yield. The decreasing trend of biodiesel yield was observed with any further increase in the methanol/oil ratio from the initial level. The figure shows that increasing the catalyst concentration (curve C) in the reaction raises the initial response yield of the blend. When the catalyst percentage was raised after the optimal yield point, a pattern toward lower yield was seen. This reaction yield behaviour may be the result of adding too much catalyst. Excessive catalysts speed up triglyceride conversions and prevent the creation of water, which could increase the probability of a secondary reaction producing soap.

3.4. Effect of Reaction Parameters on Yield %

Utilizing the Box-Behnken Modeling approach in the Design of experiments technique, three distinct 3D graphs were created in order to explore the effects of individual variables, or processing parameters, on related parameters, or the response yield of the mix. The area contained within the smallest oval in the contour graphic provided the maximum expected values for the response yield of the mix as determined by the program. The interaction of various independent components in this procedure resulted in these elliptical outlines. As seen in Figure 4, Figure 5 and Figure 6, the two sequential variables were selected at their minimal level to create the three 3D representations. The curve surface inside the shortest ellipse in the contours graphic illustrates the interaction between each pair of processing parameters, AB, BC and AC.

3.4.1. Effect of Reaction Time on Yield %

To study the effect of reaction time (Rt) on biodiesel yielding (%), Rt was varied from 60–180 min. Initially, the yield was very high when the reaction was performed for 60 min. This may be because reactants required a certain amount of minimal time to react i.e., the activation energy. Identical outcomes were recorded by Zhao et al. [31]. On further increasing the reaction time from 60 to 120 min and then to 180 min, a decreasing trend was observed for the yield % for all the ranges of catalyst concentration and methanol/ratio. This decrease can be attributed to the fact that the transesterification process is reversible [32]. Thus, beyond the optimum point, the decrease in yield % can be observed.

3.4.2. Effect of Methanol/Oil Ratio on Yield %

To study the effect of methanol concentration on yield %, the methanol/oil ratio was varied from 20–60%. The highest yield of 91.90% was obtained using a methanol concentration of 20%. It was found that, on further raising the methanol content, there was a decrease in yield % recorded. The reversible nature of transesterification necessitates a greater alcohol concentration in order to move the reaction along [33], but in the current study, this decrease may be because adding excess methanol dissolves the glycerol into methanol. The side product glycerol formed during the reaction decreases the reactivity of methanol with the fatty acids. The same result was observed by Behera et al. [34]. The biodiesel yield decreased beyond the methanol concentration of 20:1. Recovery of excess methanol is also an energy-consuming process [35]. Thus, the optimum amount of methanol/oil ratio is required for high yield.

3.4.3. Effect of Catalyst Concentration on Yield %

The catalyst concentration (CC) having g significant effect on yield %, in this study the catalyst amount had a range from 0–2%. The results indicated that the yield % improves rapidly when raising the catalyst amount from 0–1%. At lower concentrations, it was observed that the increase in yield % was very low, but with increasing catalyst concentration there was a high-rate reaction process which increased the yield %. Beyond the optimum point, the yield % increased very slowly until, after a point, a slowdown in yield % was found. The same trend was observed by Bhatia et al. [36]. The reason behind this decrease may be the cause of soap formation due to excess catalyst, which negatively affects the yield % [37].

4. Optimization of Response Yield

In addition, the optimizing method was used on every response to get the most out of the blending in terms of responses. The optimization process was based on the quadratic model (Equation (2)) made by the RSM. The conclusion of the optimal value of processing parameters for the maximized biodiesel yield, which is 95.15%, is shown in Table 6. The experimental study was conducted relying on the optimal values of the process variables as indicated in Table 6 to verify the findings at constant temperature of 50 °C. Table 7 also showed that the projected optimized response was closer to the experimentally optimized solution with a 3.314% error. Finally, a 92% response yield for the biodiesel blends was reached employing a 21:1 methanol/oil ratio and a 1.5% w/w NaOH catalyst in 110 min at 50 °C. According to ASTM D-6751 and IS-15607, physicochemical testing of biodiesel samples was performed, results are shown in Table 8 and according to this table it is found that the biodiesel prepared on optimized parameters has physicochemical properties within the range mentioned by IS-15607.

