SciELO - Scientific Electronic Library Online

 
vol.88 issue218Product development process of a digital irrigameter: a water sizing proposal for sprinkler irrigation process in different culturesOn the resource consumption of Software quantum computing simulators author indexsubject indexarticles search
Home Pagealphabetic serial listing  

Services on Demand

Journal

Article

Indicators

Related links

  • On index processCited by Google
  • Have no similar articlesSimilars in SciELO
  • On index processSimilars in Google

Share


DYNA

Print version ISSN 0012-7353On-line version ISSN 2346-2183

Dyna rev.fac.nac.minas vol.88 no.218 Medellín July/Sept. 2021  Epub Mar 17, 2022

https://doi.org/10.15446/dyna.v88n218.90154 

Articles

Influence of fractional distillation on the yield and quality of biofuels obtained through thermal catalytic cracking of crude palm oil

Influencia de la destilación fraccionada en el rendimiento y la calidad de los biocombustibles obtenidos mediante el craqueo térmico catalítico del aceite de palma crudo

Silvio Alex Pereira da Mota a   b  
http://orcid.org/0000-0003-0823-4779

Andréia de Andrade Mancio da Mota a   b  
http://orcid.org/0000-0002-7565-6480

Nélio Teixeira Machado a   c  
http://orcid.org/0000-0001-9800-3565

a Programa de Pós-Graduação em Engenharia de Recursos Naturais da Amazônia, Faculdade de Engenharia Química, Universidade Federal do Pará, Belém, Brasil. andreia@ufpa.br, machado@ufpa.br

b Programa de Pós-Graduação em Química/ICE/UNIFESSPA, Marabá, Brasil. silviomota@unifesspa.edu.br, andreia.ammota@gmail.com

c Faculdade de Engenharia Sanitária e Ambiental, Instituto de Tecnologia, Universidade Federal do Pará, Belém, Brasil. machado@ufpa.br


Abstract

This paper investigates the influence of the fractional distillation on the yield and quality of biofuels obtained in a laboratory unit (Vigreux column) at atmospheric pressure, producing three distilled fractions: (1) green gasoline, (2) green aviation kerosene, and (3) green diesel. The quality of the distilled fractions was evaluated through physical-chemical analysis, FTIR spectroscopy, and GC-MS analysis. The fractional distillation of the crude biofuel provided biofuels in the form of distilled fractions with most values of physical-chemical properties within the limits established by national and international regulatory agencies and with experimental distillation curves similar to standard distillation curves. GC-MS analysis showed that the three distilled fractions had higher contents of hydrocarbons than oxygenated compounds and contained hydrocarbons characteristic of petroleum derivatives such as gasoline, aviation kerosene, and diesel.

Keywords: pilot scale; bio-oil; distilled fractions

Resumen

Este trabajo investiga la influencia de la destilación fraccionada en el rendimiento y la calidad de los biocombustibles obtenidos en una unidad de laboratorio (columna Vigreux) a presión atmosférica, produciendo tres fracciones destiladas: (1) gasolina verde, (2) queroseno verde de aviación y (3) diesel verde. La calidad de las fracciones destiladas se evaluó mediante análisis físico-químico, espectroscopia FTIR y análisis GC-MS. La destilación fraccionada del biocombustible crudo proporcionó biocombustibles en forma de fracciones destiladas con la mayoría de valores de propiedades físico-químicas dentro de los límites establecidos por las agencias reguladoras nacionales e internacionales y con curvas de destilación experimentales similares a las curvas de destilación estándar. El análisis GC-MS mostró que las tres fracciones destiladas tenían mayores contenidos de hidrocarburos que los compuestos oxigenados y contenían hidrocarburos característicos de derivados del petróleo como gasolina, queroseno de aviación y diesel.

Palabras clave: escala piloto; bio-aceite; fracciones destiladas

1. Introduction

The concern to reduce dependence on fossil fuels, as well as to reduce pollution caused by the emission of greenhouse gases (GHGs) has stimulated the development of studies on the production of fuels from renewable resources [1,2]. In this sense, biofuels are considered promising alternatives, as they have a chemical composition similar to traditional fossil fuels and low pollutant emissions [2,3].

Biofuels can be obtained from different sources of biomass, including cellulosic biomass, biomass derived from sugar and starch, and biomass based on triglycerides [1,2]. Biomass based on triglycerides includes edible vegetable oils [4,5], used vegetable oils [6], animal fat [7], and non-edible vegetable oils [8,9], and can be converted into biofuels via thermal cracking [10,11] or catalytic cracking (also known as thermal catalytic cracking) [12-20], which is a simple and economical method with products that resemble petroleum fuels [21-23].

Several types of biomass based on triglycerides have been used for the production of biofuels through cracking [4-6,8,14,24]. Among the biomasses used for this purpose, it is possible to highlight the palm oil tree (Elaeis guineenses, Jacq.), “dendezeiro” (in Portuguese), which is an oilseed with a high yield of crude palm oil per hectare [25].

Thermal catalytic cracking products of biomass based on triglycerides include gas, organic liquid product (OLP), coke, and water [26-28]. The OLP consists of oxygenated compounds and hydrocarbons, which correspond to the boiling-point range of gasoline, aviation kerosene and diesel [9,27,29-31]. Some components, such as oxygenated compounds, confer undesirable characteristics to the OLP, limiting its use as a direct substitute for liquid transport fuel [32-34]. Therefore, various types of distillation, including atmospheric pressure distillation, has been applied to the fractionation and upgrading of OLP [20,31,33,35-43] in order to provide biofuels with better physical-chemical property values than those presented by the feed or within the limits specified by regulatory agencies of each country [42]. Due to the complex composition of the OLP, its boiling point starts below 100 °C and reaches the range of 250-280 °C [44]. Atmospheric distillation has the advantages of being an uncomplicated and mature separation process widely used in the petrochemical industry, presenting itself as a technical and economically viable way to separate the complex composition of OLP into biofuels similar to gasoline, aviation kerosene and diesel [3].

