Captcha Page

We apologize for the inconvenience...

To ensure we keep this website safe, please can you confirm you are a human by ticking the box below.

If you are unable to complete the above request please contact us using the below link, providing a screenshot of your experience.

https://ioppublishing.org/contacts/

  • Open access
  • Published: 15 September 2022

Optimization and kinetics studies of biodiesel synthesis from Jatropha curcas oil under the application of eco-friendly microwave heating technique: an environmentally benign and sustainable bio-waste management approach

  • Kassian T. T. Amesho 1 , 2 ,
  • Yuan-Chung Lin 1 , 3 , 4 , 2 ,
  • Chin-En Chen 1 , 2 ,
  • Pei-Cheng Cheng 1 , 2 &
  • Vinoth Kumar Ponnusamy 5 , 6 , 2  

Sustainable Environment Research volume  32 , Article number:  41 ( 2022 ) Cite this article

3258 Accesses

9 Citations

Metrics details

This study attempts to synthesize biodiesel as a green liquid fuel from Jatropha curcas oil (JCO) utilizing waste eggshell (WES) as an effective and excellent sustainable source of the heterogeneous catalyst under the application of environmentally benign microwave heating technique. After preparing the CaO-based catalyst, diverse characterization techniques such as X-Ray Diffraction, Energy Dispersive Spectroscopy, Scanning Electron Microscopy, Fourier Transform Infrared, and Brunauer–Emmett–Teller were employed for characterization measurements. Under different optimized conditions, a considerable high biodiesel yield of 92% was attained while employing the following reaction conditions; methanol to oil molar ratio of 9:1, 5 wt% catalyst loading, 165 min reaction time with a microwave power of 800 W, and a 65 °C reaction temperature. The developed catalyst had significantly retained its reusability up to the 5 th cycle of reuse. The catalysed transesterification process's activation energy of 38.5 kJ mol −1 demonstrated that the reaction is chemically controlled. Therefore, the WES has been utilized as a renewable base heterogeneous catalyst for novel biodiesel synthesis from JCO, which can be applied in diesel engines to lessen air pollution, specifically pollutant emissions from diesel vehicles. The results of this study are not for academic purposes only. They can also serve as models for industrial biodiesel production by exploiting bio-waste as catalysts and non-edible oils as feedstocks in microwave heating systems as environmentally friendly chemistry systems. In addition, our study uses non-consumable oil feedstock and bio-waste materials in an economical way to produce biofuel while contributing to environmental sustainability and sustainable bio-waste management. This approach extends to the existing state-of-the-art research.

1 Introduction

The fast depletion of fossil-fuel-based resources and global warming as a result of climate change ramifications are the major driving forces towards the development of renewable and more environmentally benign energy resources. The forever fast-growing population has also prompted this, along with booming industrialization across the globe, which have similarly led to a serious shortage, uncertainty and rigorous ecological impact of fossil fuel-based energy utilization [ 1 , 2 ]. In response to this, biodiesel has consequently been identified as one of the most promising green energy that can replace the unsustainable petrol diesel utilisation due to its vast environmental benefits [ 3 , 4 ].

Biodiesel as a green fuel has been broadly considered as an eco-friendly source of energy, owing to its advantage in producing less CO 2 , SOx, and hydrocarbon emissions, which contribute to environmental sustainability [ 5 ]. Conversely, over 95% of biodiesel is mainly synthesized from most notably edible oil crops (such as palm, rapeseed, soybean and sunflower), which compromise food security [ 6 ]. In view of this, various stakeholders such as government agencies, scientists, industries, and policymakers have aimed for the exploration of economical and non-edible oil-producing energy crops like Jatropha curcas for research.

Being an energy crop, J. curcas de-oiled seed cake can be used as a feedstock for the production of bio-oil via the pyrolysis process [ 7 ]. The J. curcas bio-oil has diverse potential for industrial applications. Among others, J. curcas bio-oil can also be used as bio dielectric fluids in the Electric Discharge Machining process. It is also reported that J. curcas bio-oil is an excellent substitute of the conventional hydrocarbon oils [ 7 , 8 ].

Meanwhile, shell waste, which is identified as inevitable food waste materials formed in food preparation or consumption, has become a global environmental problem, specifically concerning the issue of waste management. In this regard, eggshells waste is generated from various food industries such as egg processing companies, and enormous amounts of waste shells that are discarded to the landfills or incinerated, have virtually become a significant cause for organic pollution.

Eggshells are a classic example of food waste emanating from the food processing industry that still contains recyclable parts. By 2030, global egg production will increase to around 90 Mt. Because eggshells are considered worthless, most of this waste is usually disposed of in landfills without being converted into valuable materials. However, managing these wastes requires appropriate strategies that take into account the increasing cost of disposal, environmental concerns associated with the risk of spreading pathogens, odours, and where they are disposed of [ 8 , 9 , 10 ]. In addition, eggshells as an agro-waste have the potential to cause environmental pollution according to European Union regulations. This is because the large amounts of the discarded eggshells contribute to food wastage which causes significant damage to the environment by increasing global carbon footprint when buried, which is one of the prime greenhouse gases contributing to the global warming [ 11 ]. Therefore, it is essential to find an alternative technique to transform eggshells into valued materials for further utilization.

Essentially, eggshells are considered high-value natural materials as opposed to waste, considering their most efficient calcified shells, comprised of over 96% calcium carbonate (CaCO 3 ). Eggshells are predominantly rich in CaCO 3 that mostly appears in the mineral calcite form, a better-decontaminated and thermodynamically persistent form of CaCO 3 , which has a much lower level of contaminations. In view of this, eggshells could offer enormous opportunities for applications in the form of limestone (CaCO 3 ), if not lime (CaO), in a wide range of applications, through significant reusability as a substitute in the production of cement. In addition, eggshells signify an excellent substitute which is vastly plentiful, low-cost, and more sustainable source of filler for the rubber industry [ 4 , 6 , 11 ]. CaO, on the other hand, has a high bonding density with H 2 O and CO 2 , so it is easily degraded during the aqueous reaction [ 12 , 13 ]. CaO also demonstrated improved stability and catalytic efficiency in biodiesel synthesis processes due to its high basic strength, low cost and inadequate dissolvability in methanol. CaO is also the cheapest and most durable heterogeneous catalyst [ 14 ]. Due to these properties, CaO was obtained from eggshells [ 5 , 15 ], scallop shells [ 16 ], cockles shells (Anadara granosa) [ 17 ], and animal bones [ 18 ]. From this perspective, reusing eggshell waste for many purposes will benefit the environment and the economy. Furthermore, the reprocessing and recycling of waste eggshells (WES) integrates the fundamental principles of circular economy, which increases resource utilization efficiency through transforming waste or by-products into resources with both cost-effective and ecological beneficial impact [ 19 ].

Many researchers have been investigating the application of the microwave (MW) technique for the synthesis of biodiesel. It has been widely demonstrated that the MW heating system is one type of clean, fast, and suitable energy source that can expedite the procedure and enhance the selectivity of particular reactions [ 4 , 5 ]. Lin et al. [ 5 ] reported the exploitation of ionic liquid as a green catalyst for the transesterification procedure in the MW heating system, giving a quick way to biodiesel synthesis from J. curcas oil (JCO). The MW heating system is equally more energy-efficient or energy-saving and more economical than the conventional heating system [ 5 ].

In the light of the above-mentioned discussion, the present study attempts to promote the efficiency and synthesis of bio-waste derived CaO-based heterogeneous catalyst and transesterification of JCO under the application of environmentally benign MW heating technique as a sustainable bio-waste management approach. The developed technique can tremendously promote the integration of the circular economy concept through an efficient biorefinery model.

2 Experimental section

2.1 materials and reagents.

WES were acquired from a fast-food establishment in Taiwan. The non-consumable oil (i.e., JCO) as a raw material was provided by the Chinese Petroleum Corporation (CPC), Taiwan. Methanol and sodium hydroxide were procured from Burdick & Jackson (UNI-ONWARD, Taiwan). The methanol and sodium hydroxide used in our experimentations were of excellent analytical grade, and the purity of all solid chemicals were at least 99.5%. The feedstock was subsequently heated to remove the moisture content, whereas all other chemicals were used as received without further decontamination.

2.2 Catalyst synthesis

2.2.1 calcination of eggshells.

The collected WES were washed thoroughly with tap water to eliminate dirt, impurities, and organic materials. Subsequently, the WES were further rinsed using distilled water at least 2–3 times and, after that, stored in a hot air oven at the temperature of 110–140 °C, for approximately 24 h. At this stage, the desiccated eggshells were removed from the oven for further processing and reduced the materials to a fine powder due to continuous crushing in an agate mortar. Lastly, the calcination was conducted in a high-temperature furnace at 1000 °C for 4 h and consequently placed in a desiccator. Thus, the acquired product was CaO, which was kept in a cool, dark desiccator at a temperature in the range of 20–38 °C.

2.3 The transesterification procedures for the JCO

In this research work, an MW synthesis reactor (PreeKem APEX, Shanghai, China) fitted with an automated stirrer, and a capacitor was employed to enable MW reactions. Figure  1 shows the experimental system for this study. The stirrer was performed at 600 rpm with a hypnotizing (magnetic) midpoint. As a result, the eggshells were exploited to develop a heterogeneous catalyst, a product that has been calcined (CaO constituent), to achieve the waste products (discarded materials) reutilizing objective.

figure 1

Schematic diagram of the experimental setup (microwave system structure). (Reprinted with permission from Lin et al. [ 5 ]). Copyright (2020), Elsevier [License Number: 5270740247480]

JCO and methanol were blended in the vessels, and subsequently, the catalyst was consequently supplemented to the vessel. The vessel was placed in an MW heating system. The samples were performed at several reaction intervals and temperatures during this time. Various catalyst loading (3–7 wt%), reaction time (120–210 min), methanol to oil molar ratio (MTOMR) (9:1), and a temperature reaction of (45, 55, 65, 75, and 85 °C) were investigated during the experiments. With the application of the Fourier Transform Infrared (FTIR) technique, the JCO conversion to biodiesel was thoroughly examined. After 32 scans, the FTIR spectra of the catalyst samples were determined to be in the range of 500–4000 cm −1 . These FTIR spectra were accordingly validated in addition to the documentation aimed at every sample. Krishnamurthy et al. [ 20 ] quantified fatty acid methyl esters (FAME) in microwave-mediated transesterification biodiesel using FTIR analysis and identified peaks that may be associated with the typical C = O stretching of the esters. These peaks were found in the 1800–1700 cm −1 spectral region and were familiar to both the FAME and refined oil spectra. Foroutan et al. [ 21 ] conducted FTIR analysis to determine the functional groups of the studied catalyst and produced biodiesel from edible waste oil using a calcium oxide@magnesium oxide nanocatalyst. Analysis was performed on 400–4000 cm −1 band range.

2.4 Determination of acid value and separation procedures

The JCO sample used under this investigation was found to have an acid value of 2.8 mg KOH g −1 and a saponification value of 195 mg KOH g −1 . Additionally, the JCO sample had a free fatty acid (FFA) value of 2.5 wt% with an average molecular weight of 861 g mol −1 and water content of 0.12 wt% [ 4 , 6 ]. JCO sample was found to have a pH value of 5.8.

Under this investigation, the method on the transesterification reaction was anticipated to have had happened when the quantity of methyl ester was above 90%. In view of this, the mixture and assortment at the concluding phase of the transesterification process were separated with the help of a centrifuge running at 2000 rpm for 10 min, which resulted in the composition of an upper phase characterized by methyl ester and consequently a reduced phase composed of glycerin. This has also progressed into disproportionate methanol during the methyl ester stage, which was vaporized through heating at 80 °C. Afterwards, the catalyst was efficiently withdrawn with the support of silica gel supplemented to the reaction. Therefore, biodiesel was subsequently achieved, and its content was determined with the support of Gas Chromatography (GC-6890, Agilent, USA).

2.5 Product evaluation

The evaluation procedures employed in our experiments were validated in compliance with Taiwan CNS15051 for determining the amount of methyl ester. Additionally, the methyl ester content during this study was determined using the Gas Chromatography method furnished with the flame ionization detector (GC-FID). The conditions used for GC-FID to analyze the products and other instruments and methods used for other analysis such as X-ray Diffraction (XRD), Energy Dispersive Spectroscopy (EDS), Scanning Electron Microscopy (SEM), FTIR, and Brunauer–Emmett–Teller (BET), are reported in detail in our previous study [ 4 , 5 ]. The content of methyl ester in this study was determined by employing the following Eq. ( 1 ) as reported in our preceding studies [ 4 , 5 ]:

where ΣA is the total of the peak area of FAME from C 14 to C 24:1 ; A EI is the peak area of the internal standard, methyl heptadecanoate (C 18 H 36 O 2 ); C EI is the absorption of methyl heptadecanoate (mg mL −1 ); V EI is the quantity of methyl heptadecanoate (mL −1 ); and m is the quantity of input biodiesel (g). The methyl ester yield is expressed using Eq. ( 2 ) as reported in our earlier studies [ 4 , 5 ]:

where C is the methyl ester content (%); W B is the quantity of synthesized biodiesel (g); and W oil is the quantity of the initial volume of JCO (g). Consequently, biodiesel production was investigated and measured as part of the preliminary capacity of JCO by weight. A similar approach was used and demonstrated in our previous work [ 4 , 6 ].

2.6 Physico-chemical properties of biodiesel from JCO

Table S 1 of the Supplementary Materials shows the summarized physico-chemical properties of the JCO biodiesel. The biodiesel physico-chemical characteristics are unique, predominantly its density, which can substantially influence the engine's performance [ 6 ]. This emanates from the evidence that the fuel density can effectively split up the fuel spray from the injector. For that reason, the density at 15 °C for JCO biodiesel is mainly found to be within the scope and appropriate range of 0.859–0.891 g cm −3 for fuel standard (ASTM D6751) or EN 14,214 standard [ 22 , 23 ].

It can also be observed from Table S 1 that the JCO biodiesel could have a level of acidity of 0.18 mg KOH g −1 and saponification number of 191 mg KOH g −1 that are under the normal standard range 0.5–370 for biodiesel in consistent with the ASTM standards (ASTM D6751-02). The iodine value is used to establish the iodine absorption values in g, which could be engrossed by 100 g oil. It helps to determine the amount of unsaturation of biodiesel, and thus it is beneficial to examine the oil durability [ 6 ]. It is also worth noting that a significant proportion of desaturation can lead to fuel polymerization due to the development of epoxide in the buildup of oxygen in twofold links. As it can be observed from Table S 1 , the JCO biodiesel has an iodine value of 69 g I 2 100 g −1 oil that falls further down the recommended upper threshold of 120 (EN 14,214 standard). This iodine value is comparable with the results reported by Kumar et al. [ 22 ] for JCO biodiesel (75 g I2 100 g −1 ). Additionally, other greater iodine concentrations of 119 and 122 g I 2 100 g −1 were recorded in the study conducted by Sarma et al. [ 24 ] and Nath et al. [ 25 ]. JCO biodiesel is reported to have a higher heating value (calorific value) of 40.5 MJ kg −1 [ 7 ].

3 Results and discussion

3.1 catalyst characterization, 3.1.1 xrd analysis.

We studied the eggshells before and after calcination using the XRD (HR-XRD D8 SSS, Bruker) analysis technique. Figure S 1 illustrates the XRD patterns of the uncalcinated and calcinated eggshells before (Fig. S 1 a) and after calcination (Fig. S 1 b). According to the results, the occurrence and strong peaks were located at a brag angle of (2 θ) at 29.7, 39.3, 43.2 and 47.6° for the recycled WES before calcination (Fig. S 1 a). However, strong peaks were equally noticed after calcination (Fig. S 1 b) and were detected an angle of ( 2 θ) at 29.8, 34.1, 39.4, 43.3, 47.6 and 48.5° in accordance with the bands of CaO [ 6 ].

In the earlier investigations, it has been discovered that the snail shells were calcined at 900 °C and indicated the presence of CaO with strong peaks of brag angle ( 2 θ) located at several points, coordinating from 32.2, 37.3 as well as 53.8°. Conversely, the uncalcined snail shell band was similarly described to have peaks 2 θ located at several points coordinating at 26.2, 33.2 and 45.9°, suggesting that snail shell component comprised of CaCO 3 [ 26 ]. Risso et al. [ 15 ] examined the calcination of waste shells, calcined for the duration of 4 h at 900 °C. The grinded clamshell powder had eventually been transformed into a fine white powder, firmly demonstrating that it was converted into CaO. By comparison, the bands of XRD assessment were discovered at an angle of ( 2 θ) located at several degrees (32.5, 37.6, 54.1 and 62.6°), and this attests to the presence of CaO [ 6 ]. Takeno et al. [ 14 ] described cubic crystals in calcined Silver Croaker’s stone, showing that the calcination temperature and time were adequate to decompose all the carbonates into oxides. CaO has been designated as one of the most suitable catalysts for the transesterification of soybean oil [ 27 ].

3.1.2 FTIR analysis

The FTIR analysis technique was applied to characterize eggshells before calcination (Fig. S 2 a) and after calcination at at 1000 °C for 2 h (Fig. S2b). According to the results, pre-calcination absorption bands were situated at various peaks at 1397.0, 862.9 and 710.3 cm −1 . These types of bands are essentially caused by carbonate's existence on the surface of the catalyst [ 4 ].

Lastly, the absorption bands after calcination were located at 3650.9, 1397.1, 873.9, and 710.1 cm −1 . The expanded matrix oscillations of the CaO bond are indicated by large, strong absorption peaks of approximately 873.9 and 710.1 cm −1 . Asymmetric expansion and contraction of C-O bonds in unknown carbonate species was associated with peaks of 1397.1 and 873.9 cm −1 . There is a large peak at 3650.9 cm −1 due to the creation of -OH bonds at Ca(OH) 2 after H 2 O adsorption on the metal surface by CaO. This means that water is not the cause of catalyst deactivation [ 16 ]. It may also be due to Ca(OH) 2 forming soaps in acidic oils, which reduces yields in subsequent runs. These FTIR results acquired in the current study are comparable with related results described in our previous work [ 4 , 6 ].

3.1.3 EDS analysis

In this study, the EDS analysis was conducted with a single goal to determine particular elements that might be present in the produced catalyst. In light of this, it was revealed (after calcination) that individual chemical elements such as C (6.5%), O (19.1), and Ca (74.4%) as constituents of the developed CaO-based eggshells catalyst as displayed in Table 1 .

The determined EDS results are consistent with the XRD. Rahman et al. [ 28 ] reported that the catalyst performance could be potentially strengthened by the complementary effect of different metal oxide-based catalysts. The catalyst constituents are relatively compatible with the banana peel ash recounted by Balajii and Niju [ 29 ]. SEM diagrams in Fig.  2 show the WES material structure, prior calcination (Fig.  2 a) and after calcination (Fig.  2 b) at high temperature. The specific surface area (Micromeritics, ASAP 2020) and chemical configurations of the developed catalyst were determined. It is evidently illustrated from Fig.  2 that the pore structures in the eggshell showed extensive, substantial disparities before (Fig.  2 a) and after calcination (Fig.  2 b).

figure 2

SEM image of eggshell before ( a ) and after calcination ( b ) at high temperature

After calcination, there should be significant modifications in the shells' physical structures, signifying that the transformation of CaCO 3 into CaO has happened. SEM images of the developed CaO-based catalyst showed packs of appropriately structured cubic crystals with visible edges. These images are quite similar to the ones revealed by Nisar et al. [ 18 ] for calcined waste animal bones. The calcined bones were irregular, and some stuck together. For comparison, Sirisomboonchai et al. [ 16 ] developed nano-sized CaO-based catalysts from snail shells using SEM micrographs showing significant accumulation of catalyst particles due to a high specific area and a spherical structure without scattering morphology.

3.1.4 Brunauer–Emmett–Teller (BET)-Barrett-Joyner-Halenda (BJH) measurements

The dimensions for the surface area and the absorbency depiction of the developed catalyst were investigated using N 2 physisorption with a BET technique (AUTOSORB-1C Quantachrome). As for the adsorption of nitrogen gas employing the micrometrics equipment, the general pore capacity and average pore dimension of the developed catalysts were calculated using the BJH technique. The Hammett indicators were (acquired from Sigma-Aldrich, Kaohsiungy, Taiwan) used to measure the basic strength and the basicity of the catalyst samples. The outcomes are displayed in Table 2 . The analysis was conducted at a temperature range of 800–1000 °C. In this respect, the assessment via BET suggests that the developed catalyst’s surface area had considerably declined from 7.0 to 6.2 m 2  g −1 with a typical BJH adsorption pore measurement of pore size between 19 and 31 nm. Palitsakun et al. [ 30 ] reported a much lower surface area and pore volume measurements for CaO-T catalyst with a surface area of ​​1.95 m 2  g −1 and an average pore volume of 0.019 cm 3  g −1 , while the CaO-H catalyst had a surface area of ​​23.3 m 2  g −1 with an average pore volume of 0.201 cm 3  g −1 . The CaO-P catalyst was found to have a surface area of ​​1.76 m 2  g −1 and an average pore volume of 0.0155 cm 3  g −1 . Rahman et al. [ 28 ] have found that a solid catalyst's higher surface area can improve its catalytic performance.

3.2 Optimization of transesterification over the developed catalyst and exploring the optimum reaction conditions

3.2.1 effect of catalyst loading on biodiesel yield.

The catalyst effect substantially influences biodiesel yield, and thus, it is an indispensable parameter in optimising the transesterification reaction. On the other hand, when the percentage of free fatty acids (FFAs) in the feedstock is much high, the base catalyst can react with FFAs, resulting in saponification. This could make the product partition and purification process more complex and reduce the biodiesel yield [ 5 , 6 ]. Accordingly, various catalyst loadings on biodiesel yield were studied to examine their effect on biodiesel production yield. In this regard, the results in Fig.  3 a revealed that catalyst loading significantly influenced biodiesel yield. As shown in Fig.  3 a, the biodiesel yields increased when the catalyst loading was increased from 3 to 4%.

figure 3

Effects of a catalyst loading, b reaction time, c methanol to oil molar ratio (MTOMR), and d reaction temperature on biodiesel yield. Conditions: MTOMR 9:1; catalyst loading 5 wt%; reaction time 165 min; 65 ºC; microwave power of 800 W

Notwithstanding, as the catalyst loading was increased to 5%, the biodiesel yield had consequently increased to more than 90%. Furthermore, when the catalyst loading was increased to 6%, biodiesel yield was subsequently observed not to be increasing any longer and started to decline from 91 to 87%. Other researchers highlighted that excessive catalyst loading might impede the mixing of methanol, oil and catalyst, which may further lead to the separation phase [ 4 , 6 ]. Therefore, 5 wt% catalyst loading was the best catalyst amount during this study.

3.2.2 Effect of reaction time

Figure  3 b shows the effect of reaction time on biodiesel yield. Different experimental conditions were as follows: MTOMR of 9:1, 5 wt% catalyst loading, 180 min reaction time, temperature 65 ºC and MW power of 800 W. As it can be observed from Fig.  3 b, the biodiesel yield had significantly increased as the reaction time continued to increase. Subsequently, this trend has resulted in a high biodiesel yield of 91% in 180 min. However, the biodiesel yield began to decline slightly as the time was further increased from 195 to 210 min. The possible elucidation of this result could be that as the reaction time was at 180 min, the reaction had probably already reached its equilibrium phase. In addition, the backward reaction might occur after reaching the equilibrium period because this reaction, in effect, holds a reversibility nature of the reaction and consequently declines the yield [ 28 , 31 ].

Alternatively, the highest conversion of JCO to biodiesel can be attributed to various factors such as high basic strength, high absorbency, and large surface area of the developed CaO-supported catalysts from WES materials. It is also worthy of highlighting that a very long reaction time could reduce biodiesel yield in such a way that the transesterification reaction will shift to the left and lead to the formation of soap [ 15 ]. Such soap formation can result from esters that could have probably been affected by hydrolysis by increasing reaction time further [ 31 ]. Palitsakun et al. [ 30 ] argued that methanol normally activates CaO. A small amount of CaO is converted to the parental Ca(OCH 3 ) 2 , which is more catalytically active than inactivated CaO. On the other hand, the FAME yield rose substantially from 54 to 88% in 80 min and remained steady as the reaction time intensified to 120 min, indicating that the system reached equilibrium.

3.2.3 Effect of MTOMR

The influence of MTOMR on biodiesel yield was investigated under the following conditions: MTOMR of 9:1, 5 wt% catalyst loading, 180 min reaction time, temperature 65 ºC and MW power of 800 W. MTOMR has a similar effect analogous to the loading of catalysts, specifically on the reaction rate and biodiesel yield in particular. As a result, the study on the transesterification reaction was conducted by adjusting the MTOMR while employing the enhanced catalyst loading (5 wt%) at 65 ºC (Fig.  3 c). It can be determined from the results in Fig.  3 c that as MTOMR increased from 7:1 to 9:1, a considerable increase in biodiesel harvest from 86 to 91% was notably observed. Despite that, as the MTOMR was increased to 10:1, there was an insignificant decrease in biodiesel yield (90%). When the MTOMR was further increased to 10:1, the biodiesel yields gradually decreased to 87% at 11:1.

Conversely, as the methanol concentration continued to increase, the catalyst’s elements of the catalyst and reaction agents were diluted by methanol, which could trigger a reverse reaction and result in declined biodiesel yield [ 6 ]. A comparable pattern of the results was affirmed in the study conducted by Nath et al. [ 25 ] and Basumatary et al. [ 7 ]. Their studies recorded that large amount to MTOMR above best reaction prerequisite can weaken the reaction assortments, causing submerging of catalyst’s active sites, thus declining their essential interactions for effective reaction and subsequently, the reaction rate and decreasing biodiesel yield product. Additionally, above the ideal MTOMR level, there is always a possibility of increasing biodiesel's hydrolysis to produce soap. Such an effect would trigger a substantial reduction in biodiesel yield [ 7 , 32 ]. In the present work, the most promising level for JCO transesterification to biodiesel was determined to be 9:1 of MTOMR with 5 wt% of CaO-supported catalyst generated from eggshell waste materials.

3.2.4 Effect of reaction temperature

The effect of temperature on the biodiesel yield was studied under the following diverse conditions: MTOMR of 9:1, 5 wt% catalyst loading, 180 min reaction time, temperature 65 ºC and MW power of 800 W. According to the results in Fig.  3 d, it was found that when the temperature was lower than 65 ºC, the biodiesel yield increased with increasing temperature and achieved the highest considerable yield at 65 ºC. The biodiesel yield was observed to be decreasing slightly as the temperature above 65 ºC. Various studies envisaged reaction of this nature to be thermic [ 7 ], whereas other researchers anticipated it to be marginally exothermic [ 32 , 33 ]. It is also imperative to highlight that the molecular activity increased with the increasing temperature, increasing the reaction rate. Hitherto, as the temperature continues to increase to 65 ºC, which is assumed to be the boiling temperature of methanol, the methanol will be evaporated and disappear from the reaction mix. Accordingly, the biodiesel yield was found declining. In addition, higher temperatures could speed up the saponification of triglycerides, which can cause adverse consequences on product yields [ 34 ].

3.3 Determination of transesterification process kinetics and activation energy

To investigate the transesterification reaction's kinetics, the effects of temperature and constant reaction time on the reaction were studied. The kinetics of transesterification reaction was investigated at various reaction temperatures of 45, 55, 65, 75 and 85 °C under the ideal operational conditions. As indicated in Table 3 , the linear regression equation at diverse reaction temperatures was attained, and the results are presented thereof. Regression equation was established, and it was observed to be linear with high regression coefficient (R 2  = 0.9994). The reaction's activation energy ( Ea ) was determined as 38.5 kJ mol −1 .

Additionally, a comparison of the activation energy for biodiesel production of the synthesized catalysts with several solid heterogeneous base catalysts from the previous studies has been made, signifying how the diverse Ea is attained by exploiting different feedstock (oils) in the transesterification reaction, as shown in Table 4 . Mazubert et al. [ 35 ] had reported activation energy of 9.7 kJ mol −1 while utilizing waste cooking oil (WCO) under an MW heating system. While using the same WCO as a feedstock and heterogeneous catalysts, Gupta and Rathod [ 36 ], Foroutan et al. [ 21 ] and Al-Sakkari et al. [ 11 ] have recorded high activation energy of 27, 46 and 49 kJ mol −1 . In comparison, Pavlovic et al. [ 19 ] reported very high activation energy (67; 58 kJ mol −1 ) as compared to the present study. Uzun et al. [ 37 ] used a conventional heating method and reported a very low activation energy of 12 kJ mol −1 .

3.4 Comparative study of the synthesized catalyst

The results showed a significantly high yield of 92% for biodiesel within 180 min reaction time within ideal requirements of 9:1 MTOMR and 5 wt% catalyst loading with 65 °C reaction temperature. The results in this study were compared to various preceding studies, specifically on waste biomass derivatives for catalysts of the heterogeneous nature that were utilized for the generation of biodiesel, as indicated in Table 5 . Biodiesel production from JCO while exploiting various solid base heterogeneous catalysts derived from different sources such as by Teo et al. [ 8 ] and Vyas et al. [ 38 ] have all recounted high reaction temperatures (360 and 120 °C, respectively) for the transesterification procedure with pretty much similar yields to the catalyst used under this investigation. They have reported 94 and 92% biodiesel yields after the 9 th and 5 th cycles, respectively.

In other studies, the synthesis of biodiesel while exploiting different waste biomass derived heterogeneous catalysts, Chavan et al. [ 39 ] and Teo et al. [ 8 ] have reported a diminished catalytic performance with lengthy reaction time (2.5 and 4 h). This can perhaps be ascribed to significant smaller surface areas of the various catalysts used. However, Pratika et al. [ 40 ] and Basumatary et al. [ 7 ] reported CaO-based heterogeneous catalysts with relatively outstanding catalytic performance (with 98 and 98% biodiesel yields) even though they may have low catalysts reusability or recyclability of 4 th and 3 rd cycles as compared to that of the catalyst utilized under this present study which was recycled 5 times. Teo et al. [ 8 ] have also described similar catalytic performance for JCO biodiesel production utilizing solid based heterogeneous catalysts and produced biodiesel yields in the range of 90–97%. Consistently, the current WES derived catalyst has displayed exceptional catalytic activity for the generation of biodiesel. In the current research work, EDS analysis has reported the existence of a high quantity of calcium in the WES catalysts as compared to other elements (Table S 1 ), and the Ca is an outstanding source of calcium oxide (CaO), which has been broadly utilized in several commercial enterprises as a heterogeneous catalyst. CaO has a fundamental role in the catalysis for the application of biodiesel production [ 21 , 23 ].

3.5 Reusability and leaching study of the catalyst

It is worth mentioning that the economic sustainability or viability of the transesterification procedure regarding biodiesel production is immensely influenced by the recyclability of a catalyst [ 23 ]. Reusability is one of the most essential and promising characteristics of the solid base heterogeneous catalysts that should be considered in the continuous reaction process for the low-cost biodiesel generation [ 26 , 27 ]. Figure  4 shows the results of the reusability study of the catalyst.

figure 4

Reusability of the WES derived catalyst (CaO-based catalyst). Conditions: MTOMR 9:1; catalyst loading 5 wt%; reaction time 165 min; 65 °C; microwave power of 800 W

The synthesized catalyst was regained from the reaction mixture by separation and consequently rinsed with hexane and subsequently calcinated at 1000 °C for 4 h. Thus, the restored catalyst was acquired and immediately reused in successive 5 catalytic cycles using the same experimental and reusability approach. Furthermore, the recycled catalyst was ascertained to produce a more than 90% biodiesel yield in the 1 st cycle and stayed the same at 90%. Yet even in the 3 rd cycle of the five progressive catalytic runs, the recycled catalyst performed well with an excellent biodiesel yield of more than 80%. Conversely, the results were observed to have been drastically reduced to 72% after the 5 th cycle. This may be due to the contamination of the active site with CO 2 and H 2 O and, hence, it shortened lifespan of the developed CaO-based catalyst. As a result, the CaO-based catalyst developed in our study exhibits outstanding reusability in biodiesel fuel production with a significant reduction in yield up to 5 cycles. The degree of Ca leaching can affect the lifetime of the developed CaO catalyst, and biodiesel can be contaminated with leached Ca particles [ 19 ].

4 Economic and environmental cost–benefit analysis

Waste such as eggshells can be transformed into high value-added products, contributing to sustainable economic development and, more so, to sustainable bio-waste management. Undoubtedly, it costs about US$100,000 per annum to discard the eggshell waste produced by the average egg processing plant in the United States [ 10 ]. In contrast, the industrial scale of recycling WES to CaO-based catalysts can provide economic benefits that are at least five times higher than the cost of conservative disposal methods [ 8 ]. At the same time, by minimizing the exposure and likelihood of pathogens transmission, abbreviating discarding costs and synthesizing potential CaO-based heterogeneous catalysts for economical biodiesel production, great environmental benefits can be achieved that greatly support the sustainability of the anticipated technique [ 8 ]. One of the environmental benefits of reutilization of WES is that the eggshell is made up of calcium carbonate, which is a rich source of calcium. Eggshells are collected as waste from residences, hotels, and egg processing facilities, among other places. Eggshells are still difficult to dispose of in an environmentally acceptable manner because of their limited decomposition capacity, which generates pollution. To reduce microbiological species such as molds and bacteria, such shells can be dried at high temperatures (about 80 °C). Eggshell powder can then be made by powdering dried shells in a grinder [ 29 ]. Eggshell powder has been demonstrated in studies to be effective as a calcium supplement for plants suffering from blossom-end rot disease, such as tomatoes and berries [ 6 ].

5 Conclusions

This research has successfully synthesized an economical, eco-friendly, and highly efficient bio-waste-derived CaO solid base heterogeneous catalyst for low-cost biodiesel synthesis. The developed catalyst has demonstrated outstanding catalytic performance with a biodiesel yield of 92% under the measured, optimized reaction conditions. While the kinetic study parameters and Ea of the reaction were determined to be 38.5 kJ mol −1 , culminating that it is not an energy-intensive procedure. We hypothesized that the high coefficient of determinations (R 2 ) value of 0.9994 signified the prospect and excellent suitability of this model can be applied in future. The eggshell catalyst, produced from renewable and cheap raw materials under MW-assisted technology, is an excellent and environmentally friendly technique. These capabilities are required to develop high-performance heterogeneous catalysts that can be upgraded to a commercial scale in the future or to industrial applications for large-scale biodiesel production. The results of this study are not for academic purposes only. They can also serve as prototypes for the industrial production of biodiesel using non-consumable oils as feedstock and biowaste as catalysts for MW heating systems as environmentally friendly methods for sustainable bio-waste management approaches. In addition, this study expands on existing state-of-the-art research as we know about biofuel production using non-consumable oil feedstocks and bio-waste economically.

Availability of data and materials

The datasets supporting the conclusions of this article are included within the article and supplementary materials.

Bhatia SK, Palai AK, Kumar A, Bhatia RK, Patel AK, Thakur VK, et al. Trends in renewable energy production employing biomass-based biochar. Bioresour Technol. 2021;340:125644.

Article   Google Scholar  

Bhatia SK, Gurav R, Choi YK, Lee HJ, Kim SH, Suh MJ, et al. Rhodococcus sp. YHY01 a microbial cell factory for the valorization of waste cooking oil into lipids a feedstock for biodiesel production. Fuel. 2021;301:121070.

Bhatia SK, Joo HS, Yang YH. Biowaste-to-bioenergy using biological methods – a mini-review. Energ Convers Manage. 2018;177:640–60.

Peng YP, Amesho KTT, Chen CE, Jhang SR, Chou FC, Lin YC. Optimization of biodiesel production from waste cooking oil using waste eggshell as a base catalyst under a microwave heating system. Catalysts. 2018;8:81.

Lin YC, Amesho KTT, Chen CE, Cheng PC, Chou FC. A cleaner process for green biodiesel synthesis from waste cooking oil using recycled waste oyster shells as a sustainable base heterogeneous catalyst under the microwave heating system. Sustain Chem Pharm. 2020;17:100310.

Lee M, Tsai WS, Chen ST. Reusing shell waste as a soil conditioner alternative? A comparative study of eggshell and oyster shell using a life cycle assessment approach. J Clean Prod. 2020;265:121845.

Basumatary S, Nath B, Das B, Kalita P, Basumatary B. Utilization of renewable and sustainable basic heterogeneous catalyst from Heteropanax fragrans (Kesseru) for effective synthesis of biodiesel from Jatropha curcas oil. Fuel. 2021;286:119357.

Teo SH, Islam A, Chan ES, Choong SYT, Alharthi NH, Taufiq-Yap YH, et al. Efficient biodiesel production from Jatropha curcus using CaSO 4 /Fe 2 O 3 -SiO 2 core-shell magnetic nanoparticles. J Clean Prod. 2019;208:816–26.

Adeogun AI, Ofudje AE, Idowu MA, Kareem SO. Facile development of nano size calcium hydroxyapatite based ceramic from eggshells: synthesis and characterization. Waste Biomass Valori. 2018;9:1469–73.

FAOSTAT. Crops and Livestock Products. Rome: Food and Agriculture Organization of the United Nations; 2019. http://www.fao.org/faostat/en/#data/QL .

Al-Sakkari EG, El-Sheltawy ST, Attia NK, Mostafa SR. Kinetic study of soybean oil methanolysis using cement kiln dust as a heterogeneous catalyst for biodiesel production. Appl Catal B-Environ. 2017;206:146–57.

Hsiao MC, Kuo JY, Hsieh SA, Hsieh PH, Hou SS. Optimized conversion of waste cooking oil to biodiesel using modified calcium oxide as catalyst via a microwave heating system. Fuel. 2020;266:117114.

Keneni YG, Hvoslef-Eide AK, Marchetti JM. Optimization of the production of biofuel form Jatropha oil using a recyclable anion-exchange resin. Fuel. 2020;278:118253.

Takeno ML, Mendonca IM, Barros SD, Maia PJD, Pessoa WAG, Souza MP, et al. A novel CaO-based catalyst obtained from silver croaker ( Plagioscion squamosissimus ) stone for biodiesel synthesis: waste valorization and process optimization. Renew Energ. 2021;172:1035–45.

Risso R, Ferraz P, Meireles S, Fonseca I, Vital J. Highly active Cao catalysts from waste shells of egg, oyster and clam for biodiesel production. Appl Catal A-Gen. 2018;567:56–64.

Sirisomboonchai S, Abuduwayiti M, Guan GQ, Samart C, Abliz S, Hao XG, et al. Biodiesel production from waste cooking oil using calcined scallop shell as catalyst. Energ Convers Manage. 2015;95:242–7.

Boey PL, Maniam GP, Hamid SA, Ali DMH. Utilization of waste cockle shell ( Anadara granosa ) in biodiesel production from palm olein: optimization using response surface methodology. Fuel. 2011;90:2353–8.

Nisar J, Razaq R, Farooq M, Iqbal M, Khan RA, Sayed M, et al. Enhanced biodiesel production from Jatropha oil using calcined waste animal bones as catalyst. Renew Energ. 2017;101:111–9.

Pavlovic SM, Marinkovic DM, Kostic MD, Jankovic-Castvan IM, Mojovic LV, Stankovic MV, et al. A CaO/zeolite-based catalyst obtained from waste chicken eggshell and coal fly ash for biodiesel production. Fuel. 2020;267:117171.

Krishnamurthy KN, Sridhara SN, Ananda Kumar, CS. Optimization and kinetic study of biodiesel production from Hydnocarpus wightiana oil and dairy waste scum using snail shell CaO nano catalyst. Renew Energ. 2020;146:280–96.

Foroutan R, Mohammadi R, Razeghi J, Ramavandi B. Biodiesel production from edible oils using algal biochar/CaO/K 2 CO 3 as a heterogeneous and recyclable catalyst. Renew Energ. 2021;168:1207–16.

Kumar G, Singh V, Kumar D. Ultrasonic-assisted continuous methanolysis of Jatropha curcas oil in the appearance of biodiesel used as an intermediate solvent. Ultrason Sonochem. 2017;39:384–91.

Teo SH, Islam A, Masoumi HRF, Taufiq-Yap YH, Janaun J, Chan ES, et al. Effective synthesis of biodiesel from Jatropha curcas oil using betaine assisted nanoparticle heterogeneous catalyst from eggshell of Gallus domesticus . Renew Energ. 2017;111:892–905.

Sarma AK, Kumar P, Aslam M, Chouhan APS. Preparation and characterization of Musa balbisiana Colla underground stem nano-material for biodiesel production under elevated conditions. Catal Lett. 2014;144:1344–53.

Nath B, Das B, Kalita P, Basumatary S. Waste to value addition: utilization of waste Brassica nigra plant derived novel green heterogeneous base catalyst for effective synthesis of biodiesel. J Clean Prod. 2019;239:118112.

Mansir N, Teo SH, Rabiu I, Taufiq-Yap YH. Effective biodiesel synthesis from waste cooking oil and biomass residue solid green catalyst. Chem Eng J. 2018;347:137–44.

Kouzu M, Fujimori A, Fukakusa R, Satomi N, Yahagi S. Continuous production of biodiesel by the CaO-catalyzed transesterification operated with continuously stirred tank reactor. Fuel Process Technol. 2018;181:311–7.

Rahman WU, Fatima A, Anwer AH, Athar M, Khan MZ, Khan NA, et al. Biodiesel synthesis from eucalyptus oil by utilizing waste egg shell derived calcium based metal oxide catalyst. Process Saf Environ. 2019;122:313–9.

Balajii M, Niju S. Banana peduncle – a green and renewable heterogeneous base catalyst for biodiesel production from Ceiba pentandra oil. Renew Energ. 2020;146:2255–69.

Palitsakun S, Koonkuer K, Topool B, Seubsai A, Sudsakorn K. Transesterification of Jatropha oil to biodiesel using SrO catalysts modified with CaO from waste eggshell. Catal Commun 2021;149:106233.

Laskar IB, Gupta R, Chatterjee S, Vanlalveni C, Rokhum L. Taming waste: waste Mangifera indica peel as a sustainable catalyst for biodiesel production at room temperature. Renew Energ. 2020;161:207–20.

Rathore V, Tyagi S, Newalkar B, Badoni RP. Jatropha and Karanja oil derived DMC-biodiesel synthesis: a kinetics study. Fuel. 2015;140:597–608.

Lawan I, Garba ZN, Zhou WM, Zhang MX, Yuan ZH. Synergies between the microwave reactor and CaO/zeolite catalyst in waste lard biodiesel production. Renew Energ. 2020;145:2550–60.

Tang ZE, Lim S, Pang YL, Ong HC, Lee KT. Synthesis of biomass as heterogeneous catalyst for application in biodiesel production: state of the art and fundamental review. Renew Sust Energ Rev. 2018;92:235–53.

Mazubert A, Taylor C, Aubin J, Poux M. Key role of temperature monitoring in interpretation of microwave effect on transesterification and esterification reactions for biodiesel production. Bioresour Technol. 2014;161:270–9.

