Next Article in Journal
Synthesis of Alkyl Levulinates from α-Angelica Lactone Using Methanesulfonic Acid as a Catalyst: A Sustainable and Solvent-Free Route
Next Article in Special Issue
Immobilization of Sustine® 131 onto Spent Coffee Grounds for Efficient Biosynthesis of Ethyl Hydrocinnamate
Previous Article in Journal
The Influence of Pervaporation on Ferulic Acid and Maltol in Dealcoholised Beer
Previous Article in Special Issue
Synergistic Innovations: Organometallic Frameworks on Graphene Oxide for Sustainable Eco-Energy Solutions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Efficient Impurity Removal from Model FCC Fuel in Millireactors Using Deep Eutectic Solvents

1
PPD Hrvatska d.o.o., Vjekoslava Heinzela 70, 10000 Zagreb, Croatia
2
Department of Mechanical and Thermal Process Engineering, Faculty of Chemical Engineering and Technology, University of Zagreb, Marulićev trg 19, 10000 Zagreb, Croatia
3
INA-Industrija nafte d.d., Avenija Većeslava Holjevca 10, 10002 Zagreb, Croatia
*
Author to whom correspondence should be addressed.
ChemEngineering 2024, 8(5), 102; https://doi.org/10.3390/chemengineering8050102
Submission received: 25 August 2024 / Revised: 27 September 2024 / Accepted: 30 September 2024 / Published: 9 October 2024
(This article belongs to the Collection Green and Environmentally Sustainable Chemical Processes)

Abstract

:
The goal of strict fuel quality regulations is to decrease the levels of sulfur, nitrogen, and aromatic chemicals in gasoline, thereby enhancing environmental safety. Due to the high costs of hydrodenitrification and hydrodesulfurization, many studies are looking for alternative fuel-purifying processes. The straightforward extraction approach using deep eutectic solvents (DESs) has proven to result in the removal of impurities and the enhancement of gasoline quality. Seven DESs were employed in a batch extraction process to purify the model fuel. The TbabFa-0 solvent was chosen for extraction in millireactors with different lengths, volume flows, and solvent ratios. In the millireactor, a slug regime and a laminar flow pattern were established for every process condition. For the chosen process conditions, the diffusion coefficient, volumetric mass transfer coefficient, and distribution ratio were determined. Better separation of all three key components was achieved during extraction in a millireactor using TbabFa-0. The efficiency of extraction with regenerated solvent was lowered by a maximum of 8%, showing the possibility of performing extraction in a millireactor with solvent recirculation.

1. Introduction

The rapid expansion of the oil and transportation industries needs the most efficient use of the existing fossil fuel supply. Simultaneously, tight environmental standards compel the production of cleaner fuel with extremely low sulfur content. The combustion of fossil fuels causes the release of harmful pollutants such as SOx, NOx, and COx. Emissions cause environmental contamination and harm human health [1]. The current sulfur content in the European Union, the United States, and China is capped at 10 parts per million (ppm) [2]. These regulations regarding the environment will become more difficult to meet in the future as crude oil quality deteriorates, with increasing amounts of sulfur and aromatic content, as well as acid and heavy crudes [3]. Therefore, alternative methods of removing sulfur, nitrogen, and aromatic compounds are investigated with the goal of replacing the conventional fuel purification processes, called hydrodesulphurization and hydrodenitrification. The initiative for the replacement of conventional processes was the reduction in significant costs and the achievement of deep desulfurization and denitrification. In response, an environmentally friendly and cost-effective extraction process using green solvents was investigated [2].
Extractive desulfurization is a desirable approach for fuel purification, especially if it is performed with ecologically acceptable solvents. It is typically carried out at atmospheric pressure and low temperatures and does not require extra energy input into the system. Despite many advantages, it is important to consider the potential drawbacks that might arise during the implementation of this methodology. Removing compounds containing S and N heteroatoms may lead to a reduction in the number of saturated olefins, which are essential constituents in the petroleum blend used in commercial motor vehicles. Furthermore, the efficacy of the extraction process is affected by the solubility of the compounds requiring elimination. Consequently, the selection of an ideal solvent is critical to the success of desulfurization [4].
Numerous studies using deep eutectic solvents (DESs) were carried out in an effort to replace hazardous volatile organic solvents, which are still commonly used in the industry today [5,6,7,8]. DESs are distinguished by the fact that they are inexpensive, nonvolatile, non-hazardous, simple to produce, and regenerable. Furthermore, these solvents offer the benefit of enabling processes to be conducted at room temperature and pressure conditions, thus decreasing purification costs and fulfilling the requirements of green chemistry [9,10,11]. However, DESs’ high viscosity—which might cause mixing difficulties during extraction—is one of its key disadvantages.
Millistructured reactors can be used to solve the issues that often arise when viscous liquids are mixed. In general, millireactors have significant advantages over conventional reactors in terms of energy efficiency, reaction rate and yield, safety, reliability, and on-site production [12,13] The millireactor’s high surface-to-volume ratio enables efficient phase interaction, leading to improved fuel purification and reduced solvent usage. A continuous flow of solvents is guaranteed by the use of accurate piston pumps. Mass transfer occurs at the moment of first contact and continues across the millireactor and phase boundary. The shape of the millimixer, which is often a Y- or T-shape, determines the flow patterns in the millireactor. This is particularly evident in multiphase systems, where the flow pattern, except for the geometry of the mixer, depends on the velocities of different phases [14].
Seven DESs were used in a batch liquid extraction to purify the model, fluidized catalytic cracking gasoline, in order to investigate the potential of removing toluene, sulfur, and nitrogen compounds from the fuel. DESs with various features and physicochemical properties were examined, enabling us to identify the factors that have the greatest effect on the extraction process. The DES that yielded the most favorable results was used in a fused filament fabrication (FFF) millireactor, and its results were compared with those obtained during the bulk extraction procedure.

2. Materials and Methods

2.1. Materials

Table 1 lists the chemicals used for the DESs and the model fuel preparation. The chemicals were used without further purification except for hygroscopic components (choline chloride, ethylene glycol, and malic acid) that were dried for 8 h at 60 °C in a vacuum oven. The content of water in other chemicals was less than 1%.

