Next Article in Journal
A New Serine Protease (AsKSP) with Fibrinolytic Potential Obtained from Aspergillus tamarii Kita UCP 1279: Biochemical, Cytotoxic and Hematological Evaluation
Previous Article in Journal
Enhanced Stability and Activity of Nitrogen-Doped Carbon Nanotube-Supported Ni Catalysts for Methane Dry Reforming
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Review on Production of Ethylene Oxide from Epoxidation of Ethylene: Catalysis, Mechanism and Kinetics

by
Mahammad Ali Saritala
1,2,
Mohammed Muzammil
1,
Mohammad R. Quddus
3,
Shaikh Abdur Razzak
1,2 and
Mohammad M. Hossain
1,2,*
1
Department of Chemical Engineering, King Fahd University of Petroleum and Minerals, Dhahran 31261, Saudi Arabia
2
Interdisciplinary Research Center for Refining and Advanced Chemicals, King Fahd University of Petroleum and Minerals, Dhahran 31261, Saudi Arabia
3
Chemical Engineering, Western University, London, ON N6A 5B9, Canada
*
Author to whom correspondence should be addressed.
Catalysts 2025, 15(6), 560; https://doi.org/10.3390/catal15060560
Submission received: 14 April 2025 / Revised: 23 May 2025 / Accepted: 27 May 2025 / Published: 4 June 2025
(This article belongs to the Section Catalytic Reaction Engineering)

Abstract

This review describes the different developments in the production of ethylene oxide (EO) by epoxidation of ethylene. EO is an important chemical intermediate for the manufacture of a variety of industrial and consumer products, such as ethylene glycol, plastics, and pharmaceuticals. The conventional gas-phase epoxidation process using silver-based catalysts suffers from major drawbacks, including low selectivity and high carbon dioxide emissions. This review underlines emerging solutions for efficiency and sustainability improvement in EO production. Major developments in catalyst design, including novel silver-based hybrid nanostructures, Mn-N4GP catalysts, and chemical looping epoxidation processes, are presented. It also discusses developments in reaction kinetics, including catalyst surface optimization and the use of dopants. The article also outlines catalyst deactivation challenges, cost, and scalability and describes future research directions on renewable feedstocks, reducing energy consumption and most importantly environmental impact. These innovations are oriented toward a more sustainable and economical route for large-scale manufacturing of ethylene oxide.

1. Introduction

Ethylene oxide (EO) ranks as the third-most important ethylene derivative. Its energy content should be emphasized as a value addition to its other applications. It is extensively used in several industries, including producing glycols and polymers, essential chemicals for several consumer products. EO is also an intermediate in the petrochemical industry, with around 70 percent of its output used in the production of mono-, di– and triethylene glycols [1]. EO is extensively employed as an essential feedstock in various industries, such as plastics, solvents, detergents, and textiles. It has a critical role in the production of consumer goods like glycol derivatives, ethoxylates, and polymers, apart from ethylene glycol and several key fine petroleum and chemical intermediates [2,3,4]. In addition, it is an essential feedstock for the manufacture of various key commodity chemicals, including ethylene glycol, glycol ethers, and epoxy resins [5]. This also serves as an important intermediate needed by many commercially manufactured products, which include coolants, polyesters, detergents, surfactants, lubricants, paints, cosmetics, and pharmaceuticals [6]. It is an inflammable, colorless gas that has a smell that is sweet in nature. Ethylene oxide (EO) is one of the most widely produced chemicals, with an annual market size exceeding 34 million tons, also the level of global demand for EO and its derivatives that had been reached by the end of 2021 [6,7,8]. Increasing EO demand is underpinned by increasing consumption in various industries, including pharmaceuticals, personal care, detergents, automotive, agrochemicals, food and beverages, textiles, and other end-user industries [9]. Growing academic interest mirrors EO’s industrial importance, as shown by the rising number of ethylene epoxidation publications over the last two decades (Figure 1).
Ethylene oxide (EO) has traditionally been manufactured by the direct oxidation of ethylene in the presence of molecular oxygen (O2) via silver catalysis. The reaction of the conventional process proceeds at high temperatures (392 to 500 °F) and high-pressure conditions (10 to 30 bar). Although this process is well established, it strongly depends on fossil fuels, leading to a high environmental impact, as it emits approximately 1.9 tons of CO2 for every ton of EO produced [1,10,11]. Very recently, Yang et al. [12] surveyed the latest advances in heterogeneous catalyst architectures for ethylene epoxidation—covering metal nanoparticles, single-atom sites, bimetals, and the role of novel promoters—and highlighted emerging electrocatalytic and photocatalytic routes to EO with near-atom efficiency [12]. However, EO has poor selectivity to the epoxidation reaction, obtaining only ca. 85%, since it gives rise to very high amounts of CO2 as by-products, this process being the second-largest contributor to CO2 emissions after ammonia synthesis in chemical processes [1].
Various researchers have tried alternatives to replace the widely practiced conventional process of epoxidation in the gaseous phase, which is associated with CO2 emissions, in order to produce a clean liquid-phase epoxidation (LPE) process. This method uses a homogeneous methyltrioxorhenium (MTO) catalyst with H2O2 as the oxidant. The LPE process demonstrated excellent catalytic performance, with nearly 100% selectivity toward ethylene oxide (EO) without emitting CO2 [13]. However, the application of MTO to this process faces two main problems. Firstly, it tends to deactivate the catalyst due to the deposition of water as a by-product in the reaction. The second problem arises because rhenium (Re) is an expensive and scarce metal, making its use economically and practically difficult. Consequently, the substitution of Re with easily available, cheap catalytic materials with high H2O2 utilization efficiency should improve the economic feasibility as well as the practical ability of the process [14].
Another disadvantage of ethylene oxide, EO, arises from the inherently high reactivity of the compound, which always tends to result in disagreeable routes, inclusive of direct burning or the isomerization reactions of acetaldehyde [15]. Products from acetaldehyde burning can be CO2 and water [16]. Moreover, ethylene itself can undergo direct combustion to further contribute to the formation of CO2 and water, as illustrated in Figure 2 [9]. The mechanism of ethylene epoxidation reaction and the nature of the active species remain under debate, despite several decades of research [17,18,19].
Despite these challenges, ethylene oxide (EO) is an essential constituent of many industrial processes, and demand for this compound continues to grow. To this end, research into epoxidation has turned its focus to the investigation of the effectiveness of various materials as catalysts in ethylene oxide production [20,21]. Catalysis, reactor design, and reaction kinetics are promising approaches to enhance efficiency and support sustainability. This review provides a comprehensive overview of recent developments regarding the creation of novel catalysts, advanced kinetic models, and improved reactor designs for addressing the environmental and economic challenges facing ethylene oxide production.