5. Conclusions

This study focused on the single-step transesterification method of biodiesel production from a blend of high-FFA WCO and low-FFA algal oil using a homogeneous base catalyst. Because WCO contains high levels of FFA and a two=-step transesterification has to be used for converting this high FFA oil to biodiesel which consumes time and energy, therefore algae oil was chosen to blend with WCO in a proportion such that the resulting oil contains 1% or less FFA. By using this process, WCO can be used in combination with algal oil to produce biodiesel using single step transesterification. Three process variables, catalyst concentration (0–2 wt%), methanol concentration (v/v) (20–60%) and reaction time (60–180 min) at a constant reaction temperature of 50 °C, were selected. The design and optimization studies were conducted using the Box-Behnken Design Technique in Response Surface Methodology (RSM). The maximum biodiesel yield of 92% is obtained under optimized conditions of catalyst concentration of 1.5% (w/w), methanol/oil ratio of 21:1 and reaction time of 110 min. RSM can be successfully used to model the reaction in order to maximize the biodiesel yield. By using this process, one can eliminate the time and energy consuming acid catalyzed transesterification process used for FFA reduction.

Author Contributions

The main contributions of the authors are as follows: conceptualization: S.J., N.K., M.B., V.P.S. and S.M.; methodology: N.K.S., N.K., M.B. and S.M.; structuring of paper: T.M.Y.K., S.J., M.B. and V.P.S.; investigation and data collection: M.B., N.K. and S.M.; resources: V.P.S., S.J., M.B., T.M.Y.K. and S.M.; data curation: M.B., N.K. and S.M.; writing—original draft: N.K., S.J., M.B. and V.P.S.; writing—review and editing: S.J., M.B. and N.K.S.; visualization: M.B., V.P.S. and S.J.; supervision: M.B., S.J. and S.M.; project administration: M.B., S.J. and T.M.Y.K. 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.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Saini, M.; Sharma, A.; Singh, V.P.; Dwivedi, G.; Jain, S. Solar Thermal Receivers—A Review. Adv. Mater. Manuf. Energy 2022, 2, 310–325. [Google Scholar] [CrossRef]
  2. Singh, V.P.; Jain, S.; Karn, A.; Kumar, A.; Dwivedi, G.; Meena, C.S.; Dutt, N.; Ghosh, A. Recent Developments and Advancements in Solar Air Heaters: A Detailed Review. Sustainability 2022, 14, 2149. [Google Scholar] [CrossRef]
  3. Lo, S.L.Y.; How, B.S.; Leong, W.D.; Teng, S.Y.; Rhamdhani, M.A.; Sunarso, J. Techno-economic analysis for biomass supply chain: A state-of-the-art review. Renew. Sustain. Energy Rev. 2021, 135, 110164. [Google Scholar] [CrossRef]
  4. Verma, D.; Fortunati, E.; Jain, S.; Zhang, X. Biomass, Biopolymer-Based Materials, and Bioenergy: Construction, Biomedical, and Other Industrial Applications, 1st ed.; Elsevier Science: Amsterdam, The Netherlands, 2019. [Google Scholar]
  5. Singh, V.P.; Jain, S.; Karn, A.; Dwivedi, G.; Alam, T.; Kumar, A. Experimental Assessment of Variation in Open Area Ratio on Thermohydraulic Performance of Parallel Flow Solar Air Heater. Arab. J. Sci. Eng. 2022, 41, 1–17. [Google Scholar] [CrossRef]
  6. Ambaye, T.G.; Vaccari, M.; Bonilla-Petriciolet, A.; Prasad, S.; van Hullebusch, E.D.; Rtimi, S. Emerging technologies for biofuel production: A critical review on recent progress, challenges and perspectives. J. Environ. Manag. 2021, 290, 112627. [Google Scholar] [CrossRef]
  7. Jain, S.; Sharma, M. Engine performance and emission analysis using oxidatively stabilized Jatropha curcas biodiesel. Fuel 2013, 106, 152–156. [Google Scholar] [CrossRef]
  8. Singh, D.; Sharma, D.; Soni, S.; 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]
  9. Agarwal, A.K.; Gupta, J.G.; Dhar, A. Potential and challenges for large-scale application of biodiesel in automotive sector. Prog. Energy Combust. Sci. 2017, 61, 113–149. [Google Scholar] [CrossRef]