2.2. Preparation of Deep Eutectic Solvents and Model Fuel

DESs were produced by combining hydrogen bond acceptor (HBA) and hydrogen bond donor (HBD) in a specified molar ratio (Table 2). The materials were mixed in the IKA RV 10 basic rotating evaporator at a temperature of 60 °C until a uniform and clear liquid was obtained. After the formation of a transparent liquid without any crystalline substances, 30% by weight of water (w) was added to DESs ChMa and ChCit in order to decrease the viscosity.
The model fuel was prepared by the gravimetric method, containing six components (Table 3).

2.3. Physicochemical Properties of Deep Eutectic Solvents

The Brookfield DV-III Ultra Programmable Rheometer and spindle SC4-21T (Middleborou, MA, USA) were used for the determination of the DESs’ viscosities. Densities were measured with a Mettler Toledo Densito 30PX densitometer (Columbus, OH, USA). pH was measured by a WTW inoLab pH/Cond 740 multi-meter (London, UK). The polarities of DESs were estimated using Nile red as a solvatochromic probe, as described in [15]. The absorption spectra of the dye were acquired with Shimadzu spectrophotometer UV-1280 (Kyoto, Japan). The surface tension of all DESs was measured using a Data Physics Instruments goniometer, OCA 20 (Stuttgart, Germany).

2.4. Batch Extraction

The solvent that was most appropriate for liquid extraction was determined using the IKA KS 3000i control shaker (Staufen, Germany). To attain equilibrium, the mixture of the DES and FCC fuel was mixed for 24 h at 25 °C and 200 min−1. After shaking, the mixture was left for phase separation. The concentration of key components was determined by gas chromatography [16].

2.5. Chromatographic Analysis

The concentration of the model fuel components after extraction was evaluated using gas chromatography (GC). The GC used was a Shimadzu GC-2014 (Kyoto, Japan) equipped with an FID detector and a Zebron Phase ZB-1 GC column (30 m × 0.53 mm × 1.50 mm). The GC solution Software (Shimadzu, Kyoto, Japan) program settings for the analyses used in this study were identical to those previously reported in the literature [16].
The desulfurization efficiency was calculated using Equation (1).
ε = X F X R X F
X represents the mass ratio for feed or raffinate phase.

2.6. Continuous Extraction in Millireactor

Three millireactors were manufactured using FFF technology and made from polylactic acid (PLA). These millireactors were employed for continuous extraction, with each having a different length but an equal inlet section. Figure 1 and Table 4, respectively, display the experimental setup and characteristics of the milliextractor. The milliextractor received fuel and DESs continuously from two syringe pumps through a 120°-angled Y-shaped inlet, which ensured phase contact [17].
At flow rates of 100 to 1000 mL/min and at solvent ratios of 0.5, 1, and 2, continuous extraction in millireactors of various lengths was provided. (Table 4). For defining the flow pattern in the milliextractor, the Weber, Reynolds, and capillary numbers were calculated according to the following equations:
We = ρ · v 2 · L σ
Re = ρ · v · d μ
Ca = μ · v σ
Symbols in the equations represent as follows:
ρ—density, kg∙m−3
v—velocity, m∙s−1
L—length of millireactor, m
σ—surface tension
μ—dynamic viscosity, Pa∙s
The retention τ time in the millireactor (Table 5) was calculated according to the following equation:
τ = V Q R + Q E
where V represents volume of the millireactor and QR and QE represent volume flow for the raffinate and extract phases.

2.7. Volumetric Mass Transfer Coefficient Calculation

In order to quantify the performance of the extraction process, the distribution ratio (β) and volumetric mass transfer coefficient were calculated (Kra). The distribution ratio determines the quantity of the dissolved component in two liquid phases under equilibrium conditions and is determined from the formula
β i = X Ri Y Ei
where X and Y are the mass ratios in the raffinate and extract phases and subscript i denotes the solute.
The volumetric mass transfer coefficient is used to evaluate the efficiency of the extraction procedures. For the millireactor system, it may be obtained by calculating the driving force for the raffinate phase using known mass ratio and distribution ratios.
K R · a = Q R · X F X R V · Δ X ln
Δ X LN = X R Y E β X F Y F β ln X R Y E β X F Y F β

2.8. Diffusion Coefficients

For the estimation of the diffusion coefficient, D different equations were applied (Table 6).
T, M, and φ denote the values of the temperature, molar mass, and volume ratio. Indexes A and B represent components A and B.

2.9. Regeneration of DES

In order to evaluate the feasibility of solvent reuse, TbabFa-0 was regenerated by evaporation in a vacuum evaporator (Laborota 4000 Heidolph) at low pressure and at 60 °C for two and a half hours, or until all of the extracted components (thiophene, pyridine, and toluene) were removed. TbabFa-0 was selected for millireactor extraction and regeneration since it was the best of the tested DESs in terms of extracting all three components from the model fuel. To determine the effectiveness of the regeneration process, samples of the regenerated DES and pure DES were analyzed using Fourier-transformed infrared spectroscopy. The analysis was carried out using a Vertex 70 FTIR (Billerica, MA, USA) spectrophotometer equipped with a platinum detector.

3. Results and Discussion

3.1. Batch Extraction

To determine suitable solvents for the model fuel purification, batch extraction of thio-phene, pyridine, and toluene was carried out using seven DESs produced by mixing various HBAs and HBDs (Table 2). In order to prove the formation of DESs, FTIR analysis was performed on all generated DESs, as well as the components involved in their production, that can be found in the Supplementary Material, Figures S1–S7. The FTIR analysis confirms the existence of a eutectic system by showing changes in characteristic peaks when the spectrum of DESs is compared to that of their individual parts. This shift usually suggests the formation of strong hydrogen bonds between the donor and acceptor molecules. Figures S1–S7 show shifts in frequency, or peak broadening, for every DES. Furthermore, the absence of any new peaks in the provided FTIR analysis indicates a lack of chemical interaction among the DES components [24,25]. To achieve equilibrium, the extraction procedure was carried out on a laboratory shaker for 24 h at a temperature of 25 °C.
The physical properties of the DESs vary according to their composition (HBA and HBD), and their values are shown in Table 7. The density, viscosity, pH, polarity, and surface tension of DESs have been measured to determine the impact of the DESs’ physical properties on extraction efficiency. The densities of all DESs are higher than that of the model fuel, resulting in easy phase separation after extraction. A key variable that affects the hydrodynamic conditions during mixing in a batch process or the type of flow pattern in channels during a continuous operation is the viscosity of the applied DESs [26]. The used DESs show a broad range of viscosity, from 0.0168 Pa·s to 2.2415 Pa·s, which may impact hydrodynamic conditions and in turn extraction efficiency. The acidity of the DESs has significant effects on extraction efficiency; hence, DESs with varying acidities were used. DESs prepared with acid have a pH value below one that is high acidic, while ChEg-0 is slightly acidic due to the presence of hydroxyl groups in ethylene glycol, and BGly-0 is mildly basic (pH value 8.50). The most polar DES (ChMa-30 49.52 kcal mol−1) and the least polar DES (BGly-0 51.37 kcal mol−1) vary in polarity by 1.85 kcal mol−1, which is not significant. Generally, the high polarity of DESs impacts their capacity to separate various components. As the compounds extracted from the fuel differ significantly in polarity, it is expected that this will affect the extraction efficiency. Regardless of differences in the physicochemical characteristics of the applied DESs, pyridine, with the highest polarity of the three extracted components [27], is the easiest to separate from the fuel. Applying DESs with a pH lower than one almost completely removes pyridine. Extraction with acidic DESs prepared with choline chloride as a HBA (ChMA and CHCIT) achieved the highest efficiency values for pyridine removal. DESs that contain the –COOH group enhance hydrogen bonding interaction between the DES and pyridine due to increased dipole strength, and it is already known that the hydrogen bond between the nitrogen compound and the DES is the main driving force for mass transfer in the extraction process [22].
Also, even though they had a high viscosity, DESs made without water and DESs that were less acidic and polar worked better for extracting thiophene (Table 7 and Figure 2). The presence of water in DESs has a negative effect on the extraction of thiophene, while on the other hand, the addition of water has no effect on the extraction of pyridine from the model fuel [4]. Also, HBAs have a strong influence on the extraction of thiophene from the model fuel, i.e., for this specific purification process, tetrabutilammonium bromide is more favored as HBA than choline chloride or betaine. The formation of hydrogen bonds between the extracted component and the DES is the main driving force for the extractive desulfurization, as pointed out by Li et al. [28].
TbabFa-0 was the only DES tested that was effective in removing toluene from the model fuel, while the other DESs failed to provide any results in the extraction procedure. Extraction efficiency with TbabFa-0 was around 23%. This research demonstrates that HBA have a greater impact on extraction than HBD. DESs with formic acid as a HBD have one C=O bond, which may enhance the extraction process due to the π-π interaction between C=O bound to TbabFa-0 and toluene [29,30].
Although the extraction efficiency could be limited by the physicochemical properties of solvents (viscosity, surface tension, and polarity of the DES), results of the batch extractions show that solvent selection and their affinity to dissolve some components play a key role in the extraction efficiency (Figure 2).
By summarizing the above results from the batch extractions, conducted with different solvents, the most suitable solvent for the extraction process was TbabFa-0. Therefore, the experiments in the millireactor were performed using TbabFa-0.

3.2. Extraction in Millireactor

The concept of minimizing chemical processes has multiple advantages and has gained significant interest in the extraction field. Using microreactors or millireactors in the extraction process has many benefits, such as the ability to lower the amount of reagents needed and keep a steady laminar flow at high shear rates, which can ensure the suitable flow. Additionally, the high surface-to-volume ratio of narrow channels can result in exceptional efficiency for heat and mass transfer [31].
TbabFa-0 was selected for further research to investigate the potential for enhancing the extraction of thiophene, pyridine, and toluene from the model fuel using millireactors. This was due to its capacity to separate all three components from the fuel during batch extraction, thereby providing a potential for improvement in all three components. The hydrodynamic conditions, optimal retention time, and flow regime were all characterized in millireactors. The effect of the DES-to-feed mass ratio and the millireactor length on extraction performance was investigated.

3.2.1. The Flow Behavior

The flow behavior was determined by observation during experiments (Figure 3) and confirmed by values of dimensionless parameter groups such as the capillary, Weber, and Reynolds numbers (Table 8).
During all the millireactor experiments, a slug flow regime occurs when the capillary and Weber numbers for both phases are low, namely at Ca < 0.01 and We < 1 [32,33]. The measured values of Ca = 1.5·10−1–1.5·10−5 and We = 4.03·10−5–2.7·10−3 show that surface tension is stronger than viscous and inertial forces in the systems that were studied. This results in the formation of a slug flow regime inside the millireactor that ensures a uniform interfacial surface and thus efficient mass transfer [34].
The Reynolds numbers in examined process conditions were less than 100 (ranging from 7.17·10−4 to 2.68), indicating laminar flow. All of the aforementioned facts support the observation of a slug flow for all process conditions in the experiments (Figure 3). The movement of fluid elements within the narrow channel generates internal vortices that enhance mass transfer [17,32]. As a result, the slug flow regime in the millireactor is expected to offer superior mass transfer compared to other flow regimes [35].

3.2.2. Influence of Retention Time on Extraction Efficiency

The retention time in the continuous extraction process was varied by modifying the flow rates of solvents and by changing millireactor lengths (Table 5). Figure 4 shows the extraction efficiency of pyridine, thiophene, and toluene as a function of retention time in a defined slug flow regime. Extending the retention time from 1.6 to 26.5 min improved the extraction efficiency. Shorter retention times led to lower extraction efficiency, as the retention time became insufficient for achieving stationary conditions. Equilibrium was attained after 14 min for all extracted components. This retention time was achieved in a reactor with a length of 50 cm and at a volumetric flow rate of 100 μL/min for the FCC gasoline and 66 μL/min for TbabFa-0 (S = 1). Since the used apparatus could not achieve lower flow rates, extending the millireactor’s length is a more effective way to increase the retention time. Specifically, Figure 4 demonstrates a trend in increasing of the extraction process efficiency for all three components, irrespective of the way used to achieve the retention time.
The results show that, similar to batch extraction, TbabaFa-0 extraction is more successful in extracting pyridine, followed by thiophene and toluene. More than 96% of the pyridine, 75% of the toluene, and 14% of the toluene were removed after 14 min of extraction in the millireactor. The extraction efficiency of pyridine, thiophene, and toluene was increased by 11.0%, 54.5%, and 37.6%, respectively, as compared to batch extraction. Furthermore, this improvement was achieved in a much shorter time period. The microfluidic system exhibits superior performance, particularly when slug flow exists, as evidenced by studies that compared batch and extraction in microchannels for various systems [34,35,36,37]. A similar pattern may be seen in the case of a millireactor, where the dimensions are such that they maintain the flow and distribution of forces inside the reactor. This enhancement is attributed to the internal circulations that are characteristic of the slug flow, which significantly improve convection within each liquid phase [35]. In addition to the more intense mixing, the millireactor’s higher pressure (approximately 2 bars) [38] in comparison to the batch reactor may be another contributing factor.