2. Role of Catalysts in the Process of Epoxidation of Ethylene

This reaction has been at the center of chemical synthesis: the catalytic conversion of ethylene to ethylene oxide, sustained by continuous efforts in developing highly effective and selective catalysts. These developments have introduced different types of catalyst systems over time, each presenting different ways to improve performance. Starting from the electronic structure of Mn-N4GP (Mn-N4 porphyrin-like graphene) to the stability of Ag–h-BN (silver supported on hexagonal boron nitride) (Table 1, Entry 2) up to synergistic effects in Ag–CuO (Ag supported on CuO) and AgNiO2 (perovskite-type silver nickel oxide) systems, catalyst design advancements have considerably improved EO production. Perovskites, hybrid nanostructures, and silver-based catalysts have been explored over time for maximum EO production through a balancing act between selectivity and reactivity. The structural, electrical, and surface properties of each catalyst affect its performance and control the creation of products and intermediate stabilization. This review of the literature will be focused on the developments made so far in ethylene epoxidation catalyst design, emphasizing their special contributions to the enhancement of EO selectivity and reduction of undesirable by-products.
According to [22], the Mn-N4GP catalyst has the best ethylene epoxidation properties due to its unique electronic structure and high selectivity (Table 1, Entry 1). Its active site, Mn=O, can efficiently transfer oxygen to ethylene. It forms the target molecule ethylene oxide as the main product, but minimizes the production of by-products such as acetaldehyde and five-membered ring species, with energy barriers of 0.56 and 0.46 eV, respectively, according to kinetic studies, with a formation rate for ethylene oxide 105 and 104 times faster. DOS (density of states) surface oxide reconstruction analysis confirmed the Mn=O site’s efficient electron transfer to activate ethylene’s π-bond, making Mn-N4GP a selective and efficient catalyst for green ethylene oxide production under mild conditions [23,24].
The Ag–h-BN catalyst manifests high activity and selectivity for ethylene epoxidation, owing to the unique interaction of Ag atoms with reactants. Anchored at the B-vacancy defect in the boron nitride nanosheet (as shown in Figure 3), the Ag atom is the active site that favors ethylene adsorption (−1.21 eV) over oxygen [25]. A low energy barrier for ethylene oxide formation (0.58 eV) and resistance to isomerization to acetaldehyde ensure selective partial oxidation [26]. Molecular dynamic simulations also confirm that the catalyst is structurally stable at high temperatures and thus suitable for large-scale applications [27,28]. Silver-based catalysts remain the cornerstone of ethylene oxide production, with continued improvements in selectivity through the use of promoters such as chlorine and cesium significantly enhancing catalytic performance [29].
Silver’s catalytic efficiency arises from stabilizing oxygen species and providing selective sites. However, pristine Ag surfaces show low EO selectivity under industrial conditions due to the dominance of the OMC–DH (oxometallacycle–dehydrogenation) pathway, leading to combustion products. High oxygen pressures enhance EO selectivity via Ag surface oxides (AgOx) (Table 1, Entry 5), where positively charged Ag atoms suppress dehydrogenation. While facets like Ag(100) (Table 1, Entry 3) are more active, they favor OMC–DH and reduce selectivity. The surface oxidation and the structures of Ag, including nanoparticles and a partially oxidized surface, should be optimized to improve ethylene epoxidation performance [30,31].
According to a study by [32], silver’s remarkable activity in ethylene epoxidation is a consequence of the moderate strength of its oxygen bonding, which allows O2 to dissociate effectively, but prevents overoxidation. Ag(100) surfaces are particularly efficient in stabilizing the OMC intermediate (as shown in Figure 4), lowering the activation energy for ethylene oxide (EO) formation. In contrast, Cu binds oxygen too strongly to form oxide phases, while Au exhibits poor O2 activation and favors acetaldehyde (AA) formation. Scaling relations between oxygen-binding energy and selectivity reveal that Ag has an optimal balance of kinetics and thermodynamics, which together make it the best catalyst for EO production [33,34].
Silver catalyst systems are highly selective towards ethylene oxide (EO) because of their intermediate oxygen-binding affinity, which enables the stabilization of key reaction intermediates like the oxometallacycle (OMC). Oxygen-rich surfaces like Ag1.83O(111) and AgO_p(4 × 4) (4 × 4 gene of silver under O2-rich conditions) exhibit especially high selectivity, and the strong adsorption of oxygen is important in the stabilization of intermediates and the reduction of activation energy barriers for EO synthesis [35,36]. Electronic structure analyses, including Bader charge analysis, show that enhanced charge transfer and more stable Ag–O bonding are directly linked to improved EO selectivity. AgO_p(4 × 4) has the highest charge transfer (Q0 = −0.98 |e|), which is in line with its better performance [37]. Also, thermodynamic modeling indicates that various silver oxide phases may coexist in industrial conditions, and both oxygen coverage and surface oxygen species character are significant parameters in determining product selectivity. In addition, the cooperation of Langmuir–Hinshelwood and Eley–Rideal mechanisms over these surfaces enhances both selectivity and reactivity, which further upholds the excellent standing of silver-based catalysts in industrial ethylene epoxidation [38].
Metal dopants in Ag-catalyzed ethylene epoxidation optimize EO selectivity by engineering oxophilicity. Dopants like Cs, Cu, and Co recreate catalytic sites where oxygen binds to the dopant and ethylene to Ag, reducing EO formation energy barriers while limiting by-product stabilization [39]. DFT (density functional theory) calculations show dopants tune oxygen adsorption energy (ΔGo*) to balance reactivity and selectivity, breaking pristine Ag’s scaling relations and placing doped systems at the selectivity volcano’s peak. Dual dopants, such as Co–Cu or Cs–Re, further enhance selectivity through synergistic stabilization of intermediates and efficient EO desorption [40,41,42].
Silver supported on α-alumina is the benchmark catalyst for ethylene epoxidation due to its unique oxidative properties [43]. The low surface area of α-alumina (around 1–8 m2/g) minimizes the density of hydroxyl groups, which could otherwise catalyze the overoxidation of EO to CO2 and H2O, improving EO selectivity. Additionally, silver particles supported on high-surface-area α-alumina (around 30 m2/g) exhibit enhanced stability against sintering under reaction conditions, reducing particle growth via mechanisms like Ostwald ripening (see Figure 5) [43]. This stability is attributed to the larger interparticle distances afforded by higher-surface-area supports, which limit silver particle coalescence and maintain catalytic performance over extended operation periods [16,44].
Electrocatalysts like RuO2 and BaOx/IrO2 excel in halide-mediated ethylene epoxidation by creating active oxygen sites for selective EO production. Halide-modified RuO2 favors epoxide-like ethylene adsorption, minimizing CO2 formation, while BaOx-modified IrO2 boosts faradaic efficiency above 85% by inhibiting hypochlorous acid cleavage. Their stability and ability to modulate oxygen species make these catalysts highly promising for efficient, large-scale EO production, highlighting the importance of tailored structures and surface properties [45].
A study by [46] revealed that strontium ferrite perovskites facilitate chemical looping epoxidation by balancing oxygen capacity and release (see Figure 6). The cubic SrFeO3 phase provides high oxygen capacity, while the layered Sr3Fe2O7 phase enables faster oxygen release and uptake. Together, they supply lattice oxygen to Ag catalysts, which selectively convert ethylene to EO while minimizing CO2 formation. A 1:1 SrFeO3:Sr3Fe2O7 ratio optimizes oxygen release kinetics, quadrupling EO yield (1.6%) compared to SrFeO3 alone. This highlights the importance of tailored catalyst design in enhancing chemical looping epoxidation efficiency [46,47].
CeO2-modified SrFeO3 acts as the oxygen carrier, while Ag constitutes the active catalytic site. Silver prevents total combustion because it stabilizes the crucial OMC intermediate and allows only the selective oxidation of ethylene to EO. CeO2 enhances the rate of oxygen release and uptake during reduction and reoxidation; thus, it significantly promotes the performance of the catalyst. Such cooperation by CeO2 is effective in providing an oxygen bridge: transferring oxygen rapidly between the gas phase and the SrFeO3 lattice [48]. Consequently, μO2 (chemical potential of oxygen) increases with enhanced phase cooperation, effectively resulting in greater EO selectivity and maintaining constant performance after numerous redox cycles of catalysts. The prepared catalyst design achieves the maximum limit of desired performance above and beyond state-of-the-art traditional direct gas-phase epoxidation processes and gives improved selectivity than unmodified SrFeO3 [48,49]. In Figure 7, the SEM–EDS analysis illustrates the distribution of silver and cerium in both (CeO2)impr(SrFeO3) (CeO2 impregnated into SrFeO3) and (CeO2)ss(SrFeO3) (CeO2 incorporated via solid-state synthesis with SrFeO3) samples, revealing a uniform spread of strontium, iron, and oxygen across the material surfaces.
The Ag–SrFeO3 catalyst is highly efficient for ethylene oxidation to ethylene oxide (EO) in the CLE (chemical looping epoxidation) process. SrFeO3 acts as a lattice oxygen reservoir, while silver catalyzes the selective reaction. The Ag–SrFeO3 interface facilitates oxygen transfer, and higher calcination temperatures (e.g., 650 °C) improve EO selectivity by forming larger Ag particles that minimize overoxidation. Ceria (CeO2) doping enhances oxygen transport and catalyst regeneration, achieving 60% EO selectivity and 15% ethylene conversion [15].
The Ag–CuO hybrid nanocatalyst demonstrates enhanced performance due to the synergistic effect between silver and copper oxide domains. The unique structural configuration, where Ag emerges on the CuO surface, provides active interfaces for selective epoxidation. Chlorine acts as a promoter by influencing the adsorption–desorption energies of intermediates and products [50]. It reduces the interaction strength of EO with the catalyst surface, decreasing the probability of its overoxidation to CO2. Additionally, chlorine promotes the pseudo-direct pathway, increasing EO selectivity while driving AA formation through the less favorable OMC pathway. These combined effects explain the experimentally observed higher EO/CO2 ratio and maximum EO selectivity of 60% under optimized conditions [51,52].
A study by [53] revealed that under sun radiation, plasmonic silver (Ag) nanoparticles are essential for increasing ethylene epoxidation. LSPR (localized surface plasmon resonance) action of Ag nanoparticles generates hot, energetic electrons that enhance the reactivity of molecular oxygen while allowing efficient visible light absorption. The generated hot electrons can stabilize reactive intermediates, while O2 is reduced to promote the production of EO [54]. Plasmonic Ag reduces the activation temperature compared with conventional thermal catalysts, saving energy consumption and improving selectivity, with higher EO yields because the Ag catalyst surface is inhibiting all the oxidation pathways. Solar-powered technology provides EO without the use of high pressures or heat from fossil fuels; hence, it is energy-efficient and environmentally friendly [55,56].
The epoxidation reaction is aided by AgNiO2 nanoparticles in two ways. While the nickel oxide component (Ni2+) stabilizes the oxygen species and stops overoxidation to CO2, the silver component serves as the active site for the selective oxidation of ethylene to EO. The AgNiO2 structure partially breaks down into metallic silver (Ag) and nickel oxide (NiO) at high temperatures (over 150 °C), which lowers the selectivity of EO. Nevertheless, the delafossite structure offers a stable interface for ethylene and oxygen activation at ambient temperature, resulting in effective EO synthesis. Lattice oxygen is essential for facilitating the catalytic process without causing appreciable structural degradation, and the synergistic interaction between Ag and Ni guarantees a balance between reactivity and selectivity [57].
The oxidation state and kind of oxygen species on the Ag surface determine how well the Ag–α-Al2O3 catalyst performs in ethylene epoxidation. Silver particles produce surface and subsurface oxygen species that actively contribute to the process when exposed to severe oxidative conditions. Atomic oxygen encourages non-selective pathways that result in combustion products, whereas dioxygen species (Ag4-O2) are found to be extremely selective for EO formation [58]. Furthermore, dynamic oxidation-reduction cycles on the Ag catalyst surface control the ratio of selective to non-selective oxygen species. High catalytic activity is maintained by the α-AlO3 support, which also minimizes unwanted interactions and guarantees the Ag nanoparticles’ thermal stability. Controlled surface oxidation is necessary for optimal EO selectivity in order to stabilize Ag4-O2 species without causing an excessive amount of strongly bound oxygen species to develop [59,60].
Titanium silicalite 1 (TS-1) contains two different active sites, so it has great activity and selectivity in oxidative hydration. By creating and maintaining TiOOH intermediates, isolated and dinuclear titanium sites lower the activation energy barriers for the synthesis of EO and promote the epoxidation of ethylene [61]. Because they stabilize the OMC intermediate and allow for the catalytic conversion to ethylene glycol, hydrolyzed titanium sites with Ti−OH groups are crucial for the hydration of EO. These hydrolyzed sites are necessary for efficient tandem catalysis, because according to structural optimizations, they have lower energy barriers for hydration than ideal titanium sites. This dual-purpose mechanism brings to the fore the importance of structural tailoring in TS-1 to maximize epoxidation and hydration for industrial applications [62,63]. The comparative data presented in Table 1 summarize various catalysts investigated for ethylene epoxidation, highlighting their mechanisms, structural types, operating conditions, stability, and selectivity toward ethylene oxide. The table also outlines the impact of promoters and dopants, offering insights into how each catalyst’s design influences its performance.
Based on these observations, we can say that the selectivity of ethylene epoxidation towards ethylene oxide (EO) is influenced by a number of factors. Active species like Mn=O in Mn–N4GP and Ag4-O2 in Ag-based catalysts significantly enhance EO selectivity. Dioxygen species favor EO formation, but atomic oxygen favors combustion. Structural stability provided by support materials such as α-Al2O3 and h-BN restricts the formation of by-products, while promoters such as chlorine and dopants such as Cs and Cu facilitate the reaction towards more selective pathways. All these together highlight the significance of optimizing catalyst design in order to obtain maximum EO yield.
Table 1. Comparative summary of catalysts for ethylene epoxidation: mechanisms, structures, and performance parameters.
Table 1. Comparative summary of catalysts for ethylene epoxidation: mechanisms, structures, and performance parameters.
Entry
No.
CatalystMechanismsKey
Feature
EO
Selectivity
Catalyst Structure/TypeOperating ConditionsStabilityPromoters/DopantReferences
1Mn-N4 Porphyrin-like graphene (Mn-N4GP)Two-step: (1) N2O reduction to form Mn=O site, (2) ethylene epoxidation via alkoxide radical pathwayHigh selectivity and stability; thermodynamically favorable Mn=O site105× higher than acetaldehyde, 104× higher than 5MR speciesMn single-atom on porphyrin-like grapheneDFT modelled, mildHigh (DFT shows stability)None[22]
2Ag on B-vacancy h-BN (Ag–h-BN)Trimolecular Langmuir–Hinshelwood mechanism with -CH2CH2OOCH2CH2- intermediateSingle-atom catalyst; strong selectivity; thermally stableHigh, dominant over acetaldehyde (activation barrier: EO 0.44 eV vs. AC much higher)Ag single-atom on B-vacancy h-BNDFT modeled, ambientHigh (embedded single atom)None[25]
3Ag metal (Ag(111), Ag(100))OMC (oxometallacycle) mechanism and dominant OMC-dehydrogenation pathLow EO selectivity under low oxygen pressure; selectivity improves with Ag(100) facet<30% on Ag(111), ~65% on Ag(100)Metallic Ag surfacesHigh temp (500 K), 20 barModerate (metal surface oxidation sensitive)Cs, Cl, alkali metals[32]
4Group IB metals (Ag, Cu, Au)Three stages: (i) O2 dissociation, (ii) Oxometallacycle formation, (iii) EO or AA formationAg offers best balance of O- and C-binding; Cu oxidizes easily; Au poor O2 dissociationAg(100) > Ag(111) > Au/Cu (EO formation barrier higher on Au/Cu)Bulk Group IB metals (Ag, Cu, Au)Various surfaces and O2 pressuresLow for Cu/Au due to poor O2 handlingNone[34]
5AgOx (surface oxide Ag structures)Multiple: Langmuir–Hinshelwood, Eley–Rideal; involves OMC intermediateDiverse mechanisms; phase-dependent activity; AgO_p(4 × 4) surface enables ERVaries with surface; LH dominant in most AgOx, ER in AgO p(4 × 4)Ag surface oxidesVaried O2 pressure & surface coveragePhase-dependentNone[64]
6Doped Ag catalysts (with metal dopants)Cocatalytic mechanism optimizing oxophilicity; modifies O-affinity via dopantsImproved selectivity beyond pristine Ag by tuning O-binding without altering C-bindingHigher than pristine Ag (specific numbers not stated)Ag with metal dopants (Co, Cu, etc.)DFT modeled, optimized O-bindingEnhanced by dopantsCo, Cu, Cs, Cl[39]
7Iridium single-atom on α-MnO2 (Ir1–α-MnO2)π-coordination enabled OMC intermediate formationMolecular-like catalysis; π-interaction between Ir and ethylene enhances selectivity~99%Ir single-atom on α-MnO2~200–250 °C, ambient pressureHighNone[65]
8Ag on α-aluminaStandard Ag-catalyzed epoxidation; affected by particle size and support structureStability via particle size and interparticle distance; Ostwald ripening dominant~80% (industrial with Cl/alkali promoters)Ag on α-Al2O₃Industrial ~230–270 °CGood if particle sintering avoidedCs, Cl[43]
9Electrochemical systems (e.g., RuO2 mediated with Cl)Electrochemical halide-mediated pathway; forms hypobromite intermediatesGreen method; improved selectivity via isolated OMC sitesHigh with halide mediationRuO2 electrode with halide mediation~Room temp, electrochemicalGood (electrolyte optimized)Cl, Br[66]
10Ag on strontium ferrite perovskite (SrFeO3/Sr3Fe2O7 mix)Chemical looping with lattice oxygen donationHigh oxygen capacity; safe oxygen delivery; phase composition influences selectivityUp to 25%, it improved to 60% with ceria dopingAg on SrFeO3/Sr3Fe2O7 perovskiteChemical looping, cyclic oxidation/reductionImproved with RP-phase mixingNone[46]
11Ru single-atom on beta zeoliteHeterogeneous oxidation via isolated Ru active centersHighly dispersed Ru atoms on zeolite provide high activityHighRu single-atom on Beta Zeolite~200–300 °C (lab-scale)High (zeolite structure stabilizes)None[65]
12CoCu Co-doped Ag(111) with oxygen reconstructionOMC mechanism with enhanced desorption and reactivity via dual dopantsDual metal dopants (Co, Cu) optimize oxygen affinity, enhancing EO formation and desorptionUp to 89.5%CoCu-doped Ag(111)~300–400 °CImproved by dual dopantsCo, Cu[67]
13Pt/ZSM-5 with fluorine promotionLow temperature oxidation mechanism enhanced by F-modified acid sitesF-doping enhances low temperature activity and selectivityImproved vs. unmodified Pt/ZSM-5 (quantitative data not provided)Pt/ZSM-5 with F~Low tempImproved with FF[65]
14Ag–ZSM-5Room temperature catalytic oxidation of ethyleneSupports selective EO formation at mild conditions30–40% depending on support acidity and Ag loadingAg–ZSM-5Room temperatureStable under mildNone[65]
15Electrochemical RuO2 with ClHalide-mediated pathway via Cl-inhibited sites leading to EO productionInhibits combustion path; enables EO pathwayHigh selectivityRuO2 with ClElectrochemicalGood with halide controlCl[68]
In conclusion, ethylene epoxidation requires catalysts to achieve high EO selectivity and process efficiency. The significance of adjusting surface characteristics, oxygen dynamics, and electronic structures to improve catalytic performance is highlighted by advancements in silver-based catalysts, hybrid nanostructures, and perovskite systems. New techniques, including chemical looping, plasmonic nanostructures, and doped systems, show promise for increasing EO yields while lessening their negative effects on the environment. More sustainable and scalable EO production methods will surely be fueled by ongoing research in catalyst development and mechanism elucidation.