  10. Abomohra, A.E.-F.; Elsayed, M.; Esakkimuthu, S.; El-Sheekh, M.; Hanelt, D. Potential of fat, oil and grease (FOG) for biodiesel production: A critical review on the recent progress and future perspectives. Prog. Energy Combust. Sci. 2020, 81, 100868. [Google Scholar] [CrossRef]
  11. Garg, A.; Dwivedi, G.; Baredar, P.; Jain, S. Current Status, Synthesis, and Characterization of Biodiesel. In Biodiesel Technology and Applications; Wiley: Hoboken, NJ, USA, 2021. [Google Scholar]
  12. Jain, S.; Sharma, M.P. Oxidation, Thermal, and Storage Stability Studies of Jatropha Curcas Biodiesel. ISRN Renew. Energy 2012, 2012, 1–15. [Google Scholar] [CrossRef]
  13. Singh, V.P.; Jain, S.; Karn, A.; Dwivedi, G.; Kumar, A.; Mishra, S.; Sharma, N.K.; Bajaj, M.; Zawbaa, H.M.; Kamel, S. Heat transfer and friction factor correlations development for double pass solar air heater artificially roughened with perforated multi-V ribs. Case Stud. Therm. Eng. 2022, 39, 102461. [Google Scholar] [CrossRef]
  14. Goyal, P.; Sharma, M.P.; Jain, S. Optimization of transesterification of Jatropha curcas Oil to Biodiesel using Response Surface Methodology and its Adulteration with Kerosene. J. Mater. Environ. Sci. 2013, 4, 277–284, 2013. [Google Scholar]
  15. Goh, B.H.H.; Chong, C.T.; Ge, Y.; Ong, H.C.; Ng, J.-H.; Tian, B.; Ashokkumar, V.; Lim, S.; Seljak, T.; Józsa, V. Progress in utilisation of waste cooking oil for sustainable biodiesel and biojet fuel production. Energy Convers. Manag. 2020, 223, 113296. [Google Scholar] [CrossRef]
  16. Borges, M.; Díaz, L. Recent developments on heterogeneous catalysts for biodiesel production by oil esterification and transesterification reactions: A review. Renew. Sustain. Energy Rev. 2012, 16, 2839–2849. [Google Scholar] [CrossRef]
  17. Narula, V.; Thakur, A.; Uniyal, A.; Kalra, S.; Jain, S. Process parameter optimization of low temperature transesterification of algae-Jatropha Curcas oil blend. Energy 2017, 119, 983–988. [Google Scholar] [CrossRef]
  18. Mohadesi, M.; Aghel, B.; Maleki, M.; Ansari, A. The use of KOH/Clinoptilolite catalyst in pilot of microreactor for biodiesel production from waste cooking oil. Fuel 2020, 263, 116659. [Google Scholar] [CrossRef]
  19. Hsiao, M.-C.; Kuo, J.-Y.; Hsieh, S.-A.; Hsieh, P.-H.; Hou, S.-S. Optimized conversion of waste cooking oil to biodiesel using modified calcium oxide as catalyst via a microwave heating system. Fuel 2020, 266, 117114. [Google Scholar] [CrossRef]
  20. Aghel, B.; Mohadesi, M.; Ansari, A.; Maleki, M. Pilot-scale production of biodiesel from waste cooking oil using kettle limescale as a heterogeneous catalyst. Renew. Energy 2019, 142, 207–214. [Google Scholar] [CrossRef]
  21. Degfie, T.A.; Mamo, T.T.; Mekonnen, Y.S. Optimized Biodiesel Production from Waste Cooking Oil (WCO) using Calcium Oxide (CaO) Nano-catalyst. Sci. Rep. 2019, 9, 1–8. [Google Scholar] [CrossRef] [Green Version]
  22. Komintarachat, C.; Chuepeng, S. Catalytic enhancement of calcium oxide from green mussel shell by potassium chloride impregnation for waste cooking oil-based biodiesel production. Bioresour. Technol. Rep. 2020, 12, 100589. [Google Scholar] [CrossRef]
  23. Mardhiah, H.H.; Ong, H.C.; Masjuki, H.; Lim, S.; Lee, H. A review on latest developments and future prospects of heterogeneous catalyst in biodiesel production from non-edible oils. Renew. Sustain. Energy Rev. 2017, 67, 1225–1236. [Google Scholar] [CrossRef]
  24. Singh, B.; Jain, S.; Gangil, B. Effectiveness of Homogeneous and Heterogeneous Catalyst on Biodiesel Yield: A Review. Adv. Clean Energy Technol. 2021, 375–385. [Google Scholar] [CrossRef]
  25. Mohadesi, M.; Aghel, B.; Maleki, M.; Ansari, A. Production of biodiesel from waste cooking oil using a homogeneous catalyst: Study of semi-industrial pilot of microreactor. Renew. Energy 2019, 136, 677–682. [Google Scholar] [CrossRef]