Gupta AR, Rathod VK. Waste cooking oil and waste chicken eggshells derived solid base catalyst for the biodiesel production: optimization and kinetics. Waste Manage. 2018;79:169–78.

Uzun BB, Kilic M, Ozbay N, Putun AE, Putun E. Biodiesel production from waste frying oils: optimization of reaction parameters and determination of fuel properties. Energy. 2012;44:347–51.

Vyas AP, Subrahmanyam N, Patel PA. Production of biodiesel through transesterification of Jatropha oil using KNO 3 /Al 2 O 3 solid catalyst. Fuel. 2009;88:625–8.

Chavan SB, Kumbhar RR, Madhu D, Singh B, Sharma YC. Synthesis of biodiesel from Jatropha curcas oil using waste eggshell and study of its fuel properties. RSC Adv. 2015;5:63596–604.

Pratika RA, Wijaya K, Trisunaryanti W. Hydrothermal treatment of SO 4 /TiO 2 and TiO 2 /CaO as heterogeneous catalysts for the conversion of Jatropha oil into biodiesel. J Environ Chem Eng. 2021;9:106547.

Download references

Acknowledgements

This work was supported by the Ministry of Science and Technology (MOST), Taiwan, under the grant NSC 100-2221-E-110-015-MY2.

This work was supported by Ministry of Science and Technology (MOST) (Project No. NSC 100–2221-E-110–015-MY2).

Author information

Authors and affiliations.

Institute of Environmental Engineering, National Sun Yat-Sen University, Kaohsiung, 804, Taiwan

Kassian T. T. Amesho, Yuan-Chung Lin, Chin-En Chen & Pei-Cheng Cheng

Center for Emerging Contaminants Research, National Sun Yat-Sen University, Kaohsiung, 804, Taiwan

Kassian T. T. Amesho, Yuan-Chung Lin, Chin-En Chen, Pei-Cheng Cheng & Vinoth Kumar Ponnusamy

Department of Public Health, Kaohsiung Medical University, Kaohsiung, 807, Taiwan

Yuan-Chung Lin

Doctoral Degree Program in Toxicology, College of Pharmacy, Kaohsiung Medical University, Kaohsiung, 807, Taiwan

Department of Medicinal and Applied Chemistry, Kaohsiung Medical University, Kaohsiung, 807, Taiwan

Vinoth Kumar Ponnusamy

Department of Chemistry, National Sun Yat-sen University, Kaohsiung, 804, Taiwan

You can also search for this author in PubMed   Google Scholar

Contributions

Kassian T.T. Amesho: Writing-original draft. Yuan-Chung Lin: Conceptualization, Funding acquisition, Methodology, Supervision. Chin-En Chen: Data curation, Investigation, Formal analysis/Software. Pei-Cheng Cheng: Data curation, Investigation. Vinoth Kumar Ponnusamy: Writing- review & editing. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Yuan-Chung Lin .

Ethics declarations

Competing interests.

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Additional information

Publisher’s note.

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary Information

Additional file 1: table s1..

Physico-chemical properties of the Jatropha curcas biodiesel (ASTM D6751, EN 14214) from [1, 2]. Fig. S1. XRD analysis of eggshells (a) before and (b) after calcination. Fig. S2. FTIR analysis of eggshells before (a) and after calcination (b).

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/ .

Reprints and permissions

About this article

Cite this article.

Amesho, K.T.T., Lin, YC., Chen, CE. et al. Optimization and kinetics studies of biodiesel synthesis from Jatropha curcas oil under the application of eco-friendly microwave heating technique: an environmentally benign and sustainable bio-waste management approach. Sustain Environ Res 32 , 41 (2022). https://doi.org/10.1186/s42834-022-00151-w

Download citation

Received : 22 February 2022

Accepted : 22 August 2022

Published : 15 September 2022

DOI : https://doi.org/10.1186/s42834-022-00151-w

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

  • Waste eggshell valorization
  • Jatropha curcas oil
  • Activation energy
  • Catalyst reusability
  • CaO-based catalyst

Sustainable Environment Research

ISSN: 2468-2039

  • Submission enquiries: Access here and click Contact Us
  • General enquiries: [email protected]

experimental study of biodiesel synthesis

ISSN 0216-0455 (print) and ISSN 2527-3825 (online)

CITATION ANALYSIS

SCOPUS Clarivate Analytics Google Scholar

  • Focus & Scope
  • Author Guidelines
  • Author Fees
  • Online Submission
  • Editorial Team
  • Publication Ethics
  • Screening for Plagiarism
  • Statement of Originality
  • Order Journal
  • Visitor Statistics

template icon

  • For Readers
  • For Authors
  • For Librarians

Yaasinta Lafasha Nabila Department of Chemistry, Faculty of Mathematics and Natural Sciences, Tanjungpura University, Jl. Ahmad Yani Pontianak, Kalimantan Barat 79124 Indonesia

experimental study of biodiesel synthesis

Andi Hairil Alimuddin https://scholar.google.co.id/citations?user=iVxMuTkAAAAJ&hl=id Department of Chemistry, Faculty of Mathematics and Natural Sciences, Tanjungpura University, Jl. Ahmad Yani Pontianak, Kalimantan Barat 79124 Indonesia

experimental study of biodiesel synthesis

  • Other Journals
  • Announcements
  • Indexing & Abstracting
  • Journal History

Synthesis of Biodiesel in Low-Grade Palm Oil using Geopolymer-ZnO Catalyst

experimental study of biodiesel synthesis

This study aimed to synthesize biodiesel (fatty acid methyl ester) from low-grade palm oil using geopolymer-ZnO catalyst. The activity of catalyst was tested by mixing low-grade palm oil and methanol in a mole ratio of 1:10, with varying catalyst concentrations of 1%, 3%, and 5% at a temperature of 67 o C and different time intervals. Subsequently, the progression of the reaction was monitored using thin-layer chromatography (TLC). The results showed that catalyst ratios of 1:0, 1:1, and 1:2 provided complete conversion at concentrations of 3% and 5%. Geopolymer-ZnO catalyst reuse test was carried out using a concentration of 5%, followed by washing with n-hexane. After drying at 100 o C, the sample was reused in biodiesel synthesis. The reuse test findings showed optimum results at 1:2 variation and 2 hours of reaction time with total methyl ester conversion. In addition, the majority of the experiments performed were carried out using a 1:1 variation with a maximum of 3 repetitions, which consistently showed total conversion to methyl ester. Catalyst used was then characterized using FTIR, XRD, and XRF instruments, with the analysis results confirming that it was geopolymer-ZnO. Methyl ester obtained was analyzed using GC-MS, and the findings showed that the main compositions included methyl oleate (47.35%) and methyl palmitate (40.13%).

Abdullah, M. M. A., Liew, Y. M., Heah, C. Y., & Muhammad, F. M. T. (2018). Clay-Based Materials in Geopolymer Technology. In R. O. . Saleh, H and Rahman (Ed.), Cement Based Materials (pp. 239–264). interchopen. https://doi.org/http://doi.org/10.5772/interchopen.74438

Alimuddin, A. H, Thamrin, U., Nelly, W., Rudiyansyah, Yopa, E. P., Ismail, A., & Yudi, Y. (2017). Synthesis and Characterization of Sn/Zeolite and Catalytic Activity Test in the Esterification Reaction of Slude Oil. AIP Conference Proceedings , 1823(1):020052. https://doi.org/https://doi.org/10.1063/1.4978125

Aqhilla, N., Edy, S., & Ahmad, F. (2017). Pembuatan Biodiesel dari Minyak Nyamplung dengan Menggunakan Katalis Basa Na2SiO3/Fe3O4. JOM FTEKNIK , 4 (2), 1–5.

Arifah, M., Sri, H., & Y. (2016). Katalisis CPO menjadi Biodiesel dengan Katalis ZnO Sintesis Presipitan NaOH. Jom FTEKNIK , 3 (1), 1–8.

Astar, I. (2017). Variasi Abu Tandan Kosong Kelapa Sawit terhadap Karakteristik Quintinite serta Uji Aktivitas Katalitik pada Reaksi Esterifikasi dan Transesterifikasi CPO secara Simultan . Tesis, Program Studi S2 Kimia, FMIPA, Universitas Tanjungpura.

BPH Migas. (2017). Konsumsi BBM Nasional per Tahun – Konsumsi BBM JBU, JBKP, JBT Dari Tahun 2006-2017 . www.bphmigas.go.id/konsumsi-bbm-nasional

Bustaman, S. (2009). Strategi Pengembangan Industri Biodiesel Berbasis Kelapa di Maluku. Jurnal Litbang Pertanian , 28 (2), 46–53.

Kondamudi, N., Mohapatra, S. K., & Misra, M. (2011). Quintinite as A Bifuncsional Heterogeneous Catalyst for Biodiesel Synthesis. Applied Catalysis A General , 393 (1), 36-43.

Lee, D. W., Park, Y.M., & Lee, K. Y. (2009). Heterogeneous Base Catalysts for Transesterification in Biodiesel Synthesis. Catal Surv Asia , 13 (2), 63–77. https://doi.org/10.1007/s10563-009-9068-6

Lestari, P. P. (2018). Pengaruh Nanokatalis ZnO/CaO terhadap Biodiesel dari Minyak Biji Alpukat. Jurnal Kimia Saintek Dan Pendidikan , II (1), 1–8.

Nugraha, I & Umi, K. (2017). Sintesis dan Karakterisasi Material Komposit Kaolin-ZVI (Zero Valent Iron) serta Uji Aplikasinya sebagai Adsorben Kation Cr (VI). Jurnal Kimia Valensi , 3 (1), 59–70. https://doi.org/http://dx.doi.org/10.15408/jkv.v0i0.4650

Nugraha, M.W., Edy, S., & Zuchra, H. (2017). Sintesis Biodiesel dari Minyak Biji Nyamplung (Calophyllum inophyllum) menggunakan Katalis Geopolimer dari Abu Sekam Padi dan Kaolin. JOM FTEKNIK , 4 (2), 1–5.

Poggetto, G. D., Antonio, D. A., Ignazio, B., Simona, P., Cristina, L., & Michelina, C. (2021). FT-IR Study Thermal Analysis and Evaluation of the Antibacterial Activity of a MK-Geopolymer Mortar Using Glass Waste as Fine Aggregate. Polymers , 13 , 2970. https://doi.org/https://doi.org/10.3390/polym13172970

Prasetya, F. A., Sukmana, N.C., Roziqin, R., Deqi, N.F., Afifi, A.J., & Anggraini, U. (2018). Sintesis dan Karakterisasi Beton Geopolimer Berbasis Abu Layang sebagai Media Tanam. Jurnal Teknologia , 1 (1), 21–26.

Priyambudi, I.Y. & Amirullah, M. R. F. (2017). Aplikasi Katalis Heterogen Fe2O3/CaO pada Pembuatan Biodiesel dari Minyak Kelapa Sawit . Brawijaya University.

Putri, P. C. A., & Edy, P. (2020). Transesterifikasi Minyak Kelapa Sawit menggunakan Katalis Kalsium Oksida (CaO) menjadi Biodiesel. Jurnal Metana , 16 (2), 75–80.

Rustan M., Subaer, & Irhamsyah. (2015). Studi Tentang Pengaruh Nanopartikel ZnO (Seng Oksida) terhadap Kuat Tekan Geopolimer Berbahan Dasar Metakaolin, Jurnal Sains dan Pendidikan Fisika (JSPF). Jurnal Sains Dan Pendidikan Fisika , 11 (3), 286–291.

Sa’adah, A. F., Akhmad, F., & Bambang, J. (2017). Peramalan Penyediaan dan Konsumsi Bahan Bakar Minyak Indonesia dengan Model Sintesis Dinamik . 17 (2), 118–137.

Soerawidjaja. (2006). Fondasi-fondasi Ilmiah dan Keteknikan dari Teknologi Pembuatan Biodiesel. Seminar Nasional ”Biodiesel Sebagai Energi Alternatif Masa Depan, UGM, Yogyakarta .

Tambak, M. N. (2019). Regenerasi Katalis Zeolit Alam dan CaO dari Cangkang Telur Ayam pada Produksi Biodiesel Berbahan Baku Minyak Kelapa Sawit . Skripsi, Universitas Sumatera Utara, Medan.

Usman, T., Rudiansyah, Nelly, W., Astar, I., Yudistira, Y., Alimuddin, A.H., & Rahmalia, W. (2019). Tin-Empty Palm Bunch Ash Impregnated Zeolite as Suitable Catalyst for Simultaneous Transesterification-Esterification Reaction of Palm Oil. Seminar Nasional Material , 1–8.

Valerio, T. L., Guilherme, A. R. M., Leticia, F. G., Aline, V., Everson, P. B., & Paulo, R. P. R. (2019). Studi of the Nb2O3 Insertion in ZnO to Dye-sensitized Solar Cells. Materials Research , 22 ((Suppl. 1)), e20180864.

Wahyuni N., Imelda H. S., & Dina, A. (2019). Isoterm Adsorpsi Fenol oleh Lempung Alam, Jurnal Teknologi Lingkungan Lahan Basah. Jurnal Teknologi Lingkungan Lahan Basah , 7 (1), 29–37.

Yunita, Nurlina, & Syahbanu, I. (2020). Sintesis Nanopartikel Zink Oksida (ZnO) dengan Penambahan Ekstrak Klorofil sebagai Capping Agent. Positron , 10 (2), 123–130.

Zulkifli, M., Subaer, & Nurhayati. (2019). Sintesis Geopolimer Berbahan Dasar Lempung Laterite dengan Penambahan ZnO sebagai Komposit Fungsional,. Jurnal Sains Dan Pendidikan Fisika (JSPF) , 15 (3), 86–89.

Article Metrics

experimental study of biodiesel synthesis

  • There are currently no refbacks.

agriTECH has been Indexed by:

Emerging Source Citation Index

agriTECH (print ISSN 0216-0455 ; online ISSN 2527-3825 ) is published by Faculty of Agricultural Technology , Universitas Gadjah Mada in colaboration with Indonesian Association of Food Technologies.

website statistics

U.S. flag

An official website of the United States government

The .gov means it’s official. Federal government websites often end in .gov or .mil. Before sharing sensitive information, make sure you’re on a federal government site.

The site is secure. The https:// ensures that you are connecting to the official website and that any information you provide is encrypted and transmitted securely.

  • Publications
  • Account settings

Preview improvements coming to the PMC website in October 2024. Learn More or Try it out now .

  • Advanced Search
  • Journal List
  • PMC10881653

Logo of springeropen

Recent advances in transesterification for sustainable biodiesel production, challenges, and prospects: a comprehensive review

Sabah mohamed farouk.

1 Chemical Engineering Department, Egyptian Academy for Engineering and Advanced Technology (EA&EAT), affiliated to the Ministry of Military Production, Km. 3 Cairo Belbeis Desert Rd., Cairo Governorate, 3066 Egypt

Aghareed M. Tayeb

2 Faculty of Engineering, Minia University, Misr Aswan Agricultural Rd., EL MAHATTA, Menia Governorate, 2431384 Egypt

Shereen M. S. Abdel-Hamid

Randa m. osman.

3 Chemical Engineering and Pilot Plant Department, National Research Centre (NRC), 33 El Bohouth St., Dokki, 12622 Giza Governorate Egypt

Associated Data

Data available on request from the authors.

Biodiesel, a renewable and sustainable alternative to fossil fuels, has garnered significant attention as a potential solution to the growing energy crisis and environmental concerns. The review commences with a thorough examination of feedstock selection and preparation, emphasizing the critical role of feedstock quality in ensuring optimal biodiesel production efficiency and quality. Next, it delves into the advancements in biodiesel applications, highlighting its versatility and potential to reduce greenhouse gas emissions and dependence on fossil fuels. The heart of the review focuses on transesterification, the key process in biodiesel production. It provides an in-depth analysis of various catalysts, including homogeneous, heterogeneous, enzyme-based, and nanomaterial catalysts, exploring their distinct characteristics and behavior during transesterification. The review also sheds light on the transesterification reaction mechanism and kinetics, emphasizing the importance of kinetic modeling in process optimization. Recent developments in biodiesel production, including feedstock selection, process optimization, and sustainability, are discussed, along with the challenges related to engine performance, emissions, and compatibility that hinder wider biodiesel adoption. The review concludes by emphasizing the need for ongoing research, development, and collaboration among academia, industry, and policymakers to address the challenges and pursue further research in biodiesel production. It outlines specific recommendations for future research, paving the way for the widespread adoption of biodiesel as a renewable energy source and fostering a cleaner and more sustainable future.

Introduction

Fossil fuels are responsible for most of the world’s energy needs, endangering the environment. The consumption of fossil fuel products like coal and petroleum has increased because of ongoing globalization and industrialization (Touqeer et al. 2020 ). Around 580 TJ of energy is needed annually for the entire planet, and a staggering 80% of that energy is supplied by burning traditional fossil fuels. Greater energy use is required by expanding the population and improving accessibility, which increases the depletion of fossil fuel reserves (Martchamadol and Kumar 2012 ). Finding an alternate energy source, like biodiesel, is essential for improving energy security for economic development (Oh et al. 2002 ; Um and Kim 2009 ). Biodiesel is a renewable, sustainable, and biodegradable fuel with low greenhouse gas emissions (Sharma and Singh 2009 ; Lee et al. 2011 ). Oils from plants and animals, as well as other lipids such as triacylglycerides (TAGs), can be converted into biodiesel (Hoekman and Robbins 2012 ). The process of making biodiesel, called transesterification or alcoholysis, typically requires the use of catalysts such as acids, bases, and enzymes (Ong et al. 2014 ). Catalysts can exist in either a homogeneous or heterogeneous phase. A homogeneous catalyst undergoes alcoholysis in the same phase, often liquid, as the reactants. In contrast, a heterogeneous catalyst is in a different phase, often one that is not liquid, from the reactants (Ruhul et al. 2015 ). Nanomaterials have become increasingly important in enhancing the production of biodiesel. The weight of the catalyst, the reaction temperature, and the oil-to-alcohol ratio are all reduced when a nanocatalyst is used (Dhawane et al. 2018 ; Ghosh and Halder 2022 ). It has already been proven that nanocatalysts play a key role in increasing response velocity. Additionally, it significantly lowers the reactants’ activation energy. This study addresses the need to keep up with the rapidly changing field of biodiesel production. By conducting thorough surveys and reviews, researchers can stay informed about the latest advancements and challenges. This approach helps focus research efforts by identifying promising methodologies and comparing their performance. The production of biodiesel encounters several challenges, including costly feedstocks, environmental concerns, and the quest for improved efficiency. Through a meticulous review of relevant literature, researchers can identify these challenges and propose potential solutions. This review serves as a valuable resource, synthesizing information from different sources to provide a comprehensive understanding of the current state of biodiesel production technology. Policymakers and industry leaders rely on up-to-date insights to make informed decisions that shape the future of the biodiesel industry. Concise surveys and reviews effectively communicate the latest information, enabling informed decision-making. Additionally, this review promotes collaboration and knowledge sharing within the biodiesel research community, facilitating the development of new and better technologies. It also educates the public about biodiesel production and its potential benefits, potentially increasing support and encouraging the adoption of biodiesel as a sustainable alternative to petroleum-based fuels. Literary, and in detail, this paper offers a comprehensive review of biodiesel production, encompassing key topics such as feedstock selection, the transesterification process, the recovery and reusability of nanoparticles, the benefits and challenges of biodiesel use in engines, and the significance of techno-economic analysis. It begins with various feedstocks, including used oils, animal fats, algae, and edible and inedible vegetable oils, which are chosen for their potential to produce biodiesel. The review then highlights biodiesel as a promising alternative to fossil fuels, exploring its environmental benefits, renewability, and potential to reduce dependence on fossil fuels. The transesterification process, a crucial step in biodiesel production, is discussed in detail, covering reaction mechanisms, catalysts, influencing factors, and challenges associated with optimization. Additionally, various methods and strategies, including solid supports, magnetic separation, centrifugation, and surface modification, have been explored for the recovery and reusability of nanoparticles as catalysts in biodiesel production. These approaches offer promising prospects for enhancing the efficiency and sustainability of biodiesel synthesis while reducing costs and environmental impacts. The benefits of using biodiesel in engines, including improved lubricity, reduced emissions, and potential performance enhancements, are discussed. Lastly, this review emphasizes the importance of techno-economic analysis in assessing the financial viability and sustainability of biodiesel production, considering factors such as capital investment, operating costs, feedstock prices, and governmental policies. Overall, this comprehensive review serves as a valuable resource for researchers, engineers, and policymakers in the biodiesel industry.

Feedstocks for biodiesel production

Biodiesel can be synthesized from various raw materials, including vegetable oils, animal fats, algae, and waste cooking oil. Each feedstock has its own advantages and challenges regarding availability, cost, sustainability, and energy content. The purpose of using different raw materials is to diversify the sources of renewable energy and reduce dependence on fossil fuels. At least 80% of the current costs of producing biodiesel come from the feedstock (Sayed and El-Gharbawy 2016 ). Given the global food crisis, almost 95% of biodiesel produced worldwide is manufactured from edible oils, which are viewed as unnecessary (Azizian and Kramer 2005 ). As a result, making biodiesel from inexpensive, non-edible oil is the current trend (Balat 2011 ). A few feedstocks, including palm, jatropha, microalgae, coconut tallow, and used cooking oil, stand out for their high productivity in the manufacturing of biodiesel (Gui et al. 2008 ). Figure  1 represents some different feedstocks for biodiesel production.

An external file that holds a picture, illustration, etc.
Object name is 11356_2024_32027_Fig1_HTML.jpg

The choice of raw materials has a significant impact on the production of biodiesel in terms of yield, quality, cost, and environmental sustainability. The fatty acid profile of the raw material plays a crucial role in determining the properties and performance of biodiesel. Oils with high levels of saturated fatty acids produce biodiesel with better cold flow properties, while oils with high levels of unsaturated fatty acids produce biodiesel with better oxidative stability (Suzihaque et al. 2022 ). FFA content significantly impacts the reaction rate and biodiesel yield. Higher FFA content requires pre-treatment steps, such as esterification, to reduce acidity and prevent soap formation, which can hinder the reaction and reduce biodiesel purity (Javidialesaadi and Raeissi 2013 ). The amount of water in the raw material can also affect the transesterification reaction. Excess water can hydrolyze the fatty acid methyl esters produced during the reaction, reducing biodiesel yield. Therefore, feedstocks with low water content are preferred (Atadashi et al. 2012 ). The overall quality of the raw material, including its purity and the presence of impurities, can influence biodiesel production. Impurities like phospholipids, gums, and waxes can interfere with the reaction and complicate downstream processing (Manzanera et al. 2008 ). The environmental impact of raw material sourcing should be considered. Using waste cooking oil or non-edible oils reduces the competition for land with food production and minimizes the environmental footprint of biodiesel production. The price of the raw material is a major factor affecting the economics of biodiesel production. Utilizing inexpensive feedstocks, such as waste cooking oil or locally produced oils, can significantly reduce production costs. The choice of raw material may necessitate adjustments in processing conditions, such as reaction temperature, catalyst concentration, and reaction time, to optimize biodiesel yield and quality. Byproducts generated during biodiesel production, such as glycerol, can be utilized for various applications, adding value to the overall process and reducing waste. The sourcing of raw materials can have social and economic impacts on local communities. Promoting the use of locally available or waste-based feedstocks can stimulate agricultural economies and create employment opportunities (Srivastava et al. 2018 ; Suzihaque et al. 2022 ). In conclusion, the selection of raw materials for biodiesel production requires a careful assessment of their fatty acid profile, FFA content, water content, overall quality, environmental sustainability, cost-effectiveness, and potential for byproduct utilization. By considering these factors, biodiesel production can be optimized for yield, quality, cost, and environmental sustainability. Devaraj et al. ( 2020 ) created biodiesel from used frying oil. At 75 °C, 1 wt% catalyst concentration, 1:6 oil-to-methanol molar ratio, 350 rpm, and 90 min, 97% of the possible biodiesel output was produced. In the pilot plant, biodiesel was also produced under these process conditions, with a 97% yield. Hamed et al. ( 2021 ) converted Afia waste cooking oil (AWCO) into biodiesel fuel. At the ideal working conditions of 60 °C reaction temperature, 3 h reaction time, and 0.4 catalyst concentration, the maximum conversion and yield of biodiesel are 97.54 and 94.935%, respectively. Roy et al. ( 2021 ) reported on the conversion of used frying oil into biodiesel to manage liquid waste. The optimum conditions for the transesterification of WFO were 1:16 oil to methanol by weight, 3 wt% catalyst, 65 °C reaction temperature, and 35 min reaction time. FAME conversion (99.5%) and 96% yield are achieved at this optimal reaction setting. Karmakar and Halder ( 2021 ) researched the transesterification reaction used in supercritical settings to produce biodiesel fuel from fish waste oil. According to calculations, the experimental yield of biodiesel generation under ideal conditions was 94.6%. Al Hatrooshi et al. ( 2020 ) created the fatty acid methyl ester from waste shark liver oil (WSLO). At a methanol-to-WSLO ratio of 10.3 M, a reaction duration of 6.5 h, a temperature of 60 °C, and a catalyst concentration of 5.9 wt%, acid-catalyzed WSLO transesterification achieved 99% FAME conversion. A biodiesel yield of 99.73% was achieved at 60 °C for 1.5 h using a 1.2% catalyst and a 6:1 methanol:oil ratio. In a batch-stirred reactor, Nisa et al. ( 2020 ) conducted an experiment to produce biodiesel from the microalgae Spirulina sp. using 1 wt% (w/w) of palm oil as a co-solvent for methanol and potassium hydroxide. At 60 °C, with a methanol-to-palm oil molar ratio of 10:1 and a palm oil-to-microalgae weight ratio of 5:1, the best biodiesel production of 85.28% was generated. Olubunmi et al. ( 2020 ) converted the bio-oil from dairy scum waste into biodiesel. Using a 9:1 methanol-to-oil molar ratio, 40 min for the reaction, 65 °C for the reaction temperature, and 300 rpm for the mixing speed, the maximum biodiesel yield of 94.8% was achieved. In India, Jain and Sharma ( 2010 ) utilized Jatropha oil for the manufacture of biodiesel under the ideal conditions of a 3:7 (v/v) methanol-to-oil ratio and a 1% (w/w) catalyst concentration for mixing at 400 rpm with H 2 SO 4 and NaOH. The transesterification of pretreated JCO yielded the highest yield of 90.1%. Abdulrahman ( 2017 ) made a biodiesel fuel from used cooking oil and chicken fat. The highest conversion to ester was attained at a methanol-to-oil ratio of roughly 7:1 at 60 °C and an 83% yield, also Al-Mawaali et al. ( 2023 ) utilized discarded animal fats and used cooking oil as a cheap source of feedstock to produce biodiesel. With NaOH acting as a catalyst, the discarded frying oil produced the highest yield of synthesized biodiesel (80.6%), followed by a mixture of waste cooking oil and animal fats (79.3%). Finally, using multiple raw materials as feedstock for biodiesel synthesis helps to diversify energy sources, reduce dependency on fossil fuels, and promote sustainability. Every feedstock source has benefits and drawbacks, and the choice of feedstock is influenced by factors like cost, sustainability, availability, and the needs of the biodiesel production process. Continued R&D efforts are aimed at enhancing the efficiency, cost-effectiveness, and sustainability of biodiesel production from diverse feedstocks.

Feedstock preparation

  • Cleaning and drying: The feedstock is cleaned to get rid of pollutants, water, and debris. It could entail processes like sedimentation or filtration. The feedstock is dried to eliminate any last traces of moisture after cleaning (Sait et al. 2022 ).
  • Pretreatment: Pretreatment of feedstocks may be necessary in some situations to enhance their quality and appropriateness for the manufacture of biodiesel. Degumming, acid esterification, or neutralization are some examples of pretreatment techniques used to get rid of pollutants, gums, and free fatty acids (Singh et al. 2011 ).

The most popular method of treatment for reducing FFAs is acid esterification. One efficient method for lowering FFAs is glycerolysis, also known as glycerol esterification (Elgharbawy et al. 2021 ). Hayyan et al. ( 2011 ) treated the high FFA percentage in palm oil using sulfuric acid. They used H 2 SO 4 by 0.75 wt%, a methanol-to-oil molar ratio of 8:1, 60 min of reaction time, and 60 °C to successfully reduce the FFA content from 23 wt% to less than 2 wt%. Kara et al. ( 2018 ) examined how various methanol-to-oil molar ratios affected the final FFA% throughout the esterification reaction while maintaining constant values for other parameters. Within 160 min of the reaction, the ideal conditions were reached at a molar ratio of 15:1. The highest conversion of 92.6% was achieved, while the FFA content decreased from 21 to 1.5%. Sadaf et al. ( 2018 ) investigated how three acids HCl, H 2 SO 4 , and H 3 PO 4 affected used cooking oil containing 2.75 wt% FFAs. At 60 °C and a 2.5:1 methanol-to-oil molar ratio, the FFA dropped to 0.33 wt%, indicating that H 2 SO 4 was the most effective catalyst. Sousa et al. ( 2010 ) treated castor oil with a non-catalytic glycerolysis procedure with a glycerol-to-oil mass ratio of 1:1 and an FFA of 2.5 wt% for 2 h at 120 °C. They were able to reduce the FFAs from 2.5 to 0.2 wt%.

Triglyceride content in different feedstocks

Triglycerides are the primary component of biodiesel and are present in various amounts in the feedstocks used in the process. Mono- and diglycerides, free fatty acids, phosphatides, sterols, fatty alcohols, fat-soluble vitamins, and other compounds are among the minor constituents. The amount of triglycerides in the feedstock has a significant impact on the output and caliber of biodiesel produced. Most dietary fats and oils are mostly composed of triglycerides.

Vegetable oils

Various vegetable oils, including canola, sunflower, soybean, and palm oils, are frequently utilized as feedstocks in the manufacturing of biodiesel. High concentrations of triglycerides, which are made up of three fatty acid chains joined to a glycerol backbone, are found naturally in these oils. Vegetable oils, therefore, have a high triglyceride concentration, usually more than 90% by weight. Vegetable oils are ideal feedstocks for the synthesis of biodiesel because of their high triglyceride content. Because refined vegetable oils convert pure triglycerides (TG) to FAME at a high rate and quickly, they are the ideal feedstock for producing biodiesel.

Animal fats

Fish oil, tallow, lard, and chicken fat are examples of animal fats that can be utilized as feedstocks to produce biodiesel. Comparing animal fats to vegetable oils, the former usually have a lower triglyceride concentration, usually between 70 and 85% by weight. The species of the animal, its nutrition, and the rendering techniques all affect the variation in triglyceride content. Animal fats can still be effectively turned into biodiesel even when their triglyceride level is slightly lower.

Waste oils and greases

The manufacturing of biodiesel can use recycled waste cooking oils and greases from food processing industries, restaurants, and other sources as feedstocks. Depending on their quality and place of origin, waste oils and greases might include varying amounts of triglyceride. The quality and conversion efficiency of biodiesel can be impacted by the presence of contaminants or elevated quantities of free fatty acids in these feedstocks.

Algae and microorganisms

The capacity of algae and specific microorganisms to accumulate large concentrations of lipids (fatty acids and triglycerides) makes them viable feedstocks for the generation of biodiesel. Depending on the species and growing environment, triglycerides can make up a large portion of algal oils, anywhere from 20 to 60% by weight. Table ​ Table1 1 represents the triglyceride content of different feedstocks.

Table 1

FeedstockTriglyceride contentReferences
Soybean oil96.8Li et al. ( )
Palm oil99.4Ali et al. ( )
Sunflower oil99.3Liu et al. ( )
Cottonseed oil98.1Liu et al. ( )
Waste cooking oil97.5Li et al. ( )
Chicken fat91.4Alptekin et al. ( )
Waste fish fat87.2Mrad et al. ( )
Microalgal oil99Çakırca et al. ( )

Biodiesel as a promising alternative source of biofuel

Among the emerging alternatives, biodiesel has garnered significant attention as a promising biofuel with the potential to reduce reliance on petroleum-based fuels and mitigate environmental impacts. The non-toxic, biodegradable fuel known as biodiesel is made from leftover cooking oil, animal fats, or vegetable oil, and presents several advantages over conventional diesel. Firstly, its renewable nature alleviates concerns over depleting fossil fuel resources. Unlike petroleum diesel, which is extracted from finite underground reservoirs, biodiesel can be produced from continuously replenished sources, ensuring long-term sustainability. Secondly, biodiesel boasts superior environmental credentials compared to diesel. Its burning emits significantly lower degrees of particulate matter, nitrogen oxides, sulfur oxides, and carbon monoxide, all of which contribute to air pollution and its detrimental effects on human health and the environment. Additionally, biodiesel’s lower greenhouse gas emissions make it a promising alternative for reducing the transportation sector’s contribution to climate change (Demirbas 2009 ). Moreover, biodiesel offers economic benefits, particularly for regions with abundant feedstock sources. Local production of biodiesel can stimulate agricultural economies, reduce reliance on imported petroleum, and create employment opportunities. Additionally, biodiesel’s compatibility with existing diesel engines eliminates the need for expensive infrastructure overhauls, facilitating a smoother transition toward a more sustainable fuel source (Fukuda et al. 2001 ). Despite its promise, biodiesel faces certain challenges that need to be addressed for its widespread adoption. One challenge lies in its economic viability. The production cost of biodiesel is currently higher than that of petroleum diesel, primarily due to feedstock costs and processing expenses. However, advancements in production technologies and economies of scale are expected to reduce biodiesel costs over time.

Another challenge involves the availability of suitable feedstocks. While various plant oils, such as soybean, palm, and jatropha, can be used for biodiesel production, concerns have arisen regarding land-use competition for the food industry and the potential environmental effects of broad-scope palm oil plantations. Sustainable feedstock sourcing strategies, such as utilizing waste cooking oil and cultivating oil crops on marginal lands, are crucial to addressing these concerns (Knothe 2010 ; Singh and Singh 2010 ). In conclusion, biodiesel presents a compelling alternative to conventional diesel, offering a combination of sustainability, environmental benefits, and economic potential. Addressing the current challenges related to feedstock availability and production costs will pave the way for biodiesel’s wider adoption and its significant contribution to a more sustainable energy future.

Innovations in biodiesel applications

Biodiesel, a renewable and cleaner-burning alternative to conventional diesel fuel, has gained significant attention and widespread applications in recent years. Its versatile nature allows for various uses across different sectors. In transportation, biodiesel can be blended with petroleum diesel to power vehicles, reducing emissions of greenhouse gases and air pollutants. Moreover, biodiesel finds applications in industrial settings, where it serves as a substitute for petroleum-based fuels in machinery and equipment. The growing interest in sustainable energy solutions has spurred research and development efforts, leading to advancements in biodiesel production techniques, feedstock selection, and engine compatibility. Recent studies (Mobin et al. 2022 ; Tripathi et al. 2023 ) have explored the potential of advanced feedstocks, such as algae and waste oils, to enhance biodiesel production efficiency and reduce environmental impacts. These advancements in biodiesel applications and technology are crucial steps in achieving a more sustainable and greener energy future. Figure  2 shows some of the applications of biodiesel.

An external file that holds a picture, illustration, etc.
Object name is 11356_2024_32027_Fig2_HTML.jpg

Biodiesel applications

Transesterification in biodiesel production

Alcohol and lipids react chemically to produce fatty acid alkyl esters, a process known as transesterification. Triglycerides and alcohol are transesterified to produce FAAE and glycerol. Triglycerides and alcohols combine to form diglycerides in the first stage, which are then converted to monoglycerides and glycerol, each of which yields an alkyl ester (Thangaraj et al. 2019 ). Among the factors influencing biodiesel yield in transesterification are time, temperature, type and concentration of the catalyst, kind of feedstock oil, and alcohol-to-oil ratio. It is possible to reverse the transesterification process. To shift the reaction’s equilibrium in favor of the product’s production, an excess of alcohol is therefore necessary. Alcohols with short chains, long chains, and cyclic chains are all used in this process. The availability, polarity, better reactivity, and inexpensive price of methanol and ethanol, however, make them popular choices (Avhad and Marchetti 2015 ), Fig.  3 illustrates the characteristics and features of the transesterification process used to make biodiesel, and Fig.  4 shows a schematic representation of the biodiesel synthesis path through the transesterification process.

An external file that holds a picture, illustration, etc.
Object name is 11356_2024_32027_Fig3_HTML.jpg

The characteristics and features of the transesterification process used to make biodiesel

An external file that holds a picture, illustration, etc.
Object name is 11356_2024_32027_Fig4_HTML.jpg

Schematic representation of the biodiesel synthesis path through the transesterification process

Triglycerides are changed into diglycerides, monoglycerides, and finally glycerol by a sequence of chemical events called transesterification, which is the process used to produce biodiesel. Alkali catalysts are usually used in a single phase of this process. However, a two-step procedure can be required if the feedstock has large concentrations of water or free fatty acids (FFAs). Fatty acid esters, or acid-catalyzed alcoholysis, are the initial step that turns FFAs into fatty acid esters. The process of transesterification, which turns the fatty acid esters into biodiesel, comes next. Figure  5 displays the schematic diagram for the manufacture of biodiesel in both one and two steps.

An external file that holds a picture, illustration, etc.
Object name is 11356_2024_32027_Fig5_HTML.jpg

Biodiesel manufacturing can be done in two steps ( A ) or one step ( B ) (Fattah et al. 2014 ; Mofijur et al. 2014 )

Chemical or biological catalysts can be used in transesterification reactions that are catalyzed. The classification is shown in Fig.  6 . There are both homogeneous and heterogeneous chemical catalysts. Base or acid catalysts are included in the homogeneous catalyst. The heterogeneous catalyst is made up of nano-, biomass waste-based, base, and acid–base functionalities (Thangaraj et al. 2019 ). The choice of any catalyst is influenced by the following factors: oil quality, FFA content, operating conditions, necessary catalyst activity, cost, and availability (Tacias-Pascacio et al. 2019 ). The advantages and disadvantages of different catalysts are presented in Table  2 .

An external file that holds a picture, illustration, etc.
Object name is 11356_2024_32027_Fig6_HTML.jpg

Different catalysts used in biodiesel production (Rizwanul Fattah et al. 2020 )

Table 2

Pros and cons of utilizing various catalysts in the process of transesterification

CatalystsProsConsReferences
Homogeneous:

• They can facilitate faster reaction rates and higher conversion yields compared to other catalyst types (such as heterogeneous catalysts) due to their ability to uniformly distribute throughout the reaction mixture. This homogeneous distribution enables better contact between the catalyst and reactants, resulting in improved reaction kinetics and the overall efficiency of the process

• Exhibit greater selectivity, promoting the desired transesterification reaction while minimizing unwanted side reactions. This selectivity contributes to the higher purity and quality of the biodiesel product

• Versatility and adaptability to various feedstocks. They can effectively catalyze the transesterification of different types of triglycerides, expanding the range of potential biodiesel sources

• Operate under milder reaction conditions, such as lower temperatures and pressures, which can result in energy savings and reduced production costs

• One significant drawback is their potential for leaching into the biodiesel product, which can complicate the separation and purification processes. Homogeneous catalysts, being soluble in the reaction mixture, can remain in the biodiesel phase even after the completion of the transesterification reaction. This can increase the difficulty and cost associated with removing the catalyst from the final product, leading to additional purification steps

• Some homogeneous catalysts used in biodiesel production may have limited stability and reusability. Catalyst deactivation or degradation can occur over time, reducing their effectiveness and necessitating frequent replacement or regeneration. The need for catalyst regeneration or disposal can add to the overall cost and environmental impact of the biodiesel production process

• Impurities, including water, free fatty acids, and glycerol, that are present in the feedstock may cause certain homogeneous catalysts to become sensitive. These impurities can inhibit the catalytic activity or lead to unwanted side reactions, necessitating extra pre-treatment steps to ensure high-quality biodiesel production

Meher et al. ( ); Demirbas ( ); Kumar ( )
Heterogeneous:

• Easy separation: Heterogeneous catalysts, being solid materials, can be separated from the reaction mixture with ease. This characteristic simplifies the purification process and reduces the costs associated with catalyst recovery. The solid catalysts can be easily filtered or decanted, permitting effective catalyst and biodiesel product separation

• Catalyst reusability: Heterogeneous catalysts generally exhibit good stability and can be repeatedly utilized without experiencing a noticeable decrease in activity. This reusability feature reduces the overall catalyst cost and enhances the profitability of the manufacture of biodiesel

• Tolerance to impurities: Heterogeneous catalysts often demonstrate higher tolerance to impurities commonly found in feedstocks. This tolerance eases the overall process of producing biodiesel, potentially lowering production costs, and eliminating the need for extensive pre-treatment of the feedstock

• Environmental sustainability: Heterogeneous catalysts are considered environmentally friendly due to their reduced potential for leaching into the biodiesel product and lower toxicity compared to some homogeneous catalysts. This characteristic aligns with the goal of sustainable biodiesel production

• Mass transfer limitations: Heterogeneous catalysts often exhibit slower reaction rates compared to homogeneous catalysts due to mass transfer limitations. The reactants need to diffuse through the liquid phase to reach the catalyst surface, which can result in reduced reaction kinetics and longer reaction times

• Catalyst deactivation: Heterogeneous catalysts can experience deactivation or loss of activity over time, primarily due to the formation of surface contaminants or catalyst poisoning. Impurities present in the feedstock, such as water, free fatty acids, and glycerol, can interact with the catalyst and inhibit its activity, leading to decreased performance and the need for catalyst replacement or regeneration

• Complex catalyst preparation: The preparation of heterogeneous catalysts can be more complex compared to homogeneous catalysts. It may involve multiple steps, such as synthesis, activation, and modification, which can increase the production cost and require specialized expertise

• Limited catalyst selectivity: Some heterogeneous catalysts may exhibit lower selectivity, leading to the production of unwanted byproducts or side reactions, which may have an impact on the biodiesel product's quality and purity

Freedman et al. ( ); Meher et al. ; Balat ( )
Enzymatic:

• Mild reaction conditions: Enzymes typically operate under mild reaction conditions, including lower temperatures and atmospheric pressure. This mildness reduces energy requirements and can lead to cost savings in the biodiesel production process

• High specificity and selectivity: Enzymes exhibit high specificity towards the targeted reactions, promoting the desired transesterification while minimizing side reactions. This selectivity contributes to the higher purity and quality of the biodiesel product

• Tolerance to impurities: Enzymatic catalysts often demonstrate higher tolerance to impurities present in the feedstock, such as water and free fatty acids, compared to other catalyst types. This tolerance reduces the need for extensive pre-treatment of the feedstock, simplifying the overall biodiesel production process and potentially lowering production costs

• Biodegradability and environmental sustainability: Enzymes are biodegradable and environmentally friendly catalysts. They can be derived from renewable sources and offer the advantage of reduced environmental impact compared to traditional catalysts

• Catalyst reusability: Enzymes can be immobilized on solid supports, allowing for their repeated use in multiple batches of biodiesel production. This reusability feature reduces the overall catalyst cost and contributes to the economic viability of the process

• Higher cost: Enzymes, especially those derived from microbial sources, can be expensive compared to other catalyst types. The costs of enzyme production, purification, and immobilization can significantly contribute to the overall cost of biodiesel production

• Sensitivity to Reaction Conditions: Enzymes are sensitive to reaction conditions such as temperature, pH, and water content. The optimal conditions for enzyme activity may differ from the ideal conditions for transesterification, requiring careful control and optimization of the reaction parameters. Any deviation from the optimal conditions can result in reduced enzyme activity and lower biodiesel yields

• Limited operational stability: Enzymes can experience operational stability issues during prolonged use. Factors such as enzyme denaturation, microbial contamination, and enzyme leaching from the immobilization matrix can lead to decreased enzyme activity over time, necessitating enzyme replacement or regeneration

• Longer reaction times: Enzymatic transesterification reactions often require longer reaction times compared to other catalyst types. This is primarily due to the slower reaction kinetics associated with enzymatic catalysis, which can impact the overall productivity and efficiency of the biodiesel production process

Du et al. ( ); Xie and Wang ( )

Homogeneous catalysis

Homogeneous catalysts are the ones that are most frequently used in the production of biodiesel because they are simple to use and require less time to complete a reaction. When dissolving a homogeneous catalyst, a solvent that is in the same phase as all the reactants is usually utilized (Rizwanul Fattah et al. 2020 ). Bhuana et al. ( 2020 ) used leftover beef tallow and methanol as solvents, in a KOH catalyst to create biodiesel. Ethanol effectively functioned as a co-solvent, lowering reaction time by 61.11% and functioning as a low-polarity active ester exchange agent, which prevented soap formation and increased yield by 3.08%. Karmee et al. ( 2015 ) looked at the transesterification of algal oil using methanol as the solvent. Under identical reactional conditions, HCl outperformed the homogeneous H 2 SO 4 catalyst. Rice bran oil was esterified by Arora et al. ( 2015 ) utilizing sulfuric acid as a uniform catalyst. Studies have been done on how the oil-to-methanol molar ratio (1:5 to 1:30), catalyst concentration (0.15 to 1.0 wt%), and reaction temperature (45 to 60 °C) affect the conversion of FFA. Siddiqua ( 2015 ) transesterified palm oil to create biodiesel. Further observation shows that a mixture of methanol and NaOH at a reaction temperature of 55 °C produced the highest production. Abdulsalam ( 2023 ) assessed the process of turning thevetia peruviana seed oil into biodiesel using two different catalysts (NaOH and KOH). Separate NaOH and KOH pellets were dissolved in methanol to create different catalysts with different amounts of sodium and potassium methoxide, respectively. The NaOH catalyst’s conversion yield was 81.2%, 80.4%, and 89%, whereas the KOH catalyst’s conversion yield was 96.8%, 88.4%, and 84.0. Through the calcination of leftover pineapple leaves, de Barros et al. ( 2020 ) created a unique catalyst. After a 30-min reaction period at 60 °C, 4 wt% of catalyst, a molar ratio of 1:40 for oil to methanol, and an oil-to-biodiesel conversion of above 98%, a high catalytic activity was detected. This activity is likely connected to the 85-wt% presence of alkali/alkali metals (K, Ca, and Mg). Kasirajan ( 2021 ) used two-step procedures to produce biodiesel from Chrysophyllum albidum seed, a non-edible source. Transesterification is performed on oil from Chrysophyllum albidum seeds after esterification employing a homogeneous catalyst of H 2 SO 4 . The maximum oil-to-biodiesel conversion was 99.2 wt% when the ideal situations were achieved, which included a 1:9 oil-to-methanol molar ratio, 1 wt% KOH, 500 rpm, and 40 min at 65 °C. Belkhanchi et al. ( 2021 ) showed that transesterifying used frying oils (UFO) at 18 °C for 60 min of reaction in the presence of methanol using MeOH/UFO 6:1 mol proportion and 1 wt% of KOH yields the best conversion of UFO. Jain et al. ( 2023 ) gave a description of the homogeneous base catalyst-based single-step transesterification process for producing biodiesel from waste cooking oil that contains high levels of free fatty acids and algal oil. According to the findings, a biodiesel yield of 92% may be achieved under optimal conditions, which include a 1.5% catalyst (w/w), a methanol to oil ratio of 21:1, a time of 110 min, and a temperature of 50 °C. Saeed et al. ( 2021 ) investigated S. elongata algal for biodiesel creation. To evaluate transesterification to FAME, zeolitic catalysts, KOH, and HCl were used. KOH produced the maximum biodiesel yield (99.9%), which was obtained under the ideal reaction conditions of a 1.0% catalyst, 60 °C, 4 h, and a 1:4 volume ratio between oil and methanol. Table ​ Table3 3 highlights some of the recently published research on the use of several homogeneous catalyst types for biodiesel synthesis, various feedstock sources, experimental setups, and biodiesel yields.

