Next Article in Journal
The Activity of Two Different Solutions against Selected Fungal Phyto-Pathogens
Previous Article in Journal
Precision Warriors: Nanotechnology’s Triumph in Cancer Therapy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Proceeding Paper

Recent Advances in Reactive Distillation †

by
Demi Andrei Barrientos
1,
Beatrice Fernandez
1,
Rachel Morante
1,
Hannah Ruth Rivera
1,
Karen Simeon
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), 99; https://doi.org/10.3390/ASEC2023-15278
Published: 26 October 2023
(This article belongs to the Proceedings of The 4th International Electronic Conference on Applied Sciences)

Abstract

:
Reactive distillation (RD) combines chemical reactions and separation in a single unit essential to equilibrium-limited reactions. This new technique encompasses multiple advantages over traditional processes, including lower operating costs, increased thermal energy efficiency, high product selectivity, high purity percentage, and lower environmental impact. This paper provided an overview of the features, industrial applications, and industrial perspective of advanced reactive distillation technologies (ARDTs). This study focused on five under-development ARDTs: reactive dividing wall column (R-DWC), reactive high-gravity distillation (R-HiGee), reactive heat-integrated distillation column (R-HIDiC), catalytic cyclic distillation (CCD), and membrane-assisted reactive distillation (MA-RD). The primary drivers for new RD applications are reduced number of vessels, reduced residence time and holdup volume, increased mass and heat transfer, overcoming azeotropes, and prefractionation or impurity removal. ARDT’s potential has yet to be studied, and research remains active to improve it further by investigating other RD technologies, simulation, and optimization techniques.

1. Introduction

Reactive distillation (RD) is an integrated process that combines chemical reactions and distillation in a single unit. This integration offers several advantages over conventional processes, including reduced capital and operating costs, improved energy efficiency, and reduced environmental impact. In recent years, there have been many advances in RD technology, further enhancing its potential for industrial applications. One of the most significant advances in RD technology is the development of reactive dividing-wall columns (R-DWCs). R-DWCs are columns that are divided into two sections via a vertical barrier. The reaction section is located on one side of the barrier, and the distillation section is located on the other side. This configuration allows for the reaction and separation to be carried out in a single column, while maintaining the desired separation between the products.
Another important advancement in RD technology is the development of catalytic distillation (CD). CD is a type of RD in which a catalyst promotes the reaction. The use of a catalyst can significantly improves the rate of the reaction, which can lead to increased productivity and improved selectivity.
In addition to R-DWCs and CD, there have been several other advances in RD technology in recent years. These advances include the development of new types of reactors, new separation techniques, and new control strategies. These advances have made RD an increasingly attractive option for various industrial applications. In a study, reactive distillation, or catalytic distillation, is one of the most successful examples of process intensification [1].
Reactive distillation (RD) is a process intensification technology that integrates or combines the processes of reaction and separation that both exist in a single distillation column. This integration offers several advantages over conventional processes, which includes the development of new types of reactors, such as microreactors and fluidized bed reactors, which offer improved mass transfer and reaction rates, the development of new separation techniques, such as membrane separation and extractive distillation, which can improve the selectivity of the reaction, and lastly the development of new control strategies, which can improve the efficiency and operability of RD processes. RD has been successfully applied to various reactions, including esterification, hydration, etherification, and polymerization. In recent years, there has been a growing interest in developing advanced reactive distillation technologies (ARDTs) that can further improve the performance of RD processes.
Researchers have focused on developing advanced catalysts and reactive packing materials to enhance reaction kinetics and selectivity within the distillation column. It has been demonstrated that the use of structured catalysts in reactive distillation for the synthesis of high-value chemicals resulted in improved conversion rates and selectivity [2].
Advancements in process intensification techniques, such as heat integration and intensified mass transfer, have enabled the design of more efficient reactive distillation processes. In a previous study, a novel multifunctional column design incorporating reactive distillation, extractive distillation, and heat pump technologies was proposed, leading to significant energy savings and improved process performance [2].
Reactive distillation offers the unique advantage of combining reaction and separation processes in a single distillation column unit, eliminating the need for separate reactors and distillation columns. This integration reduces equipment size, lowers capital costs, and increases process efficiency. Reactive distillation has demonstrated substantial productivity improvements, making it a valuable technology for the chemical industry [3].

2. Recent Advances in Reactive Distillation

Distillation is a prevalent separation technique in the chemical industry, albeit a significant energy consumer. Since the mid-20th century, scholarly literature and patents on enhancing reaction and separation equipment design have prioritized energy conservation and economic efficacy [4]. Reactive distillation (RD) is a process intensification technology that has successfully enhanced chemical manufacturing [5].
The most recent research in this field primarily emphasizes the design elements of the reactive distillation unit and its corresponding operational parameters. These parameters include determining the necessary number of stages, the appropriate reactant feed locations, and potentially the liquid catalyst if it is homogeneously catalyzed. Additionally, the literature explores the reactive zone within the column if it is heterogeneously catalyzed, as well as the optimal reflux ratio. In a recent study, researchers introduced a new approach for optimizing the design and operation of a complex reactive distillation process [6]. This approach takes into account various process alternatives, such as pre-/side-reactors, side-strippers, and additional columns, using a superstructure approach. The optimization is based on a detailed cost-based objective function. Certain authors have also considered the selection of catalysts and the quantity of catalysts, which are crucial determinations in the design process [7]. In a recent study, researchers proposed a novel approach for efficiently determining the minimum catalyst quantity required to function under a heterogeneous reactive distillation process [8]. This method eliminates the need for complex simulations or optimization techniques. The authors suggested that during the design phase, the catalyst quantity should be at least six times the minimum amount determined via their proposed approach. Although numerous studies have focused on the steady-state design of reactive distillation, a few have also addressed the greater complexity of operational efficiency and control.

2.1. Advanced Reactive Distillation Technologies

Advanced reactive distillation technologies (ARDTs), which combine RD with extra PI capabilities, can enhance the working conditions, let reactions and separations occur at the same time more often, and reduce the number of units as the functions are integrated. ARDTs are being made for MA-RD, RHiGee, R-HIDiC, R-DWC, and CCD [9]. Shown in Figure 1 are the typical configurations for reactive distillation processes.

2.1.1. Reactive Dividing Wall Column (R-DWC)

The reactive dividing wall column (RDWC) (Figure 2) is a single-column reactor and dividing wall column [11]. RDWC reaction systems have been proposed recently [12]. Since simulations of processes using reactive dividing wall columns (RDWCs) demonstrated energy and capital cost benefits, RDWC modeling has gained popularity [11]. The RDWC can induce many liquid-phase processes under vacuum and higher pressure. Temperature-integrated RDWCs consume less energy [13].

2.1.2. Reactive High-Gravity Distillation (R-HiGee)

Process intensification is becoming more significant in improving industrial chemical processes, opening up interesting new opportunities for integrated process synthesis and design [14]. High gravity (HiGee) distillation (Figure 3) employs revolving equipment to create a centrifugal field. Strong gravity fields (100–1000 g) alter the flooding limit, allowing dense packing materials to have huge interfacial areas. The RZB lowers the volume by 4–7 and HETP by 2–8 cm [15]. Spinning reactive strippers improve mass transfer and product concentration over static reactors. Component-based efficiency factors, instantaneous selectivity, and intra-particle diffusion modeling describe the suggested integrated process. HiGee distillation necessitates a catalyst. The two-stage counter-current rotating packed bed (RPB), an improved rotating zirconia bed (RZB), is the top RD machine employed by HiGee contractors [15].
HiGee demonstrated promise for RD and solid-catalyzed gas–liquid processes in published tests. Gas–liquid mass transfer rates and effective catalyst wetting under high gravity are the primary causes of this potential. However, continuous operation at high gravity erodes the catalyst trapped inside internals like wire meshes. RPBs’ low liquid holdup and residence times limit their use in reactive systems with high response rates, which is a major drawback. To improve liquid-phase residence lengths and RPB applicability, reactive HiGee technologies necessitate further study [15]. Only liquid-phase reactions can be employed. Rapid reactions with competing serial processes increase selectivity and decrease yield losses [10].

2.1.3. Reactive Heat-Integrated Distillation Column (R-HIDiC)

The heat-integrated distillation column (HIDiC) is a revolutionary technology for facilitating distillation that employs heat transfer (Figure 4). Internal heat integration is a technique for converting a column’s full rectifying section into a heat source. On the other hand, the stripping part functions as a heat absorber. The heat is transferred from the rectification portion to the stripping section, which operates under high pressures. HIDiC has the potential to cut primary energy usage (and thus CO2 emissions) by up to 90% [10].
The heat-integrated distillation column (HIDiC) is a good way to reduce energy use. Even though the rectifying and stripping parts differ, heat moves away from the rectifying parts to the stripping parts inside the distillation column. The compressor raises the part’s working pressure and stage temperatures [16].

2.1.4. Catalytic Cyclic Distillation (CCD)

The principle of the cyclic distillation processes was developed over many years of engineering [17]. In CCD (Figure 5), controlled cycling alternates vapor flow and liquid flow times. Rising vapor hinders liquid flow down the column during vapor flow times. Liquid flows from tray to tray during liquid flow times. No downcomers make plates cheaper, easier, and more adaptable. Timer-controlled vapor line openings connect the reboiler and the distillation tower. Distillation is superior, as each tray’s liquid is different and does not mix [18].
Changes in the vapor flow period length can make the dwell time longer. Therefore, CCD is best for systems where reactions happen slowly but not so slowly that a traditional reactor–separator process would be better. CCD may be suggested when the rate of reaction is slow to middling [19].