  26. Shadidi, B.; Najafi, G.; Zolfigol, M.A. A Review of the Existing Potentials in Biodiesel Production in Iran. Sustainability 2022, 14, 3284. [Google Scholar] [CrossRef]
  27. Verma, P.; Sharma, M. Review of process parameters for biodiesel production from different feedstocks. Renew. Sustain. Energy Rev. 2016, 62, 1063–1071. [Google Scholar] [CrossRef]
  28. Abdullah, S.H.Y.S.; Hanapi, N.H.M.; Azid, A.; Umar, R.; Juahir, H.; Khatoon, H.; Endut, A. A review of biomass-derived heterogeneous catalyst for a sustainable biodiesel production. Renew. Sustain. Energy Rev. 2017, 70, 1040–1051. [Google Scholar] [CrossRef]
  29. Jain, S.; Sharma, M.; Rajvanshi, S. Acid base catalyzed transesterification kinetics of waste cooking oil. Fuel Process. Technol. 2011, 92, 32–38. [Google Scholar] [CrossRef]
  30. Garg, A.; Jain, S. Process parameter optimization of biodiesel production from algal oil by response surface methodology and artificial neural networks. Fuel 2020, 277, 118254. [Google Scholar] [CrossRef]
  31. Zhao, S.; Niu, S.; Yu, H.; Ning, Y.; Zhang, X.; Li, X.; Zhang, Y.; Lu, C.; Han, K. Experimental investigation on biodiesel production through transesterification promoted by the La-dolomite catalyst. Fuel 2019, 257, 116092. [Google Scholar] [CrossRef]
  32. Supraja, K.V.; Behera, B.; Paramasivan, B. Optimization of process variables on two-step microwave-assisted transesterification of waste cooking oil. Environ. Sci. Pollut. Res. 2020, 27, 27244–27255. [Google Scholar] [CrossRef]
  33. Izida, T.; Silva, J.; Andrade, L.; Simionatto, E.; Scharf, D.; Lima, S. Modeling transesterification reaction kinetics using fluorescence spectroscopy to interpret biodiesel production. Chem. Eng. Sci. 2020, 211, 115292. [Google Scholar] [CrossRef]
  34. Behera, B.; Selvam S, M.; Dey, B.; Balasubramanian, P. Algal biodiesel production with engineered biochar as a heterogeneous solid acid catalyst. Bioresour. Technol. 2020, 310, 123392. [Google Scholar] [CrossRef] [PubMed]
  35. Rezania, S.; Oryani, B.; Park, J.; Hashemi, B.; Yadav, K.K.; Kwon, E.E.; Hur, J.; Cho, J. Review on transesterification of non-edible sources for biodiesel production with a focus on economic aspects, fuel properties and by-product applications. Energy Convers. Manag. 2019, 201, 112155. [Google Scholar] [CrossRef]
  36. Bhatia, S.K.; Gurav, R.; Choi, T.-R.; Kim, H.J.; Yang, S.-Y.; Song, H.-S.; Park, J.Y.; Park, Y.-L.; Han, Y.-H.; Choi, Y.-K.; et al. Conversion of waste cooking oil into biodiesel using heterogenous catalyst derived from cork biochar. Bioresour. Technol. 2020, 302, 122872. [Google Scholar] [CrossRef] [PubMed]
  37. Kumar, S.A.; Sakthinathan, G.; Vignesh, R.; Banu, J.R.; Al-Muhtaseb, A.H. Optimized transesterification reaction for efficient biodiesel production using Indian oil sardine fish as feedstock. Fuel 2019, 253, 921–929. [Google Scholar] [CrossRef]
Figure 1. Representation of the reactor’s schematics.
Figure 1. Representation of the reactor’s schematics.
Energies 16 01293 g001
Figure 2. Comparison of theoretical and measured yield percent values.
Figure 2. Comparison of theoretical and measured yield percent values.
Energies 16 01293 g002
Figure 3. Effect of CC, Rt and methanol/oil ratio on yield (%).
Figure 3. Effect of CC, Rt and methanol/oil ratio on yield (%).
Energies 16 01293 g003
Figure 4. Biodiesel yield (%) vs. methanol/oil ratio (%) and Rt (min).
Figure 4. Biodiesel yield (%) vs. methanol/oil ratio (%) and Rt (min).
Energies 16 01293 g004
Figure 5. Biodiesel yield (%) vs. CC (%) and methanol/oil ratio (%).
Figure 5. Biodiesel yield (%) vs. CC (%) and methanol/oil ratio (%).
Energies 16 01293 g005
Figure 6. Biodiesel yield (%) vs. CC (%) and reaction time (minutes).