Capunitan and Capareda [33] performed fractional distillation of bio-oil under atmospheric conditions and reduced pressure (vacuum). The results showed high yields of heavy fractions and a significant reduction in humidity and acid value. The chemical composition analysis showed that the aromatic and oxygenated compounds were distributed in the light and medium fractions (15-20%), while the phenolic compounds were concentrated in the heavy fraction (53%). The distillation process was effective in separating the components and producing a heavy fraction with improved properties and composition and which could be further utilized as feedstock for future upgrading procedures or as a blending material with other liquid fuels. Ferreira et al. [43] investigated the deacidification of OLP by laboratory-scale fractional distillation using columns of different heights, with and without reflux, and at the pilot scale. Biofuels (distillates) showed yields at the laboratory scale with and without reflux ranging from 62.15 to 76.41 wt % and from 71.65 to 89.44 wt %, respectively, and 32.68 wt % at the pilot scale. For pilot-scale distillation experiments, the acid values of the gasoline, kerosene, and light diesel fractions were 0.33, 0.42, and 0.34 mg KOH/g, respectively. The GC-MS results of the OLP showed 92.84% (by area) of hydrocarbons and 7.16% (by area) of oxygenates. The light diesel fraction contained 100% hydrocarbons with an acid value of 0.34 mg KOH/g, proving the feasibility of OLP deacidification by fractional distillation.

In this context, the objective of the present study was to investigate the influence of the fractional distillation process of OLP on yield and quality of biofuels similar to petroleum products such as gasoline, aviation kerosene, and diesel fuel.

2. Materials and methods

2.1. Organic liquid product

In this study, the feedstock used was an OLP obtained by thermal catalytic cracking of crude palm oil at a pilot plant using 20 wt % of sodium carbonate (Na2CO3) as the catalyst, as described by Mota et al. [20]. These authors characterized the OLP using physical-chemical and FTIR analyses.

2.2. Experimental apparatus and procedure

The OLP was distilled into three fractions (green gasoline, green aviation kerosene, and green diesel) using the experimental apparatus described by Mota et al. [20].

OLP was submitted to laboratory-scale fractional distillation using a Vigreux Column with three stages. To obtain three fractions, the OLP was distilled within the following distillation temperature ranges: 90-160 °C (green gasoline); 160-245 °C (green aviation kerosene); and 245-340 °C (green diesel). The distillation temperature intervals were reported to give hydrocarbon fractions with similar properties to particular petroleum products.

The fractional distillation of OLP aimed to obtain the fractions previously defined for physical-chemical characterization and analysis of composition.

The OLP was heated gradually to first distillation temperature range (90-160 °C). The distilled fraction was collected in a 250-mL separation funnel and stored in amber glass bottles. After obtaining the first distilled fraction, the temperature controller of the heating mantle was adjusted to reach the second distillation range (160-245 °C). The same procedure was followed to obtain the remaining fraction (245-340 °C). At the end of the experimental procedure, there were three distilled fractions and a bottoms product.

Eq. (1) was applied to determine the yield of the distilled fractions (green gasoline, green aviation kerosene, and green diesel), as well as the bottoms product.

where m DF is the mass of distilled fraction (g), and m OLP is the mass of OLP (g).

2.3. Characterization of the distilled fractions

2.3.1. Physical-chemical properties

Distilled fractions were physically-chemically characterized according to the official AOCS and ASTM methods for acid value (AOCS Cd 3d-63), saponification value (AOCS Cd 3-25), specific gravity at 20 °C (ASTM D854), refractive index (AOCS Cc 7-25), kinematic viscosity at 40 °C (ASTM 446 and ASTM D2515), flash point (ASTM D93), corrosiveness to copper (ASTM D130), FFA content (AOCS Ca 5a-40), carbon residue (ASTM D4530) and ester value, which is the difference between the saponification value and the acid value, as described by Paquot [45].

2.3.2. FTIR spectroscopy

Distilled fractions were analyzed by FTIR spectroscopy (Shimadzu, model: Prestige 21). The absorbance spectra were obtained within the interval 4000-400 cm−1 at a resolution of 16 cm−1 using a KBr window. The samples were dropped onto the KBr surface by micropipette in order to spread the liquid and produce a uniform layer.

2.3.3. Distillation curve

The distillation curves of the distilled fractions were obtained according to the official method (ABNT/NBR 9619) using an automatic distillation apparatus (Tanaka, model: AD6).

2.3.4. GC-MS analysis

The GC-MS analysis of the distilled fractions was performed using a gas chromatograph coupled to a mass spectrometer (Shimadzu, model: GCMS-QP2010 Plus), as described by Mota et al. [20]. The relative content of compounds in the distilled fractions was calculated by the ratio of their peak area to the total peak area of the GC-MS spectra.

3. Results and discussion

3.1. Influence of fractional distillation on the yield of biofuels

According to Table 1, the OLP produced by thermal catalytic cracking of crude palm oil with 20 wt % of Na2CO3 generated three fractions after being subjected to a fractional distillation process at the following distillation temperature (DT) ranges: 90 ≤ DT ≤ 160 °C (green gasoline), 160 ≤ DT ≤ 245 °C (green aviation kerosene), 245 ≤ DT ≤ 340 °C (green diesel). Table 1 also shows that as the DT range increased, the yield of the distilled fractions increased, indicating that cracking of crude palm oil with Na2CO3 favored the production of heavier fractions such as green aviation kerosene, green diesel, and bottoms product. Weber et al. [36] obtained similar results, performing the thermal degradation of animal fat in a pilot-scale plant at 410-450 °C using a moving bed of sodium carbonate as the catalyst and 5 wt % of water, obtaining OLP. The authors also conducted a fractional distillation of the OLP produced, yielding 66 wt % of diesel fraction and 21 wt % of gasoline fraction.

Table 1 Yield of distilled fractions. 

Source: The authors.

The total yield from the sum of the three fractions (green gasoline, green aviation kerosene, and green diesel) was 60.43 wt %, which is higher than those found in the literature. Chew & Bhatia [46], for example, investigated the effect of additives (HZSM-5 in different Si/Al ratios, beta zeolite, SBA-15, and AlSBA-15) mixed physically with Rare Earth-Y (REY) as a catalyst in the catalytic cracking of crude and used palm oil to produce biofuels. The best results in terms of yield showed that OLP consisted of 59.3 wt % and 55.3 wt % of fractions in the boiling-point ranges of gasoline, kerosene, and diesel oil for crude and used palm oil, respectively, both using HZSM-5 (Si/Al ratio = 40) with REY as catalyst.

The bottom product was the product of the fractional distillation of OLP that showed the highest yield in the present study. The bottom product contains a wide variety of valuable chemicals, making its use possible in resins, agrochemicals, fertilizers, emission control agents [33], fuels [47], carbon anodes, steel carburization, and graphite synthesis [48].