Table 3

Different types of homogeneous catalysts used for biodiesel synthesis

Type of feedstockHomogeneous catalystExperimental conditions
Temperature (°C)
M:O molar ratio
Catalyst (wt. %)
time (h)
Biodiesel yield (%)References
oilPotassium hydroxide60 °C -9:1–1%-1 h95Kamran et al. ( )
Waste shark liver oilH SO 60 °C-10:1–5.9%-6.5 h99Al Hatrooshi et al.
Cotton oilKF/bentonite120 °C-13:1–6%-6 h95da Costa and de Andrade Lima ( )
seeds oilKOH60°C-10:1–0.6%-0.5 h99Perumal and Mahendradas ( )
Waste frying vegetable oilKOH60 °C-12:1–1.5%-1.5 h97OA et al. ( )
Tall oil fatty acidsH SO 55 °C-15:1–0.5%-1 h96.76Lawer-Yolar et al. ( )
Waste cooking oilNaOH40 °C-9:1–1%-2 h98.22Abdel-Hamid et al. ( )

Heterogeneous catalysis

Heterogeneous catalysts go through different phases or states than reactants. According to Melero et al. ( 2009 ), these are the catalysts that often generate active sites when reacting with their reactants. Greater oil/alcohol ratios and greater temperatures than in homogeneous catalysis are the primary disadvantages of this catalysis. The catalyst’s improved reusability and ease of separation and purification are other advantages. Mohamed et al. ( 2020 ) prepared by quickly pyrolyzing rice straw, a heterogeneous catalyst (RS-SO 3 H) was created. The yield of biodiesel was 90.37%. in ideal conditions: 20:1 methanol: oil molar ratio with a 10% catalyst at 70 °C for 6 h. Choksi et al. ( 2021 ) created a solid acid catalyst using the sulfonation carbonization process from a palm fruit bunch. After that, the catalyst was put through esterification and transesterification processes to produce biodiesel. Utilizing a 4% catalyst, a 21:1 methanol-to-oil molar ratio, and a 60 °C temperature, an optimal yield of 88.5 wt% methyl ester was obtained in 180 min. Aghel et al. ( 2019 ) wanted to improve a pilot-scale microreactor that used kettle limescale to turn used cooking oil (WCO) into biodiesel. The produced biodiesel had a maximum conversion of 93.41% at 61.7 °C, a catalyst concentration of 8.87 wt %, a methanol-to-oil 1.7:3 volumetric ratio, and 15 min. Bhatia et al. ( 2020 ) developed a heterogeneous catalyst to initiate the transesterification of used cooking oil by pyrolyzing waste cork. The greatest conversion (98%) for the heterogeneous catalyst produced at 600 °C occurred at alcohol:oil ratios of 25:1, catalyst loadings of 1.5% w/v, and temperatures of 65 °C. Sahani et al. ( 2019 ) used a solid-base catalyst called barium cerate in the transesterification procedure to produce biodiesel from Karanja oil. To synthesize perovskite barium cerate with maximum phase purity, the calcination temperature was optimized. At 1.2 wt% catalyst, 1:19 oil-to-methanol molar ratio, 65 °C, 100 min, and 600 rpm, karanja oil methyl ester with 98.3% conversion was obtained. Kamel et al. ( 2019 ) utilized the fig leaves that had undergone calcination, KOH activation, and activation. The highest conversion to biodiesel (92.73%) was obtained from fig leaves treated with KOH under ideal conditions (2 h of heating, a 6:1 alcohol/oil molar ratio, 1% catalyst, and 400 rpm). Singh et al. ( 2023 ) produced biodiesel from Jatropha curcas oil using the transesterification technique and calcium oxide. The results of the experiment demonstrate that at a methanol/oil ratio of 12:1, 65 °C, 3 h, and a catalyst concentration of 5 wt%, a biodiesel yield of 81.6% was produced. Carbon spheres were the heterogeneous acid catalyst that Nata et al. ( 2017 ) utilized. A maximum yield of 87% was achieved at 60 °C and 1 h when WCO was used as the feedstock to make biodiesel utilizing a C–SO 3 H acid catalyst. Du et al. ( 2019 ) converted Scenedesmus quadricauda algal oil into biodiesel using a cobalt-doped CaO catalyst. Cao was obtained from eggshells and calcined at 400, 700, and 900 °C. Todorović et al. ( 2019 ) conducted research on canola oil-based potassium-supported TiO 2 for biodiesel generation. At 55 °C for 5 h, with a 6 wt% catalyst and a 54/1 methanol/oil, the highest biodiesel output of > 90% was discovered. Salinas et al. ( 2012 ) created a carbon-based MgO catalyst for castor oil transesterification utilizing the sol–gel method. With a 96.5% biodiesel output at 6 wt% catalyst loading and a 12:1 ethanol/oil ratio at 75 °C for 1 h, the MgO/UREA-800 demonstrated remarkable catalytic activity. Gardy et al. ( 2019 ) made a strong, magnetic core–shell SO 4 /Mg–Al–FeO 3 heterogeneous catalyst with the use of surface functionalization, encapsulation, and stepwise coprecipitation. Utilizing the synthesized catalyst, the transesterification reaction was carried out with the highest possible yield of 98.5% at 9:1 methanol/WCO, 95 °C, and 5 h. Table ​ Table4 4 highlights some of the recently published research on the use of several heterogeneous catalyst types for biodiesel synthesis, various feedstock sources, experimental setups, and biodiesel yields.

Table 4

Different types of heterogeneous catalysts used for biodiesel synthesis

Type of feedstockHeterogeneous catalystExperimental conditions
Temperature (°C)
M:O molar ratio Catalyst (wt. %)
time (h)
Biodiesel Yield (%)References
Soybean oilPotassium methoxide80 °C-6:1–2%-0.25 h91Celante et al. ( )
oilClay-Na CO 60 °C-12:1–2%-1.5 h94.7Takase et al. ( )
Na ZrO 65 °C-15:1–5%-3 h99.9Martínez et al. ( )
Mixture of crop mustard and edible waste oilCalcium oxide catalyst prepared from fish bones55 °C-12:1–0.3%-5 h94.95Abbas Ghazali and Marahel ( )
Soybean oil banana trunk ash (MBTA)25 °C-6:1–0.07%-6 h98.39Rajkumari and Rokhum ( )
Waste cooking oil12-molybdophosphoric acid190 °C-90:1–5%-4 h94.5Gonçalves et al. ( )
Palm oilZinc oxide supported silver nanoparticles60 °C-10:1–10%-1 h97Laskar et al. ( )
Palm fatty acid distillateTea waste65 °C-9:1–4%-1.5 h97Rashid et al. ( )

Enzyme-based catalyst

Enzyme-based catalysts are produced from living things that speed up reactions while maintaining the stability of their composition (Amini et al. 2017 ). Extracellular lipases are the enzymes that have been isolated and processed from the microbial broth. In contrast, intracellular lipase remains inside the cell or in its walls of production (Gog et al. 2012 ). One drawback of employing extracellular enzymes as catalysts is the expense and difficulty of the separation and purification procedures (Rizwanul Fattah et al. 2020 ). The efficiency of the bio-catalyzed transesterification process is influenced by the enzyme’s source and the process variables (Aransiola et al. 2014 ). Enzymatic biodiesel production also has the advantages of being simple to remove, operating at a temperature between 35 and 45 °C, producing no byproducts, and allowing catalysts to be reused (Christopher et al. 2014 ). For the transesterification of low-grade fish oil, Marín-Suárez et al. ( 2019 ) used Novozym 435 lipase; the greatest FAEE yield was 82.91 wt% after 8 h, 35 °C, an excess of ethanol, and 1% catalyst. Novozym 435 can be used for 10 continuous cycles with a maximum activity decrease of 16%. Jayaraman et al. ( 2020 ) studied used cooking oil enzymatic transesterification with the use of pancreatic lipase to make methyl ester. The best reaction conditions were discovered to be methanol as the alcohol 3:1 M ratio, 1.5% enzyme concentration (by weight of WCO), 4 h reaction duration, 60 °C, and 88% yield after numerous attempts. Fatty acid methyl ester (FAME) was produced by Choi et al. ( 2018 ) produced FAME from the oil in rice bran by just adding methanol. The 83.4% yield was reached after 12 days under ideal conditions.

Nanocatalysts in transesterification

Nanocatalysts have garnered significant interest in the production of biodiesel. Because of their special qualities, which include a large active surface area, high reusability, better catalytic efficiency, high biodiesel conversion, and sustainability, nanocatalysts can be superior to conventional catalysts (Qiu et al. 2011 ). Since they are easily removed from the final products and retain their catalytic activity even after being reused several times, nanocatalysts are widely sought (Ahmed et al. 2023 ). There are numerous ways to create nanocatalysts. Among the techniques are microwave combustion, chemical vapor deposition, impregnation, and gas condensation (Quirino et al. 2016 ; Ambat et al. 2018 ). Some of the latest works on nanocatalysts for the transesterification reaction are listed in Table  5 .

Table 5

Various nanocatalysts in biodiesel production

FeedstockCatalystExperimental conditionsBiodiesel Yield (%)References
Temperature (°C)
M:O molar ratio
Catalyst (wt.%) time (h)
Waste cooking oil

Nano

CaO

60 °C-12:1–2.5%- 2 h 94Erchamo et al. ( )
Waste cooking oilSodium oxide impregnated on carbon nanotubes (CNTs)65 °C-20:1–3%-3 h97Ibrahim et al. ( )
Used cooking oilGraphene oxide and bimetal zirconium/strontium oxide nanoparticles120 °C-4:1–0.5%-1.5 h91Madhuranthakam et al. ( )
Used frying oilNano CaO50 °C-8:1–1%-1.5 h96Degfie et al. ( )
Used frying oilNano Mgo65 °C-24:1–2%-1 h93.3Ashok et al. ( )
Sunflower oilMgO/MgAl O nano-catalyst110 °C-12:1–3%-3 h95.7Alaei et al. ( )
Sunflower oilCs/Al/Fe O nano-catalyst58 °C-12:1–1%-2 h94.8Mostafa et al. ( )
Chicken fatCaO/CuFe O 70 °C-15:1–3%-4 h94.52Seffati et al. ( )
Waste cooking oilZnCuO/N-doped graphene (NDG)180 °C-15:1–10%-8 h97.1Kuniyil et al. ( )
Olive oilMagnetite nanoparticle-immobilized lipase37 °C-12:1–1%-1 h45Amruth Maroju et al. ( )
Microalgae oilFe O /ZnMg(Al)O solid65 °C-12:1–3%-3 h94Chen et al. ( )
Olive oilMgO nanoparticles60 °C-10:1–2%-2 h80Amirthavalli and Warrier ( )
Tannery wasteCs O loaded onto a nano-magnetic core65 °C-21:1–7%-5 h97.1Booramurthy et al. ( )
Used cooking oilBifunctional magnetic nano-catalyst65 °C-12:1–4%-2 h98.2Hazmi et al. ( )
, a marine macroalgaeClay with zinc oxide as nanocatalyst55 °C-9:1–8%-0.83 h97.43Kalavathy and Baskar ( )
oilZinc-doped calcium oxide nanocatalyst55 °C-9:1–6%-1.33 h89Naveenkumar and Baskar ( )
seed oilMgO/Fe O -SiO core–shell magnetic nanocatalyst70 °C-12:1–4.9%-4.1 h99Rahimi et al. ( )

Metal-oxide nanocatalysts

The most popular nanocatalysts are those based on metal oxide, and they play a crucial role in maximizing the synthesis of biodiesel. Nanoparticles that will be employed for transesterification catalysis have been created using the oxidized forms of numerous different metals, including Mg, Zn, and Ca (Pandit et al. 2023 ). Jamil et al. ( 2021 ) created highly efficient barium oxide using catalysts made of molybdenum oxide. Optimal conditions include 12 methanol/oil, 120 min, 65 °C, and a 4.5wt% catalyst. The best yield was achieved under these conditions, which resulted in a 97.8% yield. Sahani et al. ( 2020 ) produced biodiesel with a transesterification reaction involving used cooking oil and a mixed metal oxide catalyst made of Sr–Ti. Methanol as the alcohol in an 11:1 M ratio, 1% catalyst, an 80-min reaction period, and a temperature of 65 °C with 98% FAME conversion were found to be the best reaction conditions. In a study conducted by Tayeb et al. ( 2023 ), the production of biodiesel using a CaO catalyst through the transesterification of WCO was investigated. The study determined the optimal reaction parameters to be a WCO/methanol molar ratio of 1:6, a 1% CaO nanocatalyst, a reaction temperature of 70 °C, and a reaction duration of 85 min, which resulted in a 97% biodiesel yield.

Carbon nanocatalysts

Nanocatalysts are created from carbon materials, including graphene and reduced graphene oxides (Nizami and Rehan 2018 ). Due to their diverse structural, mechanical, thermal, and biocompatibility qualities, carbon nanocatalysts are good catalysts and have advantageous applications in electrocatalytic devices such as fuel cells and other electro-processing systems. CNTs are often manufactured from graphite sheets that have been wound into cylinder forms. They have a large surface area, measure in nanometers, and are incredibly biocompatible (Rai et al. 2016 ).

Zeolite nanocatalysts

Large exterior surface areas and the hydrophobic nature of nanozeolites increase enzyme access to the substrate. Natural zeolite materials are far less frequently used in commercial industries than synthetic-based products. Commercially available synthetic zeolites such as ZSM-5, X, Y, and beta are used primarily in the production of biodiesel (Abukhadra et al. 2019 ). Using zeolites from NaY, KL, and NaZSM-5, Wu et al. ( 2013 ) produced CaO catalysts that were utilized to catalyze the transformation of methanol with soybean oil. In comparison to pure CaO, the activities of synthesized catalysts were studied. It was discovered that after being supported by zeolites, the CaO catalyst’s activity improved, with the CaO/NaY catalyst showing the greatest performance. Using the CaO/NaY catalyst, methanol-to-soybean oil 9:1 molar ratio at 65 °C with a reaction period of 3 h, and a 3% catalyst were used to produce a 95% biodiesel yield. Firouzjaee and Taghizadeh ( 2017 ) synthesized a CaO/NaY-Fe 3 O 4 nano-magnetic catalyst that was employed for the generation of biodiesel. The ideal methanol-to-oil molar ratio is 8.78, the catalyst loading is 5.19% (30% CaO loaded on the surface nanomagnetic zeolite), and the reaction period is 4 h. The maximum methyl esters obtained are 95.37%.

Distinct behavior of nanocatalysts during biodiesel production

Nanocatalysts have been widely used in biodiesel production due to their high catalytic activity, low cost, and environmental friendliness. The properties of nanocatalysts can vary depending on the preparation method, which can affect their catalytic performance. For example, the size, shape, and surface area of the catalyst particles can influence the reaction kinetics and yield of biodiesel. Recent studies have investigated the effects of different preparation methods on the properties of nanocatalysts for biodiesel production. The preparation method and calcination temperature are important factors that can affect the properties and catalytic performance of nanocatalysts for biodiesel production. Further research is needed to optimize the preparation methods and properties of nanocatalysts to improve the efficiency and sustainability of biodiesel production. We can offer general insights into the variations of nanocatalysts throughout the biodiesel production process, focusing on the following aspects.

Catalyst types: Different generations of nanocatalysts may involve distinct types of materials. For instance, first-generation nanocatalysts might include basic materials, while second- or third-generation may involve more advanced materials like metal oxides, zeolites, or other nanostructured materials.

Particle size: Advances in nanotechnology enable the control of particle size in nanocatalysts. The particle size can significantly impact catalytic activity. Smaller particle sizes may provide larger surface areas and enhanced catalytic efficiency.

Functionalization: The functionalization of nanocatalysts with specific groups or ligands can vary across generations. Functionalization can influence the catalyst’s selectivity and stability during biodiesel production.

Reusability and stability: Reusability and recovery are the two main advantages of using heterogeneous nanocatalysts in the production of biodiesel. The nanocatalyst is recovered and utilized again at each stage of these processes, which include many cycles of producing biodiesel. Nanocatalysts are often recovered via chemical means. The intended product and any byproduct may be easily and quickly recovered from the reaction mixture thanks to heterogeneous catalysts. This type of catalyst eliminates the need for a washing step. The esterification method using nanocatalysts was proposed to have several benefits, including speedier mixing of the reactants and catalyst and easy and rapid separation from the reaction mixture (Pandit et al. 2023 ).

Synthesis methods: The methods used to synthesize nanocatalysts may evolve, affecting their structure and properties. Recent advancements might include greener synthesis approaches or techniques that enhance the reproducibility of catalysts.

In addition to the aspects mentioned, the surface chemistry of nanocatalysts can also vary across generations, affecting their catalytic behavior during biodiesel production. The surface chemistry of nanocatalysts can be modified through various methods, such as surface functionalization, doping, or coating, to tune their catalytic activity, selectivity, and stability. For instance, surface functionalization with organic molecules or inorganic ions can enhance the catalyst’s selectivity for specific reactions or improve its compatibility with the reaction medium. The use of nanocatalysts in biodiesel production also presents some challenges, such as the aggregation, fouling, and leaching of active species. These issues can lead to a decrease in catalytic activity and selectivity, as well as an increase in production costs. To address these challenges, researchers are exploring various strategies, such as surface modification, stabilization techniques, and immobilization methods, to improve the stability and reusability of nanocatalysts. In summary, the distinct behavior of nanocatalysts during biodiesel production is influenced by various factors, including catalyst type, particle size, functionalization, surface chemistry, synthesis methods, and stability. The optimization of these factors can lead to more efficient, selective, and sustainable biodiesel production processes. However, further research is needed to fully understand the underlying mechanisms and to develop new generations of nanocatalysts with enhanced performance and stability.

Transesterification reaction mechanism (alcoholysis)

A large variety of exchange reactions involving oils, fats, and other reactants may be explained by the reaction mechanism. This comprises three processes: (1) transesterification, a rearrangement that yields monoglyceride, diglyceride, or other esters; (2) acidolysis, which involves exchanging fatty acids to produce specific fatty acid products; and (3) alcoholysis, which produces methyl esters in reactions with monohydric alcohols and monyl glycerol in reactions with polyhydric alcohols. Natural vegetable oils, animal fats, and food industry waste oil may all be utilized as source materials for transesterification, a process that produces biodiesel. Methanol, ethanol, propanol, butanol, and pentanol are among the alcohols that can be utilized for transesterification. Because it is a cheap, short-chain, strong polar raw material that reacts rapidly with fatty acid glycerides, methanol is the most widely used of them. Also freely soluble in methanol are base catalysts. A catalytic agent in this reaction might be an acid, base, or enzyme. Base catalysts consist of carbonate, NaOH, KOH, and potassium and sodium alkaloids. Acid catalysts might be hydrochloric, phosphoric, or sulfuric acids. The enzyme lipase is a good catalyst for the esterification of alcohols to fatty acid glycerides. Figures  7 , ​ ,8, 8 , ​ ,9, 9 , and ​ and10 10 represent continuous reversible processes for transesterification reactions; every reaction yields a distinct type of alcohol (Kang et al. 2015 ; Sait et al.  2022 ; Li et al. 2020 ; Oyekunle et al. 2023 ).

An external file that holds a picture, illustration, etc.
Object name is 11356_2024_32027_Fig7_HTML.jpg

Continuous reversible processes of transesterification reactions

An external file that holds a picture, illustration, etc.
Object name is 11356_2024_32027_Fig8_HTML.jpg

Acid-catalyzed alcoholysis reaction mechanism

An external file that holds a picture, illustration, etc.
Object name is 11356_2024_32027_Fig9_HTML.jpg

Base-catalyzed alcoholysis reaction mechanism

An external file that holds a picture, illustration, etc.
Object name is 11356_2024_32027_Fig10_HTML.jpg

Enzyme-catalyzed alcoholysis reaction mechanism

Kinetic modeling’s significance for process optimization

Kinetic models of chemical processes are powerful tools for reactor design. The kinetic models are very helpful in choosing the best reaction conditions (temperature, pressure, mixing rate, etc.) for chemical or biochemical transformations in reactors or bioreactors. This maximizes the formation of desired products with the least material investment and financial resources. This also holds true for the many techniques used to produce biodiesel, such as homogeneous, heterogeneous, enzyme catalysis, and others. One of the most important stages in the development of chemical processes for industrial applications is thought to be carefully thought-out experimental research and the subsequent creation of a kinetic model (Trejo-Zárraga et al. 2018 ). Portha et al. ( 2012 ) were able to decrease the extra ethanol used in the transesterification reaction in a continuous mode. By adjusting the temperature of the second reactor and adding methanol in stages, they were able to enhance the system’s overall performance, as demonstrated by the results of their simulation. Using triolein as a model chemical, the authors conducted experiments and discovered that it was beneficial to convert diglyceride and monoglyceride in the second reactor and the majority of triolein in the first. Additionally, their calculations suggested that to improve reaction rates at this point, it would be prudent to raise the temperature in the second reactor. Additionally, the authors computed internal concentration profiles using a reactor model that included the kinetic model. They discovered the limiting phenomenon in the overall transformation. To get a deeper comprehension of the rates of output and the inhibitory patterns seen in the transformation scheme, a kinetic model may also be strategically employed (Firdaus et al. 2016 ). For instance, a reaction scheme for the enzymatic creation of biodiesel might consider many more reaction stages and, consequently, a greater number of parameters. This adds difficulty to the kinetic model creation process, but once this model is solved, it may be utilized to construct an enzyme-catalyzed reactor and eventually optimize the process. The use of kinetic models, which can faithfully replicate the process at various reaction conditions, is helpful in the field of research and process improvement as it offers guidelines for additional experimental work and helps eliminate potentially fruitless experimental trials. Additionally, models may be utilized to foresee how composition will affect the final product’s quality. A model might forecast, for instance, how the feedstock’s water content or FFA may impact the reaction conversion and, in turn, the biodiesel’s production and quality.

Kinetics of transesterification

Few studies have dealt with kinetic modeling; most of the heterogeneous catalysis research has been on the manufacture and utilization of catalysts. To achieve reaction conditions with inherent kinetics and minimal effects, efforts have been focused on using tiny solid particles. It has been discovered that most heterogeneous transesterifications adhere to a pseudo-first-order model. For instance, Kaur and Ali ( 2014 ) discovered that the ethanolysis of Jatropha curcas L. oil, which were catalyzed by 15-Zr/CaO-700, adhered to a pseudo-first-order rate law. The Koros-Nowak test proved that the transit impacts were insignificant. Lukić et al. ( 2014 ) also discovered a first-order reversible rate law under ideal circumstances for the transesterification of sunflower oil. Table ​ Table6 6 lists some kinetic modeling studies of heterogeneous transesterification.

Table 6

List of some kinetic modeling studies of heterogeneous transesterification

FeedstockCatalystExperimental conditionsKinetic studiesReferences
Temperature (°C)
M:O molar ratio
Mixing speed (rpm)
Kinetic model rate constant ( )
activation energy ( )
Soybean oilAmberlyst A 6-OH basic ion-exchange resin50 °C-10:1–550 rpm

Eley–Rideal

 = 1.94 h.

 = 7.48 × 10  h

Jamal et al. ( )
LZr/CaO65 °C-15:1–500 rpm

Pseudo-first-order

 = 0.062 min

 = 29.8 kJ mol

Kaur and Ali ( )
Sunflower oilCaO60 °C-6:1–900 rpm

Miladinovic model

 = 0.063 dm mol  min

Tasić et al. ( )
Waste cooking oilNaOH/chitosan-Fe O 65 °C-6.5:1–500 rpm

Pseudo-first-order

 = 260.05 min

 = 21 kJ/mol

Helmi and Hemmati ( )
Sunflower oilCa(OH) 60 °C-6:1–900 rpm

Pseudo-first order

 = 0.07(1 − exp(−  /2.86); min

Stamenković et al. ( )
Used frying oilNaOH55 °C-4:1–300 rpm

Pseudo-first-order

 = 545.65 min

 = 23.61 kJ/mol

Haryanto et al. ( )
Sunflower oilCaO60 °C-6:1–900 rpm

Pseudo-first order

 = 0.07 min

Veljković et al. ( )
Canola oilMg–Co–Al–La HDL170–200 °C-16:1–900 rpm

First order

: 60.5 kJ/mol

Li et al. ( )
Waste cooking oil

CaO·ZnO

2 wt %

96 °C-10:1–300 rpm

Pseudo-first-order

 = 0.170 min

Lukić et al. ( )
Used cooking oilNano-cobalt-doped ZnO50–80 °C-3:1–136 rpm

Pseudo-second-order

 = 0.0052 min

Noreen et al. ( )
Waste cooking oilHeteropoly acid, 10 wt %70 °C-70:1–300 rpm

First order

 = 0.1062 min

 = 53.99 kJ/mol

Talebian-Kiakalaieh et al. ( )

Characterization methods for the assessment of produced biodiesels

Characterization methods for the assessment of produced biodiesel include various analytical techniques to evaluate the quality and properties of biodiesel. These methods are essential for ensuring that biodiesel meets the required standards and specifications for use as a sustainable and efficient alternative fuel source. The American Society for Testing and Materials (ASTM) is a prominent organization that provides authoritative guidelines for biodiesel testing and characterization methods.

The most common characterization methods for assessing produced biodiesel include the following.

Fatty acid methyl ester (FAME) analysis: FAME analysis is a fundamental method for biodiesel characterization, involving the determination of the fatty acid methyl ester content in biodiesel. This analysis is typically performed using gas chromatography (GC) or high-performance liquid chromatography (HPLC) to quantify individual FAME components, which provides valuable information about the biodiesel’s composition and purity.

Viscosity measurement: Viscosity is a crucial parameter for biodiesel quality assessment, as it affects the flow behavior and performance of the fuel. Dynamic viscosity measurements are commonly conducted to determine the resistance of biodiesel to flow under specific conditions, offering insights into its suitability for use in engines and transportation applications.

Oxidation stability testing: Biodiesel’s resistance to oxidation is an important characteristic that influences its shelf life and storage stability. Various methods, such as the Rancimat test and the PetroOXY test, are employed to assess the oxidation stability of biodiesel by measuring its susceptibility to oxidative degradation over time.

Cold flow properties analysis: The cold flow properties of biodiesel, including cloud point and pour point, are critical factors affecting its performance in cold weather conditions. Characterization methods such as differential scanning calorimetry (DSC) and automated cloud and pour point analyzers are utilized to determine these properties, ensuring that biodiesel remains operational at low temperatures.

Acid value determination: The acid value of biodiesel indicates its acidity level, which can impact engine components and fuel system integrity. Acid value determination involves titration methods to quantify the amount of free fatty acids present in biodiesel, enabling the assessment of its corrosiveness and potential impact on engine performance.

Calorific value measurement: Calorific value, also known as heating value, represents the energy content of biodiesel and is crucial for evaluating its combustion efficiency and heat output. Bomb calorimetry is commonly used to measure the calorific value of biodiesel, providing essential data for assessing its energy potential as a fuel source.

Sulfur content analysis: Sulfur content determination is essential for ensuring compliance with environmental regulations and assessing the environmental impact of biodiesel combustion. Techniques such as X-ray fluorescence (XRF) spectroscopy or ultraviolet fluorescence analysis are employed to measure sulfur levels in biodiesel samples.

Glycerol content quantification: Glycerol content in biodiesel must be monitored to ensure compliance with quality standards and prevent potential issues related to fuel stability and engine performance. Analytical methods like gas chromatography coupled with flame ionization detection (GC-FID) are utilized for the accurate quantification of glycerol in biodiesel products.

These characterization methods collectively provide comprehensive insights into the chemical composition, physical properties, stability, and environmental impact of produced biodiesel, supporting quality control measures and regulatory compliance within the biofuel industry.

The recent development in biodiesel production

Biodiesel, a renewable and sustainable alternative to conventional diesel fuel, has seen significant developments in recent years. These advancements have focused on improving the efficiency of biodiesel production processes, expanding feedstock options, and enhancing the overall sustainability of biodiesel as a viable energy source. One notable recent development is the use of advanced catalysts in biodiesel production. Catalysts play a crucial role in the conversion of vegetable oils or animal fats into biodiesel through a process called transesterification. Researchers have been exploring various catalysts, such as solid acid catalysts, enzyme catalysts, and heterogeneous catalysts, to improve reaction rates, reduce energy consumption, and enhance biodiesel quality. These catalysts offer advantages like higher conversion rates, milder reaction conditions, and easier separation of the catalyst from the product (Garcia-Silvera et al. 2023 ). Another significant development is the utilization of non-traditional feedstocks for biodiesel production. While conventional biodiesel feedstocks include soybean oil and rapeseed oil, researchers have been investigating alternative sources such as algae, waste cooking oil, and non-food crops like jatropha and camelina. Algae have gained attention due to their high oil content and ability to grow in various environments. The use of non-traditional feedstocks helps to reduce competition with food production and enhances the overall sustainability of biodiesel (Garg et al. 2023 ). Furthermore, efforts have been made to improve the sustainability of biodiesel production by reducing its environmental impact. This includes optimizing production processes to minimize water and energy consumption, reducing greenhouse gas emissions, and implementing waste management strategies. Additionally, researchers have been exploring the concept of “second-generation” biodiesel, which involves utilizing waste materials, such as agricultural residues and lignocellulosic biomass, to produce biodiesel. This approach not only reduces waste but also maximizes resource utilization (Makepa et al. 2023 ).

Biodiesel engine performance and emissions

Compared to petrodiesel fuel, burning biodiesel releases fewer particulates, carbon monoxide, and unburned hydrocarbons. Since biodiesel is produced using natural resources, its sulfur content is relatively low, which means that when it burns in an engine, it releases less sulfur dioxide into the atmosphere (Rayati et al. 2020 ). All biodiesels and their blends have shown the capacity to enhance gas turbine performance while lowering emissions of carbon dioxide, carbon monoxide, nitrogen oxide, and hydrocarbons under a range of operating conditions. To employ fuels in an engine, one must be aware of their characteristics for combustion. Although fossil fuel-based diesel fuel may not be entirely replaced by biodiesel, it can aid in achieving balanced energy utilization. One benefit is that biodiesel may be used in contemporary engines with little modification. Older vehicles with natural rubber gasoline lines, however, require a few modifications. Rubber fuel lines must be replaced since they will crack when used with biodiesel. On the other hand, an oil or gasoline dilution in the fuel system is possible in a modern vehicle with a DPF (diesel particulate filter). The ability of gasoline to lubricate the fuel injection system is believed to be crucial for diesel engines. The use of diesel–biodiesel mixes can thereby enhance their general lubricity. Additionally, the lower sulfur level of today’s diesel fuel could affect its lubricity because the compounds that provided lubrication are no longer present (Veza et al. 2022 ).

Techno-economic analysis

Techno-economic analysis (TEA) plays a crucial role in assessing the economic feasibility and viability of biodiesel production processes. It involves evaluating the overall costs, revenues, and profitability of biodiesel production, considering various factors such as feedstock costs, capital investment, operational expenses, and market prices. Recent studies have employed TEA to analyze and optimize biodiesel production processes, providing valuable insights for decision-making and process design. One example of TEA in biodiesel production is a study conducted by Zhang ( 2021 ), which evaluated the techno-economic performance of different feedstocks and process configurations for biodiesel production. The analysis considered factors such as feedstock availability, conversion efficiency, capital costs, operating costs, and market prices. The study highlighted the importance of feedstock selection and process optimization in achieving cost-effective biodiesel production. Another study by Tasić ( 2020 ) performed TEA for manufacturing biodiesel from used cooking oil. The analysis included the estimation of capital and operational costs, energy consumption, and environmental impacts. The study demonstrated the economic feasibility of waste cooking oil-based biodiesel production and identified critical parameters affecting the overall economics of the process. Furthermore, a study by Atabani ( 2020 ) conducted TEA for biodiesel production from microalgae. The analysis considered various scenarios, including different cultivation systems and conversion technologies. The study assessed the economic viability of microalgae-based biodiesel production, considering factors such as biomass productivity, lipid content, capital investment, and operational costs. These recent studies emphasize the importance of TEA in evaluating the economic aspects of biodiesel production. By considering a comprehensive range of factors, TEA provides valuable insights into the cost-effectiveness, profitability, and sustainability of biodiesel production processes, helping guide decision-making and process optimization.

Challenges, perspectives, and further research

The homogeneous catalyst has been thoroughly examined, and the literature has addressed several issues. However, heterogeneous catalysts are a very new field of study, and there is now a lot of research being done in this area. The literature has documented many obstacles regarding these catalysts:

  • It has been said that the primary issues with heterogeneous catalysts are instability, reduced reaction rate, and short catalyst life.
  • It has been reported that solid-base catalysts are FFA, CO 2 , and water sensitive. Through saponification, they destroy the catalyst and render it inactive.
  • Because water hydrolyzes the ionic group in solid acid catalysts, leaching and product contamination have been documented.
  • It has been documented that during enzymatic transesterification, methanol causes lipase inhibition.
  • In the case of nanocatalysts, good performance requires increasing the reaction time under very moderate working conditions. To attain normal reaction times, however, harsh working conditions must be used, which raises energy consumption.

Future research should pay attention to the following recommendations:

  • To create novel catalysts with enhanced catalytic performance, more research into waste-derived catalysts is required.
  • The creation of highly selective and active heterogeneous catalysts that may be used in industrial settings at a reasonable cost.
  • Investigating novel catalyst supports with a network of linked pores of the right size and a selected surface area.
  • Investigating the use of waste or biomass as a catalyst source to lower related costs and increase sustainability for solid catalysts that are sold commercially.
  • Improving synthesis catalyst production procedures and treatment stages to move the technology from a lab to an industrial setting.
  • Maintaining the high basic strength of the synthesis catalyst while improving the shape and sensitivity to FFA and water.
  • Additional research is needed into the manufacturing of industrial enzymatic biodiesel as a guaranteed choice for the future.
  • To give relevant information about the suitable catalyst, future research on the catalytic mechanism of partial catalysts should examine the features in depth.
  • Ways to recover and reuse nanocatalysts that are both economical and energy efficient.
  • Kinetic studies of the biodiesel production reaction using the synthesized catalyst should be carried out.

This extensive review delves into the various aspects of biodiesel production and its promise as a sustainable alternative for a greener energy future. The significance of feedstock selection and preparation is emphasized, with effective techniques discussed for optimizing biodiesel production efficiency and quality. Biodiesel has emerged as a versatile and promising alternative for transportation, industrial processes, and energy generation, demonstrating its potential to reduce greenhouse gas emissions and dependency on fossil fuels. The key process of transesterification is thoroughly examined, encompassing the utilization of diverse catalysts, including homogeneous, heterogeneous, enzyme based, and nanomaterials. The unique characteristics and performance of nanomaterials in transesterification are highlighted, offering prospects for enhanced efficiency and selectivity. Understanding the reaction mechanism and kinetics of transesterification is crucial for optimizing the production process. Kinetic modeling is identified as a valuable tool for process optimization, enabling better control and improved production efficiency. Methods for assessing the quality and properties of produced biodiesel are discussed, highlighting the importance of accurate characterization to meet quality standards and ensure compatibility with engine systems. Recent developments in biodiesel production showcase progress in feedstock selection, process optimization, and sustainability. However, challenges related to engine performance, emissions, and compatibility remain obstacles to wider biodiesel adoption. Future research should focus on addressing these challenges through innovative engine technologies, improved fuel formulations, and effective emission control strategies. Techno-economic analysis provides insights into the economic feasibility of biodiesel production, considering factors such as feedstock costs, process efficiency, and market demand. Ongoing analysis and assessment are essential for ensuring the commercial viability and scalability of biodiesel production. In conclusion, biodiesel presents a promising sustainable solution, but its advancement requires continuous research, development, and collaboration among academia, industry, and policymakers. Addressing challenges, pursuing further research, and implementing the recommendations outlined in this review will contribute to the widespread adoption of biodiesel as a renewable energy source, paving the way for a cleaner and more sustainable future.

Author contribution

All authors contributed to the study conception and design. Data collection and analysis were performed by Sabah Mohamed Farouk, Aghareed M. Tayeb, Shereen M. S. Abdel-Hamid, and Randa M. Osman.

Open access funding provided by The Science, Technology & Innovation Funding Authority (STDF) in cooperation with The Egyptian Knowledge Bank (EKB).

Data Availability

Declarations.

Not applicable.