Figure 6. Biodiesel yield (%) vs. CC (%) and reaction time (minutes).
Energies 16 01293 g006
Table 1. Properties of Waste Cooking oil & algae-oil.
Table 1. Properties of Waste Cooking oil & algae-oil.
S.NO.PropertiesWaste Cooking OilAlgae OilBlend of Waste Cooking Oil and Algae-Oil
1.Flash point (°C)256.00102.00105.85
2.FFA (%)20.000.501.00
3.Density (kg/m3)0.9190.860.8614
4.Viscosity (cst) @40 °C39.421.0021.46
Table 2. Blend oil transesterification via BBD with independent parameters.
Table 2. Blend oil transesterification via BBD with independent parameters.
ParametersSymbolUnitLevels
−101
Reaction time (Rt)AMinutes60120180
Methanol/oil ratioB%204060
Catalyst-concentration (CC)C%000102
Table 3. List of trials designed with BBD.
Table 3. List of trials designed with BBD.
Run No.Time
(in Minutes)
Methanol/Oil Ratio (%)Catalyst-Concentration
(CC) (%)
ABC
1180400
260601
3120602
460201
5120401
6120401
7180601
8120202
9120401
10180402
1160402
1260400
13180201
14120200
15120401
16120600
17120401
Table 4. Observations to the combination of waste cooking oil and algae oil being trans-esterified.
Table 4. Observations to the combination of waste cooking oil and algae oil being trans-esterified.
RunTime (Minutes)Methanol/Oil Ratio (%)Catalyst Concentration (CC) (%)Blend Biodiesel Yield (%)Error (%)
ExperimentalPredicted
ABC
118040020.826.7822.33
26060131.0126646.1132.742
312060269.2307760.14−15.116
46020191.9055689.46−2.7336
512040182.812582.36−0.5494
612040180.7142982.361.998
718060175.3333377.813.1829
812020279.2792887.719.6120
912040181.9548982.360.4918
1018040261.4285768.109.7965
116040276.9784271.02−8.3897
126040024.0601517.41−38.197
1318020179.2792864.20−23.487
1412020018.4210527.5433.1116
1512040183.3333382.36−1.1817
1612060033.7748325.37−33.129
1712040182.9268382.36−0.6882
Table 5. Evaluating the RSM modeling.
Table 5. Evaluating the RSM modeling.
S.No.ParametersValue
1.ModelSignificant
2.Lack of fitNon-Significant
3.R20.9904
4.Adjusted R20.9745
5.Predicted R20.9044
Table 6. Optimizing for optimum possible biodiesel production.
Table 6. Optimizing for optimum possible biodiesel production.
ParametersObjectiveOptimized Value RSM
Reaction time (Rt)In-range110
Methanol volumetric ratioMinimized21
Catalyst concentration (CC)In-range1.5
Biodiesel-yieldMaximized95.15
Table 7. Results of model evaluation under optimal conditions.
Table 7. Results of model evaluation under optimal conditions.
S.No.Methanol/Oil Volumetric RatioReaction-Temperature (°C)Catalyst-Loading (wt%)Reaction-Time (min)Predicted-Biodiesel Yield (%)Experimental-Biodiesel Yield (%)Error (%)
121501.511095.1592−3.423
Table 8. Physicochemical properties of biodiesel.
Table 8. Physicochemical properties of biodiesel.
S.NO.Property (Unit)IS-15607IS-15607 LimitsBiodiesel
1Flashpoint (°C)IS 1448-105.85
2Viscosity at 40 °C (cSt)IS 14486.184.20
3Water and sediment (vol%)D: 2709Max 0.050.04
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

Jain, S.; Kumar, N.; Singh, V.P.; Mishra, S.; Sharma, N.K.; Bajaj, M.; Khan, T.M.Y. Transesterification of Algae Oil and Little Amount of Waste Cooking Oil Blend at Low Temperature in the Presence of NaOH. Energies 2023, 16, 1293. https://doi.org/10.3390/en16031293

AMA Style

Jain S, Kumar N, Singh VP, Mishra S, Sharma NK, Bajaj M, Khan TMY. Transesterification of Algae Oil and Little Amount of Waste Cooking Oil Blend at Low Temperature in the Presence of NaOH. Energies. 2023; 16(3):1293. https://doi.org/10.3390/en16031293

Chicago/Turabian Style

Jain, Siddharth, Nitin Kumar, Varun Pratap Singh, Sachin Mishra, Naveen Kumar Sharma, Mohit Bajaj, and T. M. Yunus Khan. 2023. "Transesterification of Algae Oil and Little Amount of Waste Cooking Oil Blend at Low Temperature in the Presence of NaOH" Energies 16, no. 3: 1293. https://doi.org/10.3390/en16031293

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