Therefore, the results show that OLP from the thermal catalytic cracking of crude palm oil with 20 wt % Na2CO3 at 450 °C and 1 atm and at pilot scale yielded a higher amount of distilled fraction when compared to cracking OLP using commercial catalysts such as zeolites, and the bottom product formed in the fractional distillation process of the former OLP can find several applications.

3.2. Influence of fractional distillation on the quality of biofuels

3.2.1. Physical-chemical properties

As shown in Table 2, most of the values concerning the physical-chemical properties of the green gasoline fraction were considerably lower than those of the OLP. These values were also lower than those found by Xu et al. [35] and Wisniewski Jr. et al. [40] for gasoline fractions.

Table 2 Physical-chemical properties of the green gasoline fraction. FFA, free fatty acids; max, maximum; min, minimum; ANP: Agência Nacional do Petróleo, Gás Natural e Biocombustíveis, Resolution N° 40 (Specification of regular gasoline/Type A) a viscosity at 20 °C 

Source: The authors.

Table 3 shows that most of the physical-chemical properties of the green aviation kerosene fraction had values equal to or lower than those relating to OLP. Although the acid value did not reach the limit established by the ANP Nº 37 [51] and ASTM D1655 [52], the value obtained for this property was 1.68 mg KOH/g. This value is relatively low and can be further reduced by applying separation processes such as liquid-liquid extraction and adsorption [43,55] or even fractional distillation [43] in order to achieve the values specified by ANP and ASTM.

Table 3 Physical-chemical properties of the green aviation kerosene fraction.FFA, free fatty acids; max, maximum; min, minimum; ANP: Agência Nacional do Petróleo, Gás Natural e Biocombustíveis, Resolution Nº37 (Specification of Aviation kerosene). a Density at 15 °C, kg/m3. b Viscosity at 20 °C, mm2/s (max.) c Copper strip, 2 h at 100 °C (max.) d ASTM Test Method: ASTM D3242 

Source: The authors.

The flash point value was lower than that established by ANP Nº 37 [51] and ASTM D1655 [52], as shown in Table 3. This fact is due to the presence of low-molecular-weight compounds and a lack of an adequate percentage of normal paraffinic and aromatic compounds, resulting in a lower flash point for the green aviation kerosene fraction. The flash point indicates the presence of volatile components in the oil and is used to evaluate the overall flammability hazard of a material. The lower the flash point, the higher the concentration of light hydrocarbons in the material. This fact is corroborated in Section 3.2.4, which quotes the results of chromatographic analysis of the distilled fractions. Also, the specific gravity of the green aviation kerosene was in agreement with the specifications established for aviation kerosene according to ANP Nº 37 (Specification of Aviation Kerosene) [51] and ASTM D1655 (Jet A-1) [52].

Concerning Table 4, most of the values of the physical-chemical properties of the green diesel fraction are higher than those found for the OLP. However, the specific gravity and acid value of this fraction are lower than those found by Xu et al. [35] and Wisniewski Jr. et al. [40]. The values of properties such as specific gravity at 20 °C, kinematic viscosity at 40 °C and corrosiveness to copper of the green diesel fraction are within limits established for the Diesel S10 specification of ANP Nº 65 [53]. Properties such as flash point and acid value were not within limits established by that regulatory agency; however, the flash point was lower than that established by ANP Nº 65 [53] and ASTM D975 [54], which is due to the presence of low-molecular-weight compounds, a lack of an adequate percentage of normal paraffinic and aromatic compounds, as corroborated in Section 3.2.4.

Table 4 Physical-chemical properties of the green diesel fraction. FFA, free fatty acids; max, maximum; min, minimum; ANP, Agência Nacional do Petróleo, Gás Natural e Biocombustíveis, Resolution Nº 65 (Specification of Diesel S10). a Viscosity at 20 °C. b ASTM Test Method: ASTM D524. 

Source: The authors.

The flash point of a fuel is important for safety, as its regulation is necessary for legal and safe handling and storage [56]. The removal of volatile components by upgrading can quickly improve the flash point [39,57,58], making it possible for distillate fractions, especially green aviation kerosene and green diesel, to reach the limits set by regulatory agencies.

The acid value of the green diesel fraction does not reach the limit established by ANP Nº 65 [53], and the value, 5.70 mg KOH/g, is higher than that of the other distilled fractions, a result of the higher concentration of carboxylic acids in the green diesel fraction compared to the other fractions, as indicated in Table 5.

Table 5 Chemical composition of distilled fractions in terms of hydrocarbons and oxygenated compounds. 

Source: The authors.

3.2.2. FTIR spectroscopy

The FTIR spectra of the distilled fractions shown in Fig. 1 indicates the presence of bands characteristic of saturated hydrocarbons, such as characteristic bands of the axial deformation of CH within the region 3000-2840 cm-1, corresponding to hydrocarbons from normal alkanes. Associated with these bands, we verified the presence of a band at 1375 cm-1 relating to angular deformation of methyl group C-H [59], confirming of the presence of aliphatic hydrocarbons. It is also important to note that the spectra also show broad bands in the range 3077-2750 cm-1, associated with the bands 1716 cm-1 and 1722 cm-1, which confirm the presence of carboxylic acids, aldehydes, and ketones.

Source: The authors.

Figure 1 FTIR spectra of OLP and distilled fractions. 

3.2.3. Distillation curve

Figs. 2, 3, and 4 show the distillation curves of the fractions. In Fig. 2, the curve of the green gasoline fraction shows similarity to the standard distillation curve for regular gasoline (Type A) specification of ANP Nº 40 [49]. The curve shows that the DT exceeds the standard values for recovered volumes of 10% and 50%. However, the temperatures for other recovered volumes (90% and 100%) are consistent with the standard distillation curve for regular gasoline (Type A) from petroleum, according to the specification of ANP Nº 40 [49]. Wiggers et al. [31] obtained similar results to those of this study when fractionating soybean bio-oil into light distilled fractions, with similar distillation curves to those of gasoline A from petroleum.

Source: The authors.

Figure 2 Distillation curves: ♦ Green gasoline fraction; Regular gasoline specification (ANP Nº 40). 