3. Mechanisms of the Process

Epoxidation is the major process for the manufacture of ethylene oxide, an industrially important chemical. The process has undergone extensive studies, resulting in the proposal of several catalytic routes and mechanisms that are intended to provide maximum efficiency and selectivity with minimum by-products. While new catalysts such as Mn–N4 graphene, Ag-based alloy systems, and Ir single-atom catalysts have been developed through progress in materials science, each containing different mechanisms involving such intermediates as OMCs and alkoxide radicals, conventionally used silver-based catalysts prevail in industrial processes. By underlining the interaction between catalyst composition, reaction intermediates, and activation energy barriers, this literature review summarizes current knowledge of these mechanisms and gives insight into the development and improvement of ethylene epoxidation techniques.
A two-step epoxidation of ethylene into ethylene oxide has been proposed by [22] based on the Mn-coordinated porphyrin-like graphene catalyst Mn–N4GP (see Figure 8) The end-on adsorption of N2O to the Mn center is coupled with the lengthening of the N-O bond and hence charge transfer. The subsequent steps are N-O bond breaking to form a stable Mn=O species via overcoming the energy barrier of 0.77 eV, acting as an active site of ethylene epoxidation. In the second step, ethylene interacts with the Mn=O site via one of three intermediates: carbon radical, alkoxide radical, or manganaoxetane. The alkoxide radical pathway, with a 0.89 eV barrier, is the most favorable, resulting in ethylene oxide formation [69,70].
According to [9], ethylene epoxidation on the Ag atom-embedded B-vacancy defective boron nitride nanosheet follows the trimolecular Langmuir–Hinshelwood mechanism. Two ethylene molecules co-adsorb at the Ag defect site, while further physisorption of an O2 molecule creates a -CH2CH2OOCH2CH2- intermediate. This promptly cyclizes and decomposes into two ethylene oxide molecules via a low activation barrier of 0.44 eV; hence, this three-molecule path should be energetically more favorable compared to the bimolecular route because of higher and more complicated activation barriers [25,71]. Figure 9 illustrates the catalytic behavior of unpromoted silver foil during ethylene oxidation, showing time-dependent trends in product formation and EO selectivity, changes in surface oxygen species via O1s XPS spectra, and the evolving Oelec/Onucl ratio over time.
A study by [13] states that the silver-catalyzed epoxidation of ethylene occurs through a well-defined oxometallacycle (OMC) intermediate thath forms following the dissociative adsorption of O2 molecules on the silver surface, resulting in the formation of reactive oxygen species, which then react with ethylene [32]. Following the formation of the OMC intermediate, three different reaction channels are possible: (1) cyclization to form ethylene oxide (EO), (2) hydrogen transfer to form acetaldehyde (AA), and (3) a new OMC–dehydrogenation (OMC–DH) pathway [72,73,74]. The OMC–DH pathway proposed by Diao et al. has the lowest energy barrier among the three and leads to dehydrogenation of an OxoE (surface-bound oxoethylene) intermediate, which is accountable for low EO selectivity, particularly at low oxygen partial pressures [36]. These pathways are highly surface-sensitive; for instance, the Ag(100) face favors the OMC–DH pathway due to its relatively low activation energy, but even then, EO formation competes favorably owing to its good kinetics on this surface [75]. The Cu and Au surfaces, on the other hand, exhibit poor activity towards O2 dissociation and limited EO formation, further consolidating the unique applicability of Ag to selective epoxidation of ethylene [76].
This general trend for ethylene epoxidation in TM (transition metal)-doped Ag catalysts follows a co-catalytic mechanism related to the different sites for oxygen adsorption and the activation of ethylene [39]. On clean Ag surfaces, this reaction proceeds through an OMC intermediate, which can further either ring-close to EO or isomerize into AA. All these paths possess energy barriers dependent on the adsorption affinities of both oxygen and carbon. Metal dopants enhance EO selectivity by stabilizing the OMC intermediate and lowering energy barriers for EO formation, acting as cocatalysts to promote EO production while minimizing AA formation [39,77,78,79].
These Ir1–α-MnO2 (Table 1, Entry 7) single-atom catalysts catalyze the epoxidation of ethylene through a unique mechanism, as shown in Figure 10. The π-coordination among iridium, ethylene, and molecular oxygen lowers the activation energy for oxygen dissociation in such a way that atomic oxygen can directly react with ethylene to the five-membered-ring oxometallacycle intermediate, producing ethylene oxide selectively [65]. The Ir1–α-MnO2 system showed near-unity EO selectivity (∼99%) with much higher catalytic activity compared to Ag-based catalysts. The iridium center stabilizes the intermediates, and the α-MnO2 support increases catalytic efficiency because of the stabilization of oxygen species. Structural analysis indicates that iridium atoms are present in substitutional and adsorbed states, where the latter are more catalytically active. This synergy ensures exceptional performance for EO production [65,80,81,82].
Ethylene epoxidation over silver catalysts supported on α-alumina (Ag–α-Al2O3) is predominantly via the Langmuir–Hinshelwood (LH) mechanism, in which both ethylene and oxygen species adsorb on the catalyst surface before they react [60]. Selectivity for ethylene oxide (EO) formation arises from adsorbed ethylene reacting selectively with active oxygen species, such as the Ag4–O2 dioxygen species on the reconstructed p(4 × 4)-O-Ag(111) surface, to produce the oxometallacycle (OMC) intermediate and subsequently EO. Competition from side reactions with atomic oxygen can lead to combustion products (CO2 and H2O) [83]. Steady-state isotopic transient kinetic analysis (SSITKA) confirms that EO formation is predominantly via the LH pathway with gas-phase oxygen, whereas CO2 formation is primarily via a minor Mars–van Krevelen (MvK) pathway [84,85]. Furthermore, catalyst activity is very sensitive to silver particle size with an optimum at 60 nm: sintered particle sizes that are larger cause negative effects on both catalytic activity and selectivity through a reduction in surface area and alteration of active site distribution [43,86]. The gentle interaction between silver and oxygen at optimum particle size creates mild oxidation conditions favorable for selective EO formation [76].
Electrochemical ethylene epoxidation proceeds via a multistep mechanism. The most important steps concerning halide-mediated reactions have been taken into consideration. Generally, the forms of active halides such as Cl or Br facilitate the formation of halohydrin intermediates that undergo further dehydrohalogenation in order to afford ethylene oxide (EO) [68]. In this respect, oxidized halides at the anode would yield the active halide species reacting with ethylene to form halohydrins, further converted into EO at the cathode. Studies on RuO2-based catalysts (Table 1, Entry 9) show that halides selectively block diol-like ethylene configurations, favoring epoxide formation and preventing overoxidation to CO2, thereby enhancing EO selectivity [66,87]. Figure 11 illustrates the preferred adsorption configurations and corresponding reaction pathways involved in the electrocatalytic epoxidation of ethylene to ethylene oxide (EO). It compares the surface behavior of pure RuO2 with that of RuO2 modified by adsorbed chloride ions, highlighting the role of Cl in altering the reaction mechanism and potentially enhancing EO selectivity [45].
In the chemical looping epoxidation (CLE) of ethylene, silver (Ag) functions as the active catalyst, while strontium ferrite-based perovskites, such as SrFeO3 and Sr3Fe2O7, serve as oxygen carriers, enabling the selective oxidation of ethylene without the use of gaseous oxygen. The CLE process operates via two sequential steps: (1) a reduction step, in which lattice oxygen from the perovskite is released and transferred to the Ag surface, where it reacts with ethylene to form ethylene oxide (EO) through a key oxometallacycle (OMC) intermediate; and (2) a reoxidation step, wherein the oxygen-depleted perovskite is regenerated in air to restore its lattice oxygen content [46,48]. The performance of this process is highly dependent on the oxygen release rate and the redox stability of the oxygen carriers. A synergistic 1:1 combination of SrFeO3 and Sr3Fe2O7 enhances oxygen transport and EO yield due to improved lattice oxygen mobility. Moreover, supporting Ag on CeO2-modified SrFeO3 perovskites significantly boosts redox kinetics and oxygen release capacity, achieving up to 60% EO selectivity at 10% ethylene conversion over 15 cycles [88,89]. The size of Ag particles also plays a critical role, with larger particles formed at higher calcination temperatures exhibiting better EO selectivity by stabilizing the OMC intermediate and promoting efficient oxygen transfer [15,90]. This integrated CLE strategy offers a promising pathway for selective ethylene epoxidation under milder and more sustainable conditions [91].
Chlorine-promoted ethylene epoxidation over a Ag–CuO nanocatalyst goes by a parallel pathway: a so-called oxometallacycle path and a direct path. The ethylene forms an OMC intermediate that further cyclizes to form EO through the OMC path, whereas the direct path goes via a direct reaction with chemisorbed oxygen. Chlorine lowers the activation energy for both pathways, hence facilitating the formation of EO. It lowers the activation barriers by 20% for the OMC route and 56% for the direct pathway; hence, it improves the selectivity of EO. Chlorine also stabilizes ethylene oxide and prohibits further oxidation to acetaldehyde and CO2 [92,93].
In the case of solar-driven epoxidation of ethylene, the active sites include the plasmonic Ag nanoparticles, whereby LSPR is used to enhance the activation toward oxygen. The photoinduced excitation of Ag NPs into hot electrons is transferred first to the adsorbed molecular oxygen (O2). This step is the one that involves the transient superoxide species, O2, and is important in the activation and dissociation of O2 to reactive oxygen atoms. Further reaction with ethylene, C2H4, absorbed on the surface of Ag through an intermediate known as oxometallacycle (OMC) is an important form that undergoes cyclization, forming ethylene oxide, EO [53]. The plasmonic enhancement effect lowers the activation energy for O2 dissociation and provides the possibility of performing epoxidation at a lower temperature and thus preventing overoxidation to CO2 [94,95].
The O5 phase formed on the Ag(100) surface under industrial conditions promotes the epoxidation of ethylene over silver catalysts through the stabilization of the oxometallacycle intermediate by the square-pyramidal subsurface oxygen atoms (O5β), yielding ethylene oxide [96]. The O5 phase reduces the activation energy for ethylene adsorption and EO production to 0.86 eV and hence energetically is preferreds over competing pathways, such as full oxidation to CO2 or acetaldehyde [97,98]. In contrast to the Ag(111) and Ag(110) surfaces, which are known to favor combustion products, the Ag(100) surface promotes selective EO production because of its unique O5 phase and electronic properties [90,99].
Epoxidation of ethylene over delafossite-based AgNiO2 nanoparticles is a reaction where the lattice oxygen and ethylene are activated at the silver–nickel interface. Ethylene reacts with adsorbed oxygen species on the AgNiO2 surface to form the OMC intermediate, which cyclizes to afford EO without overoxidation. Oxygen species with binding energies of 528–530 eV are responsible for selective epoxidation, and the nickel component stabilizes oxygen vacancies, thus enhancing catalytic performance. The mixed oxide structure is stable at temperatures below 150 °C, ensuring a consistent supply of reactive oxygen [98].
A study by [64] found that the two major mechanisms for ethylene epoxidation over silver-based catalysts were the Langmuir–Hinshelwood (LH) mechanism and the Eley–Rideal (ER) pathway, as illustrated in Figure 12. The LH mechanism proceeds with molecular oxygen dissociated on the silver surface to yield atomic oxygen, which reacts with adsorbed ethylene to yield the oxometallacycle (OMC) intermediate—a key precursor to both ethylene oxide (EO) and acetaldehyde (AA) [35]. This pathway is a prevailing mechanism on most Ag and AgOx surfaces, such as oxygen-covered Ag(111) where microkinetic calculations confirm EO formation as a kinetically controlled process [100,101]. In contrast, the ER mechanism prevails on very oxygen-rich surfaces such as AgO_p(4 × 4) where gas-phase ethylene directly reacts with pre-adsorbed atomic oxygen in the gas phase [67]. This direct interaction significantly lowers the activation energy and enhances EO selectivity. Among the tested surfaces, Ag1.83O(111) presents the lowest energy barrier for EO formation, showing its high efficiency and highlighting the significance of surface composition and oxygen coverage in controlling the prevailing reaction mechanism and product selectivity [72,102].
The oxidation hydration of ethylene into ethylene glycol on TS-1 catalysts proceeds through the two-step tandem mechanism of ethylene epoxidation followed by hydration of the so-formed ethylene oxide [103]. Epoxidation includes the reaction of ethylene with H2O2 at isolated or dinuclear titanium sites to form ethylene oxide through Ti-hydroperoxy (TiOOH) intermediates [104]. In the hydration step, hydrolyzed titanium sites generate reactive hydroxyl groups that stabilize the oxometallacycle (OMC) intermediate, which undergoes ring opening to produce ethylene glycol. This mechanism emphasizes the need for dual-site cooperation: perfect titanium sites for epoxidation and hydrolyzed sites for EO hydration [105,106].