The authors declare no competing interests.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

  • Abbas Ghazali A, Marahel F. Determining the characteristics of biodiesel from a mixture of crop mustard and edible waste oil using heterogeneous calcium oxide catalyst prepared from fish bones and predicting the efficiency of biodiesel using Taguchi statistical program model. J Phys Theor Chem IAU Iran. 2022; 18 :37–47. doi: 10.30495/JPTC.2022.21451. [ CrossRef ] [ Google Scholar ]
  • Abdel-Hamid SMS, Tharwat KM, Gad AM, Asmaa AY, Sohair TA. Process optimization for biodiesel production and its effect on the performance of diesel engines. Chem Eng Technol. 2023; 46 :694–701. doi: 10.1002/ceat.202200157. [ CrossRef ] [ Google Scholar ]
  • Abdulrahman RK (2017) Sustainable biodiesel production from waste cooking oil and chicken fat as an alternative fuel for diesel engine. Eur Sci J 13. 10.19044/esj.2016.v13n3p235
  • Abdulsalam H. Production of biodiesel from yellow oleander (Thevetin peruviana (Pers. K Schum) seed oil. Arid Zone J Basic Appl Res. 2023; 2 :1–9. doi: 10.55639/607.201918. [ CrossRef ] [ Google Scholar ]
  • Abukhadra MR, Ibrahim SM, Yakout SM, Mohamed E. Synthesis of Na+ trapped bentonite/zeolite-P composite as a novel catalyst for effective production of biodiesel from palm oil; Effect of ultrasonic irradiation and mechanism. Energy Convers Manag. 2019; 196 :739–750. doi: 10.1016/j.enconman.2019.06.027. [ CrossRef ] [ Google Scholar ]
  • 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. doi: 10.1016/j.renene.2019.04.100. [ CrossRef ] [ Google Scholar ]
  • Ahmed M, Ahmad KA, Vo DVN, Mohammed Y(2023) Recent trends in sustainable biodiesel production using heterogeneous nanocatalysts: function of supports, promoters, synthesis techniques, reaction mechanism, and kinetics and thermodynamic studies. Energy Convers Manag 280. 10.1016/j.enconman.2023.116821
  • Al Hatrooshi AS, Eze VC, Harvey AP. Production of biodiesel from waste shark liver oil for biofuel applications. Renew Energy. 2020; 145 :99–105. doi: 10.1016/j.renene.2019.06.002. [ CrossRef ] [ Google Scholar ]
  • Alaei S, Haghighi M, Toghiani J, Rahmani Vahid B. Magnetic and reusable MgO/MgFe2O4 nanocatalyst for biodiesel production from sunflower oil: influence of fuel ratio in combustion synthesis on catalytic properties and performance. Ind Crops Prod. 2018; 117 :322–332. doi: 10.1016/j.indcrop.2018.03.015. [ CrossRef ] [ Google Scholar ]
  • Ali A, Li B, Lu Y, Zhao C. Highly selective and low-temperature hydrothermal conversion of natural oils to fatty alcohols. Green Chem. 2019; 21 :3059–3064. doi: 10.1039/C9GC01260E. [ CrossRef ] [ Google Scholar ]
  • Al-Mawaali S, Al-Balushi K, Souissi Y (2023) Sustainable biodiesel production from waste cooking oil and waste animal fats. In: Proceedings of the 8th World Congress on Civil, Structural, and Environmental Engineering. Avestia Publishing. 10.11159/iceptp23.145
  • Alptekin E, Canakci M, Sanli H. Biodiesel production from vegetable oil and waste animal fats in a pilot plant. Waste Manag. 2014; 34 :2146–2154. doi: 10.1016/j.wasman.2014.07.019. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Ambat I, Srivastava V, Sillanpää M. Recent advancement in biodiesel production methodologies using various feedstock: a review. Renew Sustain Energy Rev. 2018; 90 :356–369. doi: 10.1016/j.rser.2018.03.069. [ CrossRef ] [ Google Scholar ]
  • Amini Z, Ilham Z, Ong HC, Hoora M, Wei HC. State of the art and prospective of lipase-catalyzed transesterification reaction for biodiesel production. Energy Convers Manag. 2017; 141 :339–353. doi: 10.1016/j.enconman.2016.09.049. [ CrossRef ] [ Google Scholar ]
  • Amirthavalli V, Warrier AR (2019) Production of biodiesel from waste cooking oil using MgO nanocatalyst. In: AIP Conference Proceedings. American Institute of Physics Inc. 10.1063/1.5113448
  • Amruth Maroju P, Ganesan R, Ray Dutta J. Biofuel generation from food waste through immobilized enzymes on magnetic nanoparticles. Mater Today Proc. 2023; 72 :62–66. doi: 10.1016/j.matpr.2022.05.555. [ CrossRef ] [ Google Scholar ]
  • Aransiola EF, Ojumu TV, Oyekola OO. A review of current technology for biodiesel production: state of the art. Biomass Bioenergy. 2014; 61 :276–297. doi: 10.1016/j.biombioe.2013.11.014. [ CrossRef ] [ Google Scholar ]
  • Arora R, Toor AP, Wanchoo RK. Esterification of high free fatty acid rice bran oil: parametric and kinetic study. Chem Biochem Eng Q. 2015; 29 :617–623. doi: 10.15255/CABEQ.2014.2117. [ CrossRef ] [ Google Scholar ]
  • Ashok A, Kennedy LJ, Vijaya JJ, Aruldoss U. Optimization of biodiesel production from waste cooking oil by magnesium oxide nanocatalyst synthesized using coprecipitation method. Clean Technol Environ Policy. 2018; 20 :1219–1231. doi: 10.1007/s10098-018-1547-x. [ CrossRef ] [ Google Scholar ]
  • Atabani AE. Techno-economic analysis of biodiesel production from microalgae: a review. Renew Sustain Energy Rev. 2020; 124 :109770. doi: 10.30638/eemj.2013.191. [ CrossRef ] [ Google Scholar ]
  • Atadashi IM, Aroua MK, Abdul Aziz AR, Sulaiman NMN. The effects of water on biodiesel production and refining technologies: a review. Renew Sustain Energy Rev. 2012; 16 :3456–3470. doi: 10.1016/j.rser.2012.03.004. [ CrossRef ] [ Google Scholar ]
  • Avhad MR, Marchetti JM. A review on recent advancement in catalytic materials for biodiesel production. Renew Sustain Energy Rev. 2015; 50 :696–718. doi: 10.1016/j.rser.2015.05.038. [ CrossRef ] [ Google Scholar ]
  • Azizian H, Kramer JKG. A rapid method for the quantification of fatty acids in fats and oils with emphasis on trans fatty acids using Fourier transform near infrared spectroscopy (FT-NIR) Lipids. 2005; 40 :855–867. doi: 10.1007/s11745-005-1448-3. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Balat M. Potential alternatives to edible oils for biodiesel production – a review of current work. Energy Convers Manag. 2011; 52 :1479–1492. doi: 10.1016/j.enconman.2010.10.011. [ CrossRef ] [ Google Scholar ]
  • Belkhanchi H, Rouan M, Hammi M. Synthesis of biodiesel by transesterification of used frying oils (Ufo) through basic homogeneous catalysts (NaOH and KOH) Biointerface Res Appl Chem. 2021; 11 :12858–12868. doi: 10.33263/BRIAC115.1285812868. [ CrossRef ] [ Google Scholar ]
  • Bhatia SK, Gurav R, Choi TR (2020) Conversion of waste cooking oil into biodiesel using heterogenous catalyst derived from cork biochar. Bioresour Technol 302. 10.1016/j.biortech.2020.122872 [ PubMed ]
  • Bhuana DS, Qadariyah L, Panjaitan R, Mahfud M (2020) Optimization of biodiesel production from Chlorella sp through in-situ microwave-Assisted acid-catalyzed trans-esterification. In: IOP Conference Series: Materials Science and Engineering. Institute of Physics Publishing. 10.1088/1757-899X/732/1/012004
  • Booramurthy VK, Kasimani R, Subramanian D, Pandian S (2020) Production of biodiesel from tannery waste using a stable and recyclable nano-catalyst: an optimization and kinetic study. Fuel 260. 10.1016/j.fuel.2019.116373
  • Çakırca EE, Tekin GN, İlgen O, Akın AN. Catalytic activity of CaO-based catalyst in transesterification of microalgae oil with methanol. Energy Environ. 2019; 30 :176–187. doi: 10.1177/0958305X18787317. [ CrossRef ] [ Google Scholar ]
  • Celante D, Schenkel JVD, de Castilhos F. Biodiesel production from soybean oil and dimethyl carbonate catalyzed by potassium methoxide. Fuel. 2018; 212 :101–107. doi: 10.1016/j.fuel.2017.10.040. [ CrossRef ] [ Google Scholar ]
  • Chen Y, Liu T, He H, Liang H. Fe3O4/ZnMg(Al)O magnetic nanoparticles for efficient biodiesel production. Appl Organomet Chem. 2018; 32 :e4330. doi: 10.1002/aoc.4330. [ CrossRef ] [ Google Scholar ]
  • Choi N, No DS, Kim H. In situ lipase-catalyzed transesterification in rice bran for synthesis of fatty acid methyl ester. Ind Crops Prod. 2018; 120 :140–146. doi: 10.1016/j.indcrop.2018.04.049. [ CrossRef ] [ Google Scholar ]
  • Choksi H, Pandian S, Arumugamurthi SS. Production of biodiesel from high free fatty acid feedstock using heterogeneous acid catalyst derived from palm-fruit-bunch. Energy Sources, Part a: Recovery, Util Environ Eff. 2021; 43 :3393–3402. doi: 10.1080/15567036.2019.1623953. [ CrossRef ] [ Google Scholar ]
  • Christopher LP, Kumar H, Zambare VP. Enzymatic biodiesel: challenges and opportunities. Appl Energy. 2014; 119 :497–520. doi: 10.1016/j.apenergy.2014.01.017. [ CrossRef ] [ Google Scholar ]
  • da Costa JM, de Andrade Lima LRP (2021) Transesterification of cotton oil with ethanol for biodiesel using a KF/bentonite solid catalyst. Fuel 293. 10.1016/j.fuel.2021.120446
  • de Barros SS, Pessoa Junior WAG, Sá ISC (2020) Pineapple (Ananás comosus) leaves ash as a solid base catalyst for biodiesel synthesis. Bioresour Technol 312.10.1016/j.biortech.2020.123569 [ PubMed ]
  • Degfie TA, Mamo TT, Mekonnen YS (2019) Optimized biodiesel production from waste cooking oil (WCO) using calcium oxide (CaO) nano-catalyst. Sci Rep 9. 10.1038/s41598-019-55403-4 [ PMC free article ] [ PubMed ]
  • Demirbas A. Progress and recent trends in biodiesel fuels. Energy Convers Manag. 2009; 50 :14–34. doi: 10.1016/j.enconman.2008.09.001. [ CrossRef ] [ Google Scholar ]
  • Demirbas A (2008) A realistic fuel alternative for diesel engines.10.1007/978-1-84628-995-8
  • Devaraj K, Mani Y, Rawoof SAA. Feasibility of biodiesel production from waste cooking oil: lab-scale to pilot-scale analysis. Environ Sci Pollut Res. 2020; 27 :25828–25835. doi: 10.1007/s11356-020-09068-6. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Dhawane SH, Kumar T, Halder G. Recent advancement and prospective of heterogeneous carbonaceous catalysts in chemical and enzymatic transformation of biodiesel. Energy Convers Manag. 2018; 167 :176–202. doi: 10.1016/j.enconman.2018.04.073. [ CrossRef ] [ Google Scholar ]
  • Du W, Xu Y, Liu D, Zeng J. Comparative study on lipase-catalyzed transformation of soybean oil for biodiesel production with different acyl acceptors. J Mol Catal B Enzym. 2004; 30 :125–129. doi: 10.1016/j.molcatb.2004.04.004. [ CrossRef ] [ Google Scholar ]
  • Du L, Li Z, Ding S. Synthesis and characterization of carbon-based MgO catalysts for biodiesel production from castor oil. Fuel. 2019; 258 :116122. doi: 10.1016/j.fuel.2019.116122. [ CrossRef ] [ Google Scholar ]
  • Elgharbawy AS, Sadik WA, Sadek OM, Kasaby MA (2021) A review on biodiesel feedstocks and production technologies.10.4067/S0717-97072021000105098
  • Erchamo YS, Mamo TT, Workneh GA, Mekonnen YS (2021) Improved biodiesel production from waste cooking oil with mixed methanol–ethanol using enhanced eggshell-derived CaO nano-catalyst. Sci Rep 11. 10.1038/s41598-021-86062-z [ PMC free article ] [ PubMed ]
  • Fattah IMR, Masjuki HH, Kalam MA. Performance and emission characteristics of a CI engine fueled with Cocos nucifera and Jatropha curcas B20 blends accompanying antioxidants. Ind Crops Prod. 2014; 57 :132–140. doi: 10.1016/j.indcrop.2014.03.022. [ CrossRef ] [ Google Scholar ]
  • Firdaus MY, Brask J, Nielsen PM. Kinetic model of biodiesel production catalyzed by free liquid lipase from Thermomyces lanuginosus. J Mol Catal B Enzym. 2016; 133 :55–64. doi: 10.1016/j.molcatb.2016.07.011. [ CrossRef ] [ Google Scholar ]
  • Firouzjaee MH, Taghizadeh M. Optimization of process variables for biodiesel production using the nanomagnetic catalyst CaO/NaY-Fe3O4. Chem Eng Technol. 2017; 40 :1140–1148. doi: 10.1002/ceat.201600406. [ CrossRef ] [ Google Scholar ]
  • Freedman B, Butterfield RO, Pryde EH. Transesterification kinetics of soybean oil 1. J Am Oil Chem Soc. 1986; 63 :1375–1380. doi: 10.1007/BF02679606. [ CrossRef ] [ Google Scholar ]
  • Fukuda H, Kondo A, Noda H. Biodiesel fuel production by transesterification of oils. J Biosci Bioeng. 2001; 92 :405–416. doi: 10.1016/S1389-1723(01)80288-7. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Garcia-Silvera EE, Melendez-Mogollon I, Pérez-Arias A, de la Rosa YA (2023) Enzymatic Biodiesel Production: Challenges and Opportunities. In: Congress on Research, Development, and Innovation in Renewable Energies: Selected Papers from CIDiER 2022. Springer, pp 19–37. 10.1007/978-3-031-26813-7_2
  • Gardy J, Nourafkan E, Osatiashtiani A. A core-shell SO4/Mg-Al-Fe3O4 catalyst for biodiesel production. Appl Catal B. 2019; 259 :118093. doi: 10.1016/j.apcatb.2019.118093. [ CrossRef ] [ Google Scholar ]
  • Garg R, Sabouni R, Ahmadipour M. From waste to fuel: challenging aspects in sustainable biodiesel production from lignocellulosic biomass feedstocks and role of metal organic framework as innovative heterogeneous catalysts. Ind Crops Prod. 2023; 206 :117554. doi: 10.1016/j.indcrop.2023.117554. [ CrossRef ] [ Google Scholar ]
  • Ghosh N, Halder G. Current progress and perspective of heterogeneous nanocatalytic transesterification towards biodiesel production from edible and inedible feedstock: a review. Energy Convers Manag. 2022; 270 :116292. doi: 10.1016/j.enconman.2022.116292. [ CrossRef ] [ Google Scholar ]
  • Gog A, Roman M, Toşa M. Biodiesel production using enzymatic transesterification – current state and perspectives. Renew Energy. 2012; 39 :10–16. doi: 10.1016/j.renene.2011.08.007. [ CrossRef ] [ Google Scholar ]
  • Gonçalves MA, Lourenço Mares EK, da Luz PTS (2021) Biodiesel synthesis from waste cooking oil using heterogeneous acid catalyst: statistical optimization using linear regression model. J Renew Sustain Energy 13. 10.1063/5.0048147
  • Gui MM, Lee KT, Bhatia S. Feasibility of edible oil vs. non-edible oil vs. waste edible oil as biodiesel feedstock. Energy. 2008; 33 :1646–1653. doi: 10.1016/j.energy.2008.06.002. [ CrossRef ] [ Google Scholar ]
  • Hamed HH, Mohammed AE, Habeeb OA. Biodiesel production from waste cooking oil using homogeneous catalyst. Egypt J Chem. 2021; 64 :2827–2832. doi: 10.21608/ejchem.2021.62395.3339. [ CrossRef ] [ Google Scholar ]
  • Haryanto A, Gita AC, Saputra TW, Telaumbanua M. First order kinetics of biodiesel synthesis using used frying oil through transesterification reaction. Aceh Int J Sci Technol. 2020; 9 :1–11. doi: 10.13170/aijst.9.1.13297. [ CrossRef ] [ Google Scholar ]
  • Hayyan A, Alam MZ, Mirghani MES. Reduction of high content of free fatty acid in sludge palm oil via acid catalyst for biodiesel production. Fuel Process Technol. 2011; 92 :920–924. doi: 10.1016/j.fuproc.2010.12.011. [ CrossRef ] [ Google Scholar ]
  • Hazmi B, Rashid U, Ibrahim ML (2021) Synthesis and characterization of bifunctional magnetic nano-catalyst from rice husk for production of biodiesel. Environ Technol Innov 21. 10.1016/j.eti.2020.101296
  • Helmi M, Hemmati A (2021) Synthesis of magnetically solid base catalyst of NaOH/Chitosan-Fe3O4 for biodiesel production from waste cooking oil: optimization, kinetics and thermodynamic studies. Energy Convers Manag 248. 10.1016/j.enconman.2021.114807
  • Hoekman SK, Robbins C. Review of the effects of biodiesel on NOx emissions. Fuel Process Technol. 2012; 96 :237–249. doi: 10.1016/j.fuproc.2011.12.036. [ CrossRef ] [ Google Scholar ]
  • Ibrahim ML, Nik Abdul Khalil NNA, Islam A. Preparation of Na2O supported CNTs nanocatalyst for efficient biodiesel production from waste-oil. Energy Convers Manag. 2020; 205 :112445. doi: 10.1016/j.enconman.2019.112445. [ CrossRef ] [ Google Scholar ]
  • Jain S, Kumar N, Singh VP (2023) Transesterification of algae oil and little amount of waste cooking oil blend at low temperature in the presence of NaOH. Energies (Basel) 16. 10.3390/en16031293
  • Jain S, Sharma MP. Biodiesel production from Jatropha curcas oil. Renew Sustain Energy Rev. 2010; 14 :3140. doi: 10.1016/j.rser.2010.07.047. [ CrossRef ] [ Google Scholar ]
  • Jamal Y, Rabie A, Boulanger BO (2015) Determination of methanolysis rate constants for low and high fatty acid oils using heterogeneous surface reaction kinetic models. React Kinet Mech Catal 114:63–74. 10.1007/s11144-014-0780-5
  • Jamil F, Al-Riyami M, Al-Haj L. Waste Balanites aegyptiaca seed oil as a potential source for biodiesel production in the presence of a novel mixed metallic oxide catalyst. Int J Energy Res. 2021; 45 :17189–17202. doi: 10.1002/er.5609. [ CrossRef ] [ Google Scholar ]
  • Javidialesaadi A, Raeissi S. Biodiesel production from high free fatty acid-content oils: experimental investigation of the pretreatment step. APCBEE Proc. 2013; 5 :474–478. doi: 10.1016/j.apcbee.2013.05.080. [ CrossRef ] [ Google Scholar ]
  • Jayaraman J, Alagu K, Appavu P. Enzymatic production of biodiesel using lipase catalyst and testing of an unmodified compression ignition engine using its blends with diesel. Renew Energy. 2020; 145 :399–407. doi: 10.1016/j.renene.2019.06.061. [ CrossRef ] [ Google Scholar ]
  • Kalavathy G, Baskar G. Synergism of clay with zinc oxide as nanocatalyst for production of biodiesel from marine Ulva lactuca. Bioresour Technol. 2019; 281 :234–238. doi: 10.1016/j.biortech.2019.02.101. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Kamel DA, Farag HA, Amin NK. Utilization of Ficus carica leaves as a heterogeneous catalyst for production of biodiesel from waste cooking oil. Environ Sci Pollut Res. 2019; 26 :32804–32814. doi: 10.1007/s11356-019-06424-z. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Kamran E, Mashhadi H, Mohammadi A, Ghobadian B (2020) Biodiesel production from Elaeagnus angustifolia.L seed as a novel waste feedstock using potassium hydroxide catalyst. Biocatal Agric Biotechnol 25. 10.1016/j.bcab.2020.101578
  • Kang X, Guo X, You H. Biodiesel development in the global scene. Energy Sources Part B. 2015; 10 :155–161. doi: 10.1080/15567249.2010.535095. [ CrossRef ] [ Google Scholar ]
  • Kara K, Ouanji F, Lotfi EM. Biodiesel production from waste fish oil with high free fatty acid content from Moroccan fish-processing industries. Egypt J Pet. 2018; 27 :249–255. doi: 10.1016/j.ejpe.2017.07.010. [ CrossRef ] [ Google Scholar ]
  • Karmakar B, Halder G (2021) Accelerated conversion of waste cooking oil into biodiesel by injecting 2-propanol and methanol under superheated conditions: a novel approach. Energy Convers Manag 247. 10.1016/j.enconman.2021.114733
  • Karmee SK, Patria RD, Lin CSK. Techno-economic evaluation of biodiesel production from waste cooking oil—a case study of Hong Kong. Int J Mol Sci. 2015; 16 :4362–4371. doi: 10.3390/ijms16034362. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Kasirajan R. Biodiesel production by two step process from an energy source of Chrysophyllum albidum oil using homogeneous catalyst. S Afr J Chem Eng. 2021; 37 :161–166. doi: 10.1016/j.sajce.2021.05.011. [ CrossRef ] [ Google Scholar ]
  • Kaur N, Ali A. Kinetics and reusability of Zr/CaO as heterogeneous catalyst for the ethanolysis and methanolysis of Jatropha crucas oil. Fuel Process Technol. 2014; 119 :173–184. doi: 10.1016/j.fuproc.2013.11.002. [ CrossRef ] [ Google Scholar ]
  • Knothe G. Biodiesel and renewable diesel: a comparison. Prog Energy Combust Sci. 2010; 36 :364–373. doi: 10.1016/j.pecs.2009.11.004. [ CrossRef ] [ Google Scholar ]
  • Kumar SD (2021) Biodiesel production using homogeneous, heterogeneous, and enzyme catalysts via transesterification and esterification reactions: a critical review.  BioEnergy Res 15(2):935 961. https://doi.org/110.1007/s12155-021-10333-w [ PMC free article ] [ PubMed ]
  • Kuniyil M, Shanmukha Kumar JV, Adil SF. Production of biodiesel from waste cooking oil using ZnCuO/N-doped graphene nanocomposite as an efficient heterogeneous catalyst. Arab J Chem. 2021; 14 :102982. doi: 10.1016/j.arabjc.2020.102982. [ CrossRef ] [ Google Scholar ]
  • Laskar IB, Rokhum L, Gupta R, Chatterjee S (2020) Zinc oxide supported silver nanoparticles as a heterogeneous catalyst for production of biodiesel from palm oil. Environ Prog Sustain Energy 39. 10.1002/ep.13369
  • Lawer-Yolar G, Dawson-Andoh B, Atta-Obeng E. Synthesis of biodiesel from tall oil fatty acids by homogeneous and heterogeneous catalysis. Sustain Chem. 2021; 2 :206–221. doi: 10.3390/suschem2010012. [ CrossRef ] [ Google Scholar ]
  • Lee SB, Lee JD, Hong IK. Ultrasonic energy effect on vegetable oil based biodiesel synthetic process. J Ind Eng Chem. 2011; 17 :138–143. doi: 10.1016/j.jiec.2010.12.012. [ CrossRef ] [ Google Scholar ]
  • Li E, Xu ZP, Rudolph V. MgCoAl–LDH derived heterogeneous catalysts for the ethanol transesterification of canola oil to biodiesel. Appl Catal B. 2009; 88 :42–49. doi: 10.1016/j.apcatb.2008.09.022. [ CrossRef ] [ Google Scholar ]
  • Li F, Jiang J, Liu P. Catalytic cracking of triglycerides with a base catalyst and modification of pyrolytic oils for production of aviation fuels. Sustain Energy Fuels. 2018; 2 :1206–1215. doi: 10.1039/C7SE00505A. [ CrossRef ] [ Google Scholar ]
  • Li C, Zhang A, Liu R (2020) Principle and Processing of Biodiesel Production. Industrial Oil Plant: Application Principles and Green Technologies 127–156. 10.1007/978-981-15-4920-5_5
  • Liu Y, Sotelo-Boyás R, Murata K. Hydrotreatment of vegetable oils to produce bio-hydrogenated diesel and liquefied petroleum gas fuel over catalysts containing sulfided Ni–Mo and solid acids. Energy Fuels. 2011; 25 :4675–4685. doi: 10.1021/ef200889e. [ CrossRef ] [ Google Scholar ]
  • Lukić I, Kesić Ž, Maksimović S. Kinetics of sunflower and used vegetable oil methanolysis catalyzed by CaO· ZnO. Fuel. 2013; 113 :367–378. doi: 10.1016/j.fuel.2013.05.093. [ CrossRef ] [ Google Scholar ]
  • Lukić I, Kesić Ž, Skala D. Kinetics of heterogeneous biodiesel synthesis using supported ZnO as catalyst. Chem Eng Technol. 2014; 37 :1879–1884. doi: 10.1002/ceat.201300714. [ CrossRef ] [ Google Scholar ]
  • Madhuranthakam CMR, Kamyabi AMN, Almheiri GM, Elkamel A (2020) Sustainable Approach for the Production of Biodiesel from Waste Cooking Oil Using Static Mixer Technology. In: Proceedings of the International Conference on Industrial Engineering and Operations Management.10.46254/AN10.20200753
  • Makepa DC, Chihobo CH, Musademba D (2023) Techno-economic analysis and environmental impact assessment of biodiesel production from bio-oil derived from microwave-assisted pyrolysis of pine sawdust. Heliyon 9(11). 10.1016/j.heliyon.2023.e22261 [ PMC free article ] [ PubMed ]
  • Manzanera M, Molina-Munoz M, Gonzalez-Lopez J. Biodiesel: an alternative fuel. Recent Pat Biotechnol. 2008; 2 :25–34. doi: 10.2174/187220808783330929. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Marín-Suárez M, Méndez-Mateos D, Guadix A, Guadix EM. Reuse of immobilized lipases in the transesterification of waste fish oil for the production of biodiesel. Renew Energy. 2019; 140 :1–8. doi: 10.1016/j.renene.2019.03.035. [ CrossRef ] [ Google Scholar ]
  • Martchamadol J, Kumar S. Thailand’s energy security indicators. Renew Sustain Energy Rev. 2012; 16 :6103–6122. doi: 10.1016/j.rser.2012.06.021. [ CrossRef ] [ Google Scholar ]
  • Martínez A, Mijangos GE, Romero-Ibarra IC. A novel green one-pot synthesis of biodiesel from Ricinus communis seeds by basic heterogeneous catalysis. J Clean Prod. 2018; 196 :340–349. doi: 10.1016/j.jclepro.2018.05.241. [ CrossRef ] [ Google Scholar ]
  • Meher LC, Sagar DV, Naik SN. Technical aspects of biodiesel production by transesterification—a review. Renew Sustain Energy Rev. 2006; 10 :248–268. doi: 10.1016/j.rser.2004.09.002. [ CrossRef ] [ Google Scholar ]
  • Melero JA, Iglesias J, Morales G. Heterogeneous acid catalysts for biodiesel production: current status and future challenges. Green Chem. 2009; 11 :1285–1308. doi: 10.1039/b902086a. [ CrossRef ] [ Google Scholar ]
  • Mobin S, Alam F, Chowdhury H (2022) Environmental impact of algae-based biofuel production: A review. AIP Conference Proceedings. AIP Publishing. 2681(1). 10.1063/5.0117093
  • Mofijur M, Masjuki HH, Kalam MA. Comparative evaluation of performance and emission characteristics of Moringa oleifera and Palm oil based biodiesel in a diesel engine. Ind Crops Prod. 2014; 53 :78–84. doi: 10.1016/j.indcrop.2013.12.011. [ CrossRef ] [ Google Scholar ]
  • Mohamed RM, Kadry GA, Abdel-Samad HA, Awad ME. High operative heterogeneous catalyst in biodiesel production from waste cooking oil. Egypt J Pet. 2020; 29 :59–65. doi: 10.1016/j.ejpe.2019.11.002. [ CrossRef ] [ Google Scholar ]
  • Mostafa F, Hassankhani A, Rafiee H. Preparation and characterization of Cs/Al/Fe3O4 nanocatalysts for biodiesel production. Energy Convers Manag. 2013; 71 :62–68. doi: 10.1016/j.enconman.2013.03.022. [ CrossRef ] [ Google Scholar ]
  • Mrad N, Varuvel EG, Tazerout M, Aloui F. Effects of biofuel from fish oil industrial residue–diesel blends in diesel engine. Energy. 2012; 44 :955–963. doi: 10.1016/j.energy.2012.04.056. [ CrossRef ] [ Google Scholar ]
  • Nata IF, Putra MD, Irawan C, Lee C-K. Catalytic performance of sulfonated carbon-based solid acid catalyst on esterification of waste cooking oil for biodiesel production. J Environ Chem Eng. 2017; 5 :2171–2175. doi: 10.1016/j.jece.2017.04.029. [ CrossRef ] [ Google Scholar ]
  • Naveenkumar R, Baskar G. Biodiesel production from Calophyllum inophyllum oil using zinc doped calcium oxide (Plaster of Paris) nanocatalyst. Bioresour Technol. 2019; 280 :493–496. doi: 10.1016/j.biortech.2019.02.078. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Nisa K, Salwa DAA, Hasanah A, Widayat (2020) Biodiesel production from waste cooking oil by using zirconia catalyst. In: AIP Conference Proceedings. American Institute of Physics Inc 2217(1). 10.1063/5.0000912
  • Nizami A-S, Rehan M. Towards nanotechnology-based biofuel industry. Biofuel Res J. 2018; 5 :798–799. doi: 10.18331/BRJ2018.5.2.2. [ CrossRef ] [ Google Scholar ]
  • Noreen S, Sahar I, Masood N (2021) Thermodynamic and kinetic approach of biodiesel production from waste cooking oil using nano-catalysts.10.1515/zpc-2020-1644
  • Oa A, Ao A, Se A (2019) Process parameter estimation of biodiesel production from waste frying oil (vegetable and palm oil) using homogeneous catalyst. J Food Process Technol 10. 10.35248/2157-7110.19.10.811
  • Oh KK, Kim YS, Yoon HH, Tae BS. Pretreatment of lignocellulosic biomass using combination of ammonia recycled percolation and dilute-acid process. J Ind Eng Chem. 2002; 8 :64–70. [ Google Scholar ]
  • Olubunmi BE, Karmakar B, Aderemi OM (2020) Parametric optimization by Taguchi L9 approach towards biodiesel production from restaurant waste oil using Fe-supported anthill catalyst. J Environ Chem Eng 8. 10.1016/j.jece.2020.104288
  • Ong HC, Masjuki HH, Mahlia TMI. Engine performance and emissions using Jatropha curcas, Ceiba pentandra and Calophyllum inophyllum biodiesel in a CI diesel engine. Energy. 2014; 69 :427–445. doi: 10.1016/j.energy.2014.03.035. [ CrossRef ] [ Google Scholar ]
  • Oyekunle DT, Barasa M, Gendy EA, Tiong SK (2023) Heterogeneous catalytic transesterification for biodiesel production: Feedstock properties, catalysts and process parameters. Process Saf Environ Prot 177(11). 10.1016/j.psep.2023.07.064
  • Pandit C, Banerjee S, Pandit S (2023) Recent advances and challenges in the utilization of nanomaterials in transesterification for biodiesel production. Heliyon 9.9(4). 10.1016/j.heliyon.2023.e15475 [ PMC free article ] [ PubMed ]
  • Perumal G, Mahendradas DK. Biodiesel production from Bauhinia variegata seeds oil using homogeneous catalyst. Pet Sci Technol. 2022; 40 :857–870. doi: 10.1080/10916466.2021.2008968. [ CrossRef ] [ Google Scholar ]
  • Portha J-F, Allain F, Coupard V. Simulation and kinetic study of transesterification of triolein to biodiesel using modular reactors. Chem Eng J. 2012; 207 :285–298. doi: 10.1016/j.cej.2012.06.106. [ CrossRef ] [ Google Scholar ]
  • Qiu F, Li Y, Yang D. Heterogeneous solid base nanocatalyst: Preparation, characterization and application in biodiesel production. Bioresour Technol. 2011; 102 :4150–4156. doi: 10.1016/j.biortech.2010.12.071. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Quirino MR, Oliveira MJC, Keyson D. Synthesis of zinc aluminate with high surface area by microwave hydrothermal method applied in the transesterification of soybean oil (biodiesel) Mater Res Bull. 2016; 74 :124–128. doi: 10.1016/j.materresbull.2015.10.027. [ CrossRef ] [ Google Scholar ]
  • Rahimi T, Kahrizi D, Feyzi M, et al (2021) Catalytic performance of MgO /Fe2O3-SiO2 core-shell magnetic nanocatalyst for biodiesel production of Camelina sativa seed oil: optimization by RSM-CCD method. Ind Crops Prod 159. 10.1016/j.indcrop.2020.113065
  • Rai M, dos Santos JC, Soler MF (2016) Strategic role of nanotechnology for production of bioethanol and biodiesel.10.1515/ntrev-2015-0069
  • Rajkumari K, Rokhum L. A sustainable protocol for production of biodiesel by transesterification of soybean oil using banana trunk ash as a heterogeneous catalyst. Biomass Convers Biorefin. 2020; 10 :839–848. doi: 10.1007/s13399-020-00647-8. [ CrossRef ] [ Google Scholar ]
  • Rashid U, Ahmad J, Ibrahim ML (2019) Single-pot synthesis of biodiesel using efficient sulfonated-derived tea waste-heterogeneous catalyst. Materials 12. 10.3390/ma12142293 [ PMC free article ] [ PubMed ]
  • Rayati M, Rajabi Islami H, Shamsaie Mehrgan M. Light intensity improves growth, lipid productivity, and fatty acid profile of Chlorococcum oleofaciens (Chlorophyceae) for biodiesel production. Bioenergy Res. 2020; 13 :1235–1245. doi: 10.1007/s12155-020-10144-5. [ CrossRef ] [ Google Scholar ]
  • Rizwanul Fattah IM, Ong HC, Mahlia TMI (2020) State of the art of catalysts for biodiesel production. Front Energy Res 8. 10.3389/fenrg.2020.00101
  • Roy T, Ágarwal AK, Sharma YC. A cleaner route of biodiesel production from waste frying oil using novel potassium tin oxide catalyst: a smart liquid-waste management. Waste Manag. 2021; 135 :243–255. doi: 10.1016/j.wasman.2021.08.046. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Ruhul AM, Kalam MA, Masjuki HH, et al. State of the art of biodiesel production processes: a review of the heterogeneous catalyst. RSC Adv. 2015; 5 :101023–101044. doi: 10.1039/C5RA09862A. [ CrossRef ] [ Google Scholar ]
  • Sadaf S, Iqbal J, Ullah I, et al. Biodiesel production from waste cooking oil: an efficient technique to convert waste into biodiesel. Sustain Cities Soc. 2018; 41 :220–226. doi: 10.1016/j.scs.2018.05.037. [ CrossRef ] [ Google Scholar ]
  • Saeed A, Hanif MA, Hanif A, et al (2021) Production of biodiesel from spirogyra elongata, a common freshwater green algae with high oil content. Sustainability (Switzerland) 13. 10.3390/su132212737
  • Sahani S, Roy T, Sharma YC. Smart waste management of waste cooking oil for large scale high quality biodiesel production using Sr-Ti mixed metal oxide as solid catalyst: optimization and E-metrics studies. Waste Manag. 2020; 108 :189–201. doi: 10.1016/j.wasman.2020.04.036. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Sahani S, Roy T, Chandra Sharma Y (2019) Clean and efficient production of biodiesel using barium cerate as a heterogeneous catalyst for the biodiesel production; kinetics and thermodynamic study. J Clean Prod 237. 10.1016/j.jclepro.2019.117699
  • Sait HH, Hussain A, Bassyouni M, Ali I, Kanthasamy R, Ayodele BV, Elhenawy Y. Hydrogen-rich syngas and biochar production by non-catalytic valorization of date palm seeds. Energies. 2022; 15 (8):2727. doi: 10.3390/en15082727. [ CrossRef ] [ Google Scholar ]
  • Salinas D, Araya P, Guerrero S. Study of potassium-supported TiO2 catalysts for the production of biodiesel. Appl Catal B. 2012; 117–118 :260–267. doi: 10.1016/j.apcatb.2012.01.016. [ CrossRef ] [ Google Scholar ]
  • Sayed A, El-Gharbawy AA (2016) Production of biodiesel from non edible vegetable oil
  • 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. doi: 10.1016/j.fuel.2018.08.118. [ CrossRef ] [ Google Scholar ]
  • Sharma YC, Singh B. Development of biodiesel: current scenario. Renew Sustain Energy Rev. 2009; 13 :1646–1651. doi: 10.1016/j.rser.2008.08.009. [ CrossRef ] [ Google Scholar ]
  • Siddiqua S. Transesterification of palm oil to biodiesel and optimization of production conditions i.e. methanol, sodium hydroxide and temperature. J Energy Nat Resour. 2015; 4 :45. doi: 10.11648/j.jenr.20150403.12. [ CrossRef ] [ Google Scholar ]
  • Singh SP, Singh D. Biodiesel production through the use of different sources and characterization of oils and their esters as the substitute of diesel: a review. Renew Sustain Energy Rev. 2010; 14 :200–216. doi: 10.1016/j.rser.2009.07.017. [ CrossRef ] [ Google Scholar ]
  • Singh A, Choudhary AK, Sinha S. Analysis of vibrations in a diesel engine produced by Jatropha biodiesel using heterogeneous catalyst. Energy Environ. 2023; 34 :407–428. doi: 10.1177/0958305X211063935. [ CrossRef ] [ Google Scholar ]
  • Singh A, Hoque ME, Chuan YL. Biodiesel from low cost feedstocks: The effects of process parameters on the biodiesel yield. Biomass and Bioenergy. 2011; 35 (4):1582–1587. doi: 10.1016/j.biombioe.2010.12.024. [ CrossRef ] [ Google Scholar ]
  • Sousa LL, Lucena IL, Fernandes FAN. Transesterification of castor oil: effect of the acid value and neutralization of the oil with glycerol. Fuel Process Technol. 2010; 91 :194–196. doi: 10.1016/J.FUPROC.2009.09.016. [ CrossRef ] [ Google Scholar ]
  • Srivastava N, Srivastava M, Gupta VK. Recent development on sustainable biodiesel production using sewage sludge. 3 Biotech. 2018; 8 :1–11. doi: 10.1007/s13205-018-1264-5. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Stamenković OS, Veljković VB, Todorović ZB. Modeling the kinetics of calcium hydroxide catalyzed methanolysis of sunflower oil. Bioresour Technol. 2010; 101 :4423–4430. doi: 10.1016/j.biortech.2010.01.109. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Suzihaque MUH, Alwi H, Kalthum Ibrahim U. Biodiesel production from waste cooking oil: a brief review. Mater Today Proc. 2022; 63 :S490–S495. doi: 10.1016/j.matpr.2022.04.527. [ CrossRef ] [ Google Scholar ]
  • Tacias-Pascacio VG, Torrestiana-Sánchez B, Dal Magro L. Comparison of acid, basic and enzymatic catalysis on the production of biodiesel after RSM optimization. Renew Energy. 2019; 135 :1–9. doi: 10.1016/j.renene.2018.11.107. [ CrossRef ] [ Google Scholar ]
  • Takase M, Kipkoech R, Miller DL, Buami EK. Optimisation of the reaction conditions for biodiesel from Parkia biglobosa oil via transesterification with heterogeneous clay base catalyst. Fuel Commun. 2023; 15 :100089. doi: 10.1016/j.jfueco.2023.100089. [ CrossRef ] [ Google Scholar ]
  • Talebian-Kiakalaieh A, Amin NAS, Zarei A, Noshadi I. Transesterification of waste cooking oil by heteropoly acid (HPA) catalyst: optimization and kinetic model. Appl Energy. 2013; 102 :283–292. doi: 10.1016/j.apenergy.2012.07.018. [ CrossRef ] [ Google Scholar ]
  • Tasić MB. Techno-economic analysis of biodiesel production from waste cooking oil. Environ Sci Pollut Res. 2020; 27 :4183–4192. doi: 10.1007/s41247-018-0050-7. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Tasić MB, Miladinović MR, Stamenković OS. Kinetic modeling of sunflower oil methanolysis catalyzed by calcium-based catalysts. Chem Eng Technol. 2015; 38 :1550–1556. doi: 10.1002/ceat.201500076. [ CrossRef ] [ Google Scholar ]
  • Tayeb AM, Abdel-Hamid SMS, Osman RM, Farouk SM. Optimization of biodiesel production from waste cooking oil using nano calcium oxide catalyst. Chem Eng Technol. 2023 doi: 10.1002/ceat.202300238. [ CrossRef ] [ Google Scholar ]
  • Thangaraj B, Solomon PR, Muniyandi B. Catalysis in biodiesel production—a review. Clean Energy. 2019; 3 :2–23. doi: 10.1093/ce/zky020. [ CrossRef ] [ Google Scholar ]
  • Todorović ZB, Troter DZ, Đokić-Stojanović DR. Optimization of CaO-catalyzed sunflower oil methanolysis with crude biodiesel as a cosolvent. Fuel. 2019; 237 :903–910. doi: 10.1016/j.fuel.2018.10.056. [ CrossRef ] [ Google Scholar ]
  • Touqeer T, Mumtaz MW, Mukhtar H (2020) Fe3O4-PDA-lipase as surface functionalized nano biocatalyst for the production of biodiesel using waste cooking oil as feedstock: characterization and process optimization. Energies (Basel) 13. 10.3390/en13010177
  • Trejo-Zárraga F, de Jesús Hernández-Loyo F, Chavarría-Hernández JC, Sotelo-Boyás R (2018) Kinetics of transesterification processes for biodiesel production. Biofuels-State Dev 149–179. 10.5772/intechopen.75927
  • Tripathi G, Dubey P, Yadav P (2023) Advancement in biodiesel production methodologies using different feedstock. In: Wastewater Resource Recovery and Biological Methods. Springer, pp 323–341.10.1007/978-3-031-40198-5_15
  • Um B-H, Kim Y-S. Review: A chance for Korea to advance algal-biodiesel technology. J Ind Eng Chem. 2009; 15 :1–7. doi: 10.1016/j.jiec.2008.08.002. [ CrossRef ] [ Google Scholar ]
  • Veljković VB, Stamenković OS, Todorović ZB. Kinetics of sunflower oil methanolysis catalyzed by calcium oxide. Fuel. 2009; 88 :1554–1562. doi: 10.1016/j.fuel.2009.02.013. [ CrossRef ] [ Google Scholar ]
  • Veza I, Zainuddin Z, Tamaldin N. Effect of palm oil biodiesel blends (B10 and B20) on physical and mechanical properties of nitrile rubber elastomer. Results Eng. 2022; 16 :100787. doi: 10.1016/j.rineng.2022.100787. [ CrossRef ] [ Google Scholar ]
  • Wu H, Zhang J, Wei Q. Transesterification of soybean oil to biodiesel using zeolite supported CaO as strong base catalysts. Fuel Process Technol. 2013; 109 :13–18. doi: 10.1016/j.fuproc.2012.09.032. [ CrossRef ] [ Google Scholar ]
  • Xie W, Wang J. Enzymatic production of biodiesel from soybean oil by using immobilized lipase on Fe3O4/poly (styrene-methacrylic acid) magnetic microsphere as a biocatalyst. Energy Fuels. 2014; 28 :2624–2631. doi: 10.1021/ef500131s. [ CrossRef ] [ Google Scholar ]
  • Zhang Y. techno-economic analysis of biodiesel production from different feedstocks. Energy Convers Manag. 2021; 235 :113948. doi: 10.3390/en14051473. [ CrossRef ] [ Google Scholar ]

Advertisement

Advertisement

RSM-based comparative experimental study of sustainable biodiesel synthesis from different 2G feedstocks using magnetic nanocatalyst CaFe 2 O 4

  • Published: 22 December 2022
  • Volume 26 , pages 3097–3126, ( 2024 )

Cite this article

experimental study of biodiesel synthesis

  • A. Saravanan 1 ,
  • Ajith J. Kings 2 ,
  • L. R. Monisha Miriam 3 &
  • R. S. Rimal Isaac 4  

258 Accesses

5 Citations

9 Altmetric

Explore all metrics

Today’s demand of energy in the world of automobile provokes the researchers to strive for the easily available and cheapest renewable source of energy. Biodiesel has become one of the booming renewable sources in the world to mitigate the atmospheric pollution and the demand of fossil fuels. Oils are chosen based on their fatty acid content, availability and sustainability. A magnetic nanocatalyst CaFe 2 O 4 has been employed in the transesterification process and is characterized by various progressive techniques to confirm its compatibility. The locally available, nonedible oils such as cotton seed oil, rubber seed oil and pungai seed oil have been taken for this experimental work for efficient and sustainable biodiesel production. Multi-variant central composite design has been employed to enhance the influencing process parameters in biodiesel conversion. Each feedstock produced more than 95% of the yield which consumed very little amount of methanol and catalyst in a short period of time. In order to ensure a quick reaction and smooth stirring, the temperature is kept at 70 °C (beyond the boiling point of the solvent). The chromatography analysis was used to describe the end product samples which revealed the right proportion of saturated and unsaturated fatty acids at the proper level, resulting in better oxidation stability and combustion properties. Moreover, density, viscosity, cetene number, iodine value and other essential properties were analysed and found to be within the standards specified by EN and ASTM for use in automotive applications without modifying the engine.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Subscribe and save.