Fig. 3 indicates that the green aviation kerosene fraction reached the limits for the distillation curve of aviation kerosene according to ANP Nº 37 [51]. Ertas and Alma [60] obtained simulated distillation curves for the kerosene fraction from the bio-oil of laurel (Laurus nobilis L.) and found that they could use it as fuel when blended with commercial petroleum products.

Source: The authors.

Figure 3 Distillation curves: ♦ Green aviation kerosene fraction; Aviation kerosene specification (ANP Nº 37). 

Fig. 4 indicates that the experimental distillation curve of the green diesel fraction is different from the standard distillation curve in Diesel S10 specification (ANP Nº 65) [62] until a recovered volume of approximately 40%. In the range of 0-40% recovered volume, the DTs are above those established by ANP, indicating that there is an excessive concentration of heavy hydrocarbons in the green diesel fraction when compared to petroleum diesel. In contrast, the temperatures of 50% and 95% recovered volume are consistent with those established for the standard distillation curve. Ertas and Alma obtained simulated distillation curves for the diesel fraction from the bio-oil of laurel (L. nobilis L.) and also concluded that it could be used as fuel when mixed with petroleum products.

Source: The authors.

Figure 4 Distillation curves: ▲ Green diesel fraction; Diesel S10 specification (ANP Nº 65). 

3.2.4. GC-MS analysis

The results of GC-MS analysis are summarized in Fig. 5 and Table 5. Fig. 5 shows the presence of C8-C20 hydrocarbon chains in the three distilled fractions. However, each distilled fraction has a specific hydrocarbon range. The green gasoline fraction was composed of hydrocarbons with chains in the C8-C17 range, with higher concentrations of C10-C12 hydrocarbons. According to Farah [61] and Speight [62], gasoline consists of hydrocarbons ranging from C4 to C12. Therefore, the green gasoline fraction is consistent with the limits established in the literature of the number of carbons present in the hydrocarbon chains.

Source: The authors.

Figure 5 Carbon number distribution of the distilled fractions. 

The green aviation kerosene fraction consisted of C10-C19 hydrocarbons, in which the predominant hydrocarbons were the C11-C16 range. These results are consistent with those described by Speight [62], who reported that kerosene-type jet fuels (JP-4) are characterized by the hydrocarbons in the C4-C16 range. Finally, the green diesel fraction contained hydrocarbons chains in the range C13-C20, with the majority in the C15-C18 range. Speight [62] reports that the carbon number limit for diesel fuel is C8-C18.

According to Table 5, the chemical distributions of the classes of hydrocarbons and oxygenated compounds of the three distilled fractions were very different from those of the OLP produced in a previous study by Mancio et al. [13]. In OLP, hydrocarbons were abundant relative to oxygenated compounds. Similarly, Table 5 also shows that all the biofuels produced in the present study in the form of distilled fractions contained more hydrocarbons than oxygenated compounds. The fractions with the most significant hydrocarbon content were green gasoline and green aviation kerosene, with lower contents of oxygenated compounds in these fractions. Among the identified and quantified hydrocarbons were normal paraffinic, olefinic, naphthenic, and aromatic compounds, which according to Farah [61] and Szklo [56] are the main components present in the distilled fractions of crude oil. Regarding the content of aromatic compounds, the green gasoline fraction follows ANP Nº 40 [49], which states that regular gasoline must not contain more than 25 wt % of aromatic compounds in its composition.

The green diesel fraction contained the highest levels of oxygenated compounds of all three distilled fractions (See Table 5). It is to be expected that some oxygenates are not completely converted into hydrocarbons, since during triglyceride cracking reactions one of the dominant steps is the elimination of heavy oxygenated hydrocarbons such as carboxylic acids, aldehyde, ketones, and esters [63]. Among the oxygenated compounds, alcohols and ketones represent the two largest chemical families in the green diesel fraction. According to Oasmaa [57], the acidity of bio-oils arises from their content of oxygenated compounds, carboxylic acids being the class that promotes higher acidity. For this reason, the green diesel fraction, with 0.98 wt % of carboxylic acids, is the distilled fraction with the highest acid value, as shown in Table 4.

Therefore, the results of the GC-MS analysis clearly show that most of the distilled fractions can be used as energy sources by their application as bio-additives to fuels, due to their high hydrocarbon content and low levels of oxygenates. Such use of distilled fractions as bio-additives may result in emulsification of the fraction with their respective oil derivatives. Besides, the distilled fractions can be used as feedstock for further upgrading process in a crude oil refinery by hydrotreating processes [33].

It is important to note that the green gasoline fraction shows characteristics that could be improved, as naphthenic and aromatic compounds are present in low concentrations. According to Farah [61], these compounds influence the resistance to detonation, since their presence is favorable for the antiknock characteristics of fuels, such as gasoline, used in spark-ignition engines. However, the presence of normal paraffins implies a greater stability or resistance to oxidation, and hence greater durability of the fuel.

The presence of oxygenates in the green gasoline fraction can influence its volatility as well as its energy efficiency. According to Farah [61], oxygenates reduce the volatility of gasoline, improving the time of ignition or burning of fuel. According to Szklo [56], the presence of oxygenated compounds increases the latent heat of vaporization, i.e., the energy required to vaporize the liquid fuel.

Unlike the results obtained for the green gasoline fraction, the presence of aromatic compounds was not identified in the green aviation kerosene or green diesel fractions, resulting in the higher levels of other constituents, as seen in Table 5. According to Farah [61], the presence of paraffinic and naphthenic compounds, as well as the absence of aromatic compounds, in diesel derived from petroleum contributes to a good quality fuel. These characteristics are favorable because the presence of aromatic compounds in fossil fuels promotes a higher resistance to detonation in diesel combustion engines, and the presence of normal (linear) paraffinic promotes greater stability, i.e., resistance to oxidation. On the other hand, the presence of olefins results in a low freezing point and a high cetane number. The high content of olefins can be reduced by hydrogenation: typically, simple hydrogenation or non-destructive hydrogenation, which are used to improve product quality without causing a significant change in boiling-point range [58].

According to Table 5, the results of this study are similar or superior to those reported in the literature. The hydrocarbon contents of the green gasoline and green aviation kerosene fractions, for example, are significantly higher than those present in the light, middle and heavy fractions obtained by fractional distillation in the temperature ranges of ≤100 °C, 100-180 °C, and 180-250 °C, respectively, as described by Capunitan and Capareda [33]. The green gasoline also showed a hydrocarbon content higher than that obtained by Wisniewski et al. [40], who despite a DT range of <200 °C obtained a light bio-oil fraction with only 60.06 wt % of hydrocarbons from the reactive distillation. On the other hand, the green gasoline showed a hydrocarbon content similar to that obtained by Zhao et al. [64], who obtained a distilled fraction with 91.97% of hydrocarbons from the simple distillation at the similar temperature range of <200 °C. The green aviation kerosene and green diesel fractions obtained in the present study also showed very similar hydrocarbon contents to those described by da Silva Almeida et al. [41] and da Silva Almeida et al. [14] for kerosene and light diesel fractions obtained by fractional distillation at 175-235 °C and 235-305 °C, respectively.

4. Conclusions

The fractional distillation of OLP provides biofuels with most physical-chemical values within limits established by national and international regulatory agencies and with experimental distillation curves similar to standard curves. Besides, a single fractionation of OLP generated biofuels (green gasoline, green aviation kerosene, and green diesel) with higher contents of hydrocarbons than oxygenated compounds and contained hydrocarbons characteristic of the respective petroleum derivatives. Therefore, biofuels have a promising potential for use as novel products in the area of renewable energy, allowing the current consumer market to purchase products that can partially or entirely replace petroleum products, as do bioethanol and biodiesel.

Acknowledgments

This work was financially supported by the Brazilian governmental agencies (CAPES and Eletrobrás S/A).

References

[1] Wako, F.M., Reshad, A.S., Bhalerao, M.S. and Goud, V.V., Catalytic cracking of waste cooking oil for biofuel production using zirconium oxide catalyst, Industrial Crops and Products. 118, pp. 282-289, 2018. DOI: 10.1016/j.indcrop.2018.03.057. [ Links ]

[2] Cao, X., Li, L., Shitao, Y., Liu, S., Hailong, Y., Qiong, W. and Ragauskas, A.J., Catalytic conversion of waste cooking oils for the production of liquid hydrocarbon biofuels using in-situ coating metal oxide on SBA-15 as heterogeneous catalyst, Journal of Analytical and Applied Pyrolysis. 138, pp. 137-144, 2019. DOI: 10.1016/j.jaap.2018.12.017. [ Links ]

[3] Zhang, X.S., Yang, G.X., Jiang, H., Liu, W.J. and Ding, H.S., Mass production of chemicals from biomass-derived oil by directly atmospheric distillation coupled with co-pyrolysis, Scientific Reports. 3, pp. 1-7, 2013. DOI: 10.1038/srep01120. [ Links ]

[4] Tamunaidu, P. and Bhatia, S., Catalytic cracking of palm oil for the production of biofuels: optimization studies, Bioresour Technol. 98, pp. 3593-3601, 2007. DOI: 10.1016/j.biortech.2006.11.028. [ Links ]

[5] Buzetzki, E., Sidorová, K., Cvengrošová, Z. and Cvengroš, J., Effects of oil type on products obtained by cracking of oils and fats, Fuel Processing Technology. 92, pp. 2041-2047, 2011. DOI: 10.1016/j.fuproc.2011.06.005. [ Links ]

[6] Li, L., Quan, K., Xu, J., Liu, F., Liu, S., Yu, S., Xie, C., Zhang, B. and Ge, X., Liquid Hydrocarbon Fuels from Catalytic Cracking of Waste Cooking Oils Using Basic Mesoporous Molecular Sieves K2O/Ba-MCM-41 as Catalysts, ACS Sustainable Chemistry & Engineering. 1, pp. 1412-1416, 2013. DOI: 10.1021/sc4001548. [ Links ]

[7] Hancsók, J., Sági, D. and Valyon, J., Diesel fuel blending components from mixture of waste animal fat and light cycle oil from fluid catalytic cracking, Journal of Environmental Management. 223, pp. 92-100 2018. DOI: 10.1016/j.jenvman.2018.06.011. [ Links ]

[8] Zhao, X., Wei, L., Cheng, S. and Julson, J., Optimization of catalytic cracking process for upgrading camelina oil to hydrocarbon biofuel, Industrial Crops and Products. 77, pp. 516-526, 2015. DOI: 10.1016/j.indcrop.2015.09.019. [ Links ]

[9] Melero, J.A., Iglesias, J. and Garcia, A., Biomass as renewable feedstock in standard refinery units. Feasibility, opportunities and challenges, Energy & Environmental Science. 5, pp. 7393-7420, 2012. DOI: 10.1039/c2ee21231e. [ Links ]

[10] Meier, H.F., Wiggers, V.R., Zonta, G.R., Scharf, D.R., Simionatto, E.L. and Ender, L., A kinetic model for thermal cracking of waste cooking oil based on chemical lumps, Fuel. 144, pp. 50-59, 2015. DOI: 10.1016/j.fuel.2014.12.020. [ Links ]

[11] Seifi, H. and Sadrameli, S.M., Improvement of renewable transportation fuel properties by deoxygenation process using thermal and catalytic cracking of triglycerides and their methyl esters, Applied Thermal Engineering. 100, pp. 1102-1110, 2016. DOI: 10.1016/j.applthermaleng.2016.02.022. [ Links ]

[12] Teixeira, C.M., Fréty, R., Barbosa, C.B.M., Santos, M.R., Bruce, E.D. and Pacheco, J.G.A., Mo influence on the kinetics of jatropha oil cracking over Mo/HZSM-5 catalysts, Catalysis Today. 279, pp. 202-208, 2017. DOI: 10.1016/j.cattod.2016.06.006. [ Links ]

[13] Mancio, A.A., da Costa, K.M.B., Ferreira, C.C., Santos, M.C., Lhamas, D.E.L., da Mota, S.A.P., Leão, R.A.C., de Souza, R.O.M.A., Araújo, M.E., Borges, L.E.P. and Machado, N.T., Thermal catalytic cracking of crude palm oil at pilot scale: Effect of the percentage of Na2CO3 on the quality of biofuels, Industrial Crops and Products. 91, pp. 32-43, 2016. DOI: 10.1016/j.indcrop.2016.06.033. [ Links ]

[14] da Silva Almeida, H., Corrêa, O.A., Eid, J.G., Ribeiro, H.J., de Castro, D.A.R., Pereira, M.S., Pereira, L.M., de Andrade Mâncio, A., Santos, M.C., da Silva Souza, J.A., Borges, L.E.P., Mendonça, N.M. and Machado, N.T., Production of biofuels by thermal catalytic cracking of scum from grease traps in pilot scale, Journal of Analytical and Applied Pyrolysis. 118, pp. 20-33, 2016. DOI: 10.1016/j.jaap.2015.12.019. [ Links ]

[15] Abbasov, V., Mammadova, T., Aliyeva, N., Abbasov, M., Movsumov, N., Joshi, A., Lvov, Y. and Abdullayev, E., Catalytic cracking of vegetable oils and vacuum gasoil with commercial high alumina zeolite and halloysite nanotubes for biofuel production, Fuel. 181, pp. 55-63, 2016. DOI: 10.1016/j.fuel.2016.04.088. [ Links ]

[16] Lovás, P., Hudec, P., Hadvinová, M. and Ház, A., Conversion of rapeseed oil via catalytic cracking: Effect of the ZSM-5 catalyst on the deoxygenation process, Fuel Processing Technology. 134, pp. 223-230, 2015. DOI: 10.1016/j.fuproc.2015.01.038. [ Links ]

[17] Zheng, Z., Wang, J., Wei, Y., Liu, X., Yu, F. and Ji, J., Effect of La-Fe/Si-MCM-41 catalysts and CaO additive on catalytic cracking of soybean oil for biofuel with low aromatics, Journal of Analytical and Applied Pyrolysis. 143, pp. 104693, 2019. DOI: 10.1016/j.jaap.2019.104693. [ Links ]

[18] Yang, X., Li, X., Liu, J. and Rong, L., Ni/phosphomolybdic acid immobilized on carbon nanotubes for catalytic cracking of Jatropha oil, Chemical Physics Letters. 720, pp. 42-51, 2019. DOI: 10.1016/j.cplett.2019.02.008. [ Links ]

[19] Vu, X.H. and Armbruster, U., Catalytic cracking of triglycerides over micro/mesoporous zeolitic composites prepared from ZSM-5 precursors with varying aluminum contents, Reaction Kinetics, Mechanisms and Catalysis. 125, pp. 381-394, 2018. DOI: 10.1007/s11144-018-1415-z. [ Links ]

[20] da Mota, S.A.P., Mancio, A.A., Lhamas, D.E.L., de Abreu, D.H., da Silva, M.S., dos Santos, W.G., de Castro, D.A.R., de Oliveira, R.M., Araújo, M.E., Borges, L.E.P. and Machado, N.T., Production of green diesel by thermal catalytic cracking of crude palm oil (Elaeis guineensis, Jacq) in a pilot plant, Journal of Analytical and Applied Pyrolysis. 110, pp. 1-11, 2014. DOI: 10.1016/j.jaap.2014.06.011. [ Links ]

[21] Zhao, X., Wei, L., Cheng, S., Huang, Y., Yu, Y. and Julson, J., Catalytic cracking of camelina oil for hydrocarbon biofuel over ZSM-5-Zn catalyst, Fuel Processing Technology. 139, pp. 117-126, 2015. DOI: 10.1016/j.fuproc.2015.07.033. [ Links ]

[22] Wang, Y., Cao, Y. and Li, J., Preparation of biofuels with waste cooking oil by fluid catalytic cracking: The effect of catalyst performance on the products, Renewable Energy. 124, pp. 34-39, 2018. DOI: 10.1016/j.renene.2017.08.084. [ Links ]

[23] Zhou, N., Liu, S., Zhang, Y., Fan, L., Cheng, Y., Wang, Y., Liu, Y., Chen, P. and Ruan, R., Silicon carbide foam supported ZSM-5 composite catalyst for microwave-assisted pyrolysis of biomass, Bioresource Technology. 267, pp. 257-264, 2018. DOI: 10.1016/j.biortech.2018.07.007. [ Links ]

[24] Katikaneni, S.P.R. , Adjaye, J.D. and Bakhshi, N.N., Studies on the catalytic conversion of canola oil to hydrocarbons - influence of hybrid catalysts and steam, Energy & Fuels. 9, pp. 599-609, 1995. DOI: 10.1021/ef00052a005. [ Links ]

[25] Kuss, V.V., Kuss, A.V., da Rosa, R.G., Aranda, D.A.G. and Cruz, Y.R., Potential of biodiesel production from palm oil at Brazilian Amazon, Renewable and Sustainable Energy Reviews. 50, pp. 1013-1020, 2015. DOI: 10.1016/j.rser.2015.05.055. [ Links ]

[26] Taufiqurrahmi, N. and Bhatia, S. Catalytic cracking of edible and non-edible oils for the production of biofuels, Energy & Environmental Science. 4, pp. 1087-1112, 2011. DOI: 10.1039/c0ee00460j. [ Links ]

[27] Melero, J.A., García, A. and Clavero, M. Production of biofuels via catalytic cracking, in Luque, R. and Clark, J. Handbook of Biofuels Production, Woodhead Publishing, 2011. pp. 390-419. [ Links ]

[28] Mohanty, P., Pant, K.K., Naik, S.N., Parikh, J., Hornung, A. and Sahu, J.N., Synthesis of green fuels from biogenic waste through thermochemical route - The role of heterogeneous catalyst: A review, Renewable and Sustainable Energy Reviews. 38, pp. 131-153, 2014. DOI: 10.1016/j.rser.2014.05.011. [ Links ]

[29] Wang, S., Guo, Z., Cai, Q. and Guo, L., Catalytic conversion of carboxylic acids in bio-oil for liquid hydrocarbons production, Biomass and Bioenergy. 45, pp. 138-143, 2012. DOI: 10.1016/j.biombioe.2012.05.023. [ Links ]

[30] Yan, S., DiMaggio, C., Wang, H., Mohan, S., Kim, M., Yang, L., Salley, S.O. and Simon Ng, K.Y., Catalytic Conversion of Triglycerides to Liquid Biofuels Through Transesterification, Cracking, and Hydrotreatment Processes, Current Catalysis. 1, pp. 41-51, 2012. [ Links ]

[31] Wiggers, V.R., Meier, H.F., Wisniewski, A., Barros, A.A.C. and Maciel, M.R.W., Biofuels from continuous fast pyrolysis of soybean oil: A pilot plant study, Bioresource Technology. 100, pp. 6570-6577, 2009. DOI: 10.1016/j.biortech.2009.07.059. [ Links ]

[32] Xiu, S.N. and Shahbazi, A., Bio-oil production and upgrading research: A review, Renewable & Sustainable Energy Reviews. 16 (2012) 4406-4414. DOI: 10.1016/j.rser.2012.04.028. [ Links ]

[33] Capunitan, J.A. and Capareda, S.C., Characterization and separation of corn stover bio-oil by fractional distillation, Fuel. 112, pp. 60-73, 2013. DOI: 10.1016/j.fuel.2013.04.079. [ Links ]

[34] Nam, H., Choi, J. and Capareda, S.C., Comparative study of vacuum and fractional distillation using pyrolytic microalgae (Nannochloropsis oculata) bio-oil, Algal Research. 17, pp. 87-96, 2016. DOI: 10.1016/j.algal.2016.04.020. [ Links ]

[35] Xu, J.M., Jiang, J.C., Zhang, T.J. and Dai, W.D., Biofuel Production from Catalytic Cracking of Triglyceride Materials Followed by an Esterification Reaction in a Scale-up Reactor, Energy & Fuels. 27, pp. 255-261, 2013. DOI: 10.1021/ef3018173. [ Links ]

[36] Weber, B., Stadlbauer, E.A., Stengl, S., Hossain, M., Frank, A., Steffens, D., Schlich, E. and Schilling, G., Production of hydrocarbons from fatty acids and animal fat in the presence of water and sodium carbonate: Reactor performance and fuel properties, Fuel. 94, pp. 262-269, 2012. DOI: 10.1016/j.fuel.2011.08.040. [ Links ]

[37] Prado, C.M.R. and Antoniosi Filho, N.R., Production and characterization of the biofuels obtained by thermal cracking and thermal catalytic cracking of vegetable oils, Journal of Analytical and Applied Pyrolysis. 86, pp. 338-347, 2009. DOI: 10.1016/j.jaap.2009.08.005. [ Links ]

[38] Santos, A.L., Martins, D.U., Iha, O.K., Ribeiro, R.A., Quirino, R.L. and Suarez, P.A., Agro-industrial residues as low-price feedstock for diesel-like fuel production by thermal cracking, Bioresour Technol. 101, pp. 6157-6162, 2010. DOI: 10.1016/j.biortech.2010.02.100. [ Links ]

[39] Varuvel, E.G., Mrad, N., Tazerout, M. and Aloui, F., Assessment of liquid fuel (bio-oil) production from waste fish fat and utilization in diesel engine, Applied Energy. 100, pp. 249-257, 2012. DOI: 10.1016/j.apenergy.2012.05.035. [ Links ]

[40] Wisniewski, A., Wosniak, L., Scharf, D.R., Wiggers, V.R., Meier, H.F. and Simionatto, E.L., Upgrade of Biofuels Obtained from Waste Fish Oil Pyrolysis by Reactive Distillation, Journal of the Brazilian Chemical Society. 26, pp. 224-232, 2015. DOI: 10.5935/0103-5053.20140251. [ Links ]

[41] da Silva Almeida, H., Corrêa, O.A., Eid, J.G., Ribeiro, H.J., de Castro, D.A.R., Pereira, M.S., Pereira, L.M., de Andrade Aâncio, A., Santos, M.C., da Mota, S.A.P., da Silva Souza, J.A., Borges, L.E.P., Mendonça, N.M. and Machado, N.T., Performance of thermochemical conversion of fat, oils, and grease into kerosene-like hydrocarbons in different production scales, Journal of Analytical and Applied Pyrolysis. 120 (2016) 126-143. DOI: 10.1016/j.jaap.2016.04.017. [ Links ]

[42] Mancio, A.A., da Mota, S.A.P., Ferreira, C.C., Carvalho, T.U.S., Neto, O.S., Zamian, J.R., Araújo, M.E., Borges, L.E.P. and Machado, N.T., Separation and characterization of biofuels in the jet fuel and diesel fuel ranges by fractional distillation of organic liquid products, Fuel. 215, pp. 212-225, 2018. DOI: 10.1016/j.fuel.2017.11.029. [ Links ]

[43] Ferreira, C.C., Costa, E.C., de Castro, D.A.R., Pereira, M.S., Mâncio, A.A., Santos, M.C., Lhamas, D.E.L., da Mota, S.A.P., Leão, A.C., Duvoisin, S., Araújo, M.E., Borges, L.E.P. and Machado, N.T., Deacidification of organic liquid products by fractional distillation in laboratory and pilot scales, Journal of Analytical and Applied Pyrolysis. 127, pp. 468-489, 2017. DOI: 10.1016/j.jaap.2017.06.016. [ Links ]

[44] Wang, S., High-Efficiency Separation of Bio-Oil, 2013. DOI: 10.5772/51423. [ Links ]

[45] Paquot, C. 2.203 - Determination of the Ester Value (E.V.), in: Standard Methods for the Analysis of Oils, Fats and Derivatives (Sixth Edition), Pergamon, 1979: p. 60. [ Links ]

[46] Chew, T.L. and Bhatia, S., Effect of catalyst additives on the production of biofuels from palm oil cracking in a transport riser reactor, Bioresource Technology. 100, pp. 2540-2545, 2009. DOI: 10.1016/j.biortech.2008.12.021. [ Links ]

[47] Elkasabi, Y., Mullen, C.A., Jackson, M.A. and Boateng, A.A., Characterization of fast-pyrolysis bio-oil distillation residues and their potential applications, Journal of Analytical and Applied Pyrolysis. 114, pp. 179-186, 2015. DOI: 10.1016/j.jaap.2015.05.018. [ Links ]

[48] Elkasabi, Y., Boateng, A.A. and Jackson, M.A., Upgrading of bio-oil distillation bottoms into biorenewable calcined coke, Biomass & Bioenergy. 81, pp. 415-423, 2015. DOI: 10.1016/j.biombioe.2015.07.028. [ Links ]

[49] Agência Nacional de Petróleo, Gás Natural e Biocombustíveis − ANP. Resolucão Nº 40 de 25 de outubro de 2013. [online]. [accessed: April 10th, 2019]. Available at: http://www.anp.org. brLinks ]

[50] ASTM D4814-19a, Standard Specification for Automotive Spark-Ignition Engine Fuel, ASTM International, West Conshohocken, PA, 2019. [online]. [accessed: April 10th, 2019]. Available at: Available at: www.astm.orgLinks ]

[51] Agência Nacional de Petróleo, Gás Natural e Biocombustíveis − ANP. Resolucão Nº 37 de 01 de dezembro de 2009. [online]. [accessed: April 10th, 2019]. Available at: http://www.anp.org.brLinks ]

[52] ASTM D1655-19a, Standard Specification for Aviation Turbine Fuels, ASTM International, West Conshohocken, PA, 2019. [online]. [accessed: April 10th, 2019]. Available at: www.astm.orgLinks ]

[53] Agência Nacional de Petróleo, Gás Natural e Biocombustíveis − ANP. Resolucão Nº 65 de 09 de dezembro de 2011. [online]. [accessed: April 10th, 2019]. Available at: http://www.anp.org.brLinks ]

[54] ASTM D975-19a, Standard Specification for Diesel Fuel, ASTM International, West Conshohocken, PA , 2019. [online]. [accessed: April 10th, 2019]. Available at: Available at: www.astm.orgLinks ]

[55] Silva, J.P., Costa, A.L.H., Chiaro, S.S.X., Delgado, B.E.P.C., de Figueiredo, M.A.G. and Senna, L.F., Carboxylic acid removal from model petroleum fractions by a commercial clay adsorbent, Fuel Processing Technology. 112, pp. 57-63, 2013. DOI: 10.1016/j.fuproc.2012.07.033. [ Links ]

[56] Szklo, A.S. and Uller, V.C. Fundamentos do refino d petróleo: tecnologia e economia, Rio de Janeiro: Interciência, 2008. [ Links ]

[57] Oasmaa, A., Elliott, D.C. and Korhonen, J., Acidity of Biomass Fast Pyrolysis Bio-oils, Energy & Fuels. 24, pp. 6548-6554, 2010. DOI: 10.1021/ef100935r. [ Links ]

[58] Speight, J.G. Synthetic Fuels Handbook: Properties, Process, and Performance. New York: McGraw-Hill, 2008. [ Links ]

[59] Pavia, D.L., Lampman, G.M., Kriz, G.S. and Vyvyan, J.R. Introduction to spectroscopy. 4th Ed., Cengage Learning, Belmont, USA, 2009. [ Links ]

[60] Ertas, M. and Alma, M.H., Pyrolysis of laurel (Laurus nobilis L.) extraction residues in a fixed-bed reactor: Characterization of bio-oil and bio-char, Journal of Analytical and Applied Pyrolysis. 88, pp. 22-29, 2010. DOI: 10.1016/j.jaap.2010.02.006. [ Links ]

[61] Farah, M.A., Petróleo e seus derivados: definição, constituição, aplicação, especificações, características de qualidade. LTC, Rio de Janeiro, Brasil,2012. [ Links ]

[62] Speight, J.G., Handbook of Petroleum Product Analysis. John Wiley & Sons, Inc., Hoboken, USA, 2002. [ Links ]

[63] Idem, R.O., Katikaneni, S.P.R. and Bakhshi, N.N., Thermal Cracking of Canola Oil: Reaction Products in the Presence and Absence of Steam, Energy & Fuels. 10, pp. 1150-1162, 1996. DOI: 10.1021/ef960029h. [ Links ]

[64] Zhao, X.H., Wei, L., Cheng, S.Y., Cao, Y.H., Julson, J. and Gu, Z.R., Catalytic cracking of carinata oil for hydrocarbon biofuel over fresh and regenerated Zn/Na-ZSM-5, Applied Catalysis A-General. 507, pp. 44-55, 2015. DOI: 10.1016/j.apcata.2015.09.031. [ Links ]

S.AP. Mota, received a BSc. Eng. in Chemical Engineering in 2006, an MSc. in Chemical Engineering in 2008, and a PhD in Engineering of Natural Resources of the Amazon in 2013, all of them from the Universidade Federal do Pará (UFPA), Belém, Brazil. Currently, he is professor in the Faculty of Materials Engineering, Universidade Federal do Sul e Sudeste do Pará (Unifesspa) and Permanent Professor of the Graduate Program in Chemistry, Universidade Federal do Sul e Sudeste do Pará (Unifesspa). His research interests include: insertion of industrial waste in ceramic formulations; asphalt binder production; bio-flooring production; growth and organization of crystalline structures; biofuels engineering; synthesis and application of catalysts; synthesis and application of zeolites; use and transformation of waste materials; and synthesis and application of adsorbents.ORCID: 0000-0003-0823-4779

A.A.M. Mota, received a BSc. Eng in Food Engineering in 2009, an MSc. in Chemical Engineering in 2011, and a PhD in Engineering of Natural Resources of the Amazon in 2015, all of them from the Universidade Federal do Pará (UFPA), Belém, Brazil. Currently, she is postdoctoral researcher in the Graduate Program in Chemistry, Universidade Federal do Sul e Sudeste do Pará (Unifesspa). Her research interests include: biofuels engineering; separation processes; modeling and separation of processes; process modeling and simulation; biomass and bioenergy; and biofuel production technology.ORCID: 0000-0002-7565-6480

N.T. Machado, received a BSc. Eng in Chemical Engineering (Universidade Federal do Pará/UFPA) in 1988, an MSc. in Mechanical Engineering (COPPE-UFRJ) in 1991, and a PhD in Process Engineering (Technische Universitat Hamburg-Harburg) in 1998. Currently, he is full professor in the Faculty of Chemical Engineering, Universidade Federal do Sul e Sudeste do Pará (Unifesspa), Permanent Professor of the Doctoral Program in Engineering of Natural Resources of the Amazon (UFPA), and Permanent Professor of the Graduate Program in Chemical Engineering (UFPA). His research interests include process thermodynamics; biofuels engineering; separation processes; process modeling and simulation; biomass processing and transformation (thermo-biochemistry) engineering; supercritical fluid technology; and high pressure phase equilibrium.ORCID: 0000-0001-9800-3565

How to cite: Mota, S.A.P., Mota, A.A.M. and Machado, N.T., Influence of fractional distillation on the yield and quality of biofuels obtained through thermal catalytic cracking of crude palm oil.. DYNA, 88(218), pp. 62-71, July - September, 2021.

Received: September 08, 2020; Revised: April 30, 2021; Accepted: May 31, 2021

Creative Commons License The author; licensee Universidad Nacional de Colombia