4. Kinetics of Epoxidation of Ethylene

Knowledge of reaction kinetics is helpful in optimization toward good yield and high selectivity for the desired product in ethylene oxide production. For partial ethylene oxidation to ethylene oxide on silver-based catalysts, a number of kinetic models have been used in order to describe epoxidation [107]. The Langmuir–Hinshelwood mechanism remains the most frequently applied to competitive adsorption of ethylene, oxygen, and the selectivity promoter dichloroethane onto catalyst surfaces. These models take into account the interaction of partial and total oxidation reactions of ethylene and ethylene oxide, in which DCE (dichloroethane) is an important modifier for promoting EO selectivity through the inhibition of undesirable side reactions, such as complete oxidation [108]. Kinetic parameters, including reaction rate constants, activation energies, and adsorption constants, are essential in the design of the reactor and process optimization. Notably, models will differ in their inclusion of reverse reactions and DCE effects, leading to variations in predictive accuracy considering different operating conditions [109].
The reaction kinetics in the continuous liquid-phase epoxidation of ethylene with hydrogen peroxide and a titanium–silicate catalyst, TS-1, were studied under a wide range of experimental conditions, such as temperature, pressure, hydrogen peroxide concentration, and liquid flow rates [110]. The catalyst exhibited excellent stability and high selectivity for ethylene oxide for a long period of 150 h. The results indicated that ethylene oxide selectivity strongly depends on temperature and pressure. With increasing temperature, the conversion of ethylene increased, but the selectivity to ethylene oxide decreased because of the enhanced decomposition of hydrogen peroxide [111]. In addition, increasing the hydrogen peroxide concentration showed positive effects on both the conversion of ethylene and selectivity. Higher water concentrations, on the other hand, reduced selectivity by the promotion of secondary reactions [112]. These results indicate that the reaction conditions, particularly the balance between hydrogen peroxide and water, should be optimized in order to maximize ethylene oxide while minimizing by-products formation [113,114].
In the study of ethylene epoxidation with hydrogen peroxide in a trickle bed reactor (TBR), modeling of transient kinetics was conducted in order to enhance the understanding of dynamic behavior. Epoxidation involves the reaction between ethylene (C2H4) and hydrogen peroxide (H2O2) on the surface of a titanium–silicate (TS-1) catalyst. The key reaction mechanism involves adsorption of H2O2 on the catalyst surface followed by reaction with dissolved ethylene to form ethylene oxide (C2H4O) and water [113]. Side reactions include the methanolysis of ethylene oxide. A dynamic model, including mass balances for gas, liquid, and surface phases, was used to simulate both transient and steady-state conditions [115]. The model included axial dispersion for both the gas and liquid phases, where adsorption–desorption and reaction steps took place on the catalyst surface. Estimation of the parameters was conducted by nonlinear regression of experimental data, and it was found that some kinetic parameters—activation energy and rate constants—were very important in the prediction of the behavior of the system [116].
The kinetics of ethylene epoxidation over chloride-promoted silver catalysts is greatly influenced by chlorine coverage in a manner influencing reaction rates and product selectivity to a large extent. Kinetic studies on 35 wt% Ag–α-Al2O3 catalysts reveal that the mechanism is greatly dependent on the chlorinating effectiveness factor of Z* for the level of promotion by surface chlorides [107]. With rising Z*, the reaction order with respect to oxygen increases from approximately 0.7 to close to 1, reflecting higher dependence on oxygen activation [107]. In contrast, the reaction order with respect to ethylene drops because of suppressed ethylene adsorption, owing to higher chloride coverage. Chloride influences catalytic activity through ensemble and electronic effects, and ensemble disruption selectively poisons overall oxidation pathways, resulting in decreased CO2 production and thus enhancing EO selectivity [13,75]. In addition, electronic effects induced by chloride alter the adsorption energies of oxygen and ethylene species and impact reaction kinetic orders and product distribution. A maximum rate of ethylene oxide production is at intermediate chlorine coverage (0–15% monolayer) with a corresponding steady reduction in CO2 production [117,118]. Ethyl chloride and CO2 co-feeding studies also provide this selectivity enhancement as reversibly inhibited EO and CO2 production by co-fed ethyl chloride and preferential inhibition of combustion paths by co-fed CO2 [119]. Thus, ensemble–electronic models give a complete and correct explanation of the observed kinetics, emphasizing the pivotal and intricate role of chlorine in inducing ethylene oxide selectivity in the silver-catalyzed ethylene epoxidation reaction [120].
The kinetics of ethylene epoxidation over silver catalysts has been the focus of extensive studies and modeling to understand the mechanism of this reaction and to optimize reactor performance. In a two-site model, the reaction involves adsorption of ethylene and oxygen species onto active sites on the catalyst surface with the formation of ethylene oxide, EO, via an intermediate oxometallacycle, OME [9]. The most important findings clearly show that models such as Linic–Barteau (LB), Huš–Hellman (HH), and Stegelmann–Stoltze (SS) differ considerably in the reproduction of experimental data. Single-site models are sufficient to reproduce broader tendencies of the system, including the temperature dependence of EO production and the effect of oxygen concentration on selectivity [60]. More detailed models have to be used in order to describe the subtleties of surface oxygen interactions and EO selectivity, especially during transient states, such as multi-pulse oxidation-titration experiments [69]. Notably, the inclusion of a subsurface oxygen reservoir in the SS model has been shown to significantly enhance its ability to simulate EO production and selectivity; however, further refinements are needed to capture the rate of oxygen species decomposition accurately [97,121].

5. Conclusions

One of the more prominent roots of industrial chemical synthesis is ethylene oxide production through ethylene epoxidation, as ethylene oxide is a significant intermediary in a great many consumer goods, petrochemical intermediates, and fine chemicals. Despite the maturity of this process, there are some inherent defects to traditional gas-phase epoxidation using silver catalysts: poor selectivity, high energy use, and significant emissions of CO2.
This literature review highlights such important developments in the catalyst design and reaction process to overcome these limitations. Improved hybrid nanostructures, doped systems, and silver-based catalysts have shown less by-product formation and more selectivity. The emerging catalytic materials, through electronic structure optimization and lattice oxygen dynamics—such as Mn–N4GP and AgNiO2 nanoparticles, respectively—promise potentially viable alternatives for enhancing the reaction pathways while reducing the ecological footprint.
This literature review sheds light on how many important improvements have been made at the level of catalyst design and reaction process in surmounting the challenges cited above. Subsequent enhancements in hybrid nanostructures, doped systems, and silver-based catalysts, on the other hand, have shown lower production of by-products, resulting in increased selectivity. Designing electronic structures for optimization and lattice oxygen dynamics—linked to new materials such as Mn-N4GP and AgNiO2 nanoparticles, respectively—bring encouraging alternatives with more promising pathways and less harm to nature.
With these developments, challenges still abound. In this respect, more research is needed when factors such as catalyst deactivation, the economic feasibility of using different materials, and how to scale up new processes come into play. Further revolutionizing of the industry, therefore, may be gained with the incorporation of renewable feedstocks and deployment of even more predictive computational models for the guidance of EO production toward better sustainability and efficiency.
In the final analysis, this confluence of sophisticated catalyst design, thorough mechanistic research, and sustainable technology ushers in a new era in the synthesis of ethylene oxide. Its continuous innovation and focus on environmental stewardship have the industry well placed to meet the growing demand for EO globally while minimizing its ecological footprint.

Author Contributions

Conceptualization, M.M.H.; methodology, M.M.H., M.A.S. and M.M.; validation, M.M.H.; M.R.Q. and S.A.R.; formal analysis, X, M.M.H.; M.R.Q. and S.A.R.; investigation, M.A.S. and M.M.; resources, M.M.H.; data curation, M.A.S. and M.M.; writing—original draft preparation, M.M.H., M.A.S. and M.M.; writing—review and editing, M.M.H.; M.R.Q. and S.A.R.; visualization, M.M.H., M.A.S. and M.M.; supervision, M.M.H.; project administration, M.M.H.; funding acquisition, M.M.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Deanship of Research at King Fahd University of Petroleum and Minerals (KFUPM, grant INRC2507).

Data Availability Statement

The data used in this study are available in this paper.

Acknowledgments

The author(s) would like to acknowledge the support provided by the Deanship of Research at King Fahd University of Petroleum and Minerals (KFUPM) for funding this work through project INRC2507.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
EOethylene oxide
DFTdensity functional theory
LSPRlocalized surface plasmon resonance
CLEchemical looping epoxidation
LPEliquid phase epoxidation
MTOmethyltrioxorhenium
DOSdensity of states analysis
OMC–DHoxometallacycle–dehydrogenation
DCEdichloroethane
TMtransition metal