2.1.5. Membrane-Assisted Reactive Distillation (MA-RD)

Membranes are chemically impenetrable. Membrane separation is cheaper than distillation for azeotropic mixtures [20]. The benefits of membranes include their capacity to regulate the components in contact, their resilience to mechanical, chemical, and thermal stress, and the hugely accessible surface area per unit volume. Due to their resistance to high temperatures, acidic environments, and basic environments, inorganic membranes are recommended. Common membranes, like those made of metal, ceramic, zeolite, or carbon, use reactive split wall columns. Membranes are used in various chemical process industries. For example, carbon membranes are the most common way to separate biodiesel, but zeolite membranes are the best way to separate TAEEs [21].
Membrane-assisted reactive distillation or MA-RD (Figure 6) is recommended when an undesirable “azeotrope” is present and the components to be separated and retrieved as an exit stream have a temperature close to the issue component [13].

2.2. Reactive Distillation from a Commercial Standpoint

Reactive distillation (RD) is a process that integrates separations and reactions within a singular vessel, thereby creating a hybrid system. In the 1920s, the development of this particular approach was obtained. However, this methodology was not extensively explored until 1980 (Eastman Company), when methyl acetate was successfully synthesized. The following reactions illustrate potential applications for reactive distillation [22].

2.2.1. Esterification

The process of esterification involves the reaction between an alcohol and an acid, forming an ester. Esters are a chemical compound class with an interesting fruity aroma [22]. Esters are primarily used to produce synthetic flavors and fragrances and serve as solvents for fats, gums, oils, and resins. In addition, they have been utilized as substances that enhance the flexibility and durability of materials. Reactive distillation columns have been utilized for esterification, considered the earliest chemical mechanism. In the context of conventional methods for generating methyl acetate, the emergence of an azeotrope with a low boiling point is one of the factors that decreases the manufacturing yield of methyl acetate. The limitation above is eliminated in the context of reactive distillation, thereby enabling the collection of methyl acetate in a nearly unadulterated form. Fatty acid esters, naturally occurring compounds, have been employed in various applications, such as plastics, cosmetics, and surfactants. Additionally, it has been documented that these esters can be synthesized via reactive distillation.

2.2.2. Transesterification

In general, this reaction involves the reaction between a triglyceride and an alcohol, resulting in the production of alkyl esters and glycerol. A prime illustration can be found in biodiesel production through this process. To date, there have been no reports of industrial facilities producing biodiesel via the synthesis in RD. However, the existing literature has suggested that a pilot-scale synthesis process is a viable option. As mentioned above, this procedure occurs through the chemical reaction between vegetable oil and alcohol facilitated via an acidic or alkaline catalyst [22].
From an economic perspective, using heterogeneous catalysts proves more efficient in producing biodiesel. Transesterification is a favorable option in some instances, as it circumvents water generation and yields an additional ester, thereby increasing the overall value. In comparison to hydrolysis, this process can prove advantageous.

2.2.3. Etherification

Etherification is a chemical process that involves the creation of ethers through the reaction between an alcohol and an acid. Ethers are extremely important to the fuel industry since, like alcohol, they can boost fuel’s octane rating when incorporated in adequate amounts. This function is what provides ethers their central place in this industry. Over the past two decades, numerous model reactions utilizing RD, including ETBEs, MTBEs, and TAMEs, have been subjected to investigations. The fuel oxygenates are synthesized through the chemical reaction between isobutylene and alcohol, forming ether and water. A further approach pertains to the chemical reaction of tert-amyl alcohol (TAA) and lower alcohol, which could include ethanol or methanol [22].

2.2.4. Alkylation

Alkylation is the general name given to shifting an alkyl group from a particular molecule to a subsequent one. This process occurs when an alkyl group is moved from one molecule to another. Ethylbenzene and cumene are both substances that can be obtained by employing the alkylation process. This process involves the reaction of alkanes, belonging to the paraffin compound family, with an aromatic compound, forming premium fuel alternatives, such as cumene. Adding compounds to gasoline as a blend is a common practice to enhance its octane number and mitigate engine-related issues, such as gum deposits resulting from oxidation [22].

2.2.5. Aldol Condensation

This is a chemical reaction in which an enolate ion undergoes a reaction with a carbonyl compound, forming a β-hydroxy ketone or a β-hydroxy aldehyde. After this step, a dehydration procedure is carried out, which results in the formation of a conjugated enone. By employing catalyst selection and constant removal of the intended outcome from the reaction zone, reactive distillation (RD) can enhance selectivity in the direction of an intermediary or end product [22].

2.2.6. Dehydration

This refers to the process of eliminating water molecules. The process of obtaining acetol from glycerol has been commonly employed. Typically, this chemical reaction is conducted using various metal catalysts, namely aluminum oxide, palladium, magnesium, copper, etc. The implementation of single- and two-stage RD methodologies is progressing, particularly in regenerating non-precious and precious catalysts [22].

2.2.7. Acetylation

Several methods generate a secondary product with significant value in other industrial applications. A consequential by-product known as glycerol is obtained under biodiesel production through methanol utilization. Glycerol is a significantly effective raw material for the acetylation process. It has been claimed that a conversion rate of roughly 99% from glycerol into triacetin may be attained when this process is carried out in a reactive distillation vessel. Triacetin is an additive that is integrated into fuels for compression engines. Its primary purpose is to minimize engine knocking [22].

2.2.8. Isomerization

This is a chemical process whereby a molecule undergoes a transformation, resulting in the formation of another molecule with identical atoms but differing in their arrangement. The isomers A-isophorone and B-isophorone can be effectively separated via reactive distillation due to their substantial variations in volatilities, despite their isomeric nature [22].

2.2.9. Oligomerization

The oligomerization process involves converting monomers into macromolecular complexes using a limited degree of polymerization. The hydrolysis of acid and oligomer esters was carried out using a reactive distillation technology, and the outcomes obtained agreed with the existing industrial-related manuscripts [22].
The attainment of product purity is a paramount necessity for meeting consumer demands. If these requirements are not compromised, or a substandard product is provided, the consumer’s expectations will not be fulfilled. Therefore, it is imperative to establish precise quality parameters. These parameters exhibit variations across multiple situations. In the case of pharmaceutical products, it is necessary to provide several quality indices, such as the chemical and physical characteristics of a product, its shelf life, toxicity, and medicinal effects. Key quality indicators, such as texture, taste, and nutritional value, are important in food products. Likewise, in the case of chemical process products, it is essential to employ the final product synthesis or its purity as a metric for assessing quality [22].

2.3. Simulation and Optimization Techniques Applied to Reactive Distillation Columns

Simulation and optimization refer to achieving the optimal outcome within a given set of conditions. “Optimization” refers to the systematic procedure of identifying the conditions that yield a given function’s highest or lowest output. Process optimization involves systematically modifying a process to optimize a predetermined set of parameters while ensuring that certain constraints are not violated. Over the past 25 years, the chemical industry has experienced notable transformations due to amplified energy expenses, progressively strict environmental policies, and intensified worldwide competition in product pricing and quality. Optimization has been considered a crucial engineering tool for tackling these concerns. Alterations in the design of plants and operational protocols have been managed to minimize expenses and comply with restrictions, focusing on enhancing efficacy and augmenting financial gain. Implementing optimal operating conditions can be achieved through heightened automation under various levels, such as process, plant, and company, commonly referred to as computer-integrated manufacturing. Statistical tasks are made more manageable and economical with the help of technology and the appropriate software [22].
Several deterministic optimization techniques based on analytical programming paradigms have been presented to design equilibrium and non-equilibrium RD columns [23,24].

2.3.1. Steady-State Simulation and Optimization

In a study, the process of the esterification of methyl acetate was simulated under steady-state conditions utilizing Aspen Plus software [22]. The Aspen Plus environment specifies various operating conditions, including the NRTL property method, Radfrac module, feed location, feed condition, column configuration (namely the reaction stage and number of stages), operating pressure, feed flow rate of the components utilized, and the type of condenser and reboiler. The findings revealed that the highest level of product purity was achieved under the uppermost stage. Moreover, it was illustrated that the highest achievable yield of methyl acetate product was 95.4%. The uppermost section of the column exhibited a significant reduction in the concentrations of methanol and acetic acid, which indicates the attainment of full conversion of reactants and the consequent generation of the desired product.
In addition, the reboiler temperature was higher than the condenser temperature. According to the data, the condenser operated at a cooler temperature (57.4 degrees Celsius) than the reboiler (62.75 degrees Celsius). The temperature of the reactive zone was adjusted between 61.3 degrees Celsius and 77.8 degrees Celsius to coincide with the exothermic nature of an esterification process. The maximum condenser temperature was experimentally measured at 58 degrees Celsius, while the corresponding value from an Aspen Plus simulation was 57.4 degrees Celsius. This indicated that the experimental results agree with the simulated results to an acceptable degree.

2.3.2. Sensitivity Analysis of a Methyl Acetate RDC

Throughout its operation, reactive distillation exhibits several steady-state conditions. The phenomenon referred to is commonly recognized as the diversity of the process. Multiplicity can be classified into two distinct sorts: input multiplicity and output multiplicity. This refers to a situation in which a column produces identical results for varying process conditions. In a study, the researchers examined the phenomenon of input multiplicity, whereby identical outputs are produced under varying input conditions [22].
In conducting sensitivity analysis, the initial selection was made on the molar flow of methyl acetate, taking into account the predetermined lower and upper bounds of heat duties, which were established at 1 kW and 3 kW, respectively. In the second condition, the mean flow rate of the acetic acid in the feed was varied from 0.01 L/min to 0.08 L/min to assess the mass percentage of methyl acetate. Similarly, we computed the bottom-to-feed ratio (B/F). As illustrated in the findings of the study, it was evident that the flow rate of methyl acetate exhibits an upward trend with an increase in heat duty. The highest flow rate was recorded at a heat duty of 6820 Btu/h, with a value of 0.927 lb mol/h. It was also shown that the mole fraction of methyl acetate produced corresponds to the acetic acid flow rate. The greatest product fraction measured was 95.2% at a flow rate equal to 0.0872 cubic feet per hour. The effect of changing the D/F and B/F ratios on the final composition was also investigated. The research revealed that the optimal values for the distillate-to-feed (D/F) ratio were 0.6275 and 0.4238, respectively, which allowed for the highest possible product purity to be achieved [22].

2.3.3. Methyl Acetate RDC Optimization

The efficiency of a reactive distillation column has the potential to be optimized using a model analysis tool through the Aspen Plus simulator. Methyl acetate mass fraction analysis was the key objective of the research, with acetic acid flow rate standardization serving as a secondary variable [22]. The aim was to attain the minimum achievable product composition at the uppermost point of the column. The heat duty was constrained within fixed values ranging from 1 kW to 3 kW, corresponding to the lower and upper limits, respectively. Following the optimization process, the minimum configuration of methyl acetate was determined to be 26.99%, while the optimized heat duty required was 2 kW.
A study organized the sensitivity results and the optimization findings collected from the Aspen Plus simulation. The reboiler heat duty and the reflux ratio were tuned to the values of 2 kW and 4.69, respectively. This demonstrated a high level of correlation between the experimental and simulated results. With reboiler heat duty as the controlled variable, the optimal methyl acetate flow rate was 0.093 lb mol/h, and the optimal acetic acid product percentage was 0.96 under the standard volumetric flow rate. The optimal flow rate and composition of methyl acetate were displayed together with the resulting sensitivity curve, and the resulting sensitivity curve for column temperature fluctuations due to reflux flow was also shown [22].

3. Research Gaps

Distillation is the most widely used chemical separation method in the chemical process industries (CPIs), accounting for 90–95% of all chemical and petroleum refining separations [25]. However, there are several drawbacks to this separation process, including increased installation space, a higher energy requirement, and decreased catalyst selectivity. Since reactants are no longer converted into products once the system reaches equilibrium, the conversion constraints for the reversible reactions are hard to exceed to achieve the highest purity of the catalyst [22]. Furthermore, distillation is an inefficient process that results in higher costs. It has been estimated that distillation columns consume 3% of global energy and as much as 80% of the power in chemical manufacturing facilities [7]. With these constraints, reactive distillation was developed as an innovative intensification technique in which product reaction and separation simultaneously occur in a single column. Furthermore, process intensification (PI) is a practical method for improving energy efficiency. Through integrated design and operation, the primary goal of process intensification is to reduce heat and mass transfer inhibitors while dealing with thermodynamic constraints [22].
Studies have evaluated a synthesis of different advanced reactive distillation technologies (ARDTs) [13]. A four-step decision-making flowchart was used in the methodology, which employs kinetic reactions and thermodynamic data, such as fundamental parameters, operating windows, and reaction kinetics. This methodology aimed to direct technology selection through fundamental information while maintaining adaptability to meet the design problem’s requirements.
One of the features of ARDT in the study is the reactive dividing wall column (R-DWC). The R-DWC is a technology that combines dividing wall columns and reactive distillation. The R-DWC aims to reduce the number of vessels while dealing with multiple outlet streams. The R-DWC has a minimum of three outlet streams, which allows for removing intermediate-boiling elements and reactants within the system through boil-up and reflux and outside via the draw stream once the conversion is insufficient. The characteristics of R-DWCs incorporate impurities that can interact with different species, reduced installation costs, increased product conversion and selectivity, dilute feeds, material constraints, undesired side reactions, and the disintegration of azeotropes in the reaction. However, the need for more recent studies can be attributed to the complexity of R-DWCs and their emergence as a field of distillation research. R-DWCs are complex due to their numerous design and operational variables. Furthermore, the study lacks information concerning the amount of catalyst to be used, reaction conditions, and residence time, as these are common problems in R-DWCs. The type and amount of catalyst used, residence time, and reaction conditions must be appropriate and compatible with the separation specifications in the separation processes.
Another ARDT investigated in the study is reactive high-gravity distillation (R-HiGee). R-HiGee is a reaction and separation system with high-gravity fields. It combines enhanced gravitational forces with increased momentum, heat, and mass transfer to promote the reaction and separation of phases. R-HiGee also reduces the volume and residence time, which leads to the equipment being decreased. Packed beds were identified as potential HiGee contractors for reactive distillations in the study.
Further, as the residence time is reduced, HiGee can achieve a high selectivity percentage. HiGee is frequently used for diffusion with limited systems with substantial viscosity and fast reactions in series. Despite these benefits, HiGee remains far from being an extensively developed technology, as indicated by the absence of recent studies. This study necessitates further research to investigate the higher pressure drop during the process. Furthermore, rotating parts such as bearings, dynamic seals, and rotors must be studied as they affect long-term reliability.
Moreover, the reactive heat-integrated distillation column (RHIDiC) combines reactive distillation with internally heat-integrated distillation columns through vapor recompression and heat integration. According to the study, heat exchange occurs in several opposite heat exchangers, with the stages not always at the same elevation. Furthermore, the viability of the R-HIDiC was assessed using its boiling points at various pressures and an appropriate compression ratio. However, the researchers did not address the consumption of thermal energy of the RHIDiC, which is highly critical. The distillation columns’ heat requirements frequently account for most of the plant’s total energy cost.
Catalytic cyclic distillation (CCD) utilizes the full advantage of the interactions of reactions combined and separated by adjusting the duration of the process during a cyclic operation, such as residence time. CCD has a two-step cycle consisting of a vapor flow period, in which vapor flows higher while the reaction happens in the liquid state, and a liquid flow phase, in which the liquid moves downward without being remixed from one tray to another. With this, CCD works best for systems involving reactions with a reducing intermediate. Despite these benefits, the study requires additional research into its flexibility and the presence of several steady states. The contact of the reaction and separation results in non-linear process behavior, resulting in various steady-state column profiles with varying conversions. Due to the higher degree of integration, flexibility was reduced, which must be studied to develop the technology further.
Further, membrane-assisted reactive distillation (MARD) is a membrane module linked to a RD unit that has emerged as a novel method for hybrid process intensification in bulk chemical products to achieve higher efficiency and yield. The membrane module is mounted to a feed stream to eliminate impurities, pre-fractionate the mixture, or is connected to an exit stream to rupture the azeotropes and recover homogeneous catalysts. According to the study, MARD is commonly used when an undesirable azeotrope exists at temperatures similar to the components that will be separated and acquired as an outlet stream. However, the study could not determine whether the membrane associated with the reactive distillation improved its performance in effective energy efficiency. In addition, the hybrid process intensification needed to be thoroughly addressed, even though the combination of pervaporation membrane and reactive distillation produced a higher degree of separation.

4. Future Outlook

Reaction distillation is an advantageous process combining chemical reactions and distillation in a single operating unit. Compared with the traditional methods of handling and producing chemicals, reactive distillation offers several advantages. Reactive distillation helps energy efficiency, reduces waste products, and produces higher yields. With these advantages, researchers and engineers look forward to developing reactive distillation using new catalysts and reaction conditions that may be more compatible with distillation processes. The discovery of new catalysts will be crucial in improving the efficiency and selectivity of reactive distillation systems. Catalysts that are tailored to specific reactions will become more widespread, allowing for more complex transformations with high yields.
Reactive distillation has the potential to considerably aid the petrochemical sector. It may be used to make high-value compounds, such as methyl tert-butyl ether (MTBE), ethyl tert-butyl ether (ETBE), and biodiesel, as well as improve the efficiency of traditional methods, such as hydrocarbon isomerization. Reactive distillation may be used to make a variety of specialized compounds, such as esters, acylates, and specialty solvents. Its capacity to operate under a wide range of pressures and temperatures makes it suited for a wide range of reactions. Continued R&D efforts are required to realize the full potential of reactive distillation. This includes the creation of novel catalysts, new reactor designs, and enhanced process control methods.
Another future insight that researchers are observing is the use of reactive distillation in other industries, such as pharmaceuticals; with this, RD may help discover and synthesize more complex compounds. Reactive distillation has the potential to transform pharmaceutical production by allowing a more efficient and cost-effective synthesis of complex compounds. This process’s continuous nature makes it appropriate for the synthesis of pharmaceutical intermediates and active pharmaceutical ingredients (APIs). Furthermore, reactive distillation can improve the manufacturing of food additives, taste compounds, and specialty components. Its capacity to manufacture high-purity goods is extremely significant in this industry.
Process intensification will become an increasingly important trend, driven via the demand for sustainability and resource efficiency. To enhance process efficiency, reactor designs will become more compact, and integration with other unit processes, such as membrane separation and heat integration, will be investigated. Advances in process control and automation will be critical to the future of reactive distillation. Real-time monitoring and control technologies will allow for better reaction management, making the process safer and more dependable.
Moreover, researchers are looking into integrating RD with other unit operations, such as membrane separation, reactor systems, and absorption. RD is useful in designing a process that maximizes its driving force, which can help design a reliable process and its controller structure. Integrating RD with other unit operations can further improve process efficiency and produce better results. Researchers are also looking into integrating RD with other technologies involving distillation. With this, it may help in retrieving better results [13].
Green chemistry concepts will play a critical part in the development of reactive distillation technologies as the world evolves towards more sustainable practices. This involves the use of renewable feedstocks, the reduction of waste, and the reduction of the environmental effect of chemical manufacturing. As a result, modular and mobile reactive distillation systems will become more common. These machines are lightweight and portable, providing greater flexibility for on-site processing and decreasing the need for large, centralized facilities.
Developing reactive distillation has become a challenge, and while there is still a long way to go, this technique has the potential to transform the chemical industry and make it more sustainable and efficient. Reactive distillation offers a paradigm change in chemical manufacturing, paving the way for more sustainable and efficient procedures in a variety of sectors. Looking ahead, continuous catalyst research, process intensification, and automation will expand reactive distillation’s possibilities. This technology is positioned to play a crucial role in determining the future of chemical engineering, with possible applications spanning from petrochemicals to medicines and alignment with green chemistry principles. As challenges are overcome and knowledge deepens, reactive distillation promises to be at the forefront of innovation in the chemical industry, driving progress towards a more sustainable and efficient future.

Author Contributions

Conceptualization, E.C.R.L.; writing—original draft preparation, D.A.B., B.F., R.M., H.R.R. and K.S.; 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. Malone, M.F.; Doherty, M.F. Reactive Distillation. Ind. Eng. Chem. Res. 2000, 39, 3953–3957. [Google Scholar] [CrossRef]
  2. Hou, Z.; Geng, X.; Ding, Q.; Li, H.; Guo, Y.; Gao, X. Understanding the Quantitative Matching Principle of Reaction and Distillation for the Reactive Distillation Process of Cyclohexyl Acetate Synthesis. Ind. Eng. Chem. Res. 2023, 62, 9849–9862. [Google Scholar] [CrossRef]
  3. Sundmacher, K.; Kienle, A. Reactive Distillation: Status and Future Directions; Chemical Engineering Science; Wiley: Weinheim, Germany, 2018; Volume 177, pp. 1–18. [Google Scholar] [CrossRef]
  4. Muthia, R.; Reijneveld, A.G.; van der Ham, A.G.; Kate, A.J.T.; Bargeman, G.; Kersten, S.R.; Kiss, A.A. Novel method for mapping the applicability of reactive distillation. Chem. Eng. Process. -Process Intensif. 2018, 128, 263–275. [Google Scholar] [CrossRef]
  5. Orjuela, A.; Santaella, M.A.; Molano, P.A. Process intensification by reactive distillation. In Process Intensification in Chemical Engineering: Design Optimization and Control; Springer: Cham, Switzerland, 2016; pp. 131–181. [Google Scholar]
  6. Tsatse, A.; Oudenhoven, S.; Kate, A.T.; Sorensen, E. Optimal design and operation of reactive distillation systems based on a superstructure methodology. Chem. Eng. Res. Des. 2021, 170, 107–133. [Google Scholar] [CrossRef]
  7. 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]
  8. Kroeze, V.; Braakhuis, L.; Kate, A.J.T.; Kersten, S.R. Determination of the Minimum Catalyst Amount in the Design of Catalytic Distillation Columns. Ind. Eng. Chem. Res. 2021, 60, 10254–10264. [Google Scholar] [CrossRef]
  9. Werth, K.; Lutze, P.; Kiss, A.A.; Stankiewicz, A.I.; Stefanidis, G.D.; Górak, A. A systematic investigation of microwave-assisted reactive distillation: Influence of microwaves on separation and reaction. Chem. Eng. Process.-Process Intensif. 2015, 93, 87–97. [Google Scholar] [CrossRef]
  10. Kiss, A.A.; Jobson, M.; Gao, X. Reactive Distillation: Stepping Up to the Next Level of Process Intensification. Ind. Eng. Chem. Res. 2019, 58, 5909–5918. [Google Scholar] [CrossRef]
  11. de Villiers, R.; Kooijman, H.A.; Taylor, R. Parallel column model for reactive dividing wall column simulation. Chem. Eng. Res. Des. 2023, 194, 366–375. [Google Scholar] [CrossRef]
  12. Schröder, M.; Ehlers, C.; Fieg, G. A Comprehensive Analysis on the Reactive Dividing-Wall Column, its Minimum Energy Demand, and Energy-Saving Potential. Chem. Eng. Technol. 2016, 39, 2323–2338. [Google Scholar] [CrossRef]
  13. Pazmiño-Mayorga, I.; Jobson, M.; Kiss, A.A. Operating windows for early evaluation of the applicability of advanced reactive distillation technologies. Chem. Eng. Res. Des. 2022, 189, 485–499. [Google Scholar] [CrossRef]
  14. Avila, J.R.A.; Kong, Z.Y.; Lee, H.-Y.; Sunarso, J. Advancements in Optimization and Control Techniques for Intensifying Processes. Processes 2021, 9, 2150. [Google Scholar] [CrossRef]
  15. Garcia, G.E.C.; van der Schaaf, J.; Kiss, A.A. A review on process intensification in HiGee distillation. J. Chem. Technol. Biotechnol. 2017, 92, 1136–1156. [Google Scholar] [CrossRef]
  16. Yala, O.; Rouzineau, D.; Thery-Hetreux, R.; Meyer, M. Design and optimization of Heat Integrated Distillation Column “HIDiC”. Comput. Aided Chem. Eng. 2017, 40, 1783–1788. [Google Scholar] [CrossRef]
  17. Nielsen, R.F.; Huusom, J.K.; Abildskov, J. Driving Force Based Design of Cyclic Distillation. Ind. Eng. Chem. Res. 2017, 56, 10833–10844. [Google Scholar] [CrossRef]
  18. Kiss, A.A.; Bildea, C.S. Revive Your Columns with Cyclic Distillation. CEP Mag. 2015, 2015, 21–27. [Google Scholar]
  19. Kiss, A.A.; Maleta, V.N. Cyclic distillation technology—A new challenger in fluid separations. Chem. Eng. Trans. 2018, 69, 823–828. [Google Scholar] [CrossRef]
  20. Errico, M.; Madeddu, C.; Bindseil, M.F.; Madsen, S.D.; Braekevelt, S.; Camilleri-Rumbau, M.S. Membrane assisted reactive distillation for bioethanol purification. Chem. Eng. Process.-Process Intensif. 2020, 157, 108110. [Google Scholar] [CrossRef]
  21. Sakhre, V.; Jain, S.; Sapkal, V.S.; Agarwal, D.P. Novel process integration for biodiesel blend in membrane reactive divided wall (MRDW) column. PJCT 2016, 18, 105–112. [Google Scholar] [CrossRef]
  22. Sakhre, V. Reactive Distillation: Modeling, Simulation, and Optimization. In Distillation—Modelling, Simulation and Optimization; IntechOpen: London, UK, 2019. [Google Scholar] [CrossRef]
  23. Ciric, A.R.; Miao, P. Steady State Multiplicities in an Ethylene Glycol Reactive Distillation Column. Ind. Eng. Chem. Res. 1994, 33, 2738–2748. [Google Scholar] [CrossRef]
  24. Gumus, Z.H.; Ciric, A.R. Reactive distillation column design with vapor/liquid/liquid equilibria. Comput. Chem. Eng. 1997, 21, S983–S988. [Google Scholar] [CrossRef]
  25. Segovia-Hernández, J.G.; Hernández, S.; Petriciolet, A.B. Reactive distillation: A review of optimal design using deterministic and stochastic techniques. Chem. Eng. Process.-Process Intensif. 2015, 97, 134–143. [Google Scholar] [CrossRef]
Figure 1. Configurations of reactive distillation (RD) processes: classical RD column (left), azeotropic RD (middle), RD with optional (dashed line), and pre-and/or side-reactors (RX) that can elevate conversion (right). Reprinted with permission from [10]. Copyright 2019 American Chemical Society.
Figure 1. Configurations of reactive distillation (RD) processes: classical RD column (left), azeotropic RD (middle), RD with optional (dashed line), and pre-and/or side-reactors (RX) that can elevate conversion (right). Reprinted with permission from [10]. Copyright 2019 American Chemical Society.
Engproc 56 00099 g001
Figure 2. Configuration of a reactive dividing wall column. Reprinted with permission from [13]. Copyright 2022 Elsevier.
Figure 2. Configuration of a reactive dividing wall column. Reprinted with permission from [13]. Copyright 2022 Elsevier.
Engproc 56 00099 g002
Figure 3. Working principle of HiGee distillation in a rotating packed bed (RPB). Reprinted with permission from [15]. Copyright 2017 John Wiley & Sons, Inc.
Figure 3. Working principle of HiGee distillation in a rotating packed bed (RPB). Reprinted with permission from [15]. Copyright 2017 John Wiley & Sons, Inc.
Engproc 56 00099 g003
Figure 4. Heat-integrated distillation column (left) and reactive HiDIC configuration (right). Reprinted with permission from [10]. Copyright 2019 American Chemical Society.
Figure 4. Heat-integrated distillation column (left) and reactive HiDIC configuration (right). Reprinted with permission from [10]. Copyright 2019 American Chemical Society.
Engproc 56 00099 g004
Figure 5. Configuration of a catalytic cyclic distillation (CCD) column. Reprinted with permission from [13]. Copyright 2022 Elsevier.
Figure 5. Configuration of a catalytic cyclic distillation (CCD) column. Reprinted with permission from [13]. Copyright 2022 Elsevier.
Engproc 56 00099 g005
Figure 6. Configuration of a membrane-assisted reactive distillation (MA-RD) column. Reprinted with permission from [13]. Copyright 2022 Elsevier.
Figure 6. Configuration of a membrane-assisted reactive distillation (MA-RD) column. Reprinted with permission from [13]. Copyright 2022 Elsevier.
Engproc 56 00099 g006
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

Barrientos, D.A.; Fernandez, B.; Morante, R.; Rivera, H.R.; Simeon, K.; Lopez, E.C.R. Recent Advances in Reactive Distillation. Eng. Proc. 2023, 56, 99. https://doi.org/10.3390/ASEC2023-15278

AMA Style

Barrientos DA, Fernandez B, Morante R, Rivera HR, Simeon K, Lopez ECR. Recent Advances in Reactive Distillation. Engineering Proceedings. 2023; 56(1):99. https://doi.org/10.3390/ASEC2023-15278

Chicago/Turabian Style

Barrientos, Demi Andrei, Beatrice Fernandez, Rachel Morante, Hannah Ruth Rivera, Karen Simeon, and Edgar Clyde R. Lopez. 2023. "Recent Advances in Reactive Distillation" Engineering Proceedings 56, no. 1: 99. https://doi.org/10.3390/ASEC2023-15278

Article Metrics

Back to TopTop