3.2.3. Impact of Solvent Ratio on Extraction Efficiency

Despite the relatively inexpensive and simple production of DESs, the industrial processes of extractive desulfurization and denitrification prefer to minimize the amount of selective solvent (DES) to achieve optimal results [28]. Figure 5 presents the extraction results for a single key component using the DES TbabFa-0 at different solvent ratios. Increasing the amount of DES enhances the extraction of all key components. To identify the optimal solvent ratio, the specific mass of the extracted component was measured at various DESs-to-feed mass ratio (Table 9). The results indicate that the highest DES-to-feed mass ratio resulted in the lowest specific removal of the key component, while the lowest mass ratio provided the most effective removal per kilogram of solvent used. Although a larger amount of DES improves extraction efficiency, the results do not justify increasing the DES-to-feed mass ratio. The DES-to-feed mass ratio depends on the importance of the purification process, the cost of the secondary solvent, and the specific application [39].

3.2.4. Volumetric Mass Transfer Coefficient and Diffusion Coefficient

Figure 6 shows the volumetric mass transfer coefficients, Kr·a, for the raffinate phase in extraction with TbabFa-0 at a DES-to-feed mass ratio of one. As the retention time increases, the volumetric mass transfer coefficients for all key components increase until they reach a stationary condition. The removal of pyridine yielded the highest volumetric mass transfer coefficients, whereas toluene yielded the lowest volumetric mass transfer coefficients, which is in coincidence with extraction efficiency results. The volumetric mass transfer coefficient reflects both the efficiency of mass transfer (through the mass transfer coefficient) and the extent of the interfacial area available for that transfer. In this case, the mass transfer coefficient is the primary factor determining the value of the volumetric mass transfer coefficient, as surface tension forces outweigh inertial forces at low flow rates, resulting in a stable slug flow regime and a constant interfacial area. As the efficiency of extraction and subsequent mass transfer increases with retention time (Figure 6), the volumetric mass transfer coefficient increases.
The diffusion coefficient is crucial to the design and development of DES liquid extraction because it affects the efficiency process and is the primary mechanism of mass transfer through liquids. Therefore, different equations were used to calculate the diffusion coefficients during the extraction of key components using TbabFa-0 (Table 6). The diffusion coefficient values that were obtained are lower (Table 10) than the literature values for mass transfer in liquid–liquid systems (D = 10−9–10−10 m2∙s−1) [40]. It is known that the order of magnitude of the diffusion coefficient in liquid extraction with DESs is typically lower for many solutes than in conventional solvents [32]. DESs often have significantly higher viscosities than traditional organic solvents or water. Higher viscosity leads to lower diffusion coefficients because the solute molecules encounter more resistance as they move through the solvent. Furthermore, DESs have a complex hydrogen bonding network, which can further slow down the movement of solute molecules, reducing the diffusion coefficient compared to conventional solvents [41].
The diffusion coefficient, depending on the correlation used, takes into account the solvent’s molar mass, the key component’s molar volume, the solvent’s viscosity, and the temperature. All of the correlations used give approximately similar diffusion coefficient values (within the same order of magnitude), except for the Stokes–Einstein correlation, which gives one order of magnitude higher values. The Stokes–Einstein correlation can be used in the systems where the dissolved key component’s molecule diameter is smaller than the solvent molecule’s diameter [42,43]. However, this is not the case in the tested systems. Therefore, we can conclude that this correlation is not appropriate for the tested systems. Upon comparing the diffusion coefficients of the key components, we observe that thiophene exhibits marginally higher values than pyridine and toluene across all correlations. However, Polson’s correlation, the sole one that considers the molar mass of the extracted component, yields the highest diffusion coefficient value for pyridine, followed by thiophene, and the lowest for toluene. This calculation aligns with the experimental data from the extractions. The obtained diffusion coefficient values suggest that the equations for estimating the diffusion coefficient must consider the affinity of DESs for the dissolution of specific key components.

3.2.5. Reuse of Regenerated DES

From an ecological and economic point of view, the possibility of regeneration and reuse of DESs in the extraction process is a significant advantage compared to conventional desulfurization and denitrification procedures. The results of extracting thiophene and pyridine from the model fuel using a fresh and a regenerated DES (TbabFa-0) demonstrated that the extraction ability of the DES did not significantly change, indicating its successful regeneration. The largest reduction in extraction efficiency for all components was determined to be less than 8%, indicating the potential for the effective use of regenerated DESs (Table 11).
The amount of regeneration rounds without significant DES performance reduction varies per instance. Some researchers have successfully regenerated DESs several times without losing any qualities, while others have observed somewhat fewer cycles. This is dependent on a number of factors, including the kind of DES, the process conditions under which it is used, the kind of contamination, and the regeneration method. The literature describes effective DES regeneration utilizing various approaches. These methods included rinsing with water [44], rinsing with water followed by a distillation procedure [45], and rinsing with diethyl ether followed by a heating procedure at 105 °C [46]. Although the method of solvent regeneration used in this work is carried out at a high temperature and reduced pressure, it is very simple and quick to implement.
FTIR analysis was employed for a comparison between the fresh, after-extraction, and regenerated DES. The peak characteristics of Tbab-Fa-0 appear in all three spectra. The maxima of the absorption bands are located at the same wavenumbers as those of the fresh DES (Figure 7). The FTIR spectra are same in the functional group region, exhibiting absorption maxima at 2958 cm−1 and 2878 cm−1, which correspond to C-H stretching. The absorption maximum at 2652 cm−1 corresponds to O-H stretching, the absorption maximum at 1743 cm−1 to C=O stretching, the vibrations at 1465 cm−1 to C-H bending, the absorption maximum at 1165 cm−1 to C-O stretching, and the absorption maximum at 879 cm−1 in the fingerprint region to C-H bending. After extraction, the FTIR spectra (Figure 7) for DESs show some extra peaks that correspond to the removed components (1595, 1581, 1437, 990, 708, 668, and 603 cm−1). A couple of these peaks (990 and 708) can also be found in the FTIR spectra of the regenerated DES. This indicates that certain removed components remain even after DES regeneration. The peak at 990 cm−1 indicates that toluene was not completely removed from the regenerated DES, while the peak at 708 cm−1 is expected for toluene.
The effective and easy regeneration of the DES allows it to be reused in industrial processes, resulting in cost savings [47]. However, determining the number of cycles necessary for the effective removal of impurities is crucial in order to determine the ideal regeneration frequency, as the prolonged usage of a DES lowers its capacity.

4. Conclusions

A batch extraction was conducted with seven DESs to determine the most suitable DES for the purification of the FCC model fuel in a millireactor. Thiophene extraction is more successful with DESs that are prepared without water, less acidic, and less polar (TbabFa-0, ChEg-0, ChGly-0, and BGly-0). Pyridine, a more polar component, is extracted more effectively by more acidic and polar solvents that contain water (ChMa-30, ChCit-30, and MaFruGly-30). The DES TbabFa-0 was the only solvent that successfully extracted toluene.
A slug flow regime was observed in all of the millireactor experiments, resulting in superior outcomes when compared with batch extraction. The extraction efficacy of all key components was enhanced as a result of a longer millireactor (50 and 90 cm) and a higher DES-to-feed mass ratio. The Polson correlation yielded the best agreement between the estimated diffusion coefficient values and the extraction efficiency results. The structure of the regenerated DES did not change according to FTIR analysis, while the extraction efficiency of the regenerated TbabFa-0 was slightly lower than that of the fresh DES. These facts suggest that a circular fuel purification process could be implemented in a millireactor through the use of a straightforward method of DES regeneration.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/chemengineering8050102/s1, Table S1. Efficiency of extraction for tiophene, pyridine and toluene; Figure S1. FTIR comparison of pure components and BGly-0; Figure S2. FTIR comparison of pure components and ChCit-30; Figure S3. FTIR comparison of pure components and ChEg-0; Figure S4. FTIR comparison of pure components and ChGly-0; Figure S5. FTIR comparison of pure components and ChMa-30; Figure S6. FTIR comparison of pure components and MaFruGly-0; Figure S7. FTIR comparison of pure components and TbabFa-0.

Author Contributions

Conceptualization and methodology; A.M. and J.P.K.; investigation; A.M. and M.L.; writing—original draft preparation, A.M.; writing—review and editing, J.P.K.; visualization, J.P.K. and M.L.; supervision, J.P.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Data is contained within the article and Supplementary Materials.