References

  1. Pinaeva, L.G.; Noskov, A.S. Prospects for the Development of Ethylene Oxide Production Catalysts and Processes (Review). Pet. Chem. 2020, 60, 1191–1206. [Google Scholar] [CrossRef]
  2. Maqbool, M.; Parveen, N.; Jaffar, S.; Hassan, S.U.; Mahmood, A.; Al-Masry, W.; Kim, T.; Han, S.K.; Park, C.H.; Razzaque, S.; et al. CO2-Free Ethylene Oxide Production via Liquid-Phase Epoxidation with Fe2O3/MSM Catalyst. Chem. Asian J. 2024, 19, e202400002. [Google Scholar] [CrossRef] [PubMed]
  3. Maqbool, M.; Akhter, T.; Faheem, M.; Nadeem, S.; Park, C.H.; Mahmood, A. CO2 Free Production of Ethylene Oxide via Liquid Phase Epoxidation of Ethylene Using Niobium Oxide Incorporated Mesoporous Silica Material as the Catalyst. RSC Adv. 2023, 13, 1779–1786. [Google Scholar] [CrossRef] [PubMed]
  4. Sirijaraensre, J.; Limtrakul, J. Theoretical Investigation on Reaction Pathways for Ethylene Epoxidation on Ti-Decorated Graphene. Struct. Chem. 2018, 29, 159–170. [Google Scholar] [CrossRef]
  5. Xue, W.; Quan, L.; Liu, H.; Yu, B.; Chen, X.; Xia, B.Y.; You, B. Bromine-Enhanced Generation and Epoxidation of Ethylene in Tandem CO2 Electrolysis Towards Ethylene Oxide. Angew. Chem. Int. Ed. 2023, 62, e202311570. [Google Scholar] [CrossRef]
  6. Lockemeyer, J.R.; Lohr, T.L. Ethylene Oxide Catalysis Under Commercial Conditions—A Guide for Researchers. ChemCatChem 2023, 15, e202201511. [Google Scholar] [CrossRef]
  7. Charchi Aghdam, N.; Chen, N.; Soltan, J. Ozonative Epoxidation of Ethylene: A Novel Process for Production of Ethylene Oxide. Appl. Catal. A Gen. 2023, 661, 119239. [Google Scholar] [CrossRef]
  8. van Hoof, A.J.F.; van der Poll, R.C.J.; Friedrich, H.; Hensen, E.J.M. Dynamics of Silver Particles during Ethylene Epoxidation. Appl. Catal. B 2020, 272, 118983. [Google Scholar] [CrossRef]
  9. Pu, T.; Tian, H.; Ford, M.E.; Rangarajan, S.; Wachs, I.E. Overview of Selective Oxidation of Ethylene to Ethylene Oxide by Ag Catalysts. ACS Catal. 2019, 9, 10727–10750. [Google Scholar] [CrossRef]
  10. Aikaterini, B.; Jose Antonio, M.R. Energy Efficiency and GHG Emissions: Prospective Scenarios for the Chemical and Petrochemical Industry; Publications Office of the European Union: Luxembourg, 2017. [Google Scholar] [CrossRef]
  11. Chung, M.; Jin, K.; Zeng, J.S.; Manthiram, K. Mechanism of Chlorine-Mediated Electrochemical Ethylene Oxidation in Saline Water. ACS Catal. 2020, 10, 14015–14023. [Google Scholar] [CrossRef]
  12. Yang, H.; Li, G.; Ma, W.; Hao, B.; Zhang, Z.; Liu, Y.; Hao, Z. Recent Developments in Heterogeneous Catalyzed Epoxidation of Ethylene to Ethylene Oxide. Chem. A Eur. J. 2025, 31, e202404773. [Google Scholar] [CrossRef] [PubMed]
  13. Diao, W.; DiGiulio, C.; Schaal, M.; Ma, S.; Monnier, J. An Investigation on the Role of Re as a Promoter in AgCsRe/α-Al2O3 High-Selectivity, Ethylene Epoxidation Catalysts. J. Catal. 2015, 322, 14–23. [Google Scholar] [CrossRef]
  14. Yan, W.; Ramanathan, A.; Ghanta, M.; Subramaniam, B. Towards Highly Selective Ethylene Epoxidation Catalysts Using Hydrogen Peroxide and Tungsten- or Niobium-Incorporated Mesoporous Silicate (KIT-6). Catal. Sci. Technol. 2014, 4, 4433–4439. [Google Scholar] [CrossRef]
  15. Marek, E.J.; García-Calvo Conde, E. Effect of Catalyst Preparation and Storage on Chemical Looping Epoxidation of Ethylene. Chem. Eng. J. 2021, 417, 127981. [Google Scholar] [CrossRef]
  16. Van den Reijen, J.E.; Versluis, W.C.; Kanungo, S.; d’Angelo, M.F.; de Jong, K.P.; de Jongh, P.E. From Qualitative to Quantitative Understanding of Support Effects on the Selectivity in Silver Catalyzed Ethylene Epoxidation. Catal. Today 2019, 338, 31–39. [Google Scholar] [CrossRef]
  17. Liu, B.; Ji, J.; Zhang, B.; Huang, W.; Gan, Y.; Leung, D.Y.C.; Huang, H. Catalytic Ozonation of VOCs at Low Temperature: A Comprehensive Review. J. Hazard. Mater. 2022, 422, 126847. [Google Scholar] [CrossRef] [PubMed]
  18. Lambert, R.M.; Williams, F.J.; Cropley, R.L.; Palermo, A. Heterogeneous Alkene Epoxidation: Past, Present and Future. J. Mol. Catal. A Chem. 2005, 228, 27–33. [Google Scholar] [CrossRef]
  19. Özbek, M.O.; van Santen, R.A. The Mechanism of Ethylene Epoxidation Catalysis. Catal. Lett. 2013, 143, 131–141. [Google Scholar] [CrossRef]
  20. Shaw, M.S.; Bates, M.R.; Jones, M.D.; Ward, B.D. Metallocene Catalysts for the Ring-Opening Co-Polymerisation of Epoxides and Cyclic Anhydrides. Polym. Chem. 2022, 13, 3315–3324. [Google Scholar] [CrossRef]
  21. Kurańska, M.; Beneš, H.; Prociak, A.; Trhlíková, O.; Walterová, Z.; Stochlińska, W. Investigation of Epoxidation of Used Cooking Oils with Homogeneous and Heterogeneous Catalysts. J. Clean. Prod. 2019, 236, 117615. [Google Scholar] [CrossRef]
  22. Impeng, S.; Roongcharoen, T.; Maitarad, P.; Wu, H.; Chitpakdee, C.; Promarak, V.; Shi, L.; Namuangruk, S. High Selective Catalyst for Ethylene Epoxidation to Ethylene Oxide: A DFT Investigation. Appl. Surf. Sci. 2020, 513, 145799. [Google Scholar] [CrossRef]
  23. Yang, Y.; Mao, K.; Gao, S.; Huang, H.; Xia, G.; Lin, Z.; Jiang, P.; Wang, C.; Wang, H.; Chen, Q. O-, N-Atoms-Coordinated Mn Cofactors within a Graphene Framework as Bioinspired Oxygen Reduction Reaction Electrocatalysts. Adv. Mater. 2018, 30, e1801732. [Google Scholar] [CrossRef]
  24. Impeng, S.; Siwaipram, S.; Bureekaew, S.; Probst, M. Ethane C-H Bond ActIVation on the Fe(IV)-Oxo Species in a Zn-Based Cluster of Metal-Organic Frameworks: A Density Functional Theory Study. Phys. Chem. Chem. Phys. 2017, 19, 3782–3791. [Google Scholar] [CrossRef] [PubMed]
  25. Esrafili, M.D.; Janebi, H.; Mousavian, P. Epoxidation of Ethylene over an Ag Atom Embedded B-Vacancy Defective Boron-Nitride Nanosheet via a Trimolecular Langmuir–Hinshelwood Mechanism: A DFT Investigation. Mol. Catal. 2021, 514, 111843. [Google Scholar] [CrossRef]
  26. García-Mota, M.; Rieger, M.; Reuter, K. Ab Initio Prediction of the Equilibrium Shape of Supported Ag Nanoparticles on α-Al2O3(0 0 0 1). J. Catal. 2015, 321, 1–6. [Google Scholar] [CrossRef]
  27. Zhang, X.; Lei, J.; Wu, D.; Zhao, X.; Jing, Y.; Zhou, Z. A Ti-Anchored Ti2CO2 Monolayer (MXene) as a Single-Atom Catalyst for CO Oxidation. J. Mater. Chem. A Mater. 2016, 4, 4871–4876. [Google Scholar] [CrossRef]
  28. Esrafili, M.D.; Mousavian, P. The Strengthening Effect of a Halogen, Chalcogen or Pnicogen Bonding on Halogen–π Interaction: A Comparative Ab Initio Study. Mol. Phys. 2018, 116, 526–535. [Google Scholar] [CrossRef]
  29. Lohr, T.L.; Lockemeyer, J.R.; Bishopp, S.D.; Motagamwala, A.H.; Wells, G.J.; Wermink, T. Ethylene Oxide: A Catalyst and Process Development Success Story. Ind. Eng. Chem. Res. 2024, 63, 18221–18240. [Google Scholar] [CrossRef]
  30. Tarrat, N.; Rapacioli, M.; Cuny, J.; Morillo, J.; Heully, J.L.; Spiegelman, F. Global Optimization of Neutral and Charged 20- and 55-Atom Silver and Gold Clusters at the DFTB Level. Comput. Theor. Chem. 2017, 1107, 102–114. [Google Scholar] [CrossRef]
  31. Li, X.T.; Chen, L.; Wei, G.F.; Shang, C.; Liu, Z.P. Sharp Increase in Catalytic Selectivity in Acetylene Semihydrogenation on Pd Achieved by a Machine Learning Simulation-Guided Experiment. ACS Catal. 2020, 10, 9694–9705. [Google Scholar] [CrossRef]
  32. Chen, D.; Kang, P.L.; Liu, Z.P. Active Site of Catalytic Ethene Epoxidation: Machine-Learning Global Pathway Sampling Rules out the Metal Sites. ACS Catal. 2021, 11, 8317–8326. [Google Scholar] [CrossRef]
  33. Li, H.; Nørskov, J.K. Effects of a Conductive Support on the Bonding of Oxygen Containing Molecules to Transition Metal Oxide Surfaces. Phys. Chem. Chem. Phys. 2020, 22, 26216–26222. [Google Scholar] [CrossRef] [PubMed]
  34. Li, H.; Cao, A.; Nørskov, J.K. Understanding Trends in Ethylene Epoxidation on Group IB Metals. ACS Catal. 2021, 11, 12052–12057. [Google Scholar] [CrossRef]
  35. Liu, J.X.; Su, Y.; Filot, I.A.W.; Hensen, E.J.M. A Linear Scaling Relation for CO Oxidation on CeO2-Supported Pd. J. Am. Chem. Soc. 2018, 140, 4580–4587. [Google Scholar] [CrossRef]
  36. Wu, S.; Tatarchuk, B.J.; Adamczyk, A.J. Ethylene Oxidation on Unpromoted Silver Catalysts: Reaction Pathway and Selectivity Analysis Using DFT Calculations. Surf. Sci. 2021, 708, 121834. [Google Scholar] [CrossRef]
  37. Van Hoof, A.J.F.; Filot, I.A.W.; Friedrich, H.; Hensen, E.J.M. Reversible Restructuring of Silver Particles during Ethylene Epoxidation. ACS Catal. 2018, 8, 11794–11800. [Google Scholar] [CrossRef]
  38. Greiner, M.T.; Jones, T.E.; Johnson, B.E.; Rocha, T.C.R.; Wang, Z.J.; Armbrüster, M.; Willinger, M.; Knop-Gericke, A.; Schlögl, R. The Oxidation of Copper Catalysts during Ethylene Epoxidation. Phys. Chem. Chem. Phys. 2015, 17, 25073–25089. [Google Scholar] [CrossRef] [PubMed]
  39. Xu, H.; Zhu, L.; Nan, Y.; Xie, Y.; Cheng, D. Revisit the Role of Metal Dopants in Enhancing the Selectivity of Ag-Catalyzed Ethylene Epoxidation: Optimizing Oxophilicity of Reaction Site via Cocatalytic Mechanism. ACS Catal. 2021, 11, 3371–3383. [Google Scholar] [CrossRef]
  40. Xu, H.; Cheng, D.; Gao, Y.; Zeng, X.C. Assessment of Catalytic Activities of Gold Nanoclusters with Simple Structure Descriptors. ACS Catal. 2018, 8, 9702–9710. [Google Scholar] [CrossRef]
  41. Calle-Vallejo, F.; Loffreda, D.; Koper, M.T.M.; Sautet, P. Introducing Structural Sensitivity into Adsorption-Energy Scaling Relations by Means of Coordination Numbers. Nat. Chem. 2015, 7, 403–410. [Google Scholar] [CrossRef]
  42. Li, F.; Li, Y.; Zeng, X.C.; Chen, Z. Exploration of High-Performance Single-Atom Catalysts on Support M1/FeOx for CO Oxidation via Computational Study. ACS Catal. 2015, 5, 544–552. [Google Scholar] [CrossRef]
  43. Keijzer, P.H.; van den Reijen, J.E.; Keijzer, C.J.; de Jong, K.P.; de Jongh, P.E. Influence of Atmosphere, Interparticle Distance and Support on the Stability of Silver on α-Alumina for Ethylene Epoxidation. J. Catal. 2022, 405, 534–544. [Google Scholar] [CrossRef]
  44. Hu, S.; Li, W.X. Influence of Particle Size Distribution on Lifetime and Thermal Stability of Ostwald Ripening of Supported Particles. ChemCatChem 2018, 10, 2900–2907. [Google Scholar] [CrossRef]
  45. Zhang, P.; Wang, T.; Gong, J. Advances in Electrochemical Oxidation of Olefins to Epoxides. CCS Chem. 2023, 5, 1028–1042. [Google Scholar] [CrossRef]
  46. Gabra, S.; Marek, E.J.; Poulston, S.; Williams, G.; Dennis, J.S. The Use of Strontium Ferrite Perovskite as an Oxygen Carrier in the Chemical Looping Epoxidation of Ethylene. Appl. Catal. B 2021, 286, 119821. [Google Scholar] [CrossRef]
  47. Lau, C.Y.; Dunstan, M.T.; Hu, W.; Grey, C.P.; Scott, S.A. Large Scale in Silico Screening of Materials for Carbon Capture through Chemical Looping. Energy Environ. Sci. 2017, 10, 818–831. [Google Scholar] [CrossRef]
  48. Marek, E.J.; Gabra, S.; Dennis, J.S.; Scott, S.A. High Selectivity Epoxidation of Ethylene in Chemical Looping Setup. Appl. Catal. B 2020, 262, 118216. [Google Scholar] [CrossRef]
  49. Hossain, S.T.; Azeeva, E.; Zhang, K.; Zell, E.T.; Bernard, D.T.; Balaz, S.; Wang, R. A Comparative Study of CO Oxidation over Cu-O-Ce Solid Solutions and CuO/CeO2 Nanorods Catalysts. Appl. Surf. Sci. 2018, 455, 132–143. [Google Scholar] [CrossRef]
  50. Suarez, A.; Ramirez, A.; Bueno-Alejo, C.; Hueso, J. Silver-Copper Oxide Heteronanostructures for the Plasmonic-Enhanced Photocatalytic Oxidation of N-Hexane in the Visible-NIR Range. Materials 2019, 12, 3858. [Google Scholar] [CrossRef]
  51. Cropley, R.L.; Williams, F.J.; Urquhart, A.J.; Vaughan, O.P.H.; Tikhov, M.S.; Lambert, R.M. Efficient Epoxidation of a Terminal Alkene Containing Allylic Hydrogen Atoms: Trans-Methylstyrene on Cu{111}. J. Am. Chem. Soc. 2005, 127, 6069–6076. [Google Scholar] [CrossRef]
  52. Ramirez, A.; Hueso, J.L.; Suarez, H.; Mallada, R.; Ibarra, A.; Irusta, S.; Santamaria, J. A Nanoarchitecture Based on Silver and Copper Oxide with an Exceptional Response in the Chlorine-Promoted Epoxidation of Ethylene. Angew. Chem. Int. Ed. Engl. 2016, 55, 11158–11161. [Google Scholar] [CrossRef] [PubMed]
  53. Tang, D.; Dang, K.; Wang, J.; Chen, C.; Zhao, J.; Zhang, Y. Solar-Driven Green Synthesis of Epoxides. Sci. China Chem. 2023, 66, 3415–3425. [Google Scholar] [CrossRef]
  54. Dorchies, F.; Serva, A.; Crevel, D.; De Freitas, J.; Kostopoulos, N.; Robert, M.; Sel, O.; Salanne, M.; Grimaud, A. Controlling the Hydrophilicity of the Electrochemical Interface to Modulate the Oxygen-Atom Transfer in Electrocatalytic Epoxidation Reactions. J. Am. Chem. Soc. 2022, 144, 22734–22746. [Google Scholar] [CrossRef] [PubMed]
  55. Sayama, K. Production of High-Value-Added Chemicals on Oxide Semiconductor Photoanodes under Visible Light for Solar Chemical-Conversion Processes. ACS Energy Lett. 2018, 3, 1093–1101. [Google Scholar] [CrossRef]
  56. Hardwick, T.; Qurashi, A.; Shirinfar, B.; Ahmed, N. Interfacial Photoelectrochemical Catalysis: Solar-Induced Green Synthesis of Organic Molecules. ChemSusChem 2020, 13, 1967–1973. [Google Scholar] [CrossRef]
  57. Svintsitskiy, D.A.; Lazarev, M.K.; Slavinskaya, E.M.; Fedorova, E.A.; Kardash, T.Y.; Cherepanova, S.V.; Boronin, A.I. Room Temperature Epoxidation of Ethylene over Delafossite-Based AgNiO2 Nanoparticles. Phys. Chem. Chem. Phys. 2023, 25, 20892–20902. [Google Scholar] [CrossRef]
  58. Tang, Z.; Chen, T.; Liu, K.; Du, H.; Podkolzin, S.G. Atomic, Molecular and Hybrid Oxygen Structures on Silver. Langmuir 2021, 37, 11603–11610. [Google Scholar] [CrossRef]
  59. Prins, R. Eley–Rideal, the Other Mechanism. Top. Catal. 2018, 61, 714–721. [Google Scholar] [CrossRef]
  60. Pu, T.; Setiawan, A.; Foucher, A.C.; Guo, M.; Jehng, J.M.; Zhu, M.; Ford, M.E.; Stach, E.A.; Rangarajan, S.; Wachs, I.E. Revealing the Nature of Active Oxygen Species and Reaction Mechanism of Ethylene Epoxidation by Supported Ag/α-Al2O3 Catalysts. ACS Catal. 2024, 14, 406–417. [Google Scholar] [CrossRef]
  61. Harris, J.W.; Arvay, J.; Mitchell, G.; Delgass, W.N.; Ribeiro, F.H. Propylene Oxide Inhibits Propylene Epoxidation over Au/TS-1. J. Catal. 2018, 365, 105–114. [Google Scholar] [CrossRef]
  62. Smidstrup, S.; Pedersen, A.; Stokbro, K.; Jónsson, H. Improved Initial Guess for Minimum Energy Path Calculations. J. Chem. Phys. 2014, 140, 214106. [Google Scholar] [CrossRef] [PubMed]
  63. Kumer, A.; Sarker, M.N.; Paul, S. The Theoretical Investigation of HOMO, LUMO, Thermophysical Properties and QSAR Study of Some Aromatic Carboxylic Acids Using HyperChem Programming. Int. J. Chem. Technol. 2019, 3, 26–37. [Google Scholar] [CrossRef]
  64. Liu, J.X.; Lu, S.; Ann, S.B.; Linic, S. Mechanisms of Ethylene Epoxidation over Silver from Machine Learning-Accelerated First-Principles Modeling and Microkinetic Simulations. ACS Catal. 2023, 13, 8955–8962. [Google Scholar] [CrossRef]
  65. Yang, H.; Wang, X.; Liu, Q.; Huang, A.; Zhang, X.; Yu, Y.; Zhuang, Z.; Li, G.; Li, Y.; Peng, Q.; et al. Heterogeneous Iridium Single-Atom Molecular-like Catalysis for Epoxidation of Ethylene. J. Am. Chem. Soc. 2023, 145, 6658–6670. [Google Scholar] [CrossRef]
  66. Li, X.; Chen, G.; Liu, Y.; Lu, R.; Ma, C.; Wang, Z.; Han, Y.; Wang, D. Optimal Selection of RuO2 for Durable Oxygen Evolution Reactions in Acidic Media by Continuous Regulation of Ru–O Covalency. Energy Environ. Sci. 2025, 18, 4200–4209. [Google Scholar] [CrossRef]
  67. Xu, H.; Zhu, L.; Nan, Y.; Xie, Y.; Cheng, D. Revisit the Role of Chlorine in Selectivity Enhancement of Ethylene Epoxidation. Ind. Eng. Chem. Res. 2019, 58, 21403–21412. [Google Scholar] [CrossRef]
  68. Liang, N.N.; Choi, W.; Suk Han, D.; Park, H. Electrocatalytic Conversion of Ethylene to Ethylene Oxide Mediated by Halide Oxidation: Chloride vs. Bromide vs. Iodide. Chem. Eng. J. 2024, 494, 153042. [Google Scholar] [CrossRef]
  69. Jones, T.E.; Wyrwich, R.; Böcklein, S.; Carbonio, E.A.; Greiner, M.T.; Klyushin, A.Y.; Moritz, W.; Locatelli, A.; Menteş, T.O.; Niño, M.A.; et al. The Selective Species in Ethylene Epoxidation on Silver. ACS Catal. 2018, 8, 3844–3852. [Google Scholar] [CrossRef]
  70. Ma, W.; Yuan, H.; Wang, H.; Zhou, Q.; Kong, K.; Li, D.; Yao, Y.; Hou, Z. Identifying Catalytically Active Mononuclear Peroxoniobate Anion of Ionic Liquids in the Epoxidation of Olefins. ACS Catal. 2018, 8, 4645–4659. [Google Scholar] [CrossRef]
  71. Silbaugh, T.L.; Devlaminck, P.; Sofranko, J.A.; Barteau, M.A. Selective Oxidation of Ethanol over Ag, Cu and Au Nanoparticles Supported on Li2O/Γ-Al2O3. J. Catal. 2018, 364, 40–47. [Google Scholar] [CrossRef]
  72. Greiner, M.T.; Jones, T.E.; Klyushin, A.; Knop-Gericke, A.; Schlögl, R. Ethylene Epoxidation at the Phase Transition of Copper Oxides. J. Am. Chem. Soc. 2017, 139, 11825–11832. [Google Scholar] [CrossRef] [PubMed]
  73. Greiner, M.T.; Cao, J.; Jones, T.E.; Beeg, S.; Skorupska, K.; Carbonio, E.A.; Sezen, H.; Amati, M.; Gregoratti, L.; Willinger, M.G.; et al. Phase Coexistence of Multiple Copper Oxides on AgCu Catalysts during Ethylene Epoxidation. ACS Catal. 2018, 8, 2286–2295. [Google Scholar] [CrossRef]
  74. Carbonio, E.A.; Rocha, T.C.R.; Klyushin, A.Y.; Píš, I.; Magnano, E.; Nappini, S.; Piccinin, S.; Knop-Gericke, A.; Schlögl, R.; Jones, T.E. Are Multiple Oxygen Species Selective in Ethylene Epoxidation on Silver? Chem. Sci. 2018, 9, 990–998. [Google Scholar] [CrossRef] [PubMed]
  75. Huš, M.; Hellman, A. Ethylene Epoxidation on Ag(100), Ag(110), and Ag(111): A Joint Ab Initio and Kinetic Monte Carlo Study and Comparison with Experiments. ACS Catal. 2019, 9, 1183–1196. [Google Scholar] [CrossRef]
  76. Van den Reijen, J.E.; Kanungo, S.; Welling, T.A.J.; Versluijs-Helder, M.; Nijhuis, T.A.; de Jong, K.P.; de Jongh, P.E. Preparation and Particle Size Effects of Ag/A-Al2O3 Catalysts for Ethylene Epoxidation. J. Catal. 2017, 356, 65–74. [Google Scholar] [CrossRef]
  77. Ren, D.; Cheng, G.; Li, J.; Li, J.; Dai, W.; Sun, X.X.; Cheng, D. Effect of Rhenium Loading Sequence on Selectivity of Ag–Cs Catalyst for Ethylene Epoxidation. Catal. Lett. 2017, 147, 2920–2928. [Google Scholar] [CrossRef]
  78. Xu, H.; Cheng, D.; Cao, D.; Zeng, X.C. A Universal Principle for a Rational Design of Single-Atom Electrocatalysts. Nat. Catal. 2018, 1, 339–348. [Google Scholar] [CrossRef]
  79. Medford, A.J.; Vojvodic, A.; Hummelshøj, J.S.; Voss, J.; Abild-Pedersen, F.; Studt, F.; Bligaard, T.; Nilsson, A.; Nørskov, J.K. From the Sabatier Principle to a Predictive Theory of Transition-Metal Heterogeneous Catalysis. J. Catal. 2015, 328, 36–42. [Google Scholar] [CrossRef]
  80. Cui, X.; Li, W.; Ryabchuk, P.; Junge, K.; Beller, M. Bridging Homogeneous and Heterogeneous Catalysis by Heterogeneous Single-Metal-Site Catalysts. Nat. Catal. 2018, 1, 385–397. [Google Scholar] [CrossRef]
  81. Yang, H.; Ma, C.; Zhang, X.; Li, Y.; Cheng, J.; Hao, Z. Understanding the Active Sites of Ag/Zeolites and Deactivation Mechanism of Ethylene Catalytic Oxidation at Room Temperature. ACS Catal. 2018, 8, 1248–1258. [Google Scholar] [CrossRef]
  82. Yang, H.; Ma, C.; Li, Y.; Wang, J.; Zhang, X.; Wang, G.; Qiao, N.; Sun, Y.; Cheng, J.; Hao, Z. Synthesis, Characterization and Evaluations of the Ag/ZSM-5 for Ethylene Oxidation at Room Temperature: Investigating the Effect of Water and Deactivation. Chem. Eng. J. 2018, 347, 808–818. [Google Scholar] [CrossRef]
  83. Lamoth, M.; Jones, T.; Plodinec, M.; Machoke, A.; Wrabetz, S.; Krämer, M.; Karpov, A.; Rosowski, F.; Piccinin, S.; Schlögl, R.; et al. Nanocatalysts Unravel the Selective State of Ag. ChemCatChem 2020, 12, 2977–2988. [Google Scholar] [CrossRef]
  84. Hess, C. New Advances in Using Raman Spectroscopy for the Characterization of Catalysts and Catalytic Reactions. Chem. Soc. Rev. 2021, 50, 3519–3564. [Google Scholar] [CrossRef] [PubMed]
  85. Pu, T.; Setiawan, A.; Mosevitzky Lis, B.; Zhu, M.; Ford, M.E.; Rangarajan, S.; Wachs, I.E. Nature and Reactivity of Oxygen Species on/in Silver Catalysts during Ethylene Oxidation. ACS Catal. 2022, 12, 4375–4381. [Google Scholar] [CrossRef]
  86. Goodman, E.D.; Schwalbe, J.A.; Cargnello, M. Mechanistic Understanding and the Rational Design of Sinter- Resistant Heterogeneous Catalysts. ACS Catal. 2017, 7, 7156–7173. [Google Scholar] [CrossRef]
  87. Jirkovský, J.S.; Busch, M.; Ahlberg, E.; Panas, I.; Krtil, P. Switching on the Electrocatalytic Ethene Epoxidation on Nanocrystalline RuO2. J. Am. Chem. Soc. 2011, 133, 5882–5892. [Google Scholar] [CrossRef] [PubMed]
  88. Luo, Y.; Chen, R.; Peng, W.; Tang, G.; Gao, X. Inverse CeO2 [Sbnd]Fe2O3 Catalyst for Superior Low-Temperature CO Conversion Efficiency. Appl. Surf. Sci. 2017, 416, 911–917. [Google Scholar] [CrossRef]
  89. Zhang, D.; Kawada, T.; Yoshioka, F.; Machida, M. Oxygen Gateway Effect of CeO2/La2O2SO4 Composite Oxygen Storage Materials. ACS Omega 2016, 1, 789–798. [Google Scholar] [CrossRef]
  90. Li, J.L.; Li, Y.F.; Liu, Z.P. In Situ Structure of a Mo-Doped Pt-Ni Catalyst during Electrochemical Oxygen Reduction Resolved from Machine Learning-Based Grand Canonical Global Optimization. JACS Au 2023, 3, 1162–1175. [Google Scholar] [CrossRef]
  91. McNeil, L.A.; Mutch, G.A.; Iacoviello, F.; Bailey, J.J.; Triantafyllou, G.; Neagu, D.; Miller, T.S.; Papaioannou, E.I.; Hu, W.; Brett, D.J.L.; et al. Dendritic Silver Self-Assembly in Molten-Carbonate Membranes for Efficient Carbon Dioxide Capture. Energy Environ. Sci. 2020, 13, 1766–1775. [Google Scholar] [CrossRef]
  92. Shi, L.; Yin, Y.; Wang, S.; Sun, H. Rational Catalyst Design for N2Reduction under Ambient Conditions: Strategies toward Enhanced Conversion Efficiency. ACS Catal. 2020, 10, 6870–6899. [Google Scholar] [CrossRef]
  93. Urbiztondo, M.; Ramirez, A.; Hueso, J.L.; Santamaria, J.; Ruiz-Salvador, A.R.; Hamad, S. Unravelling the Key Factors in the Chlorine-Promoted Epoxidation of Ethylene over a Silver-Copper Oxide Nanocatalyst. Nanoscale 2022, 14, 7332–7340. [Google Scholar] [CrossRef]
  94. Chung, M.; Jin, K.; Zeng, J.S.; Ton, T.N.; Manthiram, K. Tuning Single-Atom Dopants on Manganese Oxide for Selective Electrocatalytic Cyclooctene Epoxidation. J. Am. Chem. Soc. 2022, 144, 17416–17422. [Google Scholar] [CrossRef] [PubMed]
  95. Zhao, Y.; Duan, M.; Deng, C.; Yang, J.; Yang, S.; Zhang, Y.; Sheng, H.; Li, Y.; Chen, C.; Zhao, J. Br−/BrO−-Mediated Highly Efficient Photoelectrochemical Epoxidation of Alkenes on α-Fe2O3. Nat. Commun. 2023, 14, 1943. [Google Scholar] [CrossRef] [PubMed]
  96. Jørgensen, M.; Mortensen, H.; Meldgaard, S.; Kolsbjerg, E.; Jacobsen, T.; Sørensen, K.; Hammer, B. Atomistic Structure Learning. J. Chem. Phys. 2019, 151, 054111. [Google Scholar] [CrossRef]
  97. Huš, M.; Grilc, M.; Teržan, J.; Gyergyek, S.; Likozar, B.; Hellman, A. Going Beyond Silver in Ethylene Epoxidation with First-Principles Catalyst Screening. Angew. Chem. Int. Ed. 2023, 62, e202305804. [Google Scholar] [CrossRef]
  98. Wexler, R.B.; Qiu, T.; Rappe, A.M. Automatic Prediction of Surface Phase Diagrams Using Ab Initio Grand Canonical Monte Carlo. J. Phys. Chem. C 2019, 123, 2321–2328. [Google Scholar] [CrossRef]
  99. Chen, D.; Chen, L.; Zhao, Q.C.; Yang, Z.X.; Shang, C.; Liu, Z.P. Square-Pyramidal Subsurface Oxygen [Ag4OAg] Drives Selective Ethene Epoxidation on Silver. Nat. Catal. 2024, 7, 536–545. [Google Scholar] [CrossRef]
  100. Mukda, N.; Autthanit, C.; Praserthdam, P.; Jongsomjit, B. Production of Acetaldehyde via Oxidative Dehydrogenation of Ethanol over AgLi/SiO2 Catalysts. Bull. Chem. React. Eng. Catal. 2020, 15, 714–725. [Google Scholar] [CrossRef]
  101. Christopher, P.; Linic, S. Engineering Selectivity in Heterogeneous Catalysis: Ag Nanowires as Selective Ethylene Epoxidation Catalysts. J. Am. Chem. Soc. 2008, 130, 11264–11265. [Google Scholar] [CrossRef]
  102. Özbek, M.O.; Önal, I.; Vansanten, R.A. Ethylene Epoxidation Catalyzed by Silver Oxide. ChemCatChem 2011, 3, 150–153. [Google Scholar] [CrossRef]
  103. Xu, H.; Xu, H.; Cheng, D. Resolving the Reaction Mechanism for Oxidative Hydration of Ethylene toward Ethylene Glycol by Titanosilicate Catalysts. ACS Catal. 2022, 12, 9446–9457. [Google Scholar] [CrossRef]
  104. Nie, X.; Ren, X.; Ji, X.; Chen, Y.; Janik, M.J.; Guo, X.; Song, C. Mechanistic Insight into Propylene Epoxidation with H2O2 over Titanium Silicalite-1: Effects of Zeolite Confinement and Solvent. J. Phys. Chem. B 2019, 123, 7410–7423. [Google Scholar] [CrossRef] [PubMed]
  105. Wang, B.; Chen, S.; Zhang, J.; Li, S.; Yang, B. Propagating DFT Uncertainty to Mechanism Determination, Degree of Rate Control, and Coverage Analysis: The Kinetics of Dry Reforming of Methane. J. Phys. Chem. C 2019, 123, 30389–30397. [Google Scholar] [CrossRef]
  106. Pang, J.; Zheng, M.; Sun, R.; Wang, A.; Wang, X.; Zhang, T. Synthesis of Ethylene Glycol and Terephthalic Acid from Biomass for Producing PET. Green. Chem. 2016, 18, 342–359. [Google Scholar] [CrossRef]
  107. Chen, C.J.; Harris, J.W.; Bhan, A. Kinetics of Ethylene Epoxidation on a Promoted Ag/α-Al2O3 Catalyst—The Effects of Product and Chloride Co-Feeds on Rates and Selectivity. Chem. A Eur. J. 2018, 24, 12405–12415. [Google Scholar] [CrossRef]
  108. Fazeli, A.; Naseri, A.; Eslamjamal, F. Kinetic Models of Ethylene Oxide Production on Ag Catalysts: A Review. Kinet. Catal. 2020, 61, 603–612. [Google Scholar] [CrossRef]
  109. Partopour, B.; Dixon, A.G. Resolved-Particle Fixed Bed CFD with Microkinetics for Ethylene Oxidation. AIChE J. 2017, 63, 87–94. [Google Scholar] [CrossRef]
  110. Ghanta, M.; Ruddy, T.; Fahey, D.; Busch, D.; Subramaniam, B. Is the Liquid-Phase H2O2-Based Ethylene Oxide Process More Economical and Greener than the Gas-Phase O2-Based Silver-Catalyzed Process? Ind. Eng. Chem. Res. 2013, 52, 18–29. [Google Scholar] [CrossRef]
  111. Wang, B.; Han, H.; Ge, B.; Ma, J.; Zhu, J.; Chen, S. An Efficient Hydrophobic Modification of TS-1 and Its Application in the Epoxidation of Propylene. New J. Chem. 2019, 43, 10390–10397. [Google Scholar] [CrossRef]
  112. Alvear, M.; Fortunato, M.E.; Russo, V.; Eränen, K.; Di Serio, M.; Lehtonen, J.; Rautiainen, S.; Murzin, D.; Salmi, T. Continuous Liquid-Phase Epoxidation of Ethylene with Hydrogen Peroxide on a Titanium-Silicate Catalyst. Ind. Eng. Chem. Res. 2021, 60, 9429–9436. [Google Scholar] [CrossRef]
  113. Alvear, M.; Eranen, K.; Murzin, D.Y.; Salmi, T. Study of the Product Distribution in the Epoxidation of Propylene over TS-1 Catalyst in a Trickle-Bed Reactor. Ind. Eng. Chem. Res. 2021, 60, 2430–2438. [Google Scholar] [CrossRef]
  114. Sulimov, A.V.; Danov, S.M.; Ovcharova, A.V.; Ovcharov, A.A.; Flid, V.R. Kinetics of Propylene Epoxidation with Hydrogen Peroxide Catalyzed by Extruded Titanium Silicalite in Methanol. Kinet. Catal. 2016, 57, 466–473. [Google Scholar] [CrossRef]
  115. Alvear, M.; Fortunato, M.E.; Russo, V.; Salmi, T.; Serio, M. Di Modelling of Transient Kinetics in Trickle Bed Reactors: Ethylene Oxide Production via Hydrogen Peroxide. Chem. Eng. Sci. 2022, 248, 117156. [Google Scholar] [CrossRef]
  116. Russo, V.; Santacesaria, E.; Tesser, R.; Turco, R.; Vitiello, R.; Di Serio, M. Validation of the Kinetics of the Hydrogen Peroxide Propene Oxide Process in a Dynamic Continuous Stirred Tank Reactor. Ind. Eng. Chem. Res. 2018, 57, 16201–16208. [Google Scholar] [CrossRef]
  117. Harris, J.W.; Wang, L.; Dewilde, J.F. Kinetic Modeling of Ethylene Epoxidation Kinetics as a Function of Chlorine Coverage over a Highly Promoted Silver Catalyst. Ind. Eng. Chem. Res. 2019, 58, 19061–19071. [Google Scholar] [CrossRef]
  118. Majumdar, P.; Greeley, J. Generalized Scaling Relationships on Transition Metals: Influence of Adsorbate-Coadsorbate Interactions. Phys. Rev. Mater. 2018, 2, 45801. [Google Scholar] [CrossRef]
  119. Xu, B.; Hartigan, E.M.; Feula, G.; Huang, Z.; Lumb, J.P.; Arndtsen, B.A. Simple Copper Catalysts for the Aerobic Oxidation of Amines: Selectivity Control by the Counterion. Angew. Chem. Int. Ed. 2016, 55, 15802–15806. [Google Scholar] [CrossRef]
  120. Rocha, T.C.R.; Hävecker, M.; Knop-Gericke, A.; Schlögl, R. Promoters in Heterogeneous Catalysis: The Role of Cl on Ethylene Epoxidation over Ag. J. Catal. 2014, 312, 12–16. [Google Scholar] [CrossRef]
  121. Brandão, L.; Reece, C. Non-Steady State Validation of Kinetic Models for Ethylene Epoxidation over Silver Catalysts. Catal. Sci. Technol. 2024, 14, 3596–3608. [Google Scholar] [CrossRef]
Figure 1. Timeline of the yearly number of publications focused on the subject of ethylene epoxidation (data obtained from Scopus, keyword search “ethylene”, “oxidation”).
Figure 1. Timeline of the yearly number of publications focused on the subject of ethylene epoxidation (data obtained from Scopus, keyword search “ethylene”, “oxidation”).
Catalysts 15 00560 g001
Figure 2. Ethylene itself can undergo direct combustion to further contribute to the formation of CO2 and water.
Figure 2. Ethylene itself can undergo direct combustion to further contribute to the formation of CO2 and water.
Catalysts 15 00560 g002
Figure 3. (a,b) Display the top and side views of the optimized Ag/h-BN nanosheet; (c) shows the partial density of states (PDOS) for the Ag atom and its neighboring N atoms; and (d) presents the structural configuration along with the total energy variation of Ag/h-BN over time, obtained from molecular dynamics simulations (2 ps at 500 K). The dotted line in the PDOS plot represents the Fermi level (EF) [25].
Figure 3. (a,b) Display the top and side views of the optimized Ag/h-BN nanosheet; (c) shows the partial density of states (PDOS) for the Ag atom and its neighboring N atoms; and (d) presents the structural configuration along with the total energy variation of Ag/h-BN over time, obtained from molecular dynamics simulations (2 ps at 500 K). The dotted line in the PDOS plot represents the Fermi level (EF) [25].
Catalysts 15 00560 g003
Figure 4. Suggested mechanism from previous studies illustrating ethylene epoxidation on silver catalysts through the oxometallacycle (OMC) intermediate, where ethylene oxide (EO) is the target product and carbon dioxide (CO2) and water (H2O) are the by-products of combustion [32].
Figure 4. Suggested mechanism from previous studies illustrating ethylene epoxidation on silver catalysts through the oxometallacycle (OMC) intermediate, where ethylene oxide (EO) is the target product and carbon dioxide (CO2) and water (H2O) are the by-products of combustion [32].
Catalysts 15 00560 g004
Figure 5. Particle growth of Ag–α-alumina via Ostwald ripening during ethylene epoxidation [43].
Figure 5. Particle growth of Ag–α-alumina via Ostwald ripening during ethylene epoxidation [43].
Catalysts 15 00560 g005
Figure 6. Schematic of chemical looping epoxidation using SrFeO3—δ/Sr3Fe2O7—γ oxygen carrier [46].
Figure 6. Schematic of chemical looping epoxidation using SrFeO3—δ/Sr3Fe2O7—γ oxygen carrier [46].
Catalysts 15 00560 g006
Figure 7. SEM–EDS images showing the distribution of silver and cerium in (CeO2)impr(SrFeO3) and (CeO2)ss(SrFeO3). Sr, Fe, and O were evenly distributed throughout both samples [48].
Figure 7. SEM–EDS images showing the distribution of silver and cerium in (CeO2)impr(SrFeO3) and (CeO2)ss(SrFeO3). Sr, Fe, and O were evenly distributed throughout both samples [48].
Catalysts 15 00560 g007
Figure 8. The unit cell of Mn–N4GP catalyst in top view [22].
Figure 8. The unit cell of Mn–N4GP catalyst in top view [22].
Catalysts 15 00560 g008
Figure 9. Behavior of unpromoted silver foil catalyst during ethylene oxidation at 0.3 mbar and 230 °C with a C2H4/O2 ratio of 1:2: (a) time-dependent mole fractions of CO2 (red) and EO (blue), along with EO selectivity (green); (b) sample O1s XPS spectra recorded at various time points; (c) temporal variation in the ratio of electrophilic to nucleophilic oxygen species (Oelec/Onucl) [9].
Figure 9. Behavior of unpromoted silver foil catalyst during ethylene oxidation at 0.3 mbar and 230 °C with a C2H4/O2 ratio of 1:2: (a) time-dependent mole fractions of CO2 (red) and EO (blue), along with EO selectivity (green); (b) sample O1s XPS spectra recorded at various time points; (c) temporal variation in the ratio of electrophilic to nucleophilic oxygen species (Oelec/Onucl) [9].
Catalysts 15 00560 g009
Figure 10. Schematic illustration of the epoxidation of ethylene over Ir1–α-MnO2 [65]. The red star indicates the performance of the catalyst presented in this study.
Figure 10. Schematic illustration of the epoxidation of ethylene over Ir1–α-MnO2 [65]. The red star indicates the performance of the catalyst presented in this study.
Catalysts 15 00560 g010
Figure 11. Illustrations of preferred adsorption configurations and reaction pathways for the electrocatalytic epoxidation of ethylene to EO on (a) RuO2 and (b) RuO2 with adsorbed Cl [45].
Figure 11. Illustrations of preferred adsorption configurations and reaction pathways for the electrocatalytic epoxidation of ethylene to EO on (a) RuO2 and (b) RuO2 with adsorbed Cl [45].
Catalysts 15 00560 g011
Figure 12. Langmuir–Hinshelwood and Eley–Rideal mechanisms of ethylene oxidation on Ag-based catalysts [64].
Figure 12. Langmuir–Hinshelwood and Eley–Rideal mechanisms of ethylene oxidation on Ag-based catalysts [64].
Catalysts 15 00560 g012
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

Saritala, M.A.; Muzammil, M.; Quddus, M.R.; Razzak, S.A.; Hossain, M.M. A Review on Production of Ethylene Oxide from Epoxidation of Ethylene: Catalysis, Mechanism and Kinetics. Catalysts 2025, 15, 560. https://doi.org/10.3390/catal15060560

AMA Style

Saritala MA, Muzammil M, Quddus MR, Razzak SA, Hossain MM. A Review on Production of Ethylene Oxide from Epoxidation of Ethylene: Catalysis, Mechanism and Kinetics. Catalysts. 2025; 15(6):560. https://doi.org/10.3390/catal15060560

Chicago/Turabian Style

Saritala, Mahammad Ali, Mohammed Muzammil, Mohammad R. Quddus, Shaikh Abdur Razzak, and Mohammad M. Hossain. 2025. "A Review on Production of Ethylene Oxide from Epoxidation of Ethylene: Catalysis, Mechanism and Kinetics" Catalysts 15, no. 6: 560. https://doi.org/10.3390/catal15060560

APA Style

Saritala, M. A., Muzammil, M., Quddus, M. R., Razzak, S. A., & Hossain, M. