Next Article in Journal
Nucleation and Crystal Growth: Recent Advances and Future Trends
Previous Article in Journal
Probabilistic Evaluation of Steel Column Damage under Blast Loading via Simulation Reliability Methods and Gene Expression Programming
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Proceeding Paper

Recent Advances in Extractive Distillation †

by
Althea Francesca Aquilon
1,
Diana Mae Cargullo
1,
Jeremay Onayan
1,
Jgiordana Sarno
1,
Vea Marie Molino
1 and
Edgar Clyde R. Lopez
2,3,*
1
Chemical Engineering Department, College of Engineering, Adamson University, Ermita, Manila 1000, Philippines
2
Nanotechnology Research Laboratory, Department of Chemical Engineering, University of the Philippines Diliman, Quezon City 1101, Philippines
3
Department of Chemical Engineering, University of Santo Tomas, España Blvd., Sampaloc, Manila 1015, Philippines
*
Author to whom correspondence should be addressed.
Presented at the 4th International Electronic Conference on Applied Sciences, 27 October–10 November 2023; Available online: https://asec2023.sciforum.net/.
Eng. Proc. 2023, 56(1), 11; https://doi.org/10.3390/ASEC2023-15261
Published: 26 October 2023
(This article belongs to the Proceedings of The 4th International Electronic Conference on Applied Sciences)

Abstract

:
Distillation is widely recognized as the preferred method for separation due to its operational and control benefits. Traditional distillation processes, however, cannot successfully separate azeotropic mixtures with near boiling points. Numerous special distillation processes have been developed to address this limitation. Extractive distillation, in particular, has gained significant popularity in the chemical, petrochemical, pharmaceutical, and refining industries. This review examined the state-of-the-art advances in extractive distillation. The importance of the proper selection of a solvent was discussed. Several configurations of extractive distillation processes were presented. Additionally, alternative extractive distillation systems have been elaborated. However, significant research gaps remain, such as the need for an exhaustive investigation of various control variables, the impact of certain entrainers on distillation processes, and cost comparisons across specialized distillation systems. Furthermore, process intensification strategies require additional research to solve complexity and operability issues. The integration of energy-efficient technologies, developments in renewable energy consumption, and the development of cost-effective reactive or split distillation columns will shape the future of distillation operations. These advances will help the chemical process sector achieve improved energy efficiency, lower environmental impact, and increased sustainability.

1. Introduction

The benefits of distillation in terms of operation and control make it the most preferred separation method. However, some systems with close boiling points, commonly referred to as azeotropic mixtures, cannot be separated using traditional distillation. This is where special distillation processes like azeotropic distillation, catalytic distillation, extractive distillation, pressure-swing distillation, etc., were developed [1,2]. Among these distillation methods, extractive distillation is commonly used in the chemical, petrochemical, pharmaceutical, and refining sectors.
Conventional extractive distillation involves at least two columns wherein one of the components will go overhead, and the other will go either overhead or at the bottom column, depending on the volatilities of the components. The choice of solvent is a significant additional aspect to consider. It must have a boiling point greater than the components that need to be separated. It must also be miscible and not form an azeotrope.
The high energy demand for extractive distillation technology, like other distillation methods, has an adverse economic and environmental impact [3]. Additionally, improving the extraction method is motivated by the desire to raise the cost, quality, and yield of the extraction process. Numerous improvements have been made over time, either by combining different extraction procedures into a single process, adding more useful techniques, or changing the original methodology.
The ability of extractive distillation to separate azeotropic mixtures and mixtures with low relative volatility in various fields has highlighted its significance, resulting in numerous studies. While some processes reduce energy consumption and improve the economy, they frequently complicate the process, making operability and control more challenging. Because feasibility is still a problem, this review paper will thoroughly discuss and evaluate recent advancements in extraction distillation.

2. Selection of Appropriate Solvent

The extractant or solvent is one of the most important parts of extractive distillation. One of the parameters in choosing the perfect extractant includes its effect on the azeotropic system’s relative volatility, the intermolecular forces at atmospheric temperature and pressure, and electrostatic potential. The perfect extractant can be easily chosen using these parameters; an extractant should be able to increase the relative volatility of the system; in terms of this article, an extractant that has stronger attraction to the undesired substance and weaker attraction to the desired substance is ethanol and ETBE, respectively.
The solvent to be used determines how well an extractive distillation process works. The selected solvent needs specific properties that increase the process’s feasibility by lowering energy and solvent consumption [4,5]. However, some commonly used solvents have not performed well with mixtures exhibiting azeotropes, hence the utilization of ionic liquids (ILs). A low vapor pressure, melting point, and considerable chemical and thermal stability are just a few of the remarkable physicochemical characteristics of ILs. Moreover, there is a need for green technologies using non-hazardous solvents or renewable natural resources while retaining high-quality, safe extracts.
Distinct solvents and flow rates have distinct impacts on the components’ vapor–liquid equilibrium and volatility order. This may be used to assess solvents’ performance and flow rate.
Consider the ternary mixture of acetonitrile, methanol, and benzene, which has three binary azeotropes. Possible solvents include aniline (AN) and chlorobenzene (CB). The equilibrium between vapor and liquid in the pairwise combinations of three components is shown. The vapor–liquid equilibrium curves alter upon the introduction of solvents. Methanol, benzene, and acetonitrile are volatile compounds in decreasing order of light to heavy when using AN as the solvent; methanol, acetonitrile, and benzene are volatile compounds when using CB.

2.1. Supercritical Fluid Extraction (SFE)

Advanced extraction techniques were developed to attain high quality and extraction yield. Supercritical Fluid Extraction (SFE) enhances the conventional extraction method by reducing the amount of organic solvent used. SFE starts by heating and pressurizing the fluid to its critical state before entering the reactor. The solute can readily be sorted out by gravity. Therefore, the extracted component can be collected at the bottom of the separator, and the gas can be recycled or released. Despite being more appealing than other organic solvents like acetone, ethanol, or hexane because of its capacity to surpass conventional extraction methods, SFE has a disadvantage because it needs high pressure, which increases the operating cost due to additional safety requirements.

2.2. Extractive Distillation Using Ionic Liquids (ILs)

Various research has been conducted in the study of the behavior of ILs when they are used as a solvent for extractive distillation because of their properties. The separation of benzene and cyclohexane, which have similar boiling points and are inclined to combine into an azeotropic mixture, is one of the most difficult operations in the petrochemical industry. Adsorption, pervaporation, and liquid–liquid extraction are only a few examples of the many separation procedures that have been attempted. However, extractive distillation employing organic solvents as solvents is discovered to be the most effective method.
Ionic liquids are used as solvents, especially in the separation of hydrocarbons, because they possess a non-volatile nature and high selectivity around aromatic compounds. The performance of seven ionic liquids for C6H6 and C6H12 separation by ED was investigated. The optimum solvent for this separation has been determined to be the [4bmpy][TCM]. The C6H12/C6H6, the relative volatility of this ionic liquid, is at its peak at 363.2 K, and even though it is at its lowest at 403.2 K, it is sufficient to carry out the separation satisfactorily. Adding these ionic liquids increases the C6H12/C6H6 relative volatility, reaching values approximating 15, and breaking the azeotrope. The 1-Ethyl-3-methyl-imidazolium-dicyanamide was selected as the solvent in the EDC due to the IL’s strong influence on the relative volatility of aromatic–aliphatic mixtures and its high thermal stability. According to the results, liquid–liquid extraction and extractive distillation do not separate.
Some compounds used for various industrial applications require a certain purity level as raw materials. An instance of this is ethanol, which requires 99.5% purity before it can be used for industries such as paint, cosmetics, and perfumery. The use of ILs shows the potential to decrease energy consumption. Compared with conventional extractive distillation, ionic liquid-based extractive distillation (ILED), which produces anhydrous ethanol, uses less energy. The TAC value of $0.73 × 106 illustrates how the ILED discovered during the sensitivity analysis for the least energy consumption outperformed alternative setups.
Additionally, much study has been done to compare which IL best suits a specific system. As solvents, 1-Butyl-3-methylimidazolium trifluoromethanesulfonate or [BMIM][CF3SO3], 1-Ethyl-3-methylimidazolium tetrafluoroborate or [EMIM][BF4], and 1-Ethyl-3-methylimidazolium methanesulfonate or [EMIM][MeSO3] are used. The best option among all these ILs for extractive distillation is the stripper with 1-Ethyl-3-methylimidazolium methanesulfonate, or [EMIM][MeSO3]. The researchers used three methods: an extractive distillation column, a flash tank, and a stripper. At the EDC, IL raises the relative volatility of ethanol acetate compared to ethanol. Additionally, it causes the ethanol to flow downward and the ethyl acetate to go upward. Due to the lower volatility of the IL, it just flows down the column, leaving the bottom product.
One study employed [MMIM][DMP], 1,3-dimethylimidazolium, to separate the isopropanol–water mixture. The approach using [MMIM][DMP] as a solvent reduced TAC by 7.92%, as opposed to the process using a normal solvent, it was discovered. Because of their outstanding qualities, ILs can be used as better solvents. The success of the ED processes was examined considering the effects on the economy, energy efficiency, and environment using the thermodynamic efficiency index, CO2 emissions, and TAC [6].
The separation process is generally optimized to improve the EDC’s operating conditions and design parameters. Due to a broad selection of potential solvents, extractive distillation works effectively in the chemical industry to separate binary or multiple azeotropes, but it has several drawbacks, including energy consumption. This study offered two solutions as to how to do this. The first is process intensification using heat-integrated distillation, divided wall columns, the thermal coupling of columns, and heat pump-assisted distillation. The second method involves using a green solvent with great selectivity.
Triethylene glycol (TEG) and the solvents [EMIM][Cl] were used in studies that used the t-butanol dehydration method. According to the findings, utilizing [EMIM][Cl] as an entrainer as opposed to TEG could lower the total annual cost (TAC) by 13.9% [6].

2.3. Ionic Liquid-Based Mixed Solvents

Conventional ternary extractive distillation (CTED) uses different solvents, such as EG and ionic liquids. EG and ionic liquids are the common solvents used in extractive distillation, wherein EG significantly impacts relative volatility. Energy consumption and operating costs are one of the disadvantages of extractive distillation. An entrainer was recently developed, combining ionic liquids with another solvent in one shell, which uses a dividing wall column. Conventional ternary extractive distillation (CTED) using EG and ionic liquid solvents provide a good outcome, but ionic liquid-based mixed results are better. Conventional ternary extractive distillation, dividing wall column, and extractive distillation column reduce energy consumption and operating costs, but extractive dividing wall column ternary extractive distillation is the most effective in saving energy, as it reduces up to 7.81% [7].

3. Optimized Configurations for Extractive Distillation

Since different types of systems can be separated, researchers have studied various kinds of strategies over the years, and the separation process is not a type of process that applies to all kinds of systems; therefore, all of these strategies are efficient in different ways for different kinds of systems [8]. These methods use innovative techniques to provide a more effective way of recovering desired materials without the constraints that are encountered using conventional ways [9].
Different methods for achieving energy efficiency are being developed as distillation process designs grow more energy conscious. Some of the processes used in distillation to become energy efficient are feed flashing, multi-effect distillation, vapor compression, and heat-integrated distillation column. Feed flashing incorporates a feed-splitting vessel that creates a two-phase feed directed to their respective feed plates; this reduces the energy requirement and allows for flexibility in variations in thermodynamic modeling. Heat integration can be employed in various ways, such as heat exchange between the feed and rundown or pump-around to optimize heat recovery. Multi-effect distillation integrates the condenser and reboiler of two columns with different operating pressures. This method uses the heat the high-pressure condenser releases to operate the low-pressure reboiler. This method saves 40% of energy costs. Heat-integrated distillation columns operate in a single unit that consists of a high-pressure section (top/rectifying) and a low-pressure section (bottom/stripping); this difference in pressure allows for heat transfer between the columns [10,11]. This type of method reduces reboiler and condenser duties and significantly reduces energy consumption down to 50% [12].
Other ways of lowering energy requirements are by intensifying the process of distillation. These include reactive distillation, dividing wall column utilization, and hybrid separation processes. The wall columns separate the feed, side stream, and product streams. The product is extracted at the top of the column, and the feed and side streams are on each side of the wall, preventing contamination. This intensification method reduces capital costs and energy by up to 45%. A hybrid separation method combines two distinct units into a single process. This is applied to a design wherein one unit overcomes the limitation of the other working in a mutually beneficial relationship [13]. In azeotropic distillation, commonly used methods include extractive and pressure-swing distillation, which are energy-intensive processes. In these cases, hybrid distillation is recommended.

3.1. Classic, Modified, and Intensified Extractive Distillation

Several separation techniques can be identified while concentrating on extractive distillation, which is thought to be the most researched type of distillation employed in the manufacture of bioethanol [8]. The classic extractive distillation configuration is shown in Figure 1.
There are alternatives to the classic ED that had a higher recovery by using a recycle stream in the solvent recovery column. (Figure 2). This set of configurations was the starting point for creating various complex designs. Thermally coupled systems and extractive divided wall columns (DWC) were produced from these. These types of extractive distillation have been studied since 2012, and reports and comparative analyses have been written. Figure 3 shows an example of intensified extractive divided wall configuration wherein it utilizes a reduced number of columns compared to the classic configuration, which uses three columns; the intensified DWC uses only two or one. In a two-column configuration, the separation and solvent recovery occur on the second column using an extractive DWC; the distillate is collected in the partition and flows through a recycle stream back to the first column. This configuration ensures there are no ethanol losses and provides a higher recovery. Moreover, it reduces capital costs by 23% while having the same energy consumption as a regular configuration.

3.2. Assisted Single-Column Extractive Distillation with a Heavy Entrainer

Figure 4 proposes an extractive distillation with one column and a heavy entrainer (FT-SCED) [14]. It allows for pure light and heavy compounds to be produced from the flash tank and column overheads. In this review study, we will concentrate on the proposed design in the first case, which is the separation of an ethanol and water mixture with a heavy entrainer, ethylene glycol. In comparison to extractive dividing wall columns and conventional two-column extractive distillation, it was determined that the new process had high energy efficiencies and low investment costs. A control study also revealed that the FT-SCED process could be effectively managed. When ±20% feed flow rate and ±10% composition disturbances were introduced, the purities of the product returned to its original setting for about 10 h, indicating its effectiveness.

3.3. Ternary Extractive Distillation

Conventional distillation (CD) systems are inefficient in terms of energy due to the remixing of the components [7]. Numerous research has recommended the use of thermocouples to address this issue. The first column of the distillation is then proposed to be managed by thermally coupled systems with side rectifiers and dividing-wall columns. Operating issues are unavoidable; hence, a modified extractive distillation structure with thermal couplings (TCD) was proposed to address probable issues.

3.3.1. Alternative Extractive Distillation Systems

Extending the previously mentioned ideas, which are depicted in Figure 5a, an alternative system (AD) was proposed, which is depicted in Figure 5b and does not have a thermocouple. In the alternative configuration, a second column with one feed was used in the place of the vapor–liquid interface, and a stripping section was added to the side rectifier to complete the separation process. While the second column (C2) creates the intermediate as the top stream, the first column (C1) becomes the side stream column, where the light component is eliminated as the distillate. The solvent is then recycled to help with the separation and is present in the bottom streams from the C1 and C2. Thermally coupled and conventional distillation are reported to have five parts; however, the alternative system has six sections. The goal of this additional section, designated as S5 in Figure 5b, is the same as the task of the bottom part of the TCD system, which is purification.
The new alternative system was applied to three case studies to prove its efficiency: ethanol dehydration, heptane–toluene–aniline mixture, and acetone–methanol–water system. In the first case, it was determined that AD is advantageous to CD and TCD because it saves 13.50% more energy, resulting in a 12.40% reduction in the total annual cost. Also, CO2 emissions are reduced by 10.60% compared to the CD and 3.20% compared to TCD. Similar outcomes were also seen in the other two cases, demonstrating that AD can be more energy efficient and offer a more sustainable alternative when compared to CD and TCD.
A structure that implements heat integration between columns for additional energy and CO2 emissions saving was also proposed [3]. It was implemented in AD and CD, as seen in Figure 6. The results showed that the AD system had superior potential for energy integration, since it resulted in a reduction in energy consumption compared to the CD system and a reduction compared to the conventional system with heat integration. Such an effect leads to a decreased annual total cost. When heat integration between AD columns is used, CO2 emissions are reduced by 11% compared to the CD with heat. This is equivalent to a 39% reduction compared to the CD that is not integrated for extractive distillation.

3.3.2. Three-Column Extractive Distillation

A three-column extractive distillation (TCED) is usually used to separate azeotropic mixtures. TCED processes work in two types: XYX and XXY. The difference between the two types is shown in Figure 7 and Figure 8, wherein D3 and B1 are equal for XYX, and D3 and B1 for XXY are not equal. As shown in both figures, condenser 1 has many roles.
The feed split method is developed to help the distillate flow, feed flow, and the role of the reboiler to work together, which lessens the work for condenser 1. The applicability and feasibility of the feed-split method work in cases. XYX has six cases (EG, DMSO, and cases concerning the value of δF) and two common cases for XXY (using DMSO and NMP). These different cases provide the result if the feed-split method is applicable. It was discovered that this method works in adverse outcomes for XYX and can be used for saving energy as per XXY type [15].

3.3.3. Comparison of Conventional Ternary Extractive Distillation and VRHP

Conventional ternary extractive distillation (CTED) works directly (DCED) and indirectly (ICED). These two CTEDs are usually used to separate the binary azeotrope, which is considered expensive. Based on the calculation in this study, economic and environmental ICED performance is much better than DCED. Since it is expensive, a new method known as a vapor recompression heat pump (VRHP) is being studied to improve economic efficiency and environmental performance [16]. VRHP works directly (VRHP- DCED) and indirectly (VRHP-ICED). The calculation for ICED and CED does not correlate with VRHP-DCED and VRHP-ICED, wherein six different separation sequences for both VRHP are determined. For VRHP-DCED, sequence V provides the lowest steam costs, and four sequences (I, II, III, IV) are much better in terms of annual cost than CED sequences, and four sequences (I, II, III, IV) are better for environmental efficiency. However, II’ and IV’ made VRHP-ICED more environmentally efficient than CED sequences and I,’ IV,’ and VI’ sequences [17].

3.3.4. Extractive Distillation with Two Columns and Rectifier

The dynamics of a traditional three-column configuration are compared with a non-conventional two-column configuration with a side rectifier since a conventional ternary extractive process requires three columns, but thermally coupled side stream/rectifier columns have demonstrated economic advantages. Furthermore, it is necessary to investigate dynamic controllability because process disruptions are unavoidable and may affect product quality, safety, and environmental concerns.
Both feed flow rate and feed composition disturbances are compared between the conventional and non-conventional configurations. A 1-min dead time was employed, and a disturbance of ±10% was introduced. The non-conventional process’s dynamic response was observed to be inferior to the conventional process. It is possible to conclude that, although the non-conventional process has a lower energy cost than the conventional process, it still needs to be studied because of how poorly it responds dynamically to disturbances.

3.3.5. Comparison of Heterogeneous Azeotropic Distillation and Extractive Distillation Methods for Ternary Azeotrope Ethanol/Toluene/Water Separation

Comparisons are made between two techniques for separating the ternary azeotrope ethanol/toluene/water. These processes are extractive distillation and heterogeneous azeotropic distillation. This study covered four solvents for extractive distillation, including ethylene glycol, glycerol, N-propylbenzene, and n-butylbenzene. The authors’ solvent of choice was glycerol. Due to the immiscibility of water and toluene, only two columns are needed after selecting the solvent [18].
It was discovered that ED had a lower cost of energy than HAD—to be more precise, 18.8% and 39.3%, respectively [6]. Overall, it was found that in terms of economics, ED is more desirable than partially heat-integrated HAD.

3.4. Hybrid Reactive-Extractive Distillation Configurations

Three processes were used in this study [19], namely the common process (CP), reactive extractive dividing wall process (REDWP), and reactive extractive distillation process (REDP). Process design optimization was created using a simulation software called Aspen. The diagrams of each process discussed above are shown below, respectively.
The common process (Figure 9) utilizes three distillation columns. The REDP (Figure 10) is a modified process of the classic one, where the distillation and reactor columns are combined to form a device with azeotropic distillation. The REDWP (Figure 11) uses a single column for the azeotropic and extractive distillation process. Economic optimization results showed that the REDWP outperformed the REDP and CP, and it amounted to a total cost of 15% and 9% less than the REDP and CP, respectively. In terms of environmental optimization, REDWP again stood higher against its counterparts, showing overall remarkable results regarding greenhouse gas emissions and water pollution as well as environmental benefits such as lower biological toxicity. Regarding exergy losses, REDP and REDWP showed that these processes had great thermodynamic efficiency and less exergy loss.

4. Research Gaps

Reactive distillation, which combines chemical reaction and distillation into a single unit, has emerged as a promising technology with the potential to transform chemical processes. Reactive distillation offers benefits such as enhanced yield, lower energy usage, and streamlined process flows by integrating these two usually independent processes. Despite the advances achieved in the field, there are still significant research gaps that must be addressed in order to fully realize its promise and facilitate its widespread application in industry.
Although several authors have investigated various grounds for studying, some space remains left for further research. For one, there is usually a lack of parameters to place under study. For instance, the optimization process only focused on a single factor, such as pressure [20]. Only one article [21] reported on reduced capital costs by utilizing different configuration configurations of extractive distillation columns. One could further obtain data and information on the differences in process controls, parameters, and detailed information on recovery. In [22], entrainer selection is another one that is discussed; however, there is a lack of research on how a specific entrainer affects the distillation process [14].
One study [23] showed that, even though computer-aided molecular design is a new method for efficiency, two aspects must be considered in screening solvents: price and source. Intensification and integration techniques are used for capital cost and energy consumption but are much more complex than traditional processes. Another study [14] discovered that a vapor recompression heat pump (VRHP) is better than conventional ternary extractive distillation. As specialized distillations are created, a comparison of specialized and vapor recompression heat pumps should be made regarding their financial costs and environmental performances.
With regard to process intensification, a study [2] showed that, even though some process intensification technologies increase economic efficiency and reduce energy consumption, they also make the process more complex than it would otherwise be and make it harder to operate and monitor. Another study [24] indicated that different special distillations reduce energy consumption using inorganic salts as the entrainer. More research should be conducted regarding thermally coupled distillation columns with inorganic salt entrainers.
Another major research gap in reactive distillation is the development of comprehensive kinetic models that adequately represent the complicated chemical processes that occur within the distillation column. Many existing models are very simplistic, assuming perfect performance and ignoring the influence of variables like as temperature and pressure fluctuations down the column. More precise and thorough kinetic models are required for improving reactor design, predicting product quality, and ensuring the reactive distillation process’s stability.
In reactive distillation, the intricate interplay between mass transfer and chemical reaction kinetics remains an extremely challenging field of research. Understanding how mass transfer limitations impact reaction rates and selectivity is critical when designing efficient reactive distillation columns. The influence of parameters such as catalyst distribution, phase equilibrium, and column geometry on mass transfer within the reactive distillation system should be studied.
A significant research gap exists in the behavior of multiphase systems within reactive distillation columns. Understanding fluid dynamics, especially liquid-liquid and vapor-liquid interactions, is critical for improving column design and operation. Current research frequently falls short of the required depth in investigating the intricate fluid flow patterns and phase distributions that influence the overall performance of reactive distillation systems.
The lack of established design standards for reactive distillation is a significant problem for both researchers and industry practitioners. Establishing thorough standards that take into account varied reaction systems, feed compositions, and operating circumstances is critical for guaranteeing consistency and dependability in reactive distillation process design and scale-up. To close the research gap, academia and industry must work together to establish globally applicable design concepts.
While reactive distillation has demonstrated promise in the laboratory, there is a significant research gap in understanding the issues associated with scaling these processes up for commercial applications. To aid the effective transition of reactive distillation from the laboratory to full-scale production, factors such as heat and mass transfer restrictions, potential equipment failures, and the effects of large-scale operation on reaction kinetics must be extensively explored.
To summarize, while reactive distillation has immense promise for enhancing chemical processing, various research gaps must be addressed before its full potential can be realized. Researchers encounter several hurdles, ranging from the development of accurate kinetic models to the understanding of fluid dynamics in multiphase systems. Bridging these gaps necessitates multidisciplinary cooperation, novel techniques, and a collaborative effort from academics and industry. By filling these research gaps, we can pave the road for reactive distillation to become more widely used and contribute to the development of more sustainable and efficient chemical processes.

5. Future Outlook

Technological breakthroughs are expected to affect the future of reactive distillation. Process control and automation are two important areas of growth. Artificial intelligence and machine learning algorithms in advanced control systems can optimize the operation of reactive distillation columns in real-time. This not only improves process efficiency but also increases reactive distillation’s responsiveness to changing feedstock compositions and reaction conditions. Furthermore, the use of online sensors and advanced analytics will be critical in monitoring and improving reactive distillation operations. These technologies allow for real-time data analysis, enabling for faster decision-making and proactive modifications to maintain optimal performance and product quality.
The future of reactive distillation will most likely see an increase in R&D focusing on developing and enhancing catalytic systems particularly designed for the unique conditions within distillation columns. For reactive distillation operations to be as efficient as possible, advanced catalysts with better stability, selectivity, and resistance to deactivation will be required. Furthermore, research into multifunctional catalysts capable of catalyzing numerous reactions at the same time might lead to more streamlined and varied processes. This approach is consistent with the increased emphasis on sustainability since it allows for the design of processes with fewer reaction stages and lower environmental impact.
As the global chemical industry shifts toward sustainability, the integration of renewable feedstocks is anticipated to become more important in the future of reactive distillation. Because of the adaptability of reactive distillation, it can handle a wide range of feedstock compositions, including those obtained from biomass or other renewable sources. This transition toward sustainable feedstocks is consistent with the industry’s overall trend toward green and circular chemistry. Furthermore, the development of reactive distillation techniques for converting waste streams into useful products has the potential to contribute to the circular economy. By utilizing reactive distillation’s capacity to simultaneously execute reactions and separate products, the technique may play a critical role in converting waste into a resource, minimizing environmental impact and increasing resource efficiency.
Because of its capacity to allow complicated reaction sequences and create high-purity products, reactive distillation is poised to be an important player in the future of customized and specialized chemical manufacturing. The technology’s flexibility enables the efficient synthesis of a diverse spectrum of specialized compounds with specific features. This customization is especially useful in sectors like pharmaceuticals, where precise control of reaction conditions and product purity is critical. Reactive distillation may find more use in pharmaceutical manufacture in the future, allowing for more efficient and sustainable production procedures. This shift is being driven not just by economic reasons, but also by a regulatory push in the pharmaceutical industry toward greener and more efficient production techniques.
The integration of reactive distillation with other novel technologies offers a potential development horizon. Hybrid processes that combine reactive distillation with techniques such as membrane separation or sophisticated separation technologies have the potential to improve overall process efficiency and minimize energy usage. Combining reactive distillation with membrane technology, for example, can increase selectivity by selectively permeating specific components, resulting in improved product purity. Exploring and enhancing these hybrid techniques will be critical for tackling specific difficulties and unlocking new potential for process intensification.
Ultimately, the future of reactive distillation is defined by a convergence of scientific improvements, sustainability objectives, and growing industrial application. From improved catalytic systems to the integration of renewable feedstocks and the development of hybrid processes, the reactive distillation trajectory has considerable potential for transforming the landscape of chemical processing. As the chemical industry moves toward more ecological and efficient methods, reactive distillation stands out as a versatile and strong technology capable of fulfilling future needs. The entire potential of reactive distillation may be achieved via continued research, innovation, and cooperation, contributing to a more sustainable and resilient chemical industry in the years to come.

Author Contributions

Conceptualization, E.C.R.L.; writing—original draft preparation, A.F.A., D.M.C., J.O., J.S. and V.M.M.; writing—review and editing, E.C.R.L.; supervision, E.C.R.L.; project administration, E.C.R.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Saha, A.; Bhattacharjee, S. Research and Review: Based on Distillation: Distillation Types and Advancements: A Short Summary. J. Adv. Pharm. Pract. 2019, 1, 8–16. [Google Scholar]
  2. Ma, K.; Yu, M.; Dai, Y.; Ma, Y.; Gao, J.; Cui, P.; Wang, Y. Control of an energy-saving side-stream extractive distillation process with different disturbance conditions. Sep. Purif. Technol. 2019, 210, 195–208. [Google Scholar] [CrossRef]
  3. Tututi-Avila, S.; Medina-Herrera, N.; Hahn, J.; Jiménez-Gutiérrez, A. Design of an energy-efficient side-stream extractive distillation system. Comput. Chem. Eng. 2017, 102, 17–25. [Google Scholar] [CrossRef]
  4. Yin, C.; Liu, G. Optimization of Solvent and Extractive Distillation Sequence Considering Its Integration with Reactor. Processes 2021, 9, 565. [Google Scholar] [CrossRef]
  5. Kong, Z.Y.; Segovia-Hernández, J.G.; Lee, H.-Y.; Sunarso, J. Are process-intensified extractive distillation always energetically more efficient? Chem. Eng. Process. Process Intensif. 2022, 181, 109131. [Google Scholar] [CrossRef]
  6. Ma, S.; Shang, X.; Li, L.; Song, Y.; Pan, Q.; Sun, L. Energy-saving thermally coupled ternary extractive distillation process using ionic liquids as entrainer for separating ethyl acetate-ethanol-water ternary mixture. Sep. Purif. Technol. 2019, 226, 337–349. [Google Scholar] [CrossRef]
  7. Li, G.; Liu, S.; Yu, G.; Dai, C.; Lei, Z. Extractive distillation using ionic liquids-based mixed solvents combined with dividing wall column. Sep. Purif. Technol. 2021, 269, 118713. [Google Scholar] [CrossRef]
  8. Edwiges, T.; Roda-Serrat, M.C.; Segovia-Hernández, J.G.; Sánchez-Ramírez, E.; Tronci, S.; Errico, M. Sustainable bioalcohol production: Pretreatment, separation, and control strategies leading to sustainable processes. In Biofuels and Biorefining; Elsevier: Amsterdam, The Netherlands, 2022; pp. 41–85. [Google Scholar] [CrossRef]
  9. Selvamuthukumaran, M.; Shi, J. Recent advances in extraction of antioxidants from plant by-products processing industries. Food Qual. Saf. 2017, 1, 61–81. [Google Scholar] [CrossRef]
  10. Luyben, W.L. Control of heat-integrated extractive distillation processes. Comput. Chem. Eng. 2018, 111, 267–277. [Google Scholar] [CrossRef]
  11. Li, X.; Geng, X.; Cui, P.; Yang, J.; Zhu, Z.; Wang, Y.; Xu, D. Thermodynamic efficiency enhancement of pressure-swing distillation process via heat integration and heat pump technology. Appl. Therm. Eng. 2019, 154, 519–529. [Google Scholar] [CrossRef]
  12. Kooijman, H.A.; Sorensen, E. Recent advances and future perspectives on more sustainable and energy efficient distillation processes. Chem. Eng. Res. Des. 2022, 188, 473–482. [Google Scholar] [CrossRef]
  13. Yang, A.; Kong, Z.Y.; Sunarso, J. Design and optimisation of novel hybrid side-stream reactive-extractive distillation for recovery of isopropyl alcohol and ethyl acetate from wastewater. Chem. Eng. J. 2023, 451, 138563. [Google Scholar] [CrossRef]
  14. Yan, H.; Song, X.; Xu, L.; Yuan, X. New single-column extractive distillation with heavy entrainer to separate binary azeotropic mixtures. Sep. Purif. Technol. 2023, 312, 123393. [Google Scholar] [CrossRef]
  15. Duan, C.; Li, C. Process improvement for three-column extractive distillation by feed split. Sep. Purif. Technol. 2022, 297, 121467. [Google Scholar] [CrossRef]
  16. Parhi, S.S.; Pramanik, A.; Rangaiah, G.P.; Jana, A.K. Evolutionary Algorithm Based Multiobjective Optimization of Vapor Recompressed Batch Extractive Distillation: Assessing Economic Potential and Environmental Impact. Ind. Eng. Chem. Res. 2020, 59, 5032–5046. [Google Scholar] [CrossRef]
  17. Wang, C.; Zhuang, Y.; Liu, L.; Zhang, L.; Du, J. Heat pump assisted extractive distillation sequences with intermediate-boiling entrainer. Appl. Therm. Eng. 2021, 186, 116511. [Google Scholar] [CrossRef]
  18. Zhao, L.; Lyu, X.; Wang, W.; Shan, J.; Qiu, T. Comparison of heterogeneous azeotropic distillation and extractive distillation methods for ternary azeotrope ethanol/toluene/water separation. Comput. Chem. Eng. 2017, 100, 27–37. [Google Scholar] [CrossRef]
  19. Jiao, Y.; Yin, K.; Liu, T.; Meng, F.; Li, X.; Zhong, L.; Zhu, Z.; Cui, P.; Wang, Y. Process design and mechanism analysis of reactive distillation coupled with extractive distillation to produce an environmentally friendly gasoline additive. J. Clean Prod. 2022, 369, 133290. [Google Scholar] [CrossRef]
  20. Zhang, X.; Li, X.; Li, G.; Zhu, Z.; Wang, Y.; Xu, D. Determination of an optimum entrainer for extractive distillation based on an isovolatility curve at different pressures. Sep. Purif. Technol. 2018, 201, 79–95. [Google Scholar] [CrossRef]
  21. Zhu, Z.; Ri, Y.; Jia, H.; Li, X.; Wang, Y.; Wang, Y. Process evaluation on the separation of ethyl acetate and ethanol using extractive distillation with ionic liquid. Sep. Purif. Technol. 2017, 181, 44–52. [Google Scholar] [CrossRef]
  22. Gerbaud, V.; Rodriguez-Donis, I.; Hegely, L.; Lang, P.; Denes, F.; You, X. Review of extractive distillation. Process design, operation, optimization and control. Chem. Eng. Res. Des. 2019, 141, 229–271. [Google Scholar] [CrossRef]
  23. Sun, S.; Lü, L.; Yang, A.; Wei, S.; Shen, W. Extractive distillation: Advances in conceptual design, solvent selection, and separation strategies. Chin. J. Chem. Eng. 2019, 27, 1247–1256. [Google Scholar] [CrossRef]
  24. de Jesús Hernández-Hernández, E.; Cabrera-Ruiz, J.; Hernández-Escoto, H.; Gutiérrez-Antonio, C.; Hernández, S. Simulation study of the production of high purity ethanol using extractive distillation: Revisiting the use of inorganic salts. Chem. Eng. Process. Process Intensif. 2022, 170, 108670. [Google Scholar] [CrossRef]
Figure 1. Classic extractive distillation. Reprinted with permission from [8]. Copyright (2022) Elsevier.
Figure 1. Classic extractive distillation. Reprinted with permission from [8]. Copyright (2022) Elsevier.
Engproc 56 00011 g001
Figure 2. Modified extractive distillation. Reprinted with permission from [8]. Copyright (2022) Elsevier.
Figure 2. Modified extractive distillation. Reprinted with permission from [8]. Copyright (2022) Elsevier.
Engproc 56 00011 g002
Figure 3. Intensified extractive distillation. Reprinted with permission from [8]. Copyright (2022) Elsevier.
Figure 3. Intensified extractive distillation. Reprinted with permission from [8]. Copyright (2022) Elsevier.
Engproc 56 00011 g003
Figure 4. Proposed single-column extractive distillation. Reprinted with permission from [14]. Copyright (2023) Elsevier.
Figure 4. Proposed single-column extractive distillation. Reprinted with permission from [14]. Copyright (2023) Elsevier.
Engproc 56 00011 g004
Figure 5. (a) Extractive distillation system with thermal coupling system and side rectifier (b) Proposed alternative extractive distillation system. Reprinted with permission from [3]. Copyright (2017) Elsevier.
Figure 5. (a) Extractive distillation system with thermal coupling system and side rectifier (b) Proposed alternative extractive distillation system. Reprinted with permission from [3]. Copyright (2017) Elsevier.
Engproc 56 00011 g005aEngproc 56 00011 g005b
Figure 6. Heat integration: (a) Conventional system; (b) Alternative system. Reprinted with permission from [3]. Copyright (2017) Elsevier.
Figure 6. Heat integration: (a) Conventional system; (b) Alternative system. Reprinted with permission from [3]. Copyright (2017) Elsevier.
Engproc 56 00011 g006
Figure 7. XYX Type. Reprinted with permission from [15]. Copyright (2022) Elsevier.
Figure 7. XYX Type. Reprinted with permission from [15]. Copyright (2022) Elsevier.
Engproc 56 00011 g007
Figure 8. XXY Type. Reprinted with permission from [15]. Copyright (2022) Elsevier.
Figure 8. XXY Type. Reprinted with permission from [15]. Copyright (2022) Elsevier.
Engproc 56 00011 g008
Figure 9. Common process of ethyl tert-butyl ether production. Reprinted with permission from [19]. Copyright (2022) Elsevier.
Figure 9. Common process of ethyl tert-butyl ether production. Reprinted with permission from [19]. Copyright (2022) Elsevier.
Engproc 56 00011 g009
Figure 10. Reactive extractive distillation process. Reprinted with permission from [19]. Copyright (2022) Elsevier.
Figure 10. Reactive extractive distillation process. Reprinted with permission from [19]. Copyright (2022) Elsevier.
Engproc 56 00011 g010
Figure 11. Reactive extractive dividing wall process. Reprinted with permission from [19]. Copyright (2022) Elsevier.
Figure 11. Reactive extractive dividing wall process. Reprinted with permission from [19]. Copyright (2022) Elsevier.
Engproc 56 00011 g011
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

Aquilon, A.F.; Cargullo, D.M.; Onayan, J.; Sarno, J.; Molino, V.M.; Lopez, E.C.R. Recent Advances in Extractive Distillation. Eng. Proc. 2023, 56, 11. https://doi.org/10.3390/ASEC2023-15261

AMA Style

Aquilon AF, Cargullo DM, Onayan J, Sarno J, Molino VM, Lopez ECR. Recent Advances in Extractive Distillation. Engineering Proceedings. 2023; 56(1):11. https://doi.org/10.3390/ASEC2023-15261

Chicago/Turabian Style

Aquilon, Althea Francesca, Diana Mae Cargullo, Jeremay Onayan, Jgiordana Sarno, Vea Marie Molino, and Edgar Clyde R. Lopez. 2023. "Recent Advances in Extractive Distillation" Engineering Proceedings 56, no. 1: 11. https://doi.org/10.3390/ASEC2023-15261

Article Metrics

Back to TopTop