Conflicts of Interest

Author Anamarija Mitar is employed by PPD Hrvatska d.o.o. and Marija Lukić is employed by INA-Industrija nafte d.d. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Manisalidis, I.; Stavropoulou, E.; Stavropoulos, A.; Bezirtzoglou, E. Environmental and Health Impacts of Air Pollution: A Review. Front. Public Health 2020, 8, 14. [Google Scholar] [CrossRef] [PubMed]
  2. Chandran, D.; Khalid, M.; Walvekar, R.; Mubarak, N.M.; Dharaskar, S.; Wong, W.Y.; Gupta, T.C.S.M. Deep Eutectic Solvents for Extraction-Desulphurization: A Review. J. Mol. Liq. 2019, 275, 312–322. [Google Scholar] [CrossRef]
  3. Betiha, M.A.; Rabie, A.M.; Ahmed, H.S.; Abdelrahman, A.A.; El-Shahat, M.F. Oxidative Desulfurization Using Graphene and Its Composites for Fuel Containing Thiophene and Its Derivatives: An Update Review. Egypt. J. Pet. 2018, 27, 715–730. [Google Scholar] [CrossRef]
  4. Makoś, P.; Boczkaj, G. Deep Eutectic Solvents Based Highly Efficient Extractive Desulfurization of Fuels—Eco-Friendly Approach. J. Mol. Liq. 2019, 296, 111916. [Google Scholar] [CrossRef]
  5. Abro, R.; Kiran, N.; Ahmed, S.; Muhammad, A.; Jatoi, A.S.; Mazari, S.A.; Salma, U.; Plechkova, N.V. Extractive Desulfurization of Fuel Oils Using Deep Eutectic Solvents—A Comprehensive Review. J. Environ. Chem. Eng. 2022, 10, 107369. [Google Scholar] [CrossRef]
  6. Lima, F.; Branco, L.C.; Silvestre, A.J.D.; Marrucho, I.M. Deep Desulfurization of Fuels: Are Deep Eutectic Solvents the Alternative for Ionic Liquids? Fuel 2021, 293, 120297. [Google Scholar] [CrossRef]
  7. Gabrić, B.; Sander, A.; Cvjetko Bubalo, M.; Macut, D. Extraction of S- and N-Compounds from the Mixture of Hydrocarbons by Ionic Liquids as Selective Solvents. Sci. World J. 2013, 2013, 512953. [Google Scholar] [CrossRef]
  8. Malolan, R.; Gopinath, K.P.; Vo, D.-V.N.; Jayaraman, R.S.; Adithya, S.; Ajay, P.S.; Arun, J. Green Ionic Liquids and Deep Eutectic Solvents for Desulphurization, Denitrification, Biomass, Biodiesel, Bioethanol and Hydrogen Fuels: A Review. Environ. Chem. Lett. 2021, 19, 1001–1023. [Google Scholar] [CrossRef]
  9. Marcus, Y. Applications of Deep Eutectic Solvents. In Deep Eutectic Solvents; Marcus, Y., Ed.; Springer International Publishing: Cham, Switzerland, 2019; pp. 111–151. [Google Scholar] [CrossRef]
  10. Tang, B.; Zhang, H.; Row, K.H. Application of Deep Eutectic Solvents in the Extraction and Separation of Target Compounds from Various Samples. J. Sep. Sci. 2015, 38, 1053–1064. [Google Scholar] [CrossRef]
  11. Atilhan, M.; Aparicio, S. Review and Perspectives for Effective Solutions to Grand Challenges of Energy and Fuels Technologies via Novel Deep Eutectic Solvents. Energy Fuels 2021, 35, 6402–6419. [Google Scholar] [CrossRef]
  12. Dong, Z.; Wen, Z.; Zhao, F.; Kuhn, S.; Noël, T. Scale-up of Micro- and Milli-Reactors: An Overview of Strategies, Design Principles and Applications. Chem. Eng. Sci. X 2021, 10, 100097. [Google Scholar] [CrossRef]
  13. Ehrfeld, W.; Hessel, V.; Löwe, H. Microreactors: New Technology for Modern Chemistry; Wiley: Hoboken, NJ, USA, 2000. [Google Scholar] [CrossRef]
  14. Nieuwland, P.J.; Koch, K.; van Hest, J.C.M.; Rutjes, F.P.J.T. Flow Markers in Microreactors: A Generally Applicable Chromatographic Method for Monitoring Flow Rates during Reactions. Open Chem. Eng. J. 2010, 4, 61–67. [Google Scholar] [CrossRef]
  15. Jeong, K.M.; Ko, J.; Zhao, J.; Jin, Y.; Yoo, D.E.; Han, S.Y.; Lee, J. Multi-Functioning Deep Eutectic Solvents as Extraction and Storage Media for Bioactive Natural Products That Are Readily Applicable to Cosmetic Products. J. Clean. Prod. 2017, 151, 87–95. [Google Scholar] [CrossRef]
  16. Mitar, A.; Lukić, M.; Kolesarić, P.; Prlić Kardum, J. Pročišćavanje Modelnog fcc Benzina Primjenom Niskotemperaturnih Eutektičkih Otapala u Milireaktoru. In Proceedings of the 2nd International Conference the Holistic Approach to Environment, Sisak, Croatia, 28 May 2021; pp. 369–380. [Google Scholar]
  17. Kandlikar, S.G.; King, M.R. Chapter 1—Introduction. In Heat Transfer and Fluid Flow in Minichannels and Microchannels; Kandlikar, S.G., Garimella, S., Li, D., Colin, S., King, M.R., Eds.; Elsevier Science Ltd.: Oxford, UK, 2006; pp. 1–7. [Google Scholar] [CrossRef]
  18. Seader, J.D.; Henley, E.J.; Roper, D.K. Separation Process Principles: Chemical and Biochemical Operations, 3rd ed.; John Wiley & Sons: Chichester, UK, 2011. [Google Scholar]
  19. Ghosh, R. Principles of Bioseparations Engineering; World Scientific: Singapore, 2006. [Google Scholar] [CrossRef]
  20. Bird, R.B.; Stewart, W.E.; Lightfoot, E.N. Transport Phenomena, 2nd ed.; [Nachdr.]; Wiley: New York, NY, USA; Weinheim, Germany, 2005. [Google Scholar]
  21. Perry, R.H.; Green, D.W. Perry’s Chemical Engineers’ Handbook, 7th ed.; McGraw-Hill: New York, NY, USA, 1999. [Google Scholar]
  22. Miložič, N.; Lubej, M.; Novak, U.; Žnidaršič-Plazl, P.; Plazl, I. Evaluation of Diffusion Coefficient Determination Using a Microfluidic Device. Chem. Biochem. Eng. Q. J. 2014, 28, 215–223. [Google Scholar] [CrossRef]
  23. Wilke, C.R.; Chang, P. Correlation of Diffusion Coefficients in Dilute Solutions. AIChE J. 1955, 1, 264–270. [Google Scholar] [CrossRef]
  24. Mitar, A.; Panić, M.; Kardum, J.; Halambek, J.; Sander, A.; Zagajski Kučan, K.; Radojcic Redovnikovic, I.; Radošević, K. Physicochemical Properties, Cytotoxicity, and Antioxidative Activity of Natural Deep Eutectic Solvents Containing Organic Acid. Chem. Biochem. Eng. Q. 2019, 33, 1–18. [Google Scholar] [CrossRef]
  25. Zhu, S.; Li, H.; Zhu, W.; Jiang, W.; Wang, C.; Wu, P.; Zhang, Q.; Li, H. Vibrational Analysis and Formation Mechanism of Typical Deep Eutectic Solvents: An Experimental and Theoretical Study. J. Mol. Graph. Model. 2016, 68, 158–175. [Google Scholar] [CrossRef]