  • Get 10 units per month
  • Download Article/Chapter or eBook
  • 1 Unit = 1 Article or 1 Chapter
  • Cancel anytime

Price includes VAT (Russian Federation)

Instant access to the full article PDF.

Rent this article via DeepDyve

Institutional subscriptions

experimental study of biodiesel synthesis

Similar content being viewed by others

experimental study of biodiesel synthesis

Optimization of biodiesel production from Moringa oleifera seeds oil in the presence of nano-MgO using Taguchi method

experimental study of biodiesel synthesis

Modeling and optimization of novel biodiesel production from non-edible oil with musa balbisiana root using hybrid response surface methodology along with african buffalo optimization

experimental study of biodiesel synthesis

Central Composite Design Optimization of the Extraction and Transesterification of Tiger Nut Seed Oil to Biodiesel

Explore related subjects.

  • Environmental Chemistry

Data availability

All the figures and tables contain original experimental data which are obtained from St. Xavier’s Catholic College of Engineering, Nagercoil, and Government Polytechnic College, Vanavasi, Noorul Islam centre for Higher Education, Kumaracoil, and Manonmaniam Sundaranar University, Tirunelveli, Tamil Nadu, India.

Abbreviations

Calcium ferrite

Fourier transform infrared spectroscopy

X-ray powder diffraction

Thermogravimetric/differential thermal analyser

Brunauer–Emmett–Teller analysis

Scanning electron microscope

Atomic force microscopy

Central composite design

Fatty acid methyl ester (%)

Potassium hydroxide

Sodium hydroxide

Calcium oxide

Magnesium oxide

Strontiun oxide

Manganese (II) oxide

Molybdenum oxide

Zirconium dioxide

Iron(II,III) oxide

Response surface methodology

Analysis of variance

Gas chromatography mass spectrometry

Rubber seed oil

Cotton seed oil

Pungai seed oil

Free fatty acid (%)

Unsaponified matter (%)

Joint committee on powder diffraction standards

Density functional theory

Rubber seed oil biodiesel

Cotton seed oil biodiesel

Pungai seed oil biodiesel

Aminopropyl triethoxysilane magnetite nanoparticles

Calcium oxide/gold

Magnesium ferrite

Magnesium aluminium oxide

Titanium oxide

Nanoparticles

American Society for Testing and Materials

European Standards

Aghilinategh, M., Barati, M., & Hamadanian, M. (2019). Supercritical methanol for one put biodiesel production from chlorella vulgaris microalgae in the presence of CaO/TiO 2 nano-photocatalyst and subcritical water. Biomass and Bioenergy, 123 , 34–40. https://doi.org/10.1016/j.biombioe.2019.02.011

Article   Google Scholar  

Alhassan, F. H., Rashid, U., & Taufiq-Yap, Y. H. (2015). Synthesis of waste cooking oil-based biodiesel via effectual recyclable bi-functional Fe 2 O 3 MnOSO 4 2- /ZrO 2 nanoparticle solid catalyst. Fuel, 142 , 38–45. https://doi.org/10.1016/j.fuel.2014.10.038

Article   CAS   Google Scholar  

Ali, R. M., Elkatory, M. R., & Hamad, H. A. (2020). Highly active and stable magnetically recyclable CuFe 2 O 4 as a heterogenous catalyst for efficient conversion of waste frying oil to biodiesel. Fuel, 268 , 117297. https://doi.org/10.1016/j.fuel.2020.117297

Alia, R. M., Elkatory, M. R., Hesham, A., et al. (2020). Highly active and stable magnetically recyclable CuFe 2 O 4 as a heterogenous catalyst for efficient conversion of waste frying oil to biodiesel. Fuel, 268 , 117297. https://doi.org/10.1016/j.fuel.2020.117297

Anwar, M., Rasul, M. G., & Ashwath, N. (2018). Production optimization and quality assessment of papaya (Carica papaya) biodiesel with response surface methodology. Energy Conversion and Management, 156 , 103–112. https://doi.org/10.1016/j.enconman.2017.11.004

Ban, S., Shrestha, R., Chaudhary, Y., et al. (2022). Process simulation and economic analysis of dolomite catalyst based biodiesel production from Nepalese Jatropha Curcas. Cleaner Chemical Engineering, 2 , 100029. https://doi.org/10.1016/j.clce.2022.100029

Berchmans, L. J., Karthikeyan, R., Helan, M., et al. (2011). Mechanochemical synthesis and electrochemical characterization of nano crystalline calcium ferrite. Catalysis Letters, 141 (10), 1451–1457. https://doi.org/10.1007/s10562-011-0636-9

Berger, H. (1986). Study of the Kα emission of copper. X-Ray Spectrometry, 15 (4), 241–243. https://doi.org/10.1002/xrs.1300150405

Bet-Moushoul, E., Farhadi, K., Mansourpanah, Y., et al. (2016). Application of CaO-based/Au nanoparticles as heterogeneous nanocatalysts in biodiesel production. Fuel, 164 , 119–127. https://doi.org/10.1016/j.fuel.2015.09.067

Bhuiya, M. M. K., Rasul, M. G., Khan, M. M. K., et al. (2020). Biodiesel production and characterisation of poppy (Papaver somniferum L.) seed oil methyl ester as a source of 2nd generation biodiesel feedstock. Industrial Crops and Products, 152 , 112493. https://doi.org/10.1016/j.indcrop.2020.112493

Borah, M. J., Das, A., Das, V., et al. (2019). Transesterification of waste cooking oil for biodiesel production catalysed by Zn substituted waste egg shell derived CaO nanocatalyst. Fuel, 242 , 345–354. https://doi.org/10.1016/j.fuel.2019.01.060

Candeia, R. M., Longo, B. E., Santos, I., et al. (2004). Synthesis and characterization of spinel pigment CaFe 2 O 4 obtained by the polymeric precursor method. Materials Letters, 58 , 569–572. https://doi.org/10.1016/S0167-577X(03)00563-9

Cheary, R. W., & Coelho, A. A. (1992). fundamental parameters approach to X-ray line-profile fitting. Journal of Applied Crystallography, 25 (2), 109–121. https://doi.org/10.1107/S0021889891010804

Deeba, F., Kumar, B., Arora, N., et al. (2020). Novel biobased solid acid catalyst derived from waste yeast residue for biodiesel production. Renewable Energy, 159 , 127–139. https://doi.org/10.1016/j.renene.2020.05.029

Du, L., Li, Z., Ding, S., et al. (2019). Synthesis and characterization of carbon-based MgO catalysts for biodiesel production from castor oil. Fuel, 258 , 116122. https://doi.org/10.1016/j.fuel.2019.116122

Fangfang, F., Alagumalai, A., & Mahian, O. (2021). Sustainable biodiesel production from waste cooking oil: ANN modeling and environmental factor assessment. Sustain Energy Technol Assess, 46 (2021), 102165. https://doi.org/10.1016/j.seta.2021.101265

Farooq, M., Ramli, A., & Subbarao, D. (2013). Biodiesel production from waste cooking oil using bifunctional heterogeneous solid catalysts. Journal of Cleaner Production, 59 , 131–140. https://doi.org/10.1016/j.jclepro.2013.06.015

Garcia-Martin, J. F., Ales-Alvarez, F. J., & Lopez-Barrera, M. (2019). Cetane number prediction of waste cooking oil-derived biodiesel prior to transesterification reaction using near infrared spectroscopy. Fuel, 240 , 10–15. https://doi.org/10.1016/j.fuel.2018.11.142

Gayathri, M., Shunmugam, S., Mugasundari, A. V., et al. (2018). Growth kinetic and fuel quality parameters as selective criterion for screening biodiesel producing cyanobacterial strains. Bioresource Technology, 247 , 453–462. https://doi.org/10.1016/j.biortech.2017.09.064

Article   CAS   PubMed   Google Scholar  

Ghasemzadeh, B., Matin, A. A., Biuck, H., et al. (2022). Cotton/Fe 3 O 4 @SiO 2 @H 3 PW 12 O 40 a magnetic heterogeneous catalyst for biodiesel production: Process optimization through response surface methodology. Industrial Crops and Products, 181 , 114806. https://doi.org/10.1016/j.indcrop.2022.114806

Gómez-Trejo-López, E., Gónzalez-Díaz, M., & Aguilar-Vega, M. (2021). Waste cooking oil transesterification by sulfonated polyphenylsulfone catalytic membrane: Characterization and biodiesel production yield. Renewable Energy, 182 , 1219–1227. https://doi.org/10.1016/j.renene.2021.11.003

Hara, M. (2010). Biodiesel production by amorphous carbon bearing SO 3 H, COOH and phenolic OH groups, a solid brønsted acid catalyst. Topics in Catalysis, 53 , 805–810. https://doi.org/10.1007/s11244-010-9458-z

Harsha, H. H. R., Math, M. C., & Yatish, K. V. (2017). Optimization and kinetic study of CaO nano-particles catalyzed biodiesel production from Bombax ceiba oil. Energy, 143 (15), 25–34. https://doi.org/10.1016/j.energy.2017.10.118

Haseeb, A. S. M. A., Fazal, M. A., Jahirul, M. I., et al. (2011). Compatibility of automotive materials in biodiesel: A review. Fuel, 90 (3), 922–931. https://doi.org/10.1016/j.fuel.2010.10.042

Hazmi, B., Rashid, U., & Ibrahim, M. L. (2021). Synthesis and characterization of bifunctional magnetic nano-catalyst from rice husk for production of biodiesel. Environmental Technology & Innovation, 21 , 101296. https://doi.org/10.1016/j.eti.2020.101296

Helmi, F., Helmi, M., & Hemmati, A. (2022). Phosphomolybdic acid/chitosan as acid solid catalyst using for biodiesel production from pomegranate seed oil via microwave heating system: RSM optimization and kinetic study. Renewable Energy, 189 , 881–898. https://doi.org/10.1016/j.renene.2022.02.123

Hoang, A. T. (2021a). Prediction of the density and viscosity of biodiesel and the influence of biodiesel properties on a diesel engine fuel supply system. Journal of Marine Engineering & Technology, 20 (5), 299–311. https://doi.org/10.1080/20464177.2018.1532734

Hoang, A. T. (2021b). Combustion behavior, performance and emission characteristics of diesel engine fuelled with biodiesel containing cerium oxide nanoparticles: A review. Fuel Processing Technology, 218 , 106840. https://doi.org/10.1016/j.fuproc.2021.106840

Hoang, A. T., & Le, A. T. (2018). A review on deposit formation in the injector of diesel engines running on biodiesel. Energy Sources, Part a: Recovery, Utilization, and Environmental Effects, 41 (5), 584–599. https://doi.org/10.1080/15567036.2018.1520342

Hoang, A. T., Nizeti, S., Ong, H. C., et al. (2021a). A review on application of artificial neural network (ANN) for performance and emission characteristics of diesel engine fueled with biodiesel-based fuels. Sustainable Energy of Technology Assessment, 47 , 101416. https://doi.org/10.1016/j.seta.2021.101416

Hoang, A. T., Tabatabaei, M., & Aghbashlo, M. (2019). A review of the effect of biodiesel on the corrosion behavior of metals/alloys in diesel engines. Energy Sources, Part a: Recovery, Utilization, and Environmental Effects, 42 (23), 2923–2943. https://doi.org/10.1080/15567036.2019.1623346

Hoang, A. T., Tabatabaei, M., Aghbashlo, M., et al. (2021b). Rice bran oil-based biodiesel as a promising renewable fuel alternative to petrodiesel: A review. Renewable and Sustainable Energy Reviews, 135 , 110204. https://doi.org/10.1016/j.rser.2020.110204

Hoang, A. T., Tabatabaei, M., Aghbashlo, M., et al. (2021c). Rice bran oil-based biodiesel as a promising renewable fuel alternative to petrodiesel: A review. Renewable and Sustainable Energy Reviews, 135 , 110204. https://doi.org/10.1016/j.rser.2020.110204

Hoseini, S. S., Najafi, G., Ghobadian, B., et al. (2018). Ailanthus altissima (tree of heaven) seed oil: Characterisation and optimisation of ultrasonication-assisted biodiesel production. Fuel, 220 , 621–630. https://doi.org/10.1016/j.fuel.2018.01.094

Hoseini, S. S., Najafi, G., Ghobadian, B., et al. (2019). Characterization of biodiesel production (ultrasonic-assisted) from evening-primroses (Oenothera lamarckiana) as novel feedstock and its effect on CI engine parameters. Renewable Energy, 130 , 50–60. https://doi.org/10.1016/j.renene.2018.06.042

Hosseini, S., Moradi, G. R., & Bahrami, K. (2019). Synthesis of a novel stabilized basic ionic liquid through immobilization on boehmite nanoparticles: A robust nanocatalyst for biodiesel production from soybean oil. Renewable Energy, 138 , 70–78. https://doi.org/10.1016/j.renene.2019.01.037

Ibrahim, M. M., Mahmoud, H. R., & El-Molla, S. A. (2019). Influence of support on physicochemical properties of ZrO 2 based solid acid heterogeneous catalysts for biodiesel production. Catalysis Communications, 122 , 10–15. https://doi.org/10.1016/j.catcom.2019.01.008

Jamil, F., Al-Muhatseb, A. A. H., Myint, M. T. Z., et al. (2018). Biodiesel production by valorizing waste Phoenix dactyliferaL. Kernel oil in the presence of synthesized heterogeneous metallic oxide catalyst (Mn@MgO-ZrO2). Energy Conversion and Management, 155 , 128–137. https://doi.org/10.1016/j.enconman.2017.10.064

Karimi, M. (2016). Immobilization of lipase onto mesoporous magnetic nanoparticles for enzymatic synthesis of biodiesel. Biocatalysis and Agricultural Biotechnology, 8 , 182–188. https://doi.org/10.1016/j.bcab.2016.09.009

Kesserwan, F., Ahmad, M. N., Khalil, M., et al. (2020). Hybrid CaO/Al 2 O 3 aerogel as heterogeneous catalyst for biodiesel production. Chemical Engineering Journal, 385 , 123834. https://doi.org/10.1016/j.cej.2019.123834

Kings, A. J., Miriam, L. R. M., Raj, R. E., et al. (2017). Cultivation, extraction and optimization of biodiesel production from potential microalgae Euglena sanguinea using eco-friendly natural catalyst. Energy Conversion and Management, 141 , 224–235. https://doi.org/10.1016/j.enconman.2016.08.018

Kittur, M. I., Andriyana, A., Ang, B. C., et al. (2021). Swelling of rubber in blends of diesel and cottonseed oil biodiesel. Polymer Testing, 96 , 107116. https://doi.org/10.1016/j.polymertesting.2021.107116

Knothe, G. (2009). Improving biodiesel fuel properties by modifying fatty ester composition. Energy Environ Science, 2 (7), 759–766. https://doi.org/10.1039/B903941D

Kordouli, E., Sygellou, L., Kordulis, C., et al. (2017). Probing the synergistic ratio of the NiMo/gamma-Al 2 O 3 reduced catalysts for the transformationof natural triglycerides into green diesel. Appl Catal B-Environ, 209 , 12–22. https://doi.org/10.1016/j.apcatb.2017.02.045

Krishnamurthy, K. N., Sridhara, S. N., & Kumar, C. S. A. (2020). Optimization and kinetic study of biodiesel production from Hydnocarpus wightiana oil and dairy waste scum using snail shell CaO nano catalyst. Renewable Energy, 146 , 280–296. https://doi.org/10.1016/j.renene.2019.06.161

Lam, M. K., & Lee, K. T. (2011). Mixed methanol-ethanol technology to produce greener biodiesel from waste cooking oil: A breakthrough for SO 4 2- /SnO 2 -SiO 2 catalyst. Fuel Process Technology, 92 , 1639–1645. https://doi.org/10.1016/j.fuproc.2011.04.012

Li, H., Liu, F., Helian, Y., Yang, G., et al. (2021). Inspection of various precipitant on SrO–based catalyst for transesterification: Catalytic performance, reusability and characterizations. Catalysis Today, 376 , 197–204. https://doi.org/10.1016/j.cattod.2020.06.038

Lima, A. L., Ronconi, C. M., & Mota, C. J. A. (2016). Heterogeneous basic catalysts for biodiesel production. Catalysis Science & Technology, 6 , 2877–2891. https://doi.org/10.1039/C5CY01989C

Liu, Y., Yang, X., Adamu, A., et al. (2021). Economic evaluation and production process simulation of biodiesel production from waste cooking oil. Current Research in Green and Sustainable Chemistry, 4 , 100091. https://doi.org/10.1016/j.crgsc.2021.100091

Liu, Y., Zhang, P., Fan, M., et al. (2016). Biodiesel production from soybean oil catalyzed by magnetic nanoparticle MgFe 2 O 4 @ CaO. Fuel, 164 , 314–321. https://doi.org/10.1016/j.fuel.2015.10.008

Maheshwari, P., Haider, M. B., Yusuf, M., et al. (2022). A review on latest trends in cleaner biodiesel production: Role of feedstock, production methods, and catalysts. Journal of Cleaner Production, 355 , 131588. https://doi.org/10.1016/j.jclepro.2022.131588

Mahian, O., Bellos, E., Markides, C. N., et al. (2021). Recent advances in using nanofluids in renewable energy systems and the environmental implications of their uptake. Nano Energy, 86 , 106069. https://doi.org/10.1016/j.nanoen.2021.106069

Maleki, B., Talesh, S. S. A., & Mansouri, M. (2022). Comparison of catalysts types performance in the generation of sustainable biodiesel via transesterification of various oil sources: A review study. Materials Today Sustainability, 18 , 100157. https://doi.org/10.1016/j.mtsust.2022.100157

Marwaha, A., Rosha, P., Mohapatra, S. K., et al. (2019). Biodiesel production from Terminalia bellerica using eggshell-based green catalyst: An optimization study with response surface methodology. Energy Reports, 5 , 1580–1588. https://doi.org/10.1016/j.egyr.2019.10.022

Mendonça, I. M., Paes, O. A. R. L., Maia, P. J. S., et al. (2019). New heterogeneous catalyst for biodiesel production from waste tucum~a peels (Astrocaryum aculeatum Meyer): Parameters optimization study. Renewable Energy, 130 , 103–110. https://doi.org/10.1016/j.renene.2018.06.059

Miao, C., Yang, L., Wang, Z., et al. (2018). Lipase immobilization on amino-silane modified superparamagnetic Fe 3 O 4 nanoparticles as biocatalyst for biodiesel production. Fuel, 224 , 774–782. https://doi.org/10.1016/j.fuel.2018.02.149

Mihankhah, T., Delnavaz, M., & Khaligh, N. G. (2018). Application of TiO 2 nanoparticles for ecofriendly biodiesel production from waste olive oil. International Journal of Green Energy, 15 , 69–75. https://doi.org/10.1080/15435075.2018.1423975

Miriam, L. R. M., Raj, R. E., Kings, A. J., & Visvanathan, M. A. (2017). Enhanced FAME production using green catalyst with superior profile from the isolated halophilic Aphanothece halophytica grown in raceway ponds. Energy Conversion and Management, 151 , 63–72. https://doi.org/10.1016/j.enconman.2017.08.071

Nanaki, E. A., & Koroneos, C. J. (2012). Comparative LCA of the use of biodiesel, diesel and gasoline for transportation. Journal of Cleaner Production, 20 (1), 14–19. https://doi.org/10.1016/j.jclepro.2011.07.026

Navajas, A., Campo, I., Moral, A., et al. (2018). Outstanding performance of rehydrated Mg-Al hydrotalcites as heterogeneous methanolysis catalysts for the synthesis of biodiesel. Fuel, 211 , 173–181. https://doi.org/10.1016/j.fuel.2017.09.061

Nursal, R. S., Khalid, A., Abdullah, I. S., et al. (2021). Autoignition behavior and emission of biodiesel from palm oil, waste cooking oil, tyre pyrolysis oil, algae and jatropha. Fuel, 306 , 121695. https://doi.org/10.1016/j.fuel.2021.121695

Ong, H. C., Masjuki, H. H., Mahlia, T. M., et al. (2014). Optimization of biodiesel production and engine performance from high free fatty acid Calophyllum inophyllum oil in CI diesel engine. Energy Conversion and Management, 81 , 30–40. https://doi.org/10.1016/j.enconman.2014.01.065

Palitsakun, S., Koonkuer, K., Topool, B., et al. (2021). Transesterification of Jatropha oil to biodiesel using SrO catalysts modified with CaO from waste eggshell. Catalysis Communications, 149 , 106233. https://doi.org/10.1016/j.catcom.2020.106233

Pandit, P. R., & Fulekar, M. H. (2019). Biodiesel production from microalgal biomass using CaO catalyst synthesized from natural waste material. Renewable Energy, 136 , 837–845. https://doi.org/10.1016/j.renene.2019.01.047

Roy, T., Sahani, S., & Sharma, Y. C. (2020). Study on kinetics-thermodynamics and environmental parameter of biodiesel production from waste cooking oil and castor oil using potassium modified ceria oxide catalyst. Journal of Cleaner Production, 247 , 119166. https://doi.org/10.1016/j.jclepro.2019.119166

Salim, S. M., Izriq, R., Almaky, M. M., et al. (2022). Synthesis and characterization of ZnO nanoparticles for the production of biodiesel by transesterification: Kinetic and thermodynamic studies. Fuel, 321 , 124135. https://doi.org/10.1016/j.fuel.2022.124135

Sandouqa, A., & Al-Hamamre, Z. (2019). Energy analysis of biodiesel production from jojoba seed oil. Renewable Energy, 130 , 831–842. https://doi.org/10.1016/j.renene.2018.07.015

Saxena, V., Sharma, S., & Pandey, L. M. (2018). Fe(III) doped ZnO nano-assembly as a potential heterogeneous nano-catalyst for the production of biodiesel. Materials Letters., 237 , 232–235. https://doi.org/10.1016/j.matlet.2018.11.101

Shahraki, H., Entezaria, M. H., & Goharshadi, E. K. (2015). Sono-synthesis of biodiesel from soybean oil by KF/g-Al 2 O 3 as a nano-solid-base catalyst. Ultrasonics Sonochemistry, 23 , 266–274. https://doi.org/10.1016/j.ultsonch.2014.09.010

Shankar, V., & Jambulingam, R. (2017). Waste crab shell derived CaO impregnated Na-ZSM-5 as a solid base catalyst for the transesterification of neem oil into biodiesel. Sustainable Environment Research, 27 , 273–278.

Silveira, E. G., Perez, V. H., Reyero, I., et al. (2019). Biodiesel production from heterogeneous catalysts based K 2 CO 3 supported on extruded gamma-Al 2 O 3 . Fuel, 241 , 311–318. https://doi.org/10.1016/j.fuel.2018.12.074

Sivakumar, P., Sindhanaiselvan, S., Gandhi, N. N., et al. (2013). Optimization and kinetic studies on biodiesel production from underutilized Ceiba Pentandra oil. Fuel, 103 , 693–698. https://doi.org/10.1016/j.fuel.2012.06.029

Soria-Figueroaa, E., Mena-Cervantesa, V. Y., García-Solaresa, M., et al. (2020). Statistical optimization of biodiesel production from waste cooking oil using CaO as catalyst in a Robinson-Mahoney type reactor. Fuel, 282 , 118853. https://doi.org/10.1016/j.fuel.2020.118853

Sravan, K. Y., Klai, N., Kaur, R., et al. (2018). Oleaginous yeast biomass flocculation using Bioflocculant produced in wastewater sludge and transesterification using petroleum diesel as a co-solvent. Renewable Energy, 131 , 217–228. https://doi.org/10.1016/j.renene.2018.06.066

Vahid, B. R., & Haghighi, M. (2017). Biodiesel production from sunflower oil over MgO/MgAl 2 O 4 nanocatalyst: Effect of fuel type on catalyst nanostructure and performance. Energy Convers Manag, 134 , 290–300. https://doi.org/10.1016/j.enconman.2016.12.048

Venkanna, K. B., & Chitimi, V. R. (2013). Few physical, chemical and fuel related properties of calophyllum inophyllum linn (honne) oil and its blends with diesel fuel for their use in diesel engine. Fuel, 109 , 356–361. https://doi.org/10.1016/j.fuel.2013.02.015

Woranuch, W., Ngaosuwan, K., Kiatkittipong, W., et al. (2022). Fine-tuned fabrication parameters of CaO catalyst pellets for transesterification of palm oil to biodiesel. Fuel, 323 , 124356. https://doi.org/10.1016/j.fuel.2022.124356

Yousefi, S., & Ghasemi, B. (2019). Ultrasound-assisted synthesis of porous Mg(OH) 2 nanostructures using hypersaline brine. Micro & Nano Letters, 14 (9), 1019–1023. https://doi.org/10.1049/mnl.2019.0047

Zhang, Y., Niu, S., Lu, C., et al. (2020). Catalytic performance of NaAlO 2 /γ-Al2O 3 as heterogeneous nanocatalyst for biodiesel production: Optimization using response surface methodology. Energy Conversion and Management, 203 , 112263. https://doi.org/10.1016/j.enconman.2019.112263

Zul, N. A., Ganesan, S., Hamidon, T. S., et al. (2021). A review on the utilization of calcium oxide as a base catalyst in biodiesel production. Journal of Environmental Chemical Engineering, 9 , 105741. https://doi.org/10.1016/j.jece.2021.105741

Download references

Author information

Authors and affiliations.

Department of Mechanical Engineering, Ponjesly College of Engineering, Nagercoil, 629003, India

A. Saravanan

Department of Mechanical Engineering, St. Xavier’s Catholic College of Engineering, Nagercoil, 629003, India

Ajith J. Kings

Department of Biotechnology, Udaya School of Engineering, Vellamodi, 629204, India

L. R. Monisha Miriam

Department of Nanotechnology, Noorul Islam Centre for Higher Education, Kumaracoil, 629180, India

R. S. Rimal Isaac

You can also search for this author in PubMed   Google Scholar

Contributions

AS was involved in investigation, methodology, formal analysis and writing the original draft. AJK was responsible for conceptualization, investigation, methodology, formal analysis, supervision, validation, writing and editing. LRMM took part in investigation, methodology, formal analysis, validation and writing. RSRI participated in catalyst preparation, characterization and writing.

Corresponding author

Correspondence to Ajith J. Kings .

Ethics declarations

Conflict of interest.

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Additional information

Publisher's note.

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Saravanan, A., Kings, A.J., Miriam, L.R.M. et al. RSM-based comparative experimental study of sustainable biodiesel synthesis from different 2G feedstocks using magnetic nanocatalyst CaFe 2 O 4 . Environ Dev Sustain 26 , 3097–3126 (2024). https://doi.org/10.1007/s10668-022-02761-1

Download citation

Received : 09 April 2022

Accepted : 03 November 2022

Published : 22 December 2022

Issue Date : February 2024

DOI : https://doi.org/10.1007/s10668-022-02761-1

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

  • Nonedible oils
  • Magnetic nanocatalyst
  • Optimization
  • Transesterification
  • Find a journal
  • Publish with us
  • Track your research

Information

  • Author Services

Initiatives

You are accessing a machine-readable page. In order to be human-readable, please install an RSS reader.

All articles published by MDPI are made immediately available worldwide under an open access license. No special permission is required to reuse all or part of the article published by MDPI, including figures and tables. For articles published under an open access Creative Common CC BY license, any part of the article may be reused without permission provided that the original article is clearly cited. For more information, please refer to https://www.mdpi.com/openaccess .

Feature papers represent the most advanced research with significant potential for high impact in the field. A Feature Paper should be a substantial original Article that involves several techniques or approaches, provides an outlook for future research directions and describes possible research applications.

Feature papers are submitted upon individual invitation or recommendation by the scientific editors and must receive positive feedback from the reviewers.

Editor’s Choice articles are based on recommendations by the scientific editors of MDPI journals from around the world. Editors select a small number of articles recently published in the journal that they believe will be particularly interesting to readers, or important in the respective research area. The aim is to provide a snapshot of some of the most exciting work published in the various research areas of the journal.

Original Submission Date Received: .

  • Active Journals
  • Find a Journal
  • Proceedings Series
  • For Authors
  • For Reviewers
  • For Editors
  • For Librarians
  • For Publishers
  • For Societies
  • For Conference Organizers
  • Open Access Policy
  • Institutional Open Access Program
  • Special Issues Guidelines
  • Editorial Process
  • Research and Publication Ethics
  • Article Processing Charges
  • Testimonials
  • Preprints.org
  • SciProfiles
  • Encyclopedia

energies-logo

Article Menu

experimental study of biodiesel synthesis

  • Subscribe SciFeed
  • Recommended Articles
  • Google Scholar
  • on Google Scholar
  • Table of Contents

Find support for a specific problem in the support section of our website.

Please let us know what you think of our products and services.

Visit our dedicated information section to learn more about MDPI.

JSmol Viewer

Retrofit of a marine engine to dual-fuel methane–diesel: experimental analysis of performance and exhaust emission with continuous and phased methane injection systems.

experimental study of biodiesel synthesis

1. Introduction

2. materials and methods, 3.1. comparison between continuous and phased ng injection systems, 3.2. effect of diesel injection strategy on combustion and emissions, 3.2.1. effects of diesel injection strategy on df combustion process and fuel consumption, 3.2.2. effect of diesel injection strategy on df exhaust emissions, 3.3. effect of diesel injection pressure on combustion and emissions, 3.3.1. effects of diesel injection pressure on df combustion process and fuel consumption, 3.3.2. effect of diesel injection pressure on df exhaust emissions, 3.4. evaluation of optimum point compared to diesel references, 4. conclusions.

  • Phased injection resulted in a reduction in fuel consumption, compared to continuous mode, of between 8% and 11% for the single-injection strategy; similarly, using SOGAV results in a fuel consumption reductions from 7% to 9.7% and 5.4% to 7.2% in the cases of the double and triple-injection strategies, respectively.
  • Even advancing the diesel SOI main, the DF combustion centre (MFB50) is about 8 CAD later than in the FD reference case, due to the difference between diffusive FD and premixed DF combustion mechanisms. A slight improvement is obtained by adopting multi-injection strategies that reduce the induction time.
  • Despite the difference in MFB10 and MFB50, their combustion duration (2xMFB 10–50 ) curves are almost overlapped, indicating a faster NG combustion in the late stage in the case of single injection.
  • At 100 MPa diesel injection pressure, independent of injection strategies, minimum fuel consumption was achieved between −35 and −30 CAD ATDC, with improvements of between 14.0 and 14.5% compared to the FD reference.
  • At 100 MPa diesel injection pressure, DF NO x emissions were consistently higher than the FD reference case, ranging between 12.5 and 17.2 g/kWh, with increases of 20.3% and 65% compared to the FD reference case (10.4 g/kWh), respectively.
  • DF HC emissions were an order of magnitude higher than in the diesel case due to the different combustion mechanism. A slight reduction in HC emissions was found at the most advanced SOI main , independent of the injection strategy.
  • DF engines can significantly reduce smoke emissions. Compared to the FD reference (1.43%), at the most advanced diesel SOI main , DF test points exhibited minimal levels close to zero, independent of the injection strategy.
  • Increasing diesel injection pressure reduces the DF combustion duration: at 80 MPa injection pressure, DF combustion duration was higher than in the FD reference case; at 120 MPa, this behaviour was inverted.
  • Despite the improvements in the early stage of combustion, which were linked to the better atomization of diesel spray, and consequently triggering the combustion of NG, increasing injection pressure did not result in a corresponding reduction in fuel consumption, with the exception of the single-injection strategy. This behaviour can be ascribed to a too-short duration of the main injection, when pressure is increased, which deteriorated the NG combustion stability during the early stage of combustion. The minimum BSFCeq was 219.1 g Diesel_eq /kWh, achieved with triple injection at 80 MPa injection pressure, with a reduction of about 20.7% compared to the FD reference (276.3 g Diesel_eq /kWh).
  • Higher injection pressures corresponded to an increase in NO x , while HC and CO were reduced.
  • Even though DF combustion provided a reduction in BSFC, with a consistent reduction in CO 2 emissions of up to 28.5%, the reduction in GWP was only 2.9%, due to the increase in HC emissions. This is still a promising result, considering that the optimization of DF combustion systems can significantly reduce the emission of methane hydrocarbons in the exhaust, thus making more effective the reduction in CO 2 in terms of GHG emissions.

Author Contributions

Data availability statement, acknowledgments, conflicts of interest, abbreviations.

ATDCAfter Top Dead Centre
BMEPBreak Mean Effective Pressure
CADCrank Angle Degree
CCSCarbon Capture and Storage
CNGCompressed Natural Gas
COCarbon Monoxide
CO Carbon Dioxide
DFDual Fuel
ECUEngine Control Unit
EGRExhaust Gas Recirculation
FIDFlame Ionization Detector
FDFull Diesel
GHGGreenhouse Gas
GWPGlobal Warming Potential
HCUnburned Hydrocarbons
HPDFHigh-Pressure Dual Fuel
HVOHydrotreated Vegetable Oil
IHRIntegrated Heat Release
IRInfrared
IRENAInternational Renewable Energy Agency
IMOInternational Maritime Organization
LBSILean Burn Spark Ignition
LHVLower Heating Value
LNGLiquefied Natural Gas
LPDFLow-Pressure Dual-Fuel
MGOMarine Gas Oil
NDUVNon-Dispersive Ultraviolet
NGNatural Gas
NO Nitrogen Oxides
PMParticulate Matter
ROHRRate of Heat Release
SCRSelective Catalytic Reduction
SOGAVSolenoid-Operated Gas Admission Valve
SOIStart of Injection
SO Sulphur Oxides
  • MEPC. 2023 IMO Strategy on Reduction of GHG Emissions from Ships. MEPC 80/WP.12 Annex 1. 2023. Available online: https://wwwcdn.imo.org/localresources/en/MediaCentre/PressBriefings/Documents/Resolution%20MEPC.377(80).pdf (accessed on 1 July 2024).
  • DNV. Available online: https://www.dnv.com/expert-story/maritime-impact/exploring-all-options-to-keep-decarbonization-on-course.html (accessed on 1 July 2024).
  • Inal, O.; Charpentier, J.F.; Cengiz, D. Hybrid power and propulsion systems for ships: Current status and future challenges. Renew. Sustain. Energy Rev. 2022 , 156 , 111965. [ Google Scholar ] [ CrossRef ]
  • Balcombe, P.; Brierley, J.; Lewis, C.; Skatvedt, L.; Speirs, J.; Hawkes, A.; Staffell, I. How to decarbonise international shipping: Options for fuels, technologies and policies. Energy Convers. Manag. 2019 , 182 , 72–88. [ Google Scholar ] [ CrossRef ]
  • Weisan, H.; Yishun, S.; Xuelai, Z.; Hongfen, C. Research progress of carbon capture and storage (CCS) technology based on the shipping industry. Ocean Eng. 2023 , 281 , 114929. [ Google Scholar ]
  • Xing, H.; Stuart, C.; Spence, S.; Chen, H. Alternative fuel options for low carbon maritime transportation: Pathways to 2050. J. Clean. Prod. 2021 , 297 , 126651. [ Google Scholar ] [ CrossRef ]
  • Atilhan, S.; Park, S.; El-Halwagi, M.M.; Atilhan, M.; Moore, M.; Nielsen, R.B. Green hydrogen as an alternative fuel for the shipping industry. Curr. Opin. Chem. Eng. 2021 , 31 , 100668. [ Google Scholar ] [ CrossRef ]
  • Tarhan, C.; Çil, M.A. A study on hydrogen, the clean energy of the future: Hydrogen storage methods. J. Energy Storage 2021 , 40 , 102676. [ Google Scholar ] [ CrossRef ]
  • Tornatore, C.; Marchitto, L.; Sabia, P.; De Joannon, M. Ammonia as green fuel in internal combustion engines: State-of-the-art and future perspectives. Front. Mech. Eng. 2022 , 8 , 944201. [ Google Scholar ] [ CrossRef ]
  • Shi, J.; Zhu, Y.; Feng, Y.; Yang, J.; Xia, C. A Prompt Decarbonization Pathway for Shipping: Green Hydrogen, Ammonia, and Methanol Production and Utilization in Marine Engines. Atmosphere 2023 , 14 , 584. [ Google Scholar ] [ CrossRef ]
  • Inal, O.B.; Zincir, B.; Deniz, C. Investigation on the decarbonization of shipping: An approach to hydrogen and ammonia. Int. J. Hydrogen Energy 2022 , 47 , 19888–19900. [ Google Scholar ] [ CrossRef ]
  • Rouwenhorst, K.H.; Castellanos, G. Innovation Outlook: Renewable Ammonia. Irena. 2022. Available online: https://www.irena.org/-/media/Files/IRENA/Agency/Publication/2022/May/IRENA_Innovation_Outlook_Ammonia_2022.pdf (accessed on 1 July 2024).
  • Livaniou, S.; Papadopoulos, G.A. Liquefied Natural Gas (LNG) as a Transitional Choice Replacing Marine Conventional Fuels (Heavy Fuel Oil/Marine Diesel Oil), towards the Era of Decarbonisation. Sustainability 2022 , 14 , 16364. [ Google Scholar ] [ CrossRef ]
  • Milojević, S.T. Sustainable application of natural gas as engine fuel in city buses–Benefit and restrictions. J. Appl. Eng. Sci. 2017 , 15 , 81–88. [ Google Scholar ] [ CrossRef ]
  • Arefin, M.A.; Nabi, M.N.; Akram, M.W.; Islam, M.T.; Chowdhury, M.W. A Review on Liquefied Natural Gas as Fuels for Dual Fuel Engines: Opportunities, Challenges and Responses. Energies 2020 , 13 , 6127. [ Google Scholar ] [ CrossRef ]
  • Lindstad, E.; Eskeland, G.S.; Rialland, A.; Valland, A. Decarbonizing Maritime Transport: The Importance of Engine Technology and Regulations for LNG to Serve as a Transition Fuel. Sustainability 2020 , 12 , 8793. [ Google Scholar ] [ CrossRef ]
  • IRENA. A Pathway to Decarbonise the Shipping Sector by 2050 ; International Renewable Energy Agency: Masdar City, United Arab Emirates, 2021; ISBN 978-92-9260-330-4. [ Google Scholar ]
  • MAN Energy Solution, Survey Report, Future Fuels For Power Generation. 2022. Available online: https://www.man-es.com/energy-storage/services/tools-downloads-e-s/white-papers (accessed on 1 July 2024).
  • Julien, A.; Cédric, P. Flexible production of green hydrogen and ammonia from variable solar and wind energy: Case study of Chile and Argentina. Int. J. Hydrogen Energy 2020 , 45 , 1541–1558. [ Google Scholar ] [ CrossRef ]
  • Balcombe, P.; Staffell, I.; Kerdan, I.G.; Speirs, J.F.; Brandon, N.P.; Hawkes, A.D. How can LNG-fuelled ships meet decarbonisation targets? An environmental and economic analysis. Energy 2021 , 227 , 120462. [ Google Scholar ] [ CrossRef ]
  • Ampah, J.D.; Yusuf, A.A.; Afrane, S.; Jin, C.; Liu, H. Reviewing two decades of cleaner alternative marine fuels: Towards IMO’s decarbonization of the maritime transport sector. J. Clean. Prod. 2021 , 320 , 128871. [ Google Scholar ] [ CrossRef ]
  • Transport & Environment. Modelling the Impact of Fuel EU Maritime on EU Shipping. 2023. Available online: https://te-cdn.ams3.digitaloceanspaces.com/files/FuelEU-Maritime-Impact-Assessment.pdf (accessed on 1 July 2024).
  • International Energy Agency (IEA)/Trasport/International/Shipping. Available online: https://www.iea.org/energy-system/transport/international-shipping (accessed on 1 July 2024).
  • Xing, H.; Spence, S.; Chen, H. A comprehensive review on countermeasures for CO 2 emissions from ships. Renew. Sustain. Energy Rev. 2020 , 134 , 110222. [ Google Scholar ] [ CrossRef ]
  • Nitin, A. Is LNG the solution for decarbonised shipping? J. Int. Marit. Saf. Environ. Aff. Shipp. 2022 , 6 , 158–166. [ Google Scholar ] [ CrossRef ]
  • Intergovernmental Panel on Climate Change (IPCC). AR6 Synthesis Report Climate Change 2023. Available online: https://www.ipcc.ch/report/ar6/syr/ (accessed on 1 July 2024).
  • Ushakov, S.; Stenersen, D.; Einang, P.M. Methane slip from gas fuelled ships: A comprehensive summary based on measurement data. J. Mar. Sci. Technol. 2019 , 24 , 1308–1325. [ Google Scholar ] [ CrossRef ]
  • Kuittinen, N.; Heikkilä, M.; Lehtoranta, K. Review of Methane Slip from LNG Engines. 2023. Available online: https://greenray-project.eu/wp-content/uploads/2023/04/D1.1_Review_of_methane_slip_from_LNG_engines.pdf (accessed on 1 July 2024).
  • Houghton, J.T.; Jenkins, G.J.; Ephraums, J.J. Climate Change, the IPCC scientific assessment. Contribution of Working Group I to the First Assessment Report of the Intergovernmental Panel on Climate Change. 1990. Available online: https://www.ipcc.ch/site/assets/uploads/2018/03/ipcc_far_wg_I_full_report.pdf (accessed on 1 July 2024).
  • International Energy Agency (IEA). Available online: https://www.iea.org/reports/methane-tracker-2021/methane-and-climate-change (accessed on 1 July 2024).
  • Zardoya, A.; Oregui Bengoetxea, I.; Martinez, A.L.; Lucena, I.; Orosa, J.A. Methodological Design Optimization of a Marine LNG Internal Combustion Gas Engine to Burn Alternative Fuels. J. Mar. Sci. Eng. 2023 , 11 , 1194. [ Google Scholar ] [ CrossRef ]
  • Lu, C.; Song, E.; Xu, C.; Ni, Z.; Yang, X.; Dong, Q. Analysis of Performance of Passive Pre-Chamber on a Lean-Burn Natural Gas Engine under Low Load. J. Mar. Sci. Eng. 2023 , 11 , 596. [ Google Scholar ] [ CrossRef ]
  • Desantes, J.; Novella, R.; De La Morena, J.; Pagano Lng, V. Achieving Ultra-Lean Combustion Using a Pre-Chamber Spark Ignition System in a Rapid Compression-Expansion Machine ; SAE Technical Paper 2019-01-0236; SAE International: Warrendale, PA, USA, 2019. [ Google Scholar ]
  • Castilla, A.; Carlos, E.; Couto, G.; Roso, V.; Thiriet, A.; Valle, R. A review of Prechamber Ignition Systems as Lean Combustion Technology for SI Engines. Appl. Therm. Eng. 2017 , 128 , 10.1016. [ Google Scholar ] [ CrossRef ]
  • Bui, K.Q.; Perera, L.P.; Emblemsvåg, J. Life-cycle cost analysis of an innovative marine dual-fuel engine under uncertainties. J. Clean. Prod. 2022 , 380 , 134847. [ Google Scholar ] [ CrossRef ]
  • Curran, S.; Onorati, A.; Payri, R.; Agarwal, A.K.; Arcoumanis, C.; Bae, C.; Boulouchos, K.; Chuahy, F.D.F.; Gavaises, M.; Hampson, G.J.; et al. The future of ship engines: Renewable fuels and enabling technologies for decarbonization. 2023. Int. J. Engine Res. 2023 , 25 , 85–110. [ Google Scholar ] [ CrossRef ]
  • Stenersen, D.; Ole, T. GHG and NOx Emissions from Gas Fuelled Engines. In Mapping, Verification, Reduction Technologies ; SINTEF: Trondheim, Norway, 2017; pp. 1–52. Available online: https://www.nho.no/siteassets/nox-fondet/rapporter/2018/methane-slip-from-gas-engines-mainreport-1492296.pdf (accessed on 1 July 2024).
  • Speirs, J.; Balcombe, P.; Blomerus, P.; Stettler, M.; Achurra-Gonzalez, P.; Woo, M.; Ainalis, D.; Cooper, J.; Sharafian, A.; Merida, W.; et al. Natural gas fuel and greenhouse gas emissions in trucks and ships. Prog. Energy 2020 , 2 , 012002. [ Google Scholar ] [ CrossRef ]
  • Chintala, V.; Subramanian, K.A. A comprehensive review on utilization of hydrogen in a compression ignition engine under dual fuel mode. Renew. Sustain. Energy Rev. 2017 , 70 , 472–491. [ Google Scholar ] [ CrossRef ]
  • Krivopolianskii, V.; Valberg, I.; Stenersen, D.; Ushakov, S.; Æsøy, V. Control of the combustion process and emission formation in marine gas engines. J. Mar. Sci. Technol. 2019 , 24 , 593–611. [ Google Scholar ] [ CrossRef ]
  • De Simio, L.; Iannaccone, S.; Pennino, V.; Marchitto, L. Experimental Analysis of a Single-Cylinder Large Bore Engine with External Supercharging in Diesel/CNG Dual-Fuel Mode ; SAE Technical Paper, No. 2023-24-0058; SAE International: Warrendale, PA, USA, 2023. [ Google Scholar ] [ CrossRef ]
  • Ulishney, C.J.; Dumitrescu, C.E. Effect of gas composition on the performance and emissions of a dual-fuel diesel-natural gas engine at low load conditions. Fuel 2022 , 324 , 124531. [ Google Scholar ] [ CrossRef ]
  • Commission Staff Working Document Impact Assessment Report Accompanying the Proposal for a Regulation of the European Parliament and of the Council on Methane Emissions Reduction in the Energy Sector and Amending Regulation (EU) 2019/942. Available online: https://eur-lex.europa.eu/legal-content/EN/TXT/?uri=CELEX%3A52021SC0459&qid=1639998727689 (accessed on 1 July 2024).