  26. Su, Y.; Chen, G.; Yuan, Q. Effect of Viscosity on the Hydrodynamics of Liquid Processes in Microchannels. Chem. Eng. Technol. 2014, 37, 427–434. [Google Scholar] [CrossRef]
  27. Solvent Physical Properties. Available online: https://people.chem.umass.edu/xray/solvent.html (accessed on 26 September 2024).
  28. Li, C.; Li, D.; Zou, S.; Li, Z.; Yin, J.; Wang, A.; Cui, Y.; Yao, Z.; Zhao, Q. Extraction Desulfurization Process of Fuels with Ammonium-Based Deep Eutectic Solvents. Green Chem. 2013, 15, 2793–2799. [Google Scholar] [CrossRef]
  29. Salleh, M.Z.M.; Hadj-Kali, M.K.; Hizaddin, H.F.; Ali Hashim, M. Extraction of Nitrogen Compounds from Model Fuel Using 1-Ethyl-3-Methylimidazolium Methanesulfonate. Sep. Purif. Technol. 2018, 196, 61–70. [Google Scholar] [CrossRef]
  30. Guo, Y.; Liu, X.; Li, J.; Hu, B. Optimization Study on Deep Extractive Oxidative Desulfurization with Tetrabutylammonium Bromide/Polyethylene Glycol DES. RSC Adv. 2021, 11, 31727–31737. [Google Scholar] [CrossRef]
  31. Assmann, N.; von Rohr, P.R. Extraction in Microreactors: Intensification by Adding an Inert Gas Phase. Chem. Eng. Process. Process Intensif. 2011, 50, 822–827. [Google Scholar] [CrossRef]
  32. Jin, N.; Yue, J.; Zhao, Y.; Lü, H.; Wang, C. Experimental Study and Mass Transfer Modelling for Extractive Desulfurization of Diesel with Ionic Liquid in Microreactors. Chem. Eng. J. 2021, 413, 127419. [Google Scholar] [CrossRef]
  33. Jovanović, J.; Rebrov, E.V.; Nijhuis, T.A., (Xander); Kreutzer, M.T.; Hessel, V.; Schouten, J.C. Liquid–Liquid Flow in a Capillary Microreactor: Hydrodynamic Flow Patterns and Extraction Performance. Ind. Eng. Chem. Res. 2012, 51, 1015–1026. [Google Scholar] [CrossRef]
  34. Tamagawa, O.; Muto, A. Development of Cesium Ion Extraction Process Using a Slug Flow Microreactor. Chem. Eng. J. 2011, 167, 700–704. [Google Scholar] [CrossRef]
  35. Nandagopal, M.S.G.; Antony, R.; Selvaraju, N. Comparative Study of Liquid–Liquid Extraction in Miniaturized Channels over Other Conventional Extraction Methods. Microsyst. Technol. 2016, 22, 349–356. [Google Scholar] [CrossRef]
  36. Naleini, N.; Rahimi, M.; Heydari, R. Oleuropein Extraction Using Microfluidic System. Chem. Eng. Process. Process Intensif. 2015, 92, 1–6. [Google Scholar] [CrossRef]
  37. Sahu, A.; Vir, A.B.; Molleti, L.N.S.; Ramji, S.; Pushpavanam, S. Comparison of Liquid-Liquid Extraction in Batch Systems and Micro-Channels. Chem. Eng. Process. Process Intensif. 2016, 104, 190–200. [Google Scholar] [CrossRef]
  38. Marre, S.; Adamo, A.; Basak, S.; Aymonier, C.; Jensen, K.F. Design and Packaging of Microreactors for High Pressure and High Temperature Applications. Ind. Eng. Chem. Res. 2010, 49, 11310–11320. [Google Scholar] [CrossRef]
  39. Durdun, C.; Papuc, C.; NICORESCU, V.; Gajaila, I.; Goran, G.; Petcu, C.; Stefan, G. The Influence of Solid-to-Solvent Ratio and Extraction Method on Total Phenolic Content, Flavonoid Content and Antioxidant Properties of Some Ethanolic Plant Extracts. Rev. Chim.-Buchar.-Orig. Ed. 2016, 67, 1922–1927. [Google Scholar]
  40. Cussler, E.L. Diffusion: Mass Transfer in Fluid Systems; Cambridge University Press: Cambridge, MA, USA, 1997. [Google Scholar]
  41. Han, Q.; Veríssimo, N.V.P.; Bryant, S.J.; Martin, A.V.; Huang, Y.; Pereira, J.F.B.; Santos-Ebinuma, V.C.; Zhai, J.; Bryant, G.; Drummond, C.J.; et al. Scattering Approaches to Unravel Protein Solution Behaviors in Ionic Liquids and Deep Eutectic Solvents: From Basic Principles to Recent Developments. Adv. Colloid Interface Sci. 2024, 331, 103242. [Google Scholar] [CrossRef] [PubMed]
  42. Berkowicz, S.; Perakis, F. Exploring the Validity of the Stokes–Einstein Relation in Supercooled Water Using Nanomolecular Probes. Phys. Chem. Chem. Phys. 2021, 23, 25490–25499. [Google Scholar] [CrossRef] [PubMed]
  43. Su, J.T.; Duncan, P.B.; Momaya, A.; Jutila, A.; Needham, D. The Effect of Hydrogen Bonding on the Diffusion of Water in N-Alkanes and n-Alcohols Measured with a Novel Single Microdroplet Method. J. Chem. Phys. 2010, 132, 044506. [Google Scholar] [CrossRef] [PubMed]
  44. Jiang, W.; Jia, H.; Li, H.; Zhu, L.; Tao, R.; Zhu, W.; Li, H.; Dai, S. Boric Acid-Based Ternary Deep Eutectic Solvent for Extraction and Oxidative Desulfurization of Diesel Fuel. Green Chem. 2019, 21, 3074–3080. [Google Scholar] [CrossRef]
  45. Shu, C.; Sun, T. Extractive Desulfurisation of Gasoline with Tetrabutyl Ammonium Chloride-Based Deep Eutectic Solvents. Sep. Sci. Technol. 2016, 51, 1336–1343. [Google Scholar] [CrossRef]
  46. Xu, H.; Zhang, D.; Wu, F.; Wei, X.; Zhang, J. Deep Desulfurization of Fuels with Cobalt Chloride-Choline Chloride/Polyethylene Glycol Metal Deep Eutectic Solvents. Fuel 2018, 225, 104–110. [Google Scholar] [CrossRef]
  47. Majid, M.F.; Mohd Zaid, H.F.; Kait, C.F.; Jumbri, K.; Yuan, L.C.; Rajasuriyan, S. Futuristic Advance and Perspective of Deep Eutectic Solvent for Extractive Desulfurization of Fuel Oil: A Review. J. Mol. Liq. 2020, 306, 112870. [Google Scholar] [CrossRef]
Figure 1. Schematic of experimental setup.
Figure 1. Schematic of experimental setup.
Chemengineering 08 00102 g001
Figure 2. Extraction efficiency of pyridine, thiophene, and toluene for different DESs (shaker, 24 h and 25 °C).
Figure 2. Extraction efficiency of pyridine, thiophene, and toluene for different DESs (shaker, 24 h and 25 °C).
Chemengineering 08 00102 g002
Figure 3. A photo of a slug flow within a millireactor.
Figure 3. A photo of a slug flow within a millireactor.
Chemengineering 08 00102 g003
Figure 4. Influence of retention time τ on extraction efficiency; ε obtained with TbabFa-0 extraction and S = 1 kg∙kg−1.
Figure 4. Influence of retention time τ on extraction efficiency; ε obtained with TbabFa-0 extraction and S = 1 kg∙kg−1.
Chemengineering 08 00102 g004
Figure 5. Impact of DES-to-feed mass ratio on pyridine, thiophene, and toluene extraction efficiency at a 90 cm millireactor length.
Figure 5. Impact of DES-to-feed mass ratio on pyridine, thiophene, and toluene extraction efficiency at a 90 cm millireactor length.
Chemengineering 08 00102 g005
Figure 6. Volumetric mass transfer coefficient value for TbabaFa-0 as a function of the retention time.
Figure 6. Volumetric mass transfer coefficient value for TbabaFa-0 as a function of the retention time.
Chemengineering 08 00102 g006
Figure 7. The comparison of FTIR spectra of fresh, after-extraction, and regenerated TbabFa-0. The red arrows indicate the peaks of the remaining components after extraction and purification.
Figure 7. The comparison of FTIR spectra of fresh, after-extraction, and regenerated TbabFa-0. The red arrows indicate the peaks of the remaining components after extraction and purification.
Chemengineering 08 00102 g007
Table 1. List and specifications of used chemicals.
Table 1. List and specifications of used chemicals.
ChemicalsM, gmol−1Purity, %Manufacturer
Betaine117.1598.0Lach-Ner
Choline chloride139.6299.0Acros Organics
Citric acid119.1295.00Lach-Ner
Ethylene glycol62.0799.96Lach-Ner
Formic acid46.0398.0Lach-Ner
D-fructose180.1699.00Lach-Ner
Glycerol92.0999.70BDH Prolabo
Isooctane114.2399.8BDH Prolabo
Malic acid134.0999.0+Acros Organics
n-heptane100.2199.0Carlo Erba
n-hexane86.18≥95.0Carlo Erba
Nile red318.37-Acros Organics
Pyridine79.10≥99.0Acros Organics
Tetrabutylammonium bromide322.3699.0+Acros Organics
Thiophene84.1499.0+Acros Organics
Toluene92.1499.96Lach-Ner
Table 2. Compositions and abbreviations of DESs used in this study.
Table 2. Compositions and abbreviations of DESs used in this study.
DESHBAHBDwH2O %Molar Ratio
BGly-0BetaineGlycerol01:3
ChCit-30Choline chlorideCitric acid302:1
ChEg-0Ethylene glycol01:2.5
ChGly-0Glycerol01:2
ChMa-30Malic acid301:1
MaFruGly-30Malic acidFructosa/glycerol01:1:1
TbabFa-0Tetrabutylammonium bromideFormic acid01:1
Table 3. Composition of the model fuel [7].
Table 3. Composition of the model fuel [7].
Componentsw, wt%
n-Hexane26
n-Heptane26
Isooctan26
Pyridine6
Tiophen6
Toluene10
Table 4. Characteristics of millireactor.
Table 4. Characteristics of millireactor.
MaterialPLA
Pipe inner diameter, mm2.5
Pipe length, mm300, 500, 900
Table 5. The retention time τ at different volume flow and millireactor lengths for solvent ratio S = 1.
Table 5. The retention time τ at different volume flow and millireactor lengths for solvent ratio S = 1.
QR, μL∙min1000800400400200400200100200100100
QE, μL∙min668535267267134267134671346767
L, cm5050305030905030905090
τ, min1.601.842.213.684.416.627.358.8213.2414.7126.47
Table 6. Correlations for diffusion coefficients.
Table 6. Correlations for diffusion coefficients.
CorrelationFormulaReferences
Wilke–Chang D A B = 7.4 · 10 8 φ B M B 1 / 2 T μ B V A 0.6 [18]
Stokes–Einstein D A B = 9.96 · 10 16 T μ B · V A 1 / 3 [19,20]
Siddiqi–Lucas D A B = 2.98 · 10 7 · T V A 0.5473 μ B 1.026 [21,22]
Polson D A B = 9.4 · 10 15 T μ B · M A 1 / 3 [19,20]
Scheibel D A B = 8.2 · 10 8 T μ B V A 1 / 3 1 + 3 V B V A 2 / 3 [22,23]
Table 7. Physicochemical properties of used solvents.
Table 7. Physicochemical properties of used solvents.
DESρ, g cm−3μ, Pa spHENR kcal mol−1σ, mN m−1
BGly-01.2372.24158.8351.3749.11
ChCit-301.1960.01680.9349.7041.26
ChEg-01.1190.03564.1451.2449.29
ChGly-01.9860.35106.1351.1152.26
ChMa-30 1.1910.01790.6749.5252.48
MaFruGly-30 1.295 0.0361 0.92 49.73 57.74
TbabFa-0 1.077 1.0874 0.94 51.22 40.99
FCC0.7200.0046 -21.3
ENR represents polarity of solvents.
Table 8. Values of Reynolds, capillary, and Weber numbers of the phases.
Table 8. Values of Reynolds, capillary, and Weber numbers of the phases.
PhaseReCaWeFlow Pattern
FCC2.68–1.34·1011.45·10−5–7.27·10−51.09·10−4–2.73·10−3Slug
TbabFa-07.17·10−4–3.59·10−33.01·10−2–1.50·10−14.03·10−5–1.01·10−3Slug
Table 9. The specific mass of the extracted key component ms for different DES-to-feed mass ratios.
Table 9. The specific mass of the extracted key component ms for different DES-to-feed mass ratios.
S, kg kg−1 0.512
Key ComponentsDESms, gkc kgDES−1
ThiopheneTbabFa-038.3831.2619.97
Pyridine75.4451.4127.59
Toluene26.8624.3321.48
Table 10. Diffusion coefficient calculated for extraction of pyridine, thiophene, and toluene for extraction with TbabFa-0.
Table 10. Diffusion coefficient calculated for extraction of pyridine, thiophene, and toluene for extraction with TbabFa-0.
CorrelationDAB, m2 s−1
ThiophenePyridineToluene
Wilke–Chang1.15∙10−121.12∙10−129.50∙10−13
Stokes–Einstein3.94∙10−123.88∙10−123.54∙10−12
Siddiqi–Lucas3.76∙10−133.66∙10−133.15∙10−13
Polson 3.75∙10−133.83∙10−133.64∙10−13
Scheibel1.33∙10−121.27∙10−121.02∙10−12
Table 11. Comparison of extraction efficiency of fresh and regenerated TbabFa-0.
Table 11. Comparison of extraction efficiency of fresh and regenerated TbabFa-0.
TbabFa-0ε, %
ThiophenePyridineToluene
Fresh86.648.922.59
Regenerated79.746.520.86
Deviation, %7.974.917.66
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Mitar, A.; Prlić Kardum, J.; Lukić, M. Efficient Impurity Removal from Model FCC Fuel in Millireactors Using Deep Eutectic Solvents. ChemEngineering 2024, 8, 102. https://doi.org/10.3390/chemengineering8050102

AMA Style

Mitar A, Prlić Kardum J, Lukić M. Efficient Impurity Removal from Model FCC Fuel in Millireactors Using Deep Eutectic Solvents. ChemEngineering. 2024; 8(5):102. https://doi.org/10.3390/chemengineering8050102

Chicago/Turabian Style

Mitar, Anamarija, Jasna Prlić Kardum, and Marija Lukić. 2024. "Efficient Impurity Removal from Model FCC Fuel in Millireactors Using Deep Eutectic Solvents" ChemEngineering 8, no. 5: 102. https://doi.org/10.3390/chemengineering8050102

APA Style

Mitar, A., Prlić Kardum, J., & Lukić, M. (2024). Efficient Impurity Removal from Model FCC Fuel in Millireactors Using Deep Eutectic Solvents. ChemEngineering, 8(5), 102. https://doi.org/10.3390/chemengineering8050102

Article Metrics

Back to TopTop