Click here to enlarge figure

Fuel TypeLHV
(MJ/kg)
Volumetric
Energy Density (GJ/m )
Storage
Pressure
(MPa)
Storage
Temperature (°C)
MGO42.736.60.1120
LNG5023.40.1−162
Methanol13.315.80.120
Liquid
ammonia
18.612.70.1−34
0.8620
Liquid H 1208.50.1−253
Compressed H 1207.57020
Single-Cylinder Engine Specification
Bore [mm]170
Stroke [mm]185
Single Cylinder Displacement [l]4.2
BMEP (Max) [MPa]2.52
Diesel Injection Pressure (Max) [MPa] 160
NG Injection Pressure (Max) [MPa]1.2
CR (:1)13.2
Rated Power [kW]132.5/145@1500/1800 rpm
Max. Boost (abs.) [MPa]0.48
Head LayoutCentral Injector/4 valve
Exhaust Valve Opening100 CAD ATDC
Exhaust Valve Closure−317 CAD ATDC
Intake Valve Opening313 CAD ATDC
Intake Valve Closure−127 CAD ATDC
Test Conditions
Engine speed [rpm]1500 ± 5
Engine load (BMEP) [MPa]084 ± 0.005
Boost pressure [MPa]0.15 ± 0.002
Natural gas/diesel ratio~4
Injection pressure [MPa]80/100/120
Injection strategyS/D/T
Main injection timing [CAD ATDC]−40/−20 step 5
Pre/pilot injected mass [mg/stroke] 10
Total injected mass [mg/stroke]44.5
Dwell [CAD]5
TESTDF OptimumFD Reference
BSFC [g /kWh]219276
HC [g/kWh]90.2
CO [g/kWh]680954
GWP [g /kWh] *932960
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

Marchitto, L.; De Simio, L.; Iannaccone, S.; Pennino, V.; Altieri, N. Retrofit of a Marine Engine to Dual-Fuel Methane–Diesel: Experimental Analysis of Performance and Exhaust Emission with Continuous and Phased Methane Injection Systems. Energies 2024 , 17 , 4304. https://doi.org/10.3390/en17174304

Marchitto L, De Simio L, Iannaccone S, Pennino V, Altieri N. Retrofit of a Marine Engine to Dual-Fuel Methane–Diesel: Experimental Analysis of Performance and Exhaust Emission with Continuous and Phased Methane Injection Systems. Energies . 2024; 17(17):4304. https://doi.org/10.3390/en17174304

Marchitto, Luca, Luigi De Simio, Sabato Iannaccone, Vincenzo Pennino, and Nunzio Altieri. 2024. "Retrofit of a Marine Engine to Dual-Fuel Methane–Diesel: Experimental Analysis of Performance and Exhaust Emission with Continuous and Phased Methane Injection Systems" Energies 17, no. 17: 4304. https://doi.org/10.3390/en17174304

Article Metrics

Article access statistics, further information, mdpi initiatives, follow mdpi.

MDPI

Subscribe to receive issue release notifications and newsletters from MDPI journals

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • View all journals
  • Explore content
  • About the journal
  • Publish with us
  • Sign up for alerts
  • Open access
  • Published: 24 August 2024

Green synthesis of trimetallic CuO/Ag/ZnO nanocomposite using Ziziphus spina-christi plant extract: characterization, statistically experimental designs, and antimicrobial assessment

  • Ayman K. El-Sawaf 1 , 2 ,
  • Shahira H. El-Moslamy 3 ,
  • Elbadawy A. Kamoun 4 &
  • Kaizar Hossain 5  

Scientific Reports volume  14 , Article number:  19718 ( 2024 ) Cite this article

142 Accesses

Metrics details

  • Environmental sciences
  • Materials science
  • Microbiology
  • Nanoscience and technology

In this study, Ziziphus spina christi leaves was used to synthesize a trimetallic CuO/Ag/ZnO nanocomposite by a simple and green method. Many characterizations e.g. FTIR, UV–vis DRS, SEM–EDX, TEM, XRD, zeta-size analysis, and DLS, were used to confirm green-synthesized trimetallic CuO/Ag/ZnO nanocomposite. The green, synthesized trimetallic CuO/Ag/ZnO nanocomposite exhibited a spherical dot-like structure, with an average particle size of around 7.11 ± 0.67 nm and a zeta potential of 21.5 mV. An extremely homogeneous distribution of signals, including O (79.25%), Cu (13.78%), Zn (4.42%), and Ag (2.55%), is evident on the surface of green-synthetic nanocomposite, according to EDX data. To the best of our knowledge, this is the first study to effectively use an industrially produced green trimetallic CuO/Ag/ZnO nanocomposite as a potent antimicrobial agent by employing different statistically experimental designs. The highest yield of green synthetic trimetallic CuO/Ag/ZnO nanocomposite was (1.65 mg/mL), which was enhanced by 1.85 and 5.7 times; respectively, by using the Taguchi approach in comparison to the Plackett–Burman strategy and basal condition. A variety of assays techniques were utilized to evaluate the antimicrobial capabilities of the green-synthesized trimetallic CuO/Ag/ZnO nanocomposite at a 200 µg/mL concentration against multidrug-resistant human pathogens. After a 36-h period, the tested 200 µg/mL of the green-synthetic trimetallic CuO/Ag/ZnO nanocomposite effectively reduced the planktonic viable counts of the studied bacteria, Escherichia coli and Staphylococcus aureus , which showed the highest percentage of biofilm reduction (98.06 ± 0.93 and 97.47 ± 0.65%; respectively).

Similar content being viewed by others

experimental study of biodiesel synthesis

Loading of green-synthesized cu nanoparticles on Ag complex containing 1,3,5-triazine Schiff base with enhanced antimicrobial activities

experimental study of biodiesel synthesis

Green synthesis, characterization, antibacterial, and antifungal activity of copper oxide nanoparticles derived from Morinda citrifolia leaf extract

experimental study of biodiesel synthesis

In situ facile green synthesis of Ag–ZnO nanocomposites using Tetradenia riperia leaf extract and its antimicrobial efficacy on water disinfection

Introduction.

Nanotechnology is an area of technology that studies, applies, and develops materials at the nanoscopic scale, which typically ranges from 1 to 100 nm 1 . To manufacture nanoparticles, chemical or physical processes are commonly employed. Nevertheless, both processes are challenging to scale up and require large amounts of energy, in addition to potentially dangerous substances 2 . These hazardous substances persist at the interface between nanomaterials, affecting the biocompatibility of nanomaterials. Biological biosynthesis is therefore a well-known solution for these widely used approaches 3 . The green synthesis technique has effectively utilized a wide range of candidates, including bacteria, fungi, algae, and plants. Metallic nanomaterials can be produced by reducing metal ions with the help of the bioactive compounds found in the extracts of these candidates. Because of their large surface-to-volume ratio, these nanomaterials are widely used in many different engineering and materials science domains, including medical, optical, biotechnological, microbiological, electronics, and environmental 4 .

Several areas in Egypt are home to the common medicinal plants Mentha spp., Ziziphus spina-christi, and Ocimum basilicum, which are highly valued for their anti-inflammatory, antioxidant, antibacterial, and anticancer properties 1 , 2 , 3 . Ziziphus spina-christi , a member of the Rhamnaceae family, is commonly referred to as Sidr. Together with a number of recently introduced exotic plants, it is an important cultivated tree and one of the few surviving natural tree species in Arabia. The genus Ziziphus is well-known for its therapeutic uses as an immune system booster, hypotensive, anti-inflammatory, antimicrobial, antioxidant, and liver-protective agent. Furthermore, there have been reports that the Z. spina-christi extract protects against aflatoxicosis. Mentha spp., also known as mint (genus Mentha , family Lamiaceae ), is widely used as a spice component for many different types of food worldwide and is also regularly used to make herbal tea. Because the essential oils of mint include antibacterial and antioxidant qualities, it is widely known that the leaves of the plant are still employed in traditional medicine to treat digestive problems. According to phytochemical analyses of mentha plants, the main components of the leaf extracts that reduced and stabilized nanoparticles were phenolic, flavonoid, steroid, and terpenoid 8 , 9 , 10 . The herb sweet basil ( Ocimum basilicum ) has small, pink-tinged, or white blooms and elliptic, bright green leaves, it also has a strong scent.

Many bioactive molecules, including flavonoids, alkaloids, phenolic compounds, sterols, saponins, tannins, and fatty acids, are rich in different plant-based extracts 11 . In the green synthesis of nanoparticles, these molecules act as reductants, capping agents, and stabilizing agents, keeping the resultant nanoparticles stable and preventing them from aggregating completely 12 . According to a previously published study, the crude extract of Ziziphus-Spina Christi leaves (Sider) was utilized in the green production of zinc oxide nanoparticles (38.177 nm) at the hexagonal wurtzite phase 6 . Graphene oxide was phyto-reduced in the other study employing various doses of Ziziphus spina-christi aqueous extract 13 . Furthermore, other prior investigations reported a fast and safe synthesis of selenium-doped zinc oxide nanoparticles (50 nm) in spherical shape utilizing aqueous leaf extract ( Mangifera indica ), which showed strong antimicrobial properties 14 . The green synthesis of a safe, stable, and trimetallic nanocomposite containing Cu, Ag, and Zn was achieved previously using an aqueous leaf extract of Catharanthus roseus 15 . Previously, Ocimum basilicum L. seed extract was used to generate an Ag-doped ZnO-MgO-CaO nanocomposite in a manner that is environmentally friendly 16 . Furthermore, MgO and CuO/MgO nanoparticles were produced via a green approach using an extract from the Opuntia monacantha plant 17 . Moreover, a content-based extract from the plant powder ( Ocimum basilicum ) was used in a green method for the biological synthesis of MnO 2 nanoparticles and MnO 2 @eggshell nanocomposite 18 . Also, Calotropis gigantea leaf extract was used to generate the green-manufactured binary ZnO-CuO nanocomposites, which show promising antimicrobial properties against skin-related infections 11 . Meanwhile, silver nanoparticles (Ag NPs) that are cost-effective and environmentally friendly were produced using the aqueous extract of Ziziphus spina-christi leaves for treating Fusarium wilt disease 19 . Besides, Poly(HEMA-co-FAOEME)/ZnO nanocomposites were generated as an antimicrobial agent by biosynthesizing ZnO nanoparticles using Mentha plegium L. extract 20 . ZnO, MgO, CuO, and their composite mixed oxide nanoparticles were previously manufactured utilizing a green approach and leaf extracts of medicinal plants e.g. Pisonia grandis R.Br. 21 . The antimicrobial and pro-healing abilities of silver, copper, and zinc oxide nanoparticles are widely recognized. Because of their photo-oxidizing and photo-catalytic effects on biological species, these nanoparticles are safe and biocompatible nanomaterials 22 .

To prove their antimicrobial properties, these nanoparticles may interact chemically as well as physically. As a result of these nanomaterials' interactions with microbial cells, reactive oxygen species (ROS), H 2 O 2 , and ions are released under photoinduced conditions 23 . Conversely, depending on the examined nanomaterials, physical interaction may exhibit biocidal impacts through cellular internalization, breakdown of the cell membrane, or forceful damage 24 . Furthermore, it has been suggested that ZnO-Ag NCs have the strongest capacity to break down microbial cell membranes and interact with vital DNA elements such as phosphorus and sulfur, inhibiting DNA replication 23 , 25 , 26 . The higher specific surface area to volume ratio of the nanoparticles led to the formation of more ROS, which is dependent on binding and interacting with the cell membrane and aggregating in the lipid layer 23 . The negative charge of super oxides and hydroxide ions allows them to enter microbial cells. Eventually, this may result in the cell wall breaking down, releasing its contents, and finally causing cell death 27 . Damaged electrostatic interactions lead to the mortality of pathogens when the negative charge on the surface of the cell membrane catches the positive charge on emitted ions from nanomaterials 26 , 28 . Our results can be explained by the green, synthesized trimetallic CuO/Ag/ZnO nanocomposite's high infusibility and ability to generate more released ions e.g. (Ag +1 , Zn +2 , and Cu +2 ions) 23 . Additionally, the released ions penetrated the host cell by binding to surface proteins on the cell wall. Afterward, the microbe's cells died as a result of the altered metabolism 29 .

As environmental problems throughout the world become more urgent, scientists are looking into the possibility of using nanomaterials to address these problems. Scientists have recently focused their attention on nanomaterials and nanocomposites developed from plant extracts. Our group was drawn to the trend of using plant extracts to produce nanocomposites on a large scale. When compared to alternative fabrication methods, biosynthesized nanoparticles are less costly, non-toxic, and very stable. Biomolecules produced by plants are extensive and can be used to generate nanomaterials for a variety of biological applications.

To the best of our knowledge, no previous reports of a trimetallic CuO/Ag/ZnO nanocomposite that was produced environmentally employing a Ziziphus spina christi leaf extract, have been published to date. Thus, different experimental designs, such as the Plackett–Burman and Genichi Taguchi procedures, were also targeted in this work to commercially optimize green-synthesized trimetallic CuO/Ag/ZnO nanocomposite as a strong antimicrobial ingredient.

Materials and methods

A variety of human pathogens, such as Escherichia coli (ATCC 10536), Klebsiella pneumoniae (ATCC 10031), Staphylococcus aureus (ATCC 25923), Bacillus subtilis (ATCC 11774), Candida albicans (ATCC 10231), and Candida krusei (ATCC 6258), were used to assess the antimicrobial efficacy of the green synesthetic nanocomposites. All human pathogens were received from GEBRI, SRTA-City, Alexandria, Egypt. Fresh leaves of Mentha, Ocimum basilicum, and Ziziphus spina hristi were collected from the New Borg Al-Arab City farms in Alexandria, Egypt.

Preparation of Ziziphus spina-christi plant extract

Three well-known herbal plants i.e. ( Mentha, Ziziphus spina-christi), and (Ocimum basilicum) were gathered locally for this study. Greenish-yellow leaves were collected from these plants and washed twice under running water before being thoroughly cleansed with distilled water to remove any extra remaining dirt. After carefully cleaning every leaf with a white cloth, the leaves were allowed to air dry for three hours. It was ground into a fine powder after dried for 72 h at 60 °C. Finally, 10 g of dry powder and 100 mL of double-distilled water were added to a 250-mL Erlenmeyer flask and shaken at 200 rpm at 70 °C for 30 min to extract the components. To eliminate any last bits of tiny plant debris from the recovered material, centrifugation at 6000 rpm revolutions per minute was performed using Whatman No. 1 filter paper. The filtered extract was stored at 4 °C for future experimental use 16 , 30 , 31 .

Quantification of main leaf extract constituents

With a minor modification, the Folin-Ciocalteu method was utilized to determine the total phenolic content 32 . In brief, 40 μL of plant extract was mixed with 1.8 mL of 2N Folin-Ciocalteu for 5 min at room temperature (25 °C). The resulting mixture was subsequently mixed with 1.2 mL of a 7.5% sodium carbonate solution and allowed to react for one hour in the dark at ambient temperature, where the absorbance at 765 nm was finally measured. The standard component was gallic acid (y = 0.6812x + 0.0314, R 2  = 0.9975), and the sample’s total phenolic content was expressed in milligrams of gallic acid equivalents (mg GAE/g). Rutin, a slightly modified standard substance, was utilized with the aluminum chloride method to evaluate the total flavonoid concentration 33 . Overall, 1 mL of extracts and blank (H 2 O) were mixed with 3 mL of potassium acetate (0.1 mol/L) and 2 mL of aluminum chloride solution (0.1 mol/L). After allowing the mixture to react for 20 min, 70% aqueous ethanol (v/v) was added to dilute it to a final volume of 10 mL. The standard curve was (y = 1.4715x + 0.0364) (R 2  = 0.9998). The absorbance was finally measured at 510 nm, and the results were reported as rutin equivalents (mg RE/g). The total protein content was calculated using Bradford’s (1976) method and expressed as mg/g fresh weight (FW). To prepare one gram of fresh plant tissues for protein and enzyme extractions, three milliliters of 25 mM Tris–HCl buffer (pH 6.8) and 3% polyvinylpolypyrrolidone were homogenized at 4 °C. Protein analysis was completed using the supernatant after the resulting mixture was centrifuged for an hour at 13,000 rpm at 4 °C. The concentrations in mg/g FW were determined using standard curves for each reducing sugar 34 . The extract was centrifuged at 12,000 rpm, and the supernatant was kept in the dark for a whole day to determine anthocyanins 35 . After that, an absorbance measurement at 550 nm was taken for all the samples. After computing the total anthocyanin content using a coefficient of attrition of 33,000 mM/cm, the result was expressed as µg/g final weight.

Green synthesis of trimetallic CuO/Ag/ZnO nanocomposite

A 250-mL Erlenmeyer flask was filled with 50 mL of each diluted plant extract (50%) and was agitated for 30 min, while 0.1M AgNO 3 , 0.1M Cu (NO 3 ) 2 .3H 2 O, and 0.1M Zn (CH 3 COO) 2 .2H 2 O were titrated gradually at a time. The solution was then continually stirred at 80 °C, while the pH was adjusted to (5.5, 7.0, 14) using 2M NaOH solution. A precipitate that was dark brown developed after this reaction, was stirred for two hours. The precipitate that was produced was centrifuged for 15 min at 12,000 rpm, and after being repeatedly cleaned to remove contaminants with distilled water and ethanol, the pelt was dried for two hours at 80°C. A mortar and pestle were used to grind the dried pelt into a powder, and dry weights were estimated for each plant extract. UV–visible spectroscopy (Shimadzu, Japan) was also used to identify the absorbance bands and band gaps to confirm that the nanocomposite was synthesized utilizing each of the extracted plants.

Bioassay survey

An agar-well diffusion method was used to determine the green synthesized nanocomposite’s antimicrobial sensitivity in vitro, against a range of multi-drug-resistant human pathogens. The investigated pathogens were cultured in individual culture inoculated in sterile nutrient broth containing (0.5% peptone, 0.5% NaCl, and 0.3% yeast extract). The inocula was then incubated for 24 h at 37 °C. After incubation, the individual culture suspension was utilized for the bioassay survey. A sterile well cutter was used to cut 5-mm-diameter wells on Muller Hinton agar medium (0.2% beef extract, 0.15% starch, 1.75% casein, and 1.7% agar). Subsequently, 0.1 mL of every pathogen was spread out on the agar plates, and 50 μL of the nanocomposites that synthesized at distinct pH levels (pH 5, 7.0, and 14) were added to the hollows. 20 μL of Ziziphus spina-christi extract was used as a control. The inoculation culture plates were held at 4 °C for 5 h before incubation for 48 h at 37 °C. After incubation, the inhibitory zones that formed were measured in millimeters.

Characterization of green synthesized nanocomposite

FTIR spectra were investigated using a JASCO-410 spectrometer (JASCO, Easton, MD). To further explore the surface morphology of the green synthetic nanocomposite, a SEM (Qattro, Thermo-Scientific, USA) JSM-6510LV, USA was utilized. The transmission electron microscope was additionally utilized for analyzing the nano structural features (TEM, JEM-2100F, JEOL: Japan). Thermal stability of synthesized nanocomposites was assessed using (TGA, DTA, and DSC) was estimated at 29–1000 °C utilizing a DSC-TGA device model (SDTQ 600, USA) under a N2 atmosphere (flow rate of 100 mL/min and a heating rate of 10°C/min). Furthermore, (Horiba, SZ-100, Kyoto, Japan) specimen was used to investigate the green synthetic nanocomposite’s zeta potential using dynamic light scattering (DLS). A temperature of 25 °C was maintained during a 20-min dilution and dispersion process in an ultrasonic bath for examining the green synthetic nanocomposite in a DLS machine.

Statistical optimization to maximize the yield of green-synthesized nanocomposite

Two successive experimental designs were used in this work to maximize the yield of nanocomposite’s green-synthetic reaction. The factors affecting the green synesthetic reaction were assessed using the Plackett–Burman and Taguchi designs, such as concentrations of plant extract (F1), concentrations of precursors (F2), ratio of precursors (F3), reaction agitation (F4), reaction temperature (F5), reaction pH (F6), and incubation period (F7).

Plackett–Burman design (PBD)

In several investigations, this design was utilized to evaluate the rate of green synesthetic reaction and the overall yield of dry-weight nanocomposite 36 , 37 , 38 . Green-synthetic reaction variables are used in these qualitative and quantitative screening procedures to identify the ideal parameters for maximizing the dry weight of nanocomposite products. The yield weight of green synthetic trimetallic CuO/Ag/ZnO nanocomposite was found to be affected by seven factors, which included concentrations of plant extract, concentrations of precursors, ratios of precursors, reaction agitation, temperature, reaction pH, and incubation time. These factors were selected based on previous experiments (data not shown). As indicated in Table 1 , these factors were examined at two different levels: the highest (1) and the lowest (− 1).

The response was calculated using the average green synthesized nanocomposite dry weight, and each experiment was conducted twice. A first-order polynomial model serves as the foundation for mathematical modeling of PBD, as shown in Eq. ( 1 ). In this case, Y is the dry weight of nanocomposite that were biosynthesized (response), β 0 denotes the model intercept, β i is the linear coefficient, and X i is the number of independent variables.

Furthermore, Eq. ( 2 ) was used to calculate the efficiency of each variable. In this equation, M v represents the variable main effect, Mv + and Mv− are the cell dry weights in trials where the independent variable was present at high and low levels; respectively, and N is the number of trials divided by two.

Minitab® 18.1 software was utilized to generate a set of 12 trails for statistical analysis and graph charting. All independent variables were evaluated for their impact on the response using analysis of variance (ANOVA), with a significance level of P < 0.05. The fitness of the equation was assessed using the multiple correlation coefficient (R 2 ) and adjusted R 2 .

Taguchi technique: To generate a valid result, the Taguchi technique was built up in several steps: choosing important components, creating an accurate matrix, analyzing statistical data, and finally validating using the best values. The goal of this work is to use several criteria to determine the maximum cell dry weight of the overall dry-weight nanocomposite yield (Table 2 ). The L27(3^7) Taguchi orthogonal array design was used for this optimization technique (7 factors, 3 levels, and 27 runs). An orthogonal array (signal-to-noise ratio, or "S/N") is created by first identifying the factor levels (inner array) using numbers like 1, 2, 3 etc. These levels are then compared to different combinations of noise factors in the outer array. The S/N ratio is expressed in decibels (dB).

Once the average of produced cell dry weights and the signal-to-noise (S/N) ratio (the larger the better group) are determined for each process condition as designed, the F test and ANOVA are used to examine the significance of all factors and their relationships at levels using the MINITAB 18 software. At the end of the process, a confirmation test was conducted to compare the experimental value with the results that were achieved using Taguchi's method. The S/N ratio is expressed in decibels [dB] and calculated using Eq. ( 3 ), where n is the number of observations and Y is the observed data (dry-weight nanocomposite yield).

Moreover, the predicted S/N ratio was estimated, where n is the number of parameters, S/Nm is the total mean S/N ratio, and S/Ni is the mean S/N ratio at the optimum level using Eq. ( 4 ).

Antimicrobial efficiency bioassays

The CLSI standard 39 , was followed in the agar well-diffusion method, biofilm inhibition assay, and time-kill experiment employed to evaluate the antimicrobial efficiency of green synthesized CuO/Ag/ZnO nanocomposite.

Agar well-diffusion technique

The studied human pathogens were cultured in a nutrient broth medium (0.5% peptone, 0.3% yeast extract, 0.2% beef extract, and 0.5 NaCl) to achieve the 0.5 McFarland turbidity standards. On nutrient agar plates, 100 µL of microbial cultures were spread out using sterile cotton swabs. For this experiment, three different dosages of the selected green synesthetic nanocomposite designated as (A): 50 µg/mL, (B): 100 µg/mL, and (C): 150 µg/mL were prepared. Immediately after using a sterile 6-mm cork-borer to drill each well, 50 µL of the evaluated green synesthetic nanocomposite dosage was added. After that, the agar plates were incubated for 24 h at 37°C. A ruler was used to measure the inhibitory zone, or the clean zone, in millimeters (mm) around each well 40 .

Biofilm inhibition assay

The broth microdilution method was used to estimate the minimum inhibitory concentration of the examined green synesthetic nanocomposite by calculating the lowest concentration at which no discernible growth appeared 41 . To measure MIC, multiple dosages of the selected green synesthetic nanocomposite, ranging from 50 to 250 µg/mL were generated. The tested pathogens were cultured individually in nutrient broth medium at 37 °C and 150 rpm to generate the pre-inoculums. To obtain (2 × 10 5 ) CFU/mL, each pathogen was separately inoculated into a fresh nutrient broth, which was then incubated at 37 °C and 150 rpm. The optical density (OD) at 600 nm was measured over a 6-h incubation period to determine a microbe's exponential phase. Aseptically, 100 µL of each dosage of the tested green nanocomposite was put into 900 µL of these planktonic cultures to generate treated cultures. Moreover, green synesthetic nanocomposite-free cultures were employed for developing untreated (control) cultures. The microbiological turbidity was measured spectrophotometrically to evaluate the inhibitory effects of the tested chemical compounds. The percentage of anti-biofilm in each sample was determined by using Eq. ( 5 ) to compare OD of the treated culture (T) to the corresponding untreated culture (U).

Time kill-kinetics assay

Macro-broth dilution method was employed, to reach the early logarithmic stage; every one of such pathogens was inoculated separately into nutrient broth medium and incubated for 6 h at 37 °C while being agitated at 150 rpm. The inoculum from each microbial culture (5 × 10 8 CFU/mL) was then transferred to 9 mL of freshly made nutrient broth medium. The green synesthetic nanocomposite (1 mL) was then added at a concentration of 200 µg/mL. A control growth system was prepared for each pathogen, omitting the tested formulation. After that, these tubes were shaken constantly at 150 rpm and 37 °C for the rest of the period of incubation. Under aseptic conditions, the samples were routinely taken at many time intervals (0, 6, 12, 18, 24, 30, 36, 42, and 48 h). After diluting the samples with sterile saline, 100 µL of the mixture was swabbed onto nutrient agar plates. During the incubation periods, the number of visible colonies was counted and reported as CFU/mL. The percentage of the pathogen's cells' biofilm reduction exposed to the tested green synesthetic nanocomposite with each control was determined using the logarithm of the counted colonies (Log 10 CFU/ml) for each time interval (Eq.  6 ). Furthermore, by determining the lowest dose that eliminated at least 99.9% of the initial microbial cells, the minimum bactericidal concentration (MBC) was determined.

Statistical analyses

The results of antimicrobial efficacy tests were performed in triplicate, and the mean ± standard deviation (M ± SD) was utilized to describe the results. Tukey’s multiple comparison post hoc test was utilized in the Minitab 19 program ( MINITAB version 19.1 ) to compute a one-way analysis of variance ( ANOVA ) and confirm statistical significance. A 95% confidence interval was used for statistical significance (p < 0.05).

Results and discussion

Biosynthetic of green-synthesized trimetallic nanocomposite.

Basically, nanostructures can be synthesized using sol–gel processing, hydrolysis/condensation, and wet chemical processing. These techniques are mostly costly, require exact experimental parameters (temperature, pressure, energy, and timeframe), and require toxic traditional chemicals. However, green synthesis of nanostructures is attracting a lot of interest currently, due to many significant advantages including simpler, cheaper, and more eco-friendly technique. The most promising method of synthesis is " green synthesis ," which is achieved by employing either plant extracts or specific microbes (bacteria, fungi, algae, etc.). Many studies on the synthesis of various metals nanoparticles including Zn, Mn, Cu, Au, and Ag, which have been carried out in recent years with a focus on different types of biological systems 42 . In pharmaceutical formulations, medicinal plant extracts are utilized for their bioactive ingredients, helping in the reduction and capping of metal ions through the synthesis of nanostructures 43 . For instance, zinc nitrate ionization in an aqueous solution produced Zn 2+ , which was subsequently reduced to Zn + by a phytochemical present in the extract (functional as reducing, capping, and stabilizing agents). Chemicals containing phenolic groups and hydroxyl groups may hydrolyze and generate nanostructures. Among the several kinds of metallic nanoparticles, Ag, CuO, and ZnO have attracted the attention of many scientists, due to their many applications in different scientific sectors 44 , 45 , 46 . These nanoparticles have been extensively employed in antimicrobial, antioxidant, and photocatalytic applications 44 . On the connections between these metals in plant extract-based nanocomposites, however, there is currently no information available. Due to the synergistic effect of their respective qualities, this combination usually improves the material's properties 47 .

Thus, our work contributes to the effort to find a novel material with remarkable physiological characteristics that has been produced using green techniques. Polyphenols (including flavonoids and saponins), alkaloids, proteins, phenolic acids, sugars, and terpenoids all of which are found in various plant parts—help reduce and stabilize metal ions to produce nanostructures 48 . Thus, the use of plant extracts not only saves energy, time, and steps while reducing the use of toxic chemicals, which protects the environment and human health, but also enhances the efficacy and properties of nanoparticles in the pharmaceutical and medical fields by retaining active chemical molecules on their surfaces 49 . The green synthesis of nanocomposites in aqueous plant extracts is suggested to be influenced by a variety of bioactive molecules, including proteins, polyphenols, and polysaccharides 50 . So, the examined plant extracts were evaluated by determining their constituents. A set of methods was used to test specific components of the various leaf materials of Mentha , Ocimum basilicum , and Ziziphus spina christi before the green synthesis of CuO/Ag/ZnO nanocomposites, were developed. Table 3 initially reports the results of the determinations for total protein, reducing sugar, anthocyanin, phenol, and flavonoids. According to phytochemical investigations, the main constituents of the tested leaf extracts that contributed to the stabilization and reduction of nanoparticles were protein content, reducing sugar, flavonoids, phenolics, and anthocyanin. The results showed that these constituents were richest in Ziziphus spina-christi , Ocimum basilicum , followed by Mentha spp . The results showed that the Ziziphus spina christi extract consisted of high levels of flavonoids (26.60 ± 2.25), total phenolic compounds (35.69 ± 5.38), reducing sugar (2.84 ± 0.22), anthocyanin (5.92 ± 0.05), and total protein (2.96 ± 0.27).

These aromatic plant extracts (reductants) were then titrated under shaking conditions with the precursors composed of (0.1M AgNO 3 , 0.1M Cu(NO 3 ) 2 .3H 2 O, and 0.1M Zn(CH 3 COO) 2 .2H 2 O) together to generate a green synesthetic trimetallic nanocomposite. The reaction color changed from reddish yellow (aromatic plant extracts) to dark turbid brown, indicating that the extracts of Mentha spp. (Fig.  1 IC), Ziziphus spina-christi (Fig.  1 IIC), and Ocimum basilicum (Fig.  1 IIIC) generated a green synthetic nanocomposite (Fig.  1 I, II , and IIIN). An essential technique for figuring out the electronic structure, optical activities, and physico-chemical characteristics of nanoparticles is UV–visible (UV–vis) spectroscopy 51 . The classification of nanoparticles in the size range of 2–100 nm was found to be adequate for absorption of wavelengths 200–800 52 . The absorption edge of our green synthetic nanocomposite was estimated using a spectrophotometer scanning a range of 200–500 nm, and the results were compared with the extracts employed in each case. The real absorbance was graphed from 0 to 4.0 au. using the Origin Pro software (v. 8.0, OriginLab Co., Northampton, MA, USA) to generate fitted curves.

figure 1

Findings of green synthesized nanocomposite generated from various aromatic plant extracts. UV–vis plots for the prepared nanocomposite, compared to the examined plant extracts: Mentha spp. ( I ), Ziziphus spina-christi ( II ), and Ocimum basilicum ( III ). Photos of the extract plants ( C ) and the yield nanocomposite ( N ). The chart depicts the dry weights of nanocomposites generated from various aromatic plant extracts at different pH levels ( IV ). The color of the resulting green nanocomposite synthesized with Ziziphus spina-christi at various pH levels ( V ).

The real green synthetic nanocomposite consistently displays distinct absorption peaks at 220 nm (Fig.  1 I), 240 nm (Fig.  1II ), and 260 nm (Fig.  1III ), in addition to 320 nm when its wavelength is compared to the extract's peaks. A green-generated trimetallic Cu, Zn, and Ag nanocomposite utilizing Catharanthus roseus leaf extract has shown similar results elsewhere 53 . According to reports, the absorption bands for Cu, Zn, and Ag nanocomposites have been identified at 220, 270, and 370 nm; respectively 15 . In addition, an experiment revealed the existence of zinc ions in the crystal lattices, which caused the lattices to shift, especially in consideration of their extension. The absorbance peak's strength increases as a result of this alteration after ZnO doping with Cu 54 . Furthermore, a rise in the absorbance band at 220 nm confirms the presence of copper oxide (CuO) 54 . This peak is found in a similar range by numerous other investigations that demonstrated the generation of CuO NPs. Zinc oxide (ZnO) has a further separate peak at 270 nm. Other investigations have shown that absorbance peaks at 230 and 270 nm in copper-doped ZnO nanoparticles suggest the presence of ZnO 40 , 48 , 49 , 50 . The production of Ag nanoparticles is shown by the absorbance band at 370 nm 15 . The aqueous extract of Berberis vulgaris leaf and root was used to generate nanoparticles of silver, which showed a broad peak in 380–400 nm area 53 . The results of the current study are fully consistent with all the outcomes.

Different parameters, including pH, temperature, reaction duration, and reactant concentration, can be used to optimize the green synthesis of nanoparticle morphological characterization 33 , 51 , 52 , 53 . Most of these environmental elements that influence nanoparticle synthesis should be identified. Consequently, these aspects can be efficiently addressed to maximize the yield of industrial fabrication of metallic nanoparticles 53 . The reaction's pH has significant effects on the nanoparticles' structure 61 . To be more precise, temperature and pH have an impact on how nucleation centers develop. In order to maximize the synthesis of metal nanoparticles, it is crucial to adjust the pH level since this results in the automatic growth of nucleation centers 62 . Moreover, the size and structural composition of the nanoparticles have been found to be significantly impacted by the pH of the solution 53 . Therefore, the green synthetic nanocomposite was generated at different pHs to determine which plant extract produced the heaviest dry weight of nanocomposite. To generate a green synesthetic nanocomposite of a trimetallic nanocomposite, these aromatic plant extracts (reductants) are separately adjusted at different pHs (5.5, 7, and 14). Then, the precursors that were used are added gradually and equally. For all studied aromatic plant extracts, the optimum response was seen at a pH between 5.5 and 7, as Fig.  1IV illustrates. Moreover, the heaviest dry weight of the generated nanocomposite was obtained using the Ziziphus spina-christi extract at all applicable pHs (Fig.  1V ). In brief, the largest dry weight of green synthetic nanocomposite was measured at pH-5.5 (0.29 mg/mL), followed by pH-7 (0.25 mg/mL), and pH-14 (0.05 mg/mL) was the lowest. With the exception of other extracts in all applicable screening analyses, the Ziziphus spina-christi extract produced the heaviest dry weight of the formed nanocomposite. Therefore, in all additional investigations, the Ziziphus spina-christi extract was selected for the green-generated nanocomposite.

An antimicrobial survey is carried out utilizing the green synthetic nanocomposite, which is prepared using Ziziphus spina-christi extract at all applicable pHs. When compared to the free extract (Co), the growth of the evaluated human pathogens was impacted by every nanocomposite created, as demonstrated by the plate photographs (Fig.  2 ). In brief, the widest inhibitory zone widths (Fig.  2 D) were detected at pH 7 against Bacillus subtilis (14.21 ± 1.56 mm) and Staphylococcus aureus (13.96 ± 2.33 mm). ANOVA and Tukey post-hoc tests were used to assess the mean values of the computed inhibitory zones to statistically identify the more effective versions. To find significant mean differences, Fig.  2 E then displays Tukey 's test means for each paired comparison. The adjusted confidence intervals are computed using the Tukey simultaneous tests on a 95% scale. At pH 7 intervals, the green synthesized nanocomposite is devoid of the zero line. This indicates that there are statistically significant differences between the green synthetic nanocomposite at pH 7 and the control group and other tested pHs. The results show statistically significant antimicrobial properties for the tested green synthetic nanocomposite at pH 7.

figure 2

Antimicrobial effects of green synthesized nanocomposite utilizing Ziziphus spina-christi extract at all applicable pHs ( A ): pH-5, ( B ) pH-7, and ( C ) pH-14, in comparison to ( Co ): control against ( i ) Escherichia coli , ( ii ) Klebsiella pneumoniae , ( iii ) Staphylococcus aureus , ( iv ) Bacillus subtilis , ( v ) Candida albicans , and ( vi ) Candida krusei using agar-well diffusion analysis. Photos of antimicrobial plates are shown, as well as a chart of the computed inhibition zones ( D ) and simultaneous Tukey tests for mean difference using Tukey–Kramer post-hoc analysis ( E ).

Characterization of green synthesized CuO/Ag/ZnO nanocomposite

TEM imaging of green synthesized nanocomposite is observed with an accelerating voltage of 100 kV. Figure  3 I shows the dense, spherical dot-like structure of the green, synthetic trimetallic CuO/Ag/ZnO nanocomposite. This proves the effective development of trimetallic CuO/Ag/ZnO nanocomposite, which is produced in an environmentally friendly manner. The particle size measured on TEM images is used to visualize the real size of nanoparticles, and the result is an average particle size of 7.11 ± 0.67 nm with a narrow particle size dispersion. SEM investigation shows the film surface morphology, which can be characterized as a porosity-free, soft, smooth planar structure (Fig.  3II ) . Previous studies also used the green chemistry method using extract of Ocimum basilicum L., to generate Ag/doped ZnO-MgO-CaO nanocomposite (59 nm) and spherical and triangular-shaped Ag/doped MgO-NiO-ZnO nanocomposite (30–44 nm); respectively 12 , 56 . Furthermore, the spherical-shaped of ZnO-Ag nanocomposites (26.02 ± 1 nm) were formed by utilizing a novel, simple, cost-effective, and safe method that involved the utilization of Stenotaphrum secundatum extract 13 . Likely, a green technique is employed in a prior study to prepare Ag-doped ZnO nanoparticles (60 nm) utilizing Tridax procumbens leaf extract. These nanoparticles show synergistic antimicrobial properties against a variety of human pathogens 8 . As seen in Fig.  3III , EDX mapping verification at multiple sites demonstrates the presence of signals with a highly homogenous distribution on the surface of green synthesized nanocomposite, including O (79.25%), Cu (13.78%), Zn (4.42%), and Ag (2.55%). The study's findings verify that CuO, Ag, and ZnO nanocomposite are effectively synthesized using green techniques. Prior to this, the normal stoichiometric ratio that was employed to generate the trimetallic nanoparticles was not followed, resulting in a compositional atomic ratio of (1:1.46:1.05) of (Cu:Ag:Zn). This could have been brought about by differences in the surface energy of the nanoparticles or by the specific crystallographic orientation of the metal atoms 15 . A further vital characteristic is the ability to measure charge on a surface. The molecular weight of large molecules dissolved in water can be determined using Zeta-potential analyzer. Zeta potential levels rely on a number of factors, including chemical composition and roughness 64 . Zeta potential is a measure of the strength of charge on the surface of particles 64 , 65 . The stability of an emulsion or nanosuspension can be predicted based on the absolute value of the zeta potential. In order to stabilize the nanocrystal formation (electrostatic repulsion), a high absolute value of zeta potential needs to be achieved. Higher zeta potentials of the nanomaterial suspension predicted the formation of a more stable, non-aggregating particle dispersion. Previous studies found that a suspended particle is deemed stable if its zeta potential is either higher than + 30 mV or lower than − 30 mV 65 , 66 . According to earlier studies, particles will agglomerate when zeta potential values get closer to 0 mV 67 ; nevertheless, for values larger than ± 20 mV, the particles will remain stable and suspended 65 . The green synthetic trimetallic CuO/Ag/ZnO nanocomposite has a zeta-potential of 21.5 ± 5.53 mV, as shown in Fig.  3 IV. The large absolute zeta potentials (> 20 mV) of our developed green synthetic nanocomposite suggested long-term stability by reducing vesicle aggregation, indicating that it was stable in a liquid state.

figure 3

TEM image (I) , SEM image (II) , TEM–EDX analysis (III) , and Zeta potential pattern (IV) of green synthesized trimetallic CuO/Ag/ZnO nanocomposite.

Furthermore, the thermal analysis of green synthesized trimetallic CuO/Ag/ZnO nanocomposite is characterized using TGA, DTA, DSC profiles (Fig.  4 ). DSC data provides a detailed description of the phase transition of tested nanocomposite. The Tg value is a crucial parameter to describe the stability of the lyophilized nanocomposite. The transition temperature and associated enthalpy drop have an impact on the stability of drug pharmacokinetics. A more tightly constructed nanocomposite is suggested by a greater transient enthalpy. DSC panel indicates that green synthetic trimetallic CuO/Ag/ZnO nanocomposite's transition temperature varied between 100 and 200 °C (Fig.  4 I). The characteristic endothermic peaks appear at approximately 118.84 °C, 138.44 °C, and 200.41 °C. These are caused by the release of absorbed water, the breakdown of organic molecule function groups, depolymerization, and decomposition, as well as the dehydration, phase conversion, and full combustion of the organic residue 68 , 69 . The green synthetic trimetallic CuO/Ag/ZnO nanocomposite's DTA curve (Fig.  4 II) displays three exothermic peaks at 112.75, 130.13, and 194.78°C and three endothermic peaks at 118.15, 137.89, and 201.31°C. The heat degradation process is shown in seven phases on the TGA curve in a smooth, stepwise manner (Fig.  4 III). While weight losses of 11.03, 4.12, 2.37, 4.404, 1.89, 2.404, and 6.288% accompanied the breakdown of green synthetic trimetallic CuO/Ag/ZnO nanocomposite, are detected at 79.67, 121.41, 141.83, 211.09, 259.49, 405.12, and 493.92°C. The green synthetic trimetallic CuO/Ag/ZnO nanocomposite loses weight in the initial stages due to the evaporation of adsorbed water molecules and humidity. Because of the breakdown of green synthetic trimetallic CuO/Ag/ZnO nanocomposite matrix, the largest weight losses (> 85.92%) occur at temperatures between 0 and 250 °C, because of CuO/Ag/ZnO is crystallinity-related, the final breakdown (10.57%) takes place between 260 and 500°C.

figure 4

Characterization of green synthesized trimetallic CuO/Ag/ZnO nanocomposite's DSC ( I ), DTA ( II ), and TGA ( III ) curves with FTIR ( IV ) spectrum of green synthesized nanocomposite (black spectrum), and the extract of Ziziphus spina christi (red spectrum).

FTIR spectra of Ziziphus spina christi extract and CuO/Ag/ZnO nanocomposite specimens are shown in Fig.  4 IV. The spectrum of CuO/Ag/ZnO nanocomposite exhibits peaks at around ν 700–400 cm −1 ; in contrast to the extract bonds of Ziziphus spina christi spectrum. This can be attributed to the interactions that Ag has with metal oxides like ZnO or CuO. The green synthesized spectrum of CuO/Ag/ZnO nanocomposite shows distinct peaks at ν 630 cm −1 and peaks at approximately ν 500–420 cm −1 , which are related to the stretching vibrations of CuO and Zn–O, respectively 37 , 38 , 39 . Air humidity most likely influenced the sample measurement. A spectra band of ν 3600–3500 cm −1 is where O–H bond occurs 70 . Consequently, the signal at ν 3478 cm −1 is associated with inter-hydrogen bonding that is present in both the plant extract and nanocomposite spectra represents –OH groups and water molecules. The stretching frequency of the extract's phenolic O–H, which serves as a reducing and capping ligand, is responsible for the broad peak at ν 3354–1606 cm −1 . The stretching of carbon dioxide O=C=O bonds is also responsible for the peak at ν 2351 cm −1 . The additional clear peak is especially visible at ν 1520 cm −1 which is the vibrational frequency of a C=O bond and may indicate the presence of organic residues. Furthermore, there is a peak at ν 1427 cm −1 , which could be related to the O–H bonds in carboxylic acid bending. The stretching vibration of C=O polyphenols may be explained by strong peak at ν 1392 cm −1 . Aromatic C–O and N–H stretching vibrations from phenolic groups were responsible for the strong peaks at 1268 and ν 1076 cm −1 ; respectively. The stretching of C–O bonds in primary alcohols is connected to the peak detected at ν 1113 cm −1 . Overall, FT-IR spectrum of CuO/Ag/ZnO exhibits that; it is coated with active phytoconstituents, mainly O–H, C=O, and C–N residues of alkaloids and phenolic derivatives. To stabilize the resulting CuO/Ag/ZnO nanocomposite, O–H, C=O, and C–N residues might form bonds with metals by covering their surfaces and decreasing agglomeration 69 , 71 . Many studies have reported that a variety of biomolecules found in the Ziziphus spina christi extract are responsible for the stabilization and reduction of the green synthesized nanocomposite. The existence of several functional groups linked to active phytochemicals such as phenolic acids, flavonoids, aromatic compounds, etc . is shown by FTIR analysis of trimetallic Zn/Cu/Ag NCs that are synthesized from the leaf extract 72 . These groups have been suggested to be responsible for the generation of the trimetallic nanocomposite as well as the reduction of metal precursors and subsequent stabilization 61 . The phytochemicals in the extract, primarily flavonoids and phenolic acids, may decrease metal ions by donating electrons, resulting in the generation of metal nanoparticles. Furthermore, this may prevent the particles from aggregating by binding to the surface of the nanoparticle, forming a barrier that reduces surface energy and stabilizes the particles. Further oxidation of the nanoparticles could be inhibited by the carboxyl and hydroxyl groups binding to the metal ions on their surface, protecting the structural integrity of the particles 11 , 51 . Our FTIR results clearly show the presence of flavonoids and phenolic acids, which are responsible for the development of the green synthetic trimetallic CuO/Ag/ZnO nanocomposite.

Statistical optimization of the yield of green synthesized nanocomposite

In general, green synthesized nanocomposites show promise as antibacterial and anticancer agents for safer, more effective, and inexpensive medications or drug delivery systems. The various sizes, forms, dispersions, and stability of the generated nanocomposites are associated with the presented metabolites 40 . Green synthesized procedures for nanocomposite, in particular, have a number of delicate factors 63 , 64 . Several factors that influence the yield shape and size control include the concentrations of plant extract and precursors, as well as the ratio of precursors to other reaction parameters, including temperature, pH, agitation, and incubation time 54 , 65 . Worldwide, scientific investigations are being carried out to find out more about how temperature affects nanoparticles 33 , 54 . The main element that alters the size, shape, and degree of synthesis of the nanoparticles is the temperature 36 . Temperature-dependent modifications can be made to the synthesized nanoparticles' rod, spherical, octahedral platelet, triangular, and spherical shaped structure. Additionally, when the temperature improves, the reaction response rate increases the nucleation center development 54 , 65 . Conversely, the most important variable influencing the yield, size, and shape of nanoparticles generated during the synthesis of green nanoparticles is the reaction time 3 , 62 . According to EL-Moslamy et al., reported that reaction time is critical to produce various nanoparticles and nanocomposites. Therefore, three primary parameters that influence a nanoparticle's shape and structure are temperature, pH, and reaction time 53 , 54 . Until now the utilization of Ziziphus spina christi extract to optimize the conditions of green synthesized trimetallic (CuO/Ag/ZnO) nanocomposite statistically according to regulated conditions remains unexplored. In this study, a two-step experimental strategy known as Plackett–Burman and Taguchi designs is utilized to analyze the parameters influencing the green synthesized reaction to maximize the nanocomposite's green synthesized yield.

Plackett–Burman design

The best parameters for maximizing the dry weight of nanocomposite solutions are identified by using this qualitative and quantitative screening method employing green-synthesized reaction variables. The chosen experiments are utilized to identify the essential elements for the green synesthetic nanocomposites, determine the appropriate ratio, and create a mathematical model, that could be applied to the prediction procedure. The 12 experiments involved screening several components of green-synthetic reaction and exploring each one at two different levels: high (+ 1) and low (− 1), together with a dummy factor used to assess the experiment's standard error. The experiments are completed, and green synthesized nanocomposite’s dry weights are recorded (Table 4 ). Excel 2016 and Minitab 18 are the tools utilized for statistical analysis and graph plotting. As indicated by Table 4 the nanocomposite's highest dry weight was 0.78 mg/mL (run 12) and 0.65 mg/mL (run 8); in contrast, the lowest dry weight is 0 mg/mL that recorded at runs 5 and 7.

The effect of each independent variable on the response is ascertained by analysis of variance ( ANOVA ), where P < 0.05 was deemed statistically significant. Table 4 shows the results of equation's fitness evaluation using the multiple correlation coefficient (R 2 ) and adjusted R 2 . In the overall design, the p value indicates the significance of each independent variable. Larger t-values and smaller p-values (prob > F < 0.05) are associated with greater coefficient influence on the response. The model's overall performance is also estimated using the coefficient of determination (R 2 ) and the adjusted-R 2 (adj-R 2 ) value, which ideally should agree with R 2 value (less than 2%). A stronger model with better response prediction is indicated by R 2 value closer to 1 75 , 76 . The presented data shows model R 2 and adj-R 2 values for the bio-fabrication reaction of the green synthesized nanocomposite, which are 98.58%, and 96.10%; respectively (Table 5 ). According to these findings, the model can account for 98.58% of response data variability, with a 1.42% chance that noise is to blame for the variation. Additionally, a high adj-R 2 value showed that the model was precise and that there is a strong correlation between the experimental and anticipated findings. ANOVA summary typical of experimental Plackett–Burman tests indicated that the model was highly significant, due to the low probability value (p value ~ 0.05). Regarding the green synesthetic nanocomposite's dry weight (mg/mL), each of these components showed an acceptable adjustment (Table 5 ).

As shown in Fig. 5 I, II, and IV nanocomposite's yield is affected by the minimized values of precursor concentrations (F2), precursor ratio (F3), reaction agitation (F4), and reaction temperature (F5) factors, alongside the maximized values of plant extract concentrations (F1), reaction pH (F6), and incubation period (F7). The production efficiency of green synthesized nanocomposites is statistically significantly impacted by all evaluated parameters. As illustrated in Fig.  5 V concentrations of plant extract (F1), concentrations of precursors (F2), ratio of precursors (F3), reaction agitation (F4), reaction pH (F6), and incubation time (F7); are the main factors that influence the production efficiency of green synthesized nanocomposites, more so than reaction temperature (F5). Figure  5 I illustrates the principal impacts of every variable under investigation on the nanocomposite's dry weight. These main effects describe the average differences for each variable between its low and high values. Except for F2, F3, F4, and F5 factors, which vary dramatically between high and low levels, suggesting their impact on amplifying the response at low levels. As a factor rises from a low to a high level, the response always increases when the major effect of the factor is positive (F1, F6, and F7). Because it predicts the maximum dry weight of the nanocomposite using optimal parameters (Fig.  5 III) to determine individual effectiveness, the optimizer tool in MINITAB 18.0 was utilized to solve Eq. ( 7 ). Equation ( 7 ) indicates that a first-order polynomial model that serves as the starting point for the mathematical modeling of the PBD is used to verify the reaction by calculating the average dry weight of green synthesized nanocomposite. The green synthesized nanocomposites are verified by means of the ideal conditions expected for the green reaction, and the results are compared with those recorded under the baseline settings. By using this optimization process, the nanocomposite's dry weight increases from 0.29 to 0.89 mg/mL, i.e. a 3.06-fold increase.

figure 5

Model summary of the factorial regression for the green synthetic nanocomposite (g/ml) for the investigated variables: plant extract concentrations ( F1 ), precursor concentrations ( F2 ), precursor ratio ( F3 ), reaction agitation ( F4 ), reaction temperature ( F5 ), reaction pH ( F6 ), and incubation period ( F7 ) via the following parameters: the main effect plot ( I and II ), the standardized effect using normal plot ( IV ), Pareto chart of the standardized effects ( V ), and a response optimizer with a maximum outcome and optimal values for these variables ( III ) of each variable.

Taguchi statistical method

A cost-efficient and attractive tool for optimizing and generating excellent industrial production processes has been developed by Genichi Taguchi model. According to the requirements of the experiment, several arrays included in Taguchi's model can be employed. Orthogonal arrays (OAs) have designs indicated by Ln (mP), where n signifies the total number of sections, m denotes the number of parameter levels, and P is the total number of parameters 69 , 70 . This work is the first to employ the Taguchi experimental design to statistically optimize the conditions of green synthesized trimetallic CuO/Ag/ZnO nanocomposite qualities using Ziziphus spina christi extract. Taguchi's L27 (3^7) orthogonal array design is utilized to optimize the yield of a green trimetallic CuO/Ag/ZnO nanocomposite. The yield and S/N ratio values of green trimetallic CuO/Ag/ZnO nanocomposite are determined by conducting 27 trials with seven parameters classified according to L27 (3^7) OA design, as indicated in Table 6 . The ideal combination of the responses of green trimetallic CuO/Ag/ZnO nanocomposite is designed by experimentation using Taguchi model. To identify the best combination of the evaluated factors, the data is examined using statistical techniques, including regression analysis and ANOVA . The biggest and smallest yields of green trimetallic CuO/Ag/ZnO nanocomposite values (0.04 and 1.42 mg/mL) are demonstrated in experimental No. 25 and No. 12; respectively. Table 6 displays the structure of Taguchi's orthogonal robust structure, as well as the measurement outcomes. The quality feature that deviates from the intended value is measured using S/N ratio data obtained from Taguchi method. S/N ratios vary based on the green trimetallic CuO/Ag/ZnO nanocomposite yield values (Table 6 ). The S/N ratio and green trimetallic CuO/Ag/ZnO nanocomposite yield values determined by Taguchi's equation (Eq.  3 ) are displayed in Table 6 .

The mean S/N ratio for each parameter level is reported, and Table 6 displays the S/N response table for yield of the green trimetallic CuO/Ag/ZnO nanocomposite. Both an ANOVA and an F-test can be used to assess the experimental data (Table 7 ). Our chosen model suits the experimental data well, as evidenced by its R 2 of 97.36%. So, both the model and its parameters were highly significant (P < 0.0001). The model's F-value stands at 100.33, and the significance F-value is 1.19 E−13 (Table 7 ).

It is shown that, the suggested model is adequate by the residuals found above and below zero line of the residual plot. A straight-line distribution is seen in the residual plots, which suggests the model fits the results effectively (Fig.  6 ). The end confidence level (%) and P-values of each factor indicated that F4, F5, and F6 are significant factors, followed by F7, F2, F1, and F3 (Fig.  6III ). This orthogonal array model is represented by equation No. 8, which also explains the yields of green trimetallic CuO/Ag/ZnO nanocomposite and the relationships between each of the seven elements. As illustrated in Fig.  6 I, the final rankings have the largest S/N ratio value (bigger is better) based on the ANOVA analysis of S/N ratio value and the factor level calculation of the main impact for this dry weight (Table 7 ). For the key effects obtained all through the optimization trial runs, a primary impact graphic was drawn (Fig.  6II ).

figure 6

Characteristics of Taguchi's experimental results: ( I ) the larger-the-better main effects plot for S/N ratios; ( II ) the main effects plot for means of the production efficacy of green synesthetic nanocomposite; ( III ) the p-values and confidence level (%) of each factor in the yield of green synesthetic nanocomposite, and (IV) Schematic diagram of the green synthetic trimetallic CuO/Ag/ZnO nanocomposite employing 25% diluted Ziziphus spina-christi extract (pH = 5) as reducing/capping agent and 0.25 M AgNO, Cu(NO 3 ) 2 .3H 2 O, and Zn(CH 3 COO) 2 .2H 2 O as precursors. The green synthesized trimetallic CuO/Ag/ZnO nanocomposite coated with active phytoconstituents, mainly O–H, C=O, and C–N residues of alkaloids and phenolic derivatives.

To get the highest yield of the green synthesized trimetallic CuO/Ag/ZnO NCs, this approach recommends the optimal combination of the investigated parameters. For green trimetallic CuO/Ag/ZnO nanocomposite, S/N ratio indicates that the following are the ideal conditions: F1 at level 1, F2 at level 1, F3 at level 1, F4 at level 3, F5 at level 3, F6 at level 1, and F7 at level 3. The last stage is to predict and confirm the improvement of quality profile using the ideal level of design parameters after the optimal level has been determined. According to Eq. ( 4 ), it is possible to compute the predicted S/N ratio using design parameters at their optimal level. The pH of 25% diluted Ziziphus spina-christi extract is adjusted to 5 to achieve the highest dry weight possible for green trimetallic CuO/Ag/ZnO nanocomposite. This extract is then titrated slowly using (0.25M AgNO 3 , Cu(NO 3 ) 2 .3H 2 O, and Zn(CH 3 COO) 2 .2H 2 O), which are prepared at 1:1:1 ratio. This reaction is incubated at 50 °C and agitated for 3h at 200 rpm, after this titration phase (Fig.  6 IV). The green trimetallic CuO/Ag/ZnO nanocomposite is optimized statistically under controlled conditions using Ziziphus spina christi extract, resulting in an estimated yield of 1.65 mg/ml and a predicted S/N ratio of roughly 7.79 dB. Lastly, the comparison of data using Placket Burman strategy and Taguchi approach shown that it is feasible to efficiently raise and enhance the yield of green synthetic trimetallic CuO/Ag/ZnO nanocomposite. Compared to Plackett Burman strategy and basal condition, the maximum green synthetic trimetallic CuO/Ag/ZnO nanocomposite yield (1.65 mg/mL) may be increased by 1.85 and 5.7 times; respectively by applying Taguchi strategy.

Antimicrobial potency of green trimetallic CuO/Ag/ZnO nanocomposite

Human infections with antibiotic-resistant microbes are a major cause of death worldwide 79 . Accordingly, a number of antimicrobial nanostructures have been generated recently 80 . So, our study investigated the antimicrobial qualities of different doses of optimized yield of green synthesized trimetallic CuO/Ag/ZnO nanocomposite. Initially, the antimicrobial activity of evaluated doses (50, 100, and 150 µg/mL) is evaluated using agar-well-diffusion method (Fig.  7 ). The inhibitory zone widths of tested doses of green trimetallic CuO/Ag/ZnO nanocomposite against multidrug-resistant human pathogens are determined. Generally, the largest inhibitory zone widths are recorded by using different doses of the green trimetallic CuO/Ag/ZnO nanocomposite against Gram -negative bacteria (Fig. 7 i,ii), and Gram- positive bacteria (Figs. 7 iii, and 8 iv), followed by yeast cells (Fig. 7 v,vi). There are differences in the affected doses for each human pathogen that has been studied, as seen in Fig.  7 I. The results of Table 8 demonstrate that Escherichia coli that are treated with 150 µg/mL of green trimetallic CuO/Ag/ZnO nanocomposite shows the largest inhibitory zone widths (20.68 ± 3.54 mm), followed by Klebsiella pneumoniae (19.22 ± 1.41 mm), and Staphylococcus aureus (17.14 ± 1.98 mm). Additionally, Bacillus subtilis (15.39 ± 3.52 mm), Candida albicans (14.32 ± 2.54 mm), and Candida krusei (13.29 ± 4.22 mm) show the narrowest inhibitory zones, when exposed to 150 µg/mL of green trimetallic CuO/Ag/ZnO NC. Ag-ZnO nanocomposites (75 nm) generated from fenugreek leaf extract at a dosage of 20 mg/mL are found to have antimicrobial properties against several human diseases in a previous study. In the agar diffusion method, the inhibition zone diameter for Escherichia coli is 12.5 ± 0.707 mm, for Staphylococcus aureus is 13.5 ± 0.707 mm, and for Candida albicans is 10.5 ± 0.707 mm 81 . But herein, Escherichia coli treated with 150 µg/mL of green trimetallic CuO/Ag/ZnO nanocomposite demonstrate the highest inhibitory zone widths (20.68 ± 3.54 mm), followed by Staphylococcus aureus (17.14 ± 1.98 mm) and Candida albicans (14.32 ± 2.54 mm). Our results show that Ag-ZnO nanocomposites have less impact on S. aureus and E. coli . Other reports on the antimicrobial abilities of Ag, ZnO, and CuO nanoparticles generated from various plant extracts have also been reported previously 11 , 74 , 75 , 76 . Numerous nanoparticles with antibacterial qualities have also been demonstrated in other studies, which include silica, iron oxide, copper oxide, magnesium oxide, titanium dioxide, silver, zinc oxide, and cerium dioxide. The capacity of nanomaterials to limit microbial development depends on the layers of the pathogen's cell wall or membrane structure. The synthesized nanostructure's size, shape, and core–shell morphology, which provide a high surface-area-to-volume ratio, also have an impact on the proliferation of microbes 17 , 82 , 85 .

figure 7

Antimicrobial efficacy results for tested doses of green trimetallic CuO/Ag/ZnO nanocomposite labeled ( A : 50 µg/mL, B : 100 µg/mL, C : 150 µg/mL) against various multidrug-resistant human pathogens ( i : Escherichia coli, ii : Klebsiella pneumoniae, iii : Staphylococcus aureus, iv : Bacillus subtilis, v : Candida albicans, and vi : Candida krusei ). Photographs depict an Agar-well diffusion investigation. Chart displays the computed inhibition zones ( I ), box-plot graph ( II ) displays the inhibitory value distributions corresponding to the tested doses; and simultaneous results for analyzing the overall group's difference ( III ) via Tukey–Kramer post-hoc analysis. Means that don't have the same letter differ greatly.

figure 8

Reduction in biofilm generation of the tested human pathogens using a biofilm inhibition assay. Chart shows the percentage of biofilm reduction ( I ), box-plot graph shows biofilm reduction value distributions corresponding to drug dosages via Tukey–Kramer post-hoc analysis ( II ), and simultaneous Tukey results appearing the overall group's difference ( III ).

The ANOVA , and Tukey–Kramer post-hoc analysis is employed to demonstrate the inhibitory value distributions that correspond to tested doses. Furthermore, data about the correlation between tested products and antimicrobial effectiveness is grouped using statistical clustering. So, the inhibitory effect distributions that match tested treatments are displayed on comparable interval and box plot graphs (Fig. 7 II, III). The Tukey–Kramer post-hoc results show that, out of all the treatments that are assessed, 150 µg/mL of green trimetallic CuO/Ag/ZnO nanocomposite have the highest anti-biofilm value. A boxplot that displays the significant mean differences is produced for each paired comparison using the means of Tukey's test. As can be seen in the box-plot graph (Fig.  7 II), there are significant antimicrobial variations among all tested doses of green trimetallic CuO/Ag/ZnO nanocomposite. Especially, 150 µg/mL of green trimetallic CuO/Ag/ZnO nanocomposite have the highest inhibitory values based on Tukey–Kramer post-hoc results. On 95% scale, the modified confidence intervals are computed using Tukey simultaneous tests. Due to the absence of zero line in the intervals for formulation with the highest efficacy, 150 µg/mL of green trimetallic CuO/Ag/ZnO nanocomposite, whose mean values are shown in Fig.  7 III, shows significant differences. All these results indicate that 150 µg/mL of green trimetallic CuO/Ag/ZnO nanocomposite have the strongest antimicrobial properties, compared to all doses that are tested.

Spectrophotometric antibiofilm assay is used to assess the antimicrobial efficacy of green trimetallic CuO/Ag/ZnO nanocomposite with several doses ranging from 50 to 250 µg/mL, against all tested human pathogens. The percentage of biofilm reduction is utilized to determine doses' in vitro efficacy to prevent pathogen growth (Fig.  8 ). The antimicrobial chart depicts 200 µg/mL dose's strong antagonistic antimicrobial effects against all pathogens tested (Fig.  8 I). Additionally, tested Gram -positive have the highest antimicrobial effect more than tested Gram -negative, and yeast cells. The highest percentage of antibiofilm after treatment with 200 µg/mL of green trimetallic CuO/Ag/ZnO nanocomposite are of 98.31 ± 0.98, and 97.68 ± 1.11% that are recorded against tested Gram -positive pathogens e.g. Bacillus subtilis , and Staphylococcus aureus; respectively (Table 9 ). Additionally, the modest percentage of antibiofilm are recorded against Escherichia coli (92.45 ± 1.41%), Klebsiella pneumoniae (91.07 ± 1.09%), Candida albicans (90.99 ± 0.87%), Candida krusei and (89.59 ± 0.15%), as seen in Table 9 . To statistically ascertain whether doses are more effective, the mean values of computed antibiofilm percentages are assessed using ANOVA and Tukey post-hoc test , Fig.  8 II, III. Furthermore, the correlation data between tested doses and antimicrobial effectiveness is grouped using statistical clustering. So, the inhibitory effect distributions that match the tested treatments are displayed on comparable interval and box plot graphs. A boxplot displays the significant mean differences and is produced for each paired comparison using the means of Tukey's test. Additionally, out of all doses that are assessed, 200 µg/mL of green trimetallic CuO/Ag/ZnO nanocomposite have the highest anti-biofilm value (Fig.  8 II). On a 95% scale, the modified confidence intervals are computed using Tukey simultaneous tests. There are narrow statistical differences between the recorded antibiofilm percentages intervals of 150–200, and 150–250 µg/mL doses of green trimetallic CuO/Ag/ZnO nanocomposite (pass through the zero line), as seen in Fig.  8 III. Due to the absence of zero line in the intervals for 200–250 µg/mL dose shows significant differences. So, the recorded MICs for all tested human pathogens range from 150 to 200 µg/mL (Table 9 ). An additional investigation 16 , examined the antimicrobial potential of green binary ZnO/CuO nanocomposites (irregular rod-shaped particles 7.52 nm in size) produced from Calotropis gigantea against drug-sensitive human pathogens ( Staphylococcus aureus and Escherichia coli ), multi-drug-resistant human pathogens ( Klebsiella pneumoniae , Pseudomonas aeruginosa , and methicillin-resistant S. aureus ). For S. aureus, its MICs varied between 5 and 2.5 mg/mL. Furthermore, for E. coli, P. aeruginosa, K. pneumoniae, and MRSA , the MIC values were 0.625, 0.15625, 0.625, and 0.15625 mg/mL, respectively. Therefore, our outcomes are extremely proficient, compared to earlier studies.

The 200 µg/mL of green trimetallic CuO/Ag/ZnO nanocomposite that demonstrates the highest degree of anti-microbial activity, is further focused for more antimicrobial exploration. The 200 µg/mL's time-kill kinetics are studied for every pathogen as part of time-kill analysis. Additionally, the log 10 CFU/mL levels and quantitative reduction of biofilm for each examined pathogens (treated, and untreated cells) are listed in Table 10 . The comparability of all studied human pathogens treated with 200 µg/mL of green trimetallic CuO/Ag/ZnO nanocomposite with the corresponding untreated cells are shown in Fig.  9 . As seen, there are differences in log 10 CFU/mL measurements within all tested human pathogens. Gram -positive bacteria show a significant decline in planktonic viable counts after 18 h (Fig. 9 iii, iv), however Gram -negative bacteria (Fig. 9 i, ii) and yeast cells (Fig. 9 v, vi) show a similar decline after 24 h. Among the studied bacteria, Escherichia coli , and Staphylococcus aureus show the highest percentage of biofilm reduction (98.06 ± 0.93, and 97.47 ± 0.65%; respectively), and its planktonic viable counts are effectively diminished by the tested 200 µg/mL of green trimetallic CuO/Ag/ZnO nanocomposite after 36-h period. However, the planktonic viable counts of Candida albicans (95.42 ± 1.78%) is subsequently successfully reduced after a 36-h interval (Table 10 ). The time-kill assay is also utilized to ascertain the length of 200 µg/mL of green trimetallic CuO/Ag/ZnO nanocomposite that is necessary to completely eradicate the pathogens' biofilm. The biofilms of treated Gram -positive bacteria reveal 0% CFU/ml after 52 h; however, the biofilms are destroyed by the treated yeast cells and Gram -negative bacteria after 72 and 96 h; respectively. Lastly, a promising green trimetallic CuO/Ag/ZnO nanocomposite has the potential to be used as an antimicrobial substance to suppress different human pathogens that are resistant to antibiotics.

figure 9

Growth rate reduction in cell viability for all multidrug-resistant human pathogens ( i : Escherichia coli, ii : Klebsiella pneumoniae, iii : Staphylococcus aureus, iv : Bacillus subtilis, v : Candida albicans, and vi : Candida krusei ) treated with 200 µg/mL of green trimetallic CuO/Ag/ZnO nanocomposite, as well as the untreated cells during the incubation period.

The main mechanism causing the antimicrobial effect is an interaction between the pathogenic microbes' cell wall receptors and the surface of the generated nanomaterials 33 , 54 . The green trimetallic CuO/Ag/ZnO nanocomposite might have direct contact with the negatively charged microbial membrane through ions released, due to surface oxidation, complicated porosity, or electrostatic interaction. According to Noohpisheh et al., there is a strong interaction between metallic silver and semiconductor zinc oxide that splits the cell membrane and increases antimicrobial activity 73 . Numerous investigations have indicated nanocomposites have superior antimicrobial properties, compared to their individual nanoparticle counterparts 81 . Based on previous studies, the green synthesized trimetallic CuO/Ag/ZnO nanocomposite damages the microbial wall, penetrates the cytoplasm, and causes cell death; because it generates superoxide and hydroxy, which alter membrane protein as well as enzyme activity 53 , 54 , 66 , 73 . This green synthesized trimetallic CuO/Ag/ZnO nanocomposite should therefore be used in food packaging and surgical tool coatings to prevent microbial infection and strongly inhibit the growth.

Conclusions

In conclusion, the industrial green-synthesis of trimetallic CuO/Ag/ZnO nanocomposite was achieved by the utilization of an eco-friendly, straightforward approach that involved the extraction of sustainable resources of leaves from Ziziphus spina christi . Results of FTIR and phytochemical analysis showed that this extract included large concentrations of proteins, reducing sugar, anthocyanin, flavonoids, and phenolic compounds. Subsequently, the highest feasible dry weight for the green synthetic trimetallic CuO/Ag/ZnO nanocomposite was obtained by adjusting the pH of 25% diluted Ziziphus spina-christi extract (reductants) to 5. The precursors composed of (0.25M AgNO 3 , Cu (NO 3 ) 2 .3H 2 O, and Zn (CH 3 COO) 2 .2H 2 O) were subsequently prepared at a (1:1:1) ratio and gradually titrated into this plant extract. Furthermore, distinct statistically experimental designs ( Plackett Burman and Taguchi methods) were used to scaling-up the yield of green-produced trimetallic CuO/Ag/ZnO nanocomposite. Finally, the highest green synthesized trimetallic CuO/Ag/ZnO nanocomposite yield (1.65 mg/mL) might be enhanced by 1.85 and 5.7 times; respectively, by using the Taguchi approach in comparison to Plackett–Burman strategy and basal condition. In vitro, trimetallic nanocomposites have been applied to eliminate multi-drug-resistant human pathogens to determine their antimicrobial capabilities. All examined human pathogens were found to have MICs ranging from 150 to 200 µg/mL. The biofilms of treated Gram -positive bacteria showed 0% CFU/mL after 52 h. However, the treated Gram -negative bacteria and yeast cells totally eradicated the formation of biofilms after 72 and 96 h; respectively. Overall, the green synthesized trimetallic CuO/Ag/ZnO nanocomposite on a large scale has novel opportunities for the generation of antimicrobial agents that are both very stable and effective for inhibiting and preventing the microbial growth, moreover its eco-friendly generation from plant extracts.

Data availability

The datasets used and/or analyzed during the current study available from the corresponding authors (A.K. EL-Sawaf and E.A. Kamoun) on reasonable request.

Bhole, R. et al. Superparamagnetic spherical magnetite nanoparticles: Synthesis, characterization and catalytic potential. Appl. Nanosci. 13 (9), 6003–6014. https://doi.org/10.1007/s13204-022-02532-4 (2023).

Article   ADS   CAS   Google Scholar  

Sridevi, H., Bhat, M. R., Kumar, P. S., Kumar, N. M. & Selvaraj, R. Structural characterization of cuboidal α-Fe2O3 nanoparticles synthesized by a facile approach. Appl. Nanosci. 13 (8), 5605–5613. https://doi.org/10.1007/s13204-023-02780-y (2023).

El-Moslamy, S. H., Abd-Elhamid, A. I. & El Fawal, G. Large-scale production of myco-fabricated ZnO/MnO nanocomposite using endophytic Colonstachys rosea with its antimicrobial efficacy against human pathogens. Sci. Rep. 14 (1), 1–20. https://doi.org/10.1038/s41598-024-51398-9 (2024).

Article   CAS   Google Scholar  

Shet, V. B. et al. Cocoa pod shell mediated silver nanoparticles synthesis, characterization, and their application as nanocatalyst and antifungal agent. Appl. Nanosci. 13 (6), 4235–4245. https://doi.org/10.1007/s13204-023-02873-8 (2023).

Razavi, R., Amiri, M., Alshamsi, H. A., Eslaminejad, T. & Salavati-Niasari, M. Green synthesis of Ag nanoparticles in oil-in-water nano-emulsion and evaluation of their antibacterial and cytotoxic properties as well as molecular docking. Arab. J. Chem. 14 (9), 103323. https://doi.org/10.1016/j.arabjc.2021.103323 (2021).

Abbas, H. H., Al-Saadi, N. H. & Hassan, A. H. Green synthesis of zinc oxide-nanoparticles from Ziziphus Spina-Christi leaves extract: Characterization and their protective effects against liver disturbances in adenine-exposed male rats. Trends Sci. 20 , 10. https://doi.org/10.48048/tis.2023.6909 (2023).

Article   Google Scholar  

Tilahun, E., Adimasu, Y. & Dessie, Y. Biosynthesis and optimization of ZnO nanoparticles using Ocimum lamifolium leaf extract for electrochemical sensor and antibacterial activity. ACS Omega 8 (30), 27344–27354. https://doi.org/10.1021/acsomega.3c02709 (2023).

Article   CAS   PubMed   PubMed Central   Google Scholar  

Abdelkhalek, A. & Al-Askar, A. A. Green synthesized ZnO nanoparticles mediated by Mentha spicata extract induce plant systemic resistance against Tobacco mosaic virus. Appl. Sci. 10 , 15. https://doi.org/10.3390/app10155054 (2020).

Stoyanova, D. et al. Modified approach using Mentha arvensis in the synthesis of ZnO nanoparticles—textural, structural, and photocatalytic properties. Appl. Sci. 12 , 3. https://doi.org/10.3390/app12031096 (2022).

Ahmad, N. et al. Antimicrobial efficacy of Mentha piperata-derived biogenic zinc oxide nanoparticles against UTI-resistant pathogens. Sci. Rep. 13 (1), 1–16. https://doi.org/10.1038/s41598-023-41502-w (2023).

Article   ADS   MathSciNet   CAS   Google Scholar  

Govindasamy, G. A., Mydin, R. B. S. M. N., Sreekantan, S. & Harun, N. H. Compositions and antimicrobial properties of binary ZnO–CuO nanocomposites encapsulated calcium and carbon from Calotropis gigantea targeted for skin pathogens. Sci. Rep. 11 (1), 1–14. https://doi.org/10.1038/s41598-020-79547-w (2021).

Ahamad Khan, M. et al. Phytogenically synthesized zinc oxide nanoparticles (ZnO-NPs) potentially inhibit the bacterial pathogens: In vitro studies. Toxics 11 (5), 1–23. https://doi.org/10.3390/toxics11050452 (2023).

Mahmoud, A. E. D., El-Maghrabi, N., Hosny, M. & Fawzy, M. Biogenic synthesis of reduced graphene oxide from Ziziphus spina-christi (Christ’s thorn jujube) extracts for catalytic, antimicrobial, and antioxidant potentialities. Environ. Sci. Pollut. Res. 29 (59), 89772–89787. https://doi.org/10.1007/s11356-022-21871-x (2022).

Dhivya, A., Yadav, R. & Powrnika, S. Green synthesis of selenium doped zinc oxide nanoparticles using Mangifera indica leaf extract and its photodegradation and antibacterial activities. J. Nanosci. Technol. 5 (3), 741–744. https://doi.org/10.30799/jnst.s08.19050310 (2019).

Kunwar, S. et al. Bio-fabrication of Cu/Ag/Zn nanoparticles and their antioxidant and dye degradation activities. Catalysts 13 (5), 1–19. https://doi.org/10.3390/catal13050891 (2023).

Sabouri, Z. et al. Phytoextract-mediated synthesis of Ag-doped ZnO–MgO–CaO nanocomposite using Ocimum Basilicum L seeds extract as a highly efficient photocatalyst and evaluation of their biological effects. Ceram. Int. 49 (12), 20989–20997. https://doi.org/10.1016/j.ceramint.2023.03.234 (2023).

Abbas, S. et al. Dual-functional green facile CuO/MgO nanosheets composite as an efficient antimicrobial agent and photocatalyst. Arab. J. Sci. Eng. 1 , 1. https://doi.org/10.1007/s13369-021-05741-1 (2021).

Moeen, M. et al. Green synthesis, characterization and sorption efficiency of MnO2 nanoparticles and MnO2@waste eggshell nanocomposite. J. Taibah Univ. Sci. 16 (1), 1075–1095. https://doi.org/10.1080/16583655.2022.2139483 (2022).

Abdelaziz, A. M. et al. Ziziphus spina-christi extract-stabilized novel silver nanoparticle synthesis for combating Fusarium oxysporum-causing pepper wilt disease: in vitro and in vivo studies. Arch. Microbiol. 205 (2), 1–17. https://doi.org/10.1007/s00203-023-03400-7 (2023).

Erol, I., Sivrier, M., Cigerci, I. H., Özkara, A. & Akyıl, D. ZnO-containing nanocomposites produced from Mentha pulegium L. of a new HEMA-based methacrylate copolymer: Improvement the thermal and antimicrobial effect. J. Polym. Res. 30 , 3. https://doi.org/10.1007/s10965-023-03461-8 (2023).

Suresh, J., Ragunath, L. & Hong, S. I. Biosynthesis of mixed nanocrystalline Zn-Mg-Cu oxide nanocomposites and their antimicrobial behavior. Adv. Nat. Sci. Nanosci. Nanotechnol. 10 , 4. https://doi.org/10.1088/2043-6254/ab52f5 (2019).

Baig, A. U. et al. Facile green synthesis of silver doped zno nanoparticles using tridax procumbens leaf extract and their evaluation of antibacterial activity. J. Water Environ. Nanotechnol. 5 (4), 307–320. https://doi.org/10.22090/jwent.2020.04.002 (2020).

Article   MathSciNet   CAS   Google Scholar  

Panchal, P. et al. Eco-friendly synthesis of Ag-doped ZnO/MgO as a potential photocatalyst for antimicrobial and dye degradation applications. Coord. Chem. Rev. 493 , 215283. https://doi.org/10.1016/j.ccr.2023.215283 (2023).

Vijayaram, S. et al. Applications of green synthesized metal nanoparticles—a review. Biol. Trace Elem. Res. 1 , 1. https://doi.org/10.1007/s12011-023-03645-9 (2023).

Sabouri, Z., Kazemi, M. & Sabouri, M. Biosynthesis of Ag doped MgO-NiO-ZnO nanocomposite with Ocimum Basilicum L extract and assessment of their biological and photocatalytic applications. J. Mol. Struct. 1306 , 137895. https://doi.org/10.1016/j.molstruc.2024.137895 (2024).

Makauki, E., Mtavangu, S. G., Basu, O. D., Rwiza, M. & Machunda, R. Facile biosynthesis of Ag–ZnO nanocomposites using Launaea cornuta leaf extract and their antimicrobial activity. Discov. Nano 18 , 1. https://doi.org/10.1186/s11671-023-03925-2 (2023).

Jakinala, P. et al. Green synthesis of ZnO-Ag nanocomposite using Stenotaphrum secundatum grass extract: Antibacterial activity and anticancer effect in oral squamous cell carcinoma CAL 27 cells. Inorg. Chem. Commun. 152 , 110735. https://doi.org/10.1016/j.inoche.2023.110735 (2023).

Jalil, P. J., Shnawa, B. H., Hamad, S. M., Hamad, B. S. & Ahmed, M. H. The efficiency of fabricated Ag/ZnO nanocomposite using Ruta chalepensis L. leaf extract as a potent protoscolicidal and anti-hydatid cysts agent. J. Biomater. Appl. 38 (5), 629–645. https://doi.org/10.1177/08853282231207236 (2023).

Article   CAS   PubMed   Google Scholar  

Pompapathi, K. et al. Visible-light-driven mentha spicata L.-mediated Ag-doped Bi2Zr2O7 nanocomposite for enhanced degradation of organic pollutants, electrochemical sensing, and antibacterial applications. ACS Environ. 1 , 1. https://doi.org/10.1021/acsenvironau.3c00057 (2023).

Mohanaparameswari, S. et al. Investigation of structural properties and antibacterial activity of AgO nanoparticle extract from Solanum nigrum/Mentha leaf extracts by green synthesis method. Green Process. Synth. 12 (1), 1–17. https://doi.org/10.1515/gps-2023-0080 (2023).

El-Shahir, A. A., El-Wakil, D. A., Latef, A. A. H. A. & Youssef, N. H. Bioactive compounds and antifungal activity of leaves and fruits methanolic extracts of Ziziphus spina-christi L. Plants 11 (6), 1–18. https://doi.org/10.3390/plants11060746 (2022).

Souri, M. & Shakeri, A. Optimization of total phenol and tannin content and biological activity of Dittrichia graveolens (L.) GREUTER. Curr. Bioact. Compd. 16 (2), 124–132. https://doi.org/10.2174/1573407214666180730110830 (2018).

Seddiek, A. S., Hamad, G. M., Zeitoun, A. A., Zeitoun, M. A. M. & Ali, S. Antimicrobial and antioxidant activity of some plant extracts against different food spoilage and pathogenic microbes. Eur. J. Nutr. Food Saf. 1 , 1–12. https://doi.org/10.9734/ejnfs/2020/v12i1130312 (2020).

Verma, A., Kumar, R., Kumari, S., Lata, S., & Zinzala, V. Synthesized nano ZnO and its comparative effects with ZnO and heptahydrate ZnSO4 on sweet corn (Zea mays L. saccharata). Pharma Innov. J. 10 (10), 1–7 [Online]. Available: http://www.thepharmajournal.com (2021).

Adak, T. et al. Green silver nano-particles: Synthesis using rice leaf extract, characterization, efficacy, and non-target effects. Environ. Sci. Pollut. Res. 1 , 4452–4462. https://doi.org/10.1007/s11356-020-10601-w (2020).

El-Moslamy, S. H., Yahia, I. S., Zahran, H. Y. & Kamoun, E. A. Novel biosynthesis of MnO NPs using Mycoendophyte: Industrial bioprocessing strategies and scaling-up production with its evaluation as anti-phytopathogenic agents. Sci. Rep. 13 (1), 1–21. https://doi.org/10.1038/s41598-023-28749-z (2023).

El-Moslamy, S. H., El-Morsy, E. S. M., Mohaisen, M. T., Rezk, A. H. & Abdel-Fattah, Y. R. Industrial bioprocessing strategies for cultivation of local Streptomyces violaceoruber strain SYA3 to fabricate nano-ZnO as anti-phytopathogens agent. J. Pure Appl. Microbiol. 12 (3), 1133–1145. https://doi.org/10.22207/JPAM.12.3.12 (2018).

El-Moslamy, S. H. Bioprocessing strategies for cost-effective large-scale biogenic synthesis of nano-MgO from endophytic Streptomyces coelicolor strain E72 as an anti-multidrug-resistant pathogens agent. Sci. Rep. 8 (1), 1–22. https://doi.org/10.1038/s41598-018-22134-x (2018).

CLSI. Methods for antimicrobial broth dilution and disk diffusion susceptibility testing of bacteria isolated from aquatic animals. In VET03-Clinical and Laboratory Standards Institute , vol. VET03, 2020, p. 122. Accessed: Oct. 19, 2022. [Online]. Available: www.clsi.org .

El-Moslamy, S. H., Elnouby, M. S., Rezk, A. H. & El-Fakharany, E. M. Scaling-up strategies for controllable biosynthetic ZnO NPs using cell free-extract of endophytic Streptomyces albus: characterization, statistical optimization, and biomedical activities evaluation. Sci. Rep. 13 (1), 1–22. https://doi.org/10.1038/s41598-023-29757-9 (2023).

Kakian, F., Mirzaei, E., Moattari, A., Takallu, S. & Bazargani, A. Determining the cytotoxicity of the Minimum Inhibitory Concentration (MIC) of silver and zinc oxide nanoparticles in ESBL and carbapenemase producing Proteus mirabilis isolated from clinical samples in Shiraz, Southwest Iran. BMC Res. Notes 17 (1), 1–6. https://doi.org/10.1186/s13104-023-06402-2 (2024).

Titus, D., James Jebaseelan Samuel, E., & Roopan, S. M. Nanoparticle characterization techniques. In Green Synthesis, Characterization and Applications of Nanoparticles , pp. 303–319. (Elsevier, 2019). https://doi.org/10.1016/b978-0-08-102579-6.00012-5 .

Betiha, M. A. et al. A review on different plants extract mediated silver nanoparticles: Preparation, antimicrobials, and antioxidant. Egypt. J. Chem. 65 (5), 575–589. https://doi.org/10.21608/ejchem.2021.99747.4637 (2022).

Arvanagh, F. M. et al. Anti-inflammatory and collagenation effects of zinc oxide-based nanocomposites biosynthesised with Mentha longifolia leaf extract. J. Wound Care 32 (1), 44–54. https://doi.org/10.12968/jowc.2023.32.1.44 (2023).

Article   PubMed   Google Scholar  

Consolo, V. F., Torres-Nicolini, A. & Alvarez, V. A. Mycosinthetized Ag, CuO and ZnO nanoparticles from a promising Trichoderma harzianum strain and their antifungal potential against important phytopathogens. Sci. Rep. 10 (1), 1–10. https://doi.org/10.1038/s41598-020-77294-6 (2020).

Shinde, R. S. et al. Design, fabrication, antitubercular, antibacterial, antifungal and antioxidant study of silver doped ZnO and CuO nano candidates: A comparative pharmacological study. Curr. Res. Green Sustain. Chem. 4 , 100138. https://doi.org/10.1016/j.crgsc.2021.100138 (2021).

Singh, J. et al. ‘Green’ synthesis of metals and their oxide nanoparticles: Applications for environmental remediation. J. Nanobiotechnol. 16 , 1. https://doi.org/10.1186/s12951-018-0408-4 (2018).

Jubair, N., Rajagopal, M., Chinnappan, S., Abdullah, N. B. & Fatima, A. Review on the antibacterial mechanism of plant-derived compounds against multidrug-resistant bacteria (MDR). Evid. Based Complement. Altern. Med. 1 , 1. https://doi.org/10.1155/2021/3663315 (2021).

Khosravi, R. et al. Facile green synthesis of zinc oxide nanoparticles using Thymus vulgaris extract, characterization, and mechanism of chromium photocatalytic reduction. Mater. Res. Express 6 (11), 115093. https://doi.org/10.1088/2053-1591/ab4ae7 (2019).

Article   ADS   Google Scholar  

Pekdemir, M. E., Pekdemir, S., İnci, Ş, Kırbağ, S. & Çiftci, M. Thermal, magnetic properties and antimicrobial effects of magnetic iron oxide nanoparticles treated with polygonum cognatum. Iran. J. Sci. Technol. Trans. A Sci. 45 (5), 1579–1586. https://doi.org/10.1007/s40995-021-01167-4 (2021).

Tabrizi Hafez Moghaddas, S. S., Samareh Moosavi, S. & Kazemi Oskuee, R. Green synthesis of calcium oxide nanoparticles in Linum usitatissimum extract and investigation of their photocatalytic and cytotoxicity effects. Biomass Convers. Biorefinery 14 (4), 5125–5134. https://doi.org/10.1007/s13399-022-02643-6 (2024).

Pujar, M., Kumar, K. & Kumar, P. IRJET- Recent Trends in Green Synthesis of ZnO Nanomaterials Using Plant Extracts. Int. Res. J. Eng. Technol. 08 , 152–163 (2021).

Google Scholar  

Azad, A., Zafar, H., Raza, F. & Sulaiman, M. Factors influencing the green synthesis of metallic nanoparticles using plant extracts: A comprehensive review. Pharm. Front. 05 (03), e117–e131. https://doi.org/10.1055/s-0043-1774289 (2023).

Nzilu, D. M. et al. Green synthesis of copper oxide nanoparticles and its efficiency in degradation of rifampicin antibiotic. Sci. Rep. 13 (1), 1–18. https://doi.org/10.1038/s41598-023-41119-z (2023).

Bin Chan, Y. et al. Impact of diverse parameters on the physicochemical characteristics of green-synthesized zinc oxide-copper oxide nanocomposites derived from an aqueous extract of garcinia mangostana L. Leaf. Materials 16 (15), 1. https://doi.org/10.3390/ma16155421 (2023).

Mohamed, A. A., Abu-Elghait, M., Ahmed, N. E. & Salem, S. S. Correction to: Eco-friendly mycogenic synthesis of ZnO and CuO nanoparticles for in vitro antibacterial, antibiofilm and antifungal applications (biological trace element research, (2021), 199, 7, (2788–2799), DOI: 10.1007/s12011-020-02369-4). Biol. Trace Elem. Res. 199 (7), 2800–2801. https://doi.org/10.1007/s12011-020-02391-6 (2021).

Asamoah, R. B., et al. A comparative study of antibacterial activity of CuO/Ag and ZnO/Ag nanocomposites. Adv. Mater. Sci. Eng. https://doi.org/10.1155/2020/7814324 (2020).

El Moslamy, S. H. & Kabeil, S. S. A. Bioprocess development for chlorella vulgaris cultivation and biosynthesis of anti-phytopathogens silver nanoparticles. J. Nanomater. Mol. Nanotechnol. 5 (1), 1. https://doi.org/10.4172/2324-8777.1000177 (2016).

El-Moslamy, S. H., & Abdel-Fattah, Y. R. Statistical bioprocessing strategy for cellulases production by endophytic Trichoderma harzianum utilizing lignocellulosic wastes. Biosci. Res. 15 (3), 1852–1866. Available: www.isisn.org (2018).

El-Moslamy, S. H., Elnouby, M. S., Rezk, A. H. & El-Fakharany, E. M. Scaling-up strategies for controllable biosynthetic ZnO NPs using cell free-extract of endophytic Streptomyces albus: characterization, statistical optimization, and biomedical activities evaluation. Sci. Rep. 13 , 1. https://doi.org/10.1038/s41598-023-29757-9 (2023).

S. H. EL-Moslamy, A. H. Rezk, M. F. Elkady, and H. Shokry Hassan,. Semi-industrial bio-fabrication of ZnO/MnO2 nanocomposite using endophytic streptomyces coelicolor: Characterization, statistical design, exponential pulse fed-batch fermentation, and its antimicrobial application. Arab. J. Sci. Eng. 49 , 9067–9088. https://doi.org/10.1007/s13369-024-08709-z (2024).

El-Moslamy, S. H., Yahia, I. S., Zahran, H. Y. & Kamoun, E. A. Novel biosynthesis of MnO NPs using Mycoendophyte: Industrial bioprocessing strategies, characterization, scaling-up production, and its evaluation as anti-phytopathogenic agents. Sci. Rep. 1 , 1–21. https://doi.org/10.1038/s41598-023-28749-z (2022).

Sabouri, Z., Sammak, S., Sabouri, S., Sadat, S., & Hafez, T. Green synthesis of Ag-Se doped ZnO-Co 3 O 4 -NiO fivenary nanocomposite using poly anionic cellulose and evaluation of their anticancer and photocatalyst applications. 8 , 164–176 (2024).

Mahmood, K. et al. Green synthesis of Ag@CdO nanocomposite and their application towards brilliant green dye degradation from wastewater. J. Nanostructure Chem. 12 (3), 329–341. https://doi.org/10.1007/s40097-021-00418-5 (2022).

Karvekar, O. S. et al. Biogenic synthesis of silver anchored ZnO nanorods as nano catalyst for organic transformation reactions and dye degradation. Appl. Nanosci. 12 (7), 2207–2226. https://doi.org/10.1007/s13204-022-02470-1 (2022).

Article   ADS   CAS   PubMed   PubMed Central   Google Scholar  

Vinayagam, R. et al. Green synthesized hydroxyapatite nanoadsorbent for the adsorptive removal of AB113 dye for environmental applications. Environ. Res. 212 , 1. https://doi.org/10.1016/j.envres.2022.113274 (2022).

Varadavenkatesan, T., Pai, S., Vinayagam, R. & Selvaraj, R. Characterization of silver nano-spheres synthesized using the extract of Arachis hypogaea nuts and their catalytic potential to degrade dyes. Mater. Chem. Phys. 272 , 1. https://doi.org/10.1016/j.matchemphys.2021.125017 (2021).

Abd El-Mohdy, H. L. & Aly, H. M. Characterization, properties and antimicrobial activity of radiation induced phosphorus-containing PVA hydrogels. Arab. J. Sci. Eng. 48 (1), 341–351. https://doi.org/10.1007/s13369-022-07031-w (2023).

Kumar, Y. et al. Development and characterization of defatted pumpkin seed meal and halloysite nanoclay composite films for food packaging. Packag. Technol. Sci. 36 (8), 715–727. https://doi.org/10.1002/pts.2751 (2023).

Fekri, R., Mirbagheri, S. A., Fataei, E., Ebrahimzadeh-Rajaei, G. & Taghavi, L. Green synthesis of CuO nanoparticles using Peganum harmala extract for photocatalytic and sonocatalytic degradation of reactive dye and organic compounds. Main Gr. Chem. 21 (4), 975–996. https://doi.org/10.3233/MGC-220045 (2022).

Al-Qadsy, I. et al. Antimicrobial activity of novel Ni(II) and Zn(II) Complexes with (E)-2-((5-Bromothiazol-2-yl)imino)methyl)phenol Ligand: Synthesis, characterization and molecular docking studies. Antibiotics 12 , 11. https://doi.org/10.3390/antibiotics12111634 (2023).

Aasy, N. K. A. et al. Concurrent tissue engineering for wound healing in diabetic rats utilizing dual actions of green synthesized CuO NPs prepared from two plants grown in Egypt. Int. J. Nanomed. 18 (April), 1927–1947. https://doi.org/10.2147/IJN.S397045 (2023).

Yang, Y., Waterhouse, G. I. N., Chen, Y., Sun-Waterhouse, D. & Li, D. Microbial-enabled green biosynthesis of nanomaterials: Current status and future prospects. Biotechnol. Adv. 55 , 107914. https://doi.org/10.1016/j.biotechadv.2022.107914 (2022).

Pakseresht, S. et al. Review—nanomaterials green synthesis for high-performance secondary rechargeable batteries: Approaches, challenges, and perspectives. J. Electrochem. Soc. 169 (1), 010534. https://doi.org/10.1149/1945-7111/ac4843 (2022).

El-Belely, E. F. et al. Green synthesis of zinc oxide nanoparticles (Zno-nps) using arthrospira platensis (class: Cyanophyceae) and evaluation of their biomedical activities. Nanomaterials 11 (1), 1–18. https://doi.org/10.3390/nano11010095 (2021).

Jamdagni, P., Rana, J. S., Khatri, P. & Nehra, K. Comparative account of antifungal activity of green and chemically synthesized Zinc Oxide nanoparticles in combination with agricultural fungicides. Int. J. Nano Dimens. 9 (2), 198–208 (2018).

CAS   Google Scholar  

Sharma, R. et al. Statistical optimization of process parameters for improvement of phycobiliproteins (PBPs) yield using ultrasound-assisted extraction and its kinetic study. Ultrason. Sonochem. 60 , 104762. https://doi.org/10.1016/j.ultsonch.2019.104762 (2020).

Rizal, S. et al. Preparation of palm oil ash nanoparticles: Taguchi optimization method by particle size distribution and morphological studies. Appl. Sci. 10 (3), 985. https://doi.org/10.3390/app10030985 (2020).

Chen, H. et al. Characterization and source identification of antibiotic resistance genes in the sediments of an interconnected river-lake system. Environ. Int. 137 , 105538. https://doi.org/10.1016/j.envint.2020.105538 (2020).

Haseena, A. et al. Synthesis of ribose-coated copper-based metal-organic framework for enhanced antibacterial potential of chloramphenicol against multi-drug resistant bacteria. Antibiotics 10 , 12. https://doi.org/10.3390/antibiotics10121469 (2021).

Noohpisheh, Z., Amiri, H., Farhadi, S., Mohammadi-gholami, A. Green synthesis of Ag-ZnO nanocomposites using Trigonella foenum-graecum leaf extract and their antibacterial, antifungal, antioxidant and photocatalytic properties. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 240 , 118595. https://doi.org/10.1016/j.saa.2020.118595 (2020).

Javed, B., Nadhman, A., Razzaq, A. & Mashwani, Z. U. R. One-pot phytosynthesis of nano-silver from Mentha longifolia L.: Their characterization and evaluation of photodynamic potential. Mater. Res. Express 7 , 5. https://doi.org/10.1088/2053-1591/ab903b (2020).

Krishnaraj, C., Ji, B. J., Harper, S. L. & Il Yun, S. Plant extract-mediated biogenic synthesis of silver, manganese dioxide, silver-doped manganese dioxide nanoparticles and their antibacterial activity against food- and water-borne pathogens. Bioprocess Biosyst. Eng. 39 (5), 759–772. https://doi.org/10.1007/s00449-016-1556-2 (2016).

Said, A., Abu-Elghait, M., Atta, H. M., & Salem, S. S. Antibacterial activity of green synthesized silver nanoparticles using Lawsonia inermis against common pathogens from urinary tract infection. Appl. Biochem. Biotechnol. 0123456789. https://doi.org/10.1007/s12010-023-04482-1 (2023).

Lashin, I., Hasanin, M., Hassan, S. A. M. & Hashem, A. H. Green biosynthesis of zinc and selenium oxide nanoparticles using callus extract of Ziziphus spina-christi: Characterization, antimicrobial, and antioxidant activity. Biomass Convers. Biorefinery 13 (11), 10133–10146. https://doi.org/10.1007/s13399-021-01873-4 (2023).

Download references

Acknowledgements

This project is sponsored by Prince Sattam Bin Abdulaziz University (PSAU) as part of funding for its SDG Roadmap Research Funding Program Project number PSAU-2023/SDG/44.

Author information

Authors and affiliations.

Department of Chemistry College of Science and Humanities in Al-Kharj, Prince Sattam Bin Abdulaziz University, 11942, Al-Kharj, Saudi Arabia

Ayman K. El-Sawaf

Department of Chemistry, Faculty of Science, Menoufia University, Shebin El-Kom, Egypt

Department of Bioprocess Development, Genetic Engineering and Biotechnology Research Institute (GEBRI), City of Scientific Research and Technological Applications (SRTA-City), New Borg Al-Arab City 21934, Alexandria, Egypt

Shahira H. El-Moslamy

Polymeric Materials Research Department, Advance Technology and New Materials Research Institute (ATNMRI), City of Scientific Research and Technological Applications (SRTA-City), New Borg Al-Arab City, Alexandria, 21934, Egypt

Elbadawy A. Kamoun

Department of Environmental Science, Asutosh College, University of Calcutta, 92 Shyama Prasad Mukherjee Rd, Jatin Das Park, Bhowanipore, Kolkata, W.B., India

Kaizar Hossain

You can also search for this author in PubMed   Google Scholar

Contributions

Shahira H. EL-Moslamy: Experiments, data acquisition and analysis, interpretation of data, and wrote the original draft; Elbadawy A. Kamoun: Design of the work, draft revision, revised the final draft. Ayman El-Sawaf: Characterization and funding; and Kaizar Hossain: Data acquisition and analysis. All authors have critically reviewed and approved the final draft and are responsible for the content and similarity index of the manuscript.

Corresponding author

Correspondence to Elbadawy A. Kamoun .

Ethics declarations

Competing interests.

The authors declare no competing interests.

Additional information

Publisher's note.

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/ .

Reprints and permissions

About this article

Cite this article.

El-Sawaf, A.K., El-Moslamy, S.H., Kamoun, E.A. et al. Green synthesis of trimetallic CuO/Ag/ZnO nanocomposite using Ziziphus spina-christi plant extract: characterization, statistically experimental designs, and antimicrobial assessment. Sci Rep 14 , 19718 (2024). https://doi.org/10.1038/s41598-024-67579-5

Download citation

Received : 04 May 2024

Accepted : 12 July 2024

Published : 24 August 2024

DOI : https://doi.org/10.1038/s41598-024-67579-5

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

  • Ziziphus spina christi leaves
  • Trimetallic nanocomposite
  • Green-synthesized ZnO
  • Antimicrobial activities

By submitting a comment you agree to abide by our Terms and Community Guidelines . If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

Quick links

  • Explore articles by subject
  • Guide to authors
  • Editorial policies

Sign up for the Nature Briefing: Microbiology newsletter — what matters in microbiology research, free to your inbox weekly.

experimental study of biodiesel synthesis

experimental study of biodiesel synthesis

Maintenance work is planned from 21:00 BST on Tuesday 20th August 2024 to 21:00 BST on Wednesday 21st August 2024, and on Thursday 29th August 2024 from 11:00 to 12:00 BST.

During this time the performance of our website may be affected - searches may run slowly, some pages may be temporarily unavailable, and you may be unable to log in or to access content. If this happens, please try refreshing your web browser or try waiting two to three minutes before trying again.

We apologise for any inconvenience this might cause and thank you for your patience.

experimental study of biodiesel synthesis

Physical Chemistry Chemical Physics

Computational study on the mechanism for the synthesis of the active pharmaceutical ingredients nitrofurantoin and dantrolene in both solution and mechanochemical conditions.

A combination of density functional theory (DFT) calculations and microkinetic simulations is applied to the study of the condensation in acidic media between N-acyl-hydrazides and aldehydes to produce the active pharmaceutical ingredients (API) nitrofurantoin and dantrolene. Previous experimental reports by some of us had shown that the use of ball milling conditions led to a reduction in reaction time which came associated with significant reduction of waste. This result is reproduced by the current calculations, which, moreover, provide a detailed mechanistic explanation for this behavior.

  • This article is part of the themed collection: Fundamental Basis of Mechanochemical Reactivity

Supplementary files

  • Supplementary information PDF (777K)

Article information

Download citation, permissions.

experimental study of biodiesel synthesis

D. M. Galeas, I. Tolbatov, E. Colacino and F. Maseras, Phys. Chem. Chem. Phys. , 2024, Accepted Manuscript , DOI: 10.1039/D4CP01613K

To request permission to reproduce material from this article, please go to the Copyright Clearance Center request page .

If you are an author contributing to an RSC publication, you do not need to request permission provided correct acknowledgement is given.

If you are the author of this article, you do not need to request permission to reproduce figures and diagrams provided correct acknowledgement is given. If you want to reproduce the whole article in a third-party publication (excluding your thesis/dissertation for which permission is not required) please go to the Copyright Clearance Center request page .

Read more about how to correctly acknowledge RSC content .

Social activity

Search articles by author.

This article has not yet been cited.

Advertisements

Academia.edu no longer supports Internet Explorer.

To browse Academia.edu and the wider internet faster and more securely, please take a few seconds to  upgrade your browser .

Enter the email address you signed up with and we'll email you a reset link.

  • We're Hiring!
  • Help Center

paper cover thumbnail

An Experimental Study of Optimization of Biodiesel Synthesis from Waste Cooking Oil and Effect of the Combustion Duration on Engine Performance

Profile image of Seifallah Fetni

The use of waste cooking oil (WCO) as a reagent for biodiesel synthesis ensures their transformation from harmful products into beneficial ones. The possibility of safe use as pure fuels or mixtures with diesel is a promoting and an environmental friendly alternative. This strategy is very encouraging especially for countries which have not enough space to produce vegetable oils. However, the researches in this field (WCO biodiesels) are still rare. In this work, we have synthesized biodiesel from WCO using the transesterification technique, then experimental investigations have been carried out on a four cylinder-direct injection diesel, engine equipped with a turbocharger on a test bench, according to the International norm ISO 27.020. In a first time, effects of different blends of methyl-ester/diesel in different proportions (B00, B10, B20 and B30) on engine behavior were studied and compared with petroleum diesel. In a second time, B20 blend was investigated but with variation of injection timing compared to original settings (as set by the engine manufacturer), on the same engine and following the same testing procedure. Experimental results showed that engine performances decreased with increasing amount of methyl ester in the fuel mixture. Moreover, it is found that advanced injecting B20 fuel by 2 crank angle degrees compared to that of the original injection timing, gives better performance without penalty on pollutant emissions (smoke opacity). The use of B20 accompanied with the advanced injection timing lead to a significant power increase (up to 4%) as well as an increase in torque (up to 2.8%) on conventional diesel engines compared to diesel. Emissions such as Smoke opacity remained close to the original values (without variation of injection timing).

Related Papers

IOSR Journals

The present work investigates the engine performance parameters and emissions characteristics for in-line four cylinders four stroke direct injection diesel engine using biodiesel blends without any engine modifications. A two fuel blends samples and efficient diesel are used, neat diesel (100% diesel fuel), B20 (20% biodiesel and 80% diesel fuel), and B50 (50% biodiesel and 50% diesel fuel) respectively. Engine performance test carried out at full load, with variable speeds ranging from 1000 to 2600 rpm at an interval of 200 rpm. The engine emission was measured in all tests. The study results indicated that there has been a decrease in the brake thermal efficiency by (3.5768%) for the B20 and (7.671%) for the B50. The increase in specific fuel consumption was (6.243%) for the B20 and (13.1257%) for the B50 over the entire studied speed range compared to neat diesel fuel. The engine exhaust gas emissions measures declared that a higher CO by (4.069%) for the B20 and (7.303%) for the B50. HC increased by (5.295%) for the B20 and (12.594%) for the B50. Lower CO2 was obtained by (3.56%) for the B20 and (7.778%) for the B50 emissions, and lower exhaust gas temperature by (3.163%) for the B20 and (6.369%) for the B50 compared to diesel fuel. Therefore, it can be concluded that B20 and B50 can be used in diesel engines without any engine modifications as an alternative petroleum diesel fuel.

experimental study of biodiesel synthesis

Journal of the Japan Institute of Energy

Selwin Rajadurai

Alagumalai Avinash

vikash choudhary

In the present experimental research work, used vegetable oil methyl ester (UVOME) is produced through transesterification of used vegetable oil using methanol in the presence of two different catalyst such as sodium hydroxide (NaOH) and Potassium hydroxide (KOH). Experimental investigations have been carried out to examine the combustion and emission characteristics in a direct injection transportation diesel engine running with

Lectito Journals , Vinayak Gaitonde

The compression ignition (CI) engines are most efficient and robust but they rely on depleting fossil fuel. Hence there is a speedy need to use alternative fuels that replaces diesel and at the same time engine should yield better performance. Accordingly, honge oil methyl ester (BHO) and cotton seed oil methyl ester (BCO) were selected as an alternative fuel to power CI engine in the study. In the first part, this paper aims to evaluate best fuel injection timing (IT) and injector opening pressure (IOP) for the biodiesel fuels (BDF). The combustion chamber (CC) used for the study is toriodal re-entrant (TRCC). The experimental tests showed that BHO and BCO yielded overall better performance at IT of 19° before top dead centre (bTDC) and IOP of 240 bar. In the second part, the effect of number of holes on the performance of BDF powered CI engine was studied keeping optimized IT and IOP. The six-hole injector with 0.2 mm injector orifice diameter yielded better performance compared to other injectors of different holes and size tested.

International journal of Innovative Research in Science, Engineering and Technology

Dr. Vishnu Sankar

Alternative fuels have received much attention due to the depletion of world petroleum reserves and increased environmental concerns. The desire to reach higher efficiencies, lower specific fuel consumption and reduced emissions in modern engines has become the primary focus of engine researchers and manufacturers over the past three decades. Thus processed form of waste cooking oil (Biodiesel) offers attractive green alternative fuels to compression ignition engines. Biodiesel used in the experiment is a methyl ester of free fatty acid made from waste cooking oil (WCO).The fuel properties of biodiesel are very similar to the diesel fuel, so it can work in existing infrastructure of conventional diesel engine without any modification in the engine.The present work investigates and compares the engine performance parameters such as brake power and brake specific fuel consumption and emission characteristics such as CO, CO 2, HC and NO x emissions of direct injection Kirloskar diesel engine using various blends of waste cooking oil biodiesel and diesel. The biodiesel blends B10, B15, B20, B25, B35, B50, B75 and B100 were tested and B20 was found as the optimum blend based on the experimental results as it had properties comparable to diesel and lower emissions than diesel. From the results of investigation it was found that, there has been a decrease in brake power with an increase in brake specific fuel consumption for all blends of biodiesel over the entire load range when compared to the diesel fuel. In the case of engine exhaust gas emissions; lower HC, CO and higher CO 2 and NO x emissions have been found for all biodiesel blends when compared to diesel. Moreover, reductions in sound level for all biodiesel blends have been observed when compared to diesel.B20 was found as the best fuel as it showed lower emissions, better properties and economy than diesel.

RENEWABLE ENERGY SOURCES AND TECHNOLOGIES

D. Prasanth D. Prasanth

International Journal of Engineering Research and Technology (IJERT)

IJERT Journal

https://www.ijert.org/experimental-investigation-of-castrol-oil-methyl-esters-as-biodiesel-on-compression-ignition-engine https://www.ijert.org/research/experimental-investigation-of-castrol-oil-methyl-esters-as-biodiesel-on-compression-ignition-engine-IJERTV2IS2105.pdf The methyl esters of vegetable oils, known as biodiesel are becoming increasingly popular because of their low environmental impact and potential as a green alternative fuel for diesel engine and they would not require significant modification of existing engine hardware. Methyl ester of castor (CME) derived through transesterification process. Experimental investigations have been carried out to examine properties, performance and emissions of different blends (B10, B20, and B40) of castor in comparison to diesel. The use of alternatives for the fossil fuels, such as castor, jatropha, pongamia, etc. is very essential. It has been found that biodiesel plays an important role in the automobile industry now a day. This work aims at reducing the cost of the fuel consumed by blending the biodiesel from castor oil with diesel with different proportions and testing the performance of blended diesel. Initially the engine has to run by diesel and the following characteristics of Brake Power, Total Fuel Consumption, Indicated Power, Mechanical Efficiency, Brake Thermal Efficiency, and Volumetric Efficiency are calculated. Then same procedure is followed for biodiesel blend with varying the proportion by 20%, 40%, 70%, and then by 100% biodiesel. The Characteristics are obtained and compared with diesel in a graph. Based on the performance analysis the efficiency of B-40 just decreases by 2.62% then of pure diesel, which gives an inference that the biodiesel is an acceptable alternative fuel.

Energy Conversion and Management

K. Antonopoulos , Dimitrios Rakopoulos , Evangelos Giakoumis

The increased demand for renewable energy sources and developing countries like India’s need to secure its energy supply has spurred interest in development of bio fuel production whereas the exhaust emission of the biodiesel is deteriorating the environment also. The aim of the research is analyze the emission characteristics of used vegetable oil methyl ester (UVOME) and its blends. Used vegetable oil methyl ester is derived through transesterification process in the presence of two different catalyst such as Sodium hydroxide (NaOH) or Potassium hydroxide (KOH). A single cylinder, water cooled, four stroke diesel engine was used for this work. The following fuels were tested such as diesel, B20N, B20K, (where K and N are denoted as the catalyst KOH and NaOH respectively) and observed the exhaust emission characteristics in terms of concentration of NOx, CO, HC, particulate matter and smoke density. The obtained results of diesel, used vegetable oil methyl esters and their blends with diesel by volume were compared. Compared to diesel, the emission of Carbon monoxide (CO) and Hydro Carbon (HC) emission are reduced with increase in % of BK and BN blend. Compared to B20N, the emission of HC is slightly increased in B20K. The value of Oxides of Nitrogen (NOx) is increased in B20N and B20K. Compared to B20N the emission of NOx is slightly higher in B20K. However the Particulate matter and smoke density of B20 blends are lower than the diesel; among B20N and B20K, the B20N has the lower values

Loading Preview

Sorry, preview is currently unavailable. You can download the paper by clicking the button above.

RELATED PAPERS

Int. J. Recent …

kanit wattanavichien

Pacific-Asia Conference on Mechanical Engineering

Carlos Zapanta , Elexis Edmond Lauzon

Kamal Kumar

IJAERS Journal

Sendilvelan Subramanian

Applied Energy

Dr J.M.BABU

bambang sudarmanta

jiraphon srisertpol

G. Venkata subbaiah

Dzevad Bibic , Boran Pikula

International Journal of Energy Research

C Venkataramana Reddy

Energy & Fuels

Mustafa Canakci

European Journal of Sustainable Development Research

Nagaraj Banapurmath

International Journal of Trend in Scientific Research and Development

Nitin Tenguria

AEJ - Alexandria Engineering Journal

Medhat Nemitallah

puneet verma

IJIRST - International Journal for Innovative Research in Science and Technology

Fluid Dynamics &amp; Materials Processing

Prabhahar Muthuswamy

Mohammed Saber Gad

Dimitrios Rakopoulos

ATANU K PAUL

Jitendra Malik

GJESR Journal

RELATED TOPICS

  •   We're Hiring!
  •   Help Center
  • Find new research papers in:
  • Health Sciences
  • Earth Sciences
  • Cognitive Science
  • Mathematics
  • Computer Science
  • Academia ©2024

IMAGES

  1. 4: Detailed experimental scheme for the synthesis of biodiesel from

    experimental study of biodiesel synthesis

  2. Schematic view of the experimental setup used in the synthesis of

    experimental study of biodiesel synthesis

  3. Schematic view of the experimental setup used in the synthesis of

    experimental study of biodiesel synthesis

  4. (a) Schematic representation of experimental setup used for synthesis

    experimental study of biodiesel synthesis

  5. Experiment 11- Synthesis of Biodiesel and Soap

    experimental study of biodiesel synthesis

  6. Experimental scheme for biodiesel production.

    experimental study of biodiesel synthesis

COMMENTS

  1. Modeling and simulation of biodiesel synthesis in fixed bed and packed

    In this study, modeling and simulation of biodiesel synthesis through transesterification of triglyceride (TG) over a heterogeneous catalyst in a packed bed membrane reactor (PBMR) was performed ...

  2. Experimental analysis of biodiesel synthesis from palm ...

    This work studied the transesterification reaction of palm kernel oil to produce Biodiesel FAME, using as catalyst KOH incorporated as a potassium methoxide intermediate. The catalytic tests were performed modifying representative variables such as reaction temperature (°C), methanol/oil molar ratio, and catalyst content (%KOH). The experimental data were adjusted to a linear empirical model ...

  3. Microwave-assisted biodiesel production using bio-waste ...

    The preparation and activities of different catalysts used for biodiesel synthesis are presented in Table 1, whereas the complete experimental biodiesel yields using the optimized catalyst (BP-SO ...

  4. Process optimization and simulation of biodiesel synthesis ...

    This study reports optimization and simulation of biodiesel synthesis from waste cooking oil through supercritical transesterification reaction without the use of any catalyst. Although the catalyst enhances the reaction rate but due to the presence of water contents in waste cooking oil, the use of catalyst could cause a negative impact on the ...

  5. Optimization and kinetics studies of biodiesel synthesis from

    This study attempts to synthesize biodiesel as a green liquid fuel from Jatropha curcas oil (JCO) utilizing waste eggshell (WES) as an effective and excellent sustainable source of the heterogeneous catalyst under the application of environmentally benign microwave heating technique. After preparing the CaO-based catalyst, diverse characterization techniques such as X-Ray Diffraction, Energy ...

  6. Experimental parametric analysis of biodiesel synthesis in

    This study investigates biodiesel synthesis from waste cooking oil (WCO) using a microreactor completely 3D printed in metal. Production of biodiesel from WCO will inevitably lead to higher levels of free fatty acids (FFAs), since a hydrolysis reaction occurs with the water from food, turning triglycerides into FFAs when the refined vegetable oil is used for cooking. To overcome the negative ...

  7. Towards rapid and sustainable synthesis of biodiesel: A review of

    In a microwave irradiation study conducted by Panadare and Rathod [59], the addition of 0.2 vol/vol% water to the Novozym 435-catalysed biodiesel synthesis increased the conversion from 51 to 87.5%. Excess water was reported to cause lower product yields owing to the unfavourable ester hydrolysis reaction, where biodiesel is converted to FFA.

  8. Development of microreactors applied on biodiesel synthesis: From

    Our research group has been developing studies covering the modeling and simulation of biodiesel synthesis in microscale. The influences of geometry design, mixing degree, flow rate and process variables (temperature, alcohol/oil molar ratio and catalyst concentration) over the reaction efficiency were evaluated applying computational fluid ...

  9. PDF Modeling and simulation of biodiesel synthesis in fixed bed ...

    In this study, modeling and simulation of biodiesel synthesis through transesterication of triglyceride (TG) over a heterogeneous catalyst in a packed bed membrane reactor (PBMR) was performed using a

  10. An experimental study on synthesis of ternary biodiesel through

    The novelty of this study is synthesis fuel of biodiesel from the Brassica napus seed oil (BN-B), Linum usitatissimum seed oil (LU-B) and diesel to improve the performance of SCDE and minimize the CO 2 emissions. BN-B and LU-B were produced using a 0.1 N potassium hydroxide catalyst and methanol to synthesis biodiesel from B. napus and L. usitatissimum seed oil.

  11. Experimental investigation of quaternary blends of diesel/hybrid

    This research thus fills a crucial gap by providing new insights into the potential of hybrid biodiesel blends to enhance engine performance and reduce emissions, contributing to more sustainable and efficient fuel use in the transportation sector. In the current study, biodiesel is produced from hybrid oil using a homogeneous catalyst.

  12. Transesterification reaction of sunflower oil and ethanol for biodiesel

    Biodiesel synthesis through vegetable sunflower oil and ethanol using sodium hydroxide as catalyst was carried out using experimental and numerical approaches. In the first one, biodiesel synthesis was performed in a batch system and in a microreactor, in a range of operating conditions (temperature, ethanol/oil ratio and catalyst concentration).

  13. Synthesis of Biodiesel in Low-Grade Palm Oil using Geopolymer-ZnO

    This study aimed to synthesize biodiesel (fatty acid methyl ester) from low-grade palm oil using geopolymer-ZnO catalyst. The activity of catalyst was tested by mixing low-grade palm oil and methanol in a mole ratio of 1:10, with varying catalyst concentrations of 1%, 3%, and 5% at a temperature of 67 o C and different time intervals.

  14. (PDF) An Experimental Study of Optimization of Biodiesel Synthesis from

    In this work, we have synthesized biodiesel from WCO using the transesterification technique, then experimental investigations have been carried out on a four cylinder-direct injection diesel ...

  15. Synthesis, testing, and evaluation of efficiency and emissions

    Seela et al. (2019) performed an experimental study on a diesel engine using blends of Mahua biodiesel with CeO 2 nanoscaled particles, investigating CeO 2 nanoparticles' impact on engine efficiency and emissions. ... Synthesis of tamarind biodiesel. The process commences with the acquisition of tamarind oil, obtained as the feedstock for ...

  16. Recent advances in transesterification for sustainable biodiesel

    Experimental conditions Kinetic studies References; Temperature (°C) M:O molar ratio Mixing speed (rpm) Kinetic model rate constant (k ... Haryanto A, Gita AC, Saputra TW, Telaumbanua M. First order kinetics of biodiesel synthesis using used frying oil through transesterification reaction. Aceh Int J Sci Technol. 2020; 9:1-11. doi: 10.13170 ...

  17. Kinetic Modeling Studies of Heterogeneously Catalyzed Biodiesel

    The heterogeneously catalyzed transesterification reaction for the production of biodiesel from triglycerides was investigated for reaction mechanism and kinetic constants. Three elementary reaction mechanisms Eley−Rideal (ER), Langmuir−Hinshelwood−Hougen−Watson (LHHW), and Hattori with assumptions, such as quasi-steady-state conditions for the surface species and methanol adsorption ...

  18. An Experimental Study of Optimization of Biodiesel Synthesis from Waste

    Biodiesel blends were varied from 15% to 40% (B15 to B40) by a step of 5%. Engine experiment results revealed a decrease in power and torque by 5%, for each addition of 10 % in biodiesel to the blend. In addition to that, fuel consumption was slightly increased (up to 6% for each 10% of biodiesel blend added) compared to pure diesel fuel.

  19. Biodiesel Production from Locally Sourced Restaurant Waste Cooking Oil

    Biodiesel was synthesized from locally sourced, on-campus, dining facility waste cooking oil and grease by base-catalyzed transesterification with methanol. The components and properties of the biodiesel were characterized by gas chromatography-mass spectrometry (GC-MS), Fourier transform infrared spectroscopy (FT-IR), nuclear magnetic resonance spectroscopy (NMR (1H and 13C)), inductively ...

  20. Statistical optimization of biodiesel synthesis from waste cooking oil

    The central composite design of the response surface methodology was also used to optimize the biodiesel production in order to find the best candidate for biodiesel synthesis. As a result, the experimental research will be in accordance with institutional and international requirements. 2. Materials and methods2.1. Materials

  21. Synthesis and Determination of Biodiesel: An Experiment ...

    Biodiesel production is a promising and important field of research because the relevance it gains from the rising petroleum price and its environmental advantages. This paper reviews the history ...

  22. Energies

    The objective of the research delineated in this paper was to ascertain select physicochemical attributes of second-generation biodiesel, derived from spent frying oil, as well as mixtures of this biodiesel with diesel and biodiesel concentrations of 10, 20, and 30% (v/v). The biodiesel produced is the waste frying oil methyl esters WFOME.

  23. RSM-based comparative experimental study of sustainable biodiesel

    Today's demand of energy in the world of automobile provokes the researchers to strive for the easily available and cheapest renewable source of energy. Biodiesel has become one of the booming renewable sources in the world to mitigate the atmospheric pollution and the demand of fossil fuels. Oils are chosen based on their fatty acid content, availability and sustainability. A magnetic ...

  24. Influence of Structure on Supercapacitors of Co-Vanadate-Based

    This paper reports the synthesis of different crystalline phase-pure Co-vanadates (CVOs), CoV2O6, Co2V2O7, and Co3V2O8, in the same structure. Electrochemical performance experiments were conducted, and density functional theory calculation was used to study their crystalline properties. This paper finds that the supercapacitor performance of CVOs largely depends on the crystalline structure ...

  25. The Synthesis of Biodiesel from Vegetable Oil

    The synthesis of biodiesel by transesterification of vegetable oils was carried out in this study. Two varieties of oils are used in this work, the first type is the waste oils used in frying and the second are olive-pomace oils. Waste oil residue becomes harmful to the environment.

  26. Retrofit of a Marine Engine to Dual-Fuel Methane-Diesel: Experimental

    Shipping is a highly energy-intensive sector, and fleet decarbonization initiatives can significantly reduce greenhouse gas emissions. In the short-to-medium term, internal combustion engines will continue to be used for propulsion or as electricity generators onboard ships. Natural gas is an effective solution which can be used to mitigate greenhouse gas emissions from the marine sector ...

  27. Green synthesis of trimetallic CuO/Ag/ZnO nanocomposite using ...

    In this study, Ziziphus spina christi leaves was used to synthesize a trimetallic CuO/Ag/ZnO nanocomposite by a simple and green method. Many characterizations e.g. FTIR, UV-vis DRS, SEM-EDX ...

  28. Computational study on the mechanism for the synthesis of the active

    A combination of density functional theory (DFT) calculations and microkinetic simulations is applied to the study of the condensation in acidic media between N-acyl-hydrazides and aldehydes to produce the active pharmaceutical ingredients (API) nitrofurantoin and dantrolene. Previous experimental reports by Fundamental Basis of Mechanochemical Reactivity

  29. (PDF) An Experimental Study of Optimization of Biodiesel Synthesis from

    In the present experimental research work, used vegetable oil methyl ester (UVOME) is produced through transesterification of used vegetable oil using methanol in the presence of two different catalyst such as sodium hydroxide (NaOH) and Potassium hydroxide (KOH). ... 2019 An Experimental Study of Optimization of Biodiesel Synthesis from Waste ...

  30. Biodiesel synthesis from waste coconut scum oil ...

    This study aims to develop a novel and efficient magnetic nanocatalyst for producing biodiesel from waste coconut scum oil (WCSO). In this regard, a retrievable and robust nanocatalyst, SnFe 2 O 4 /biochar derived from cigarette butts, was synthesized and applied in the transesterification of WCSO under ultrasonication. The aforementioned nanocatalyst was synthesized by sol-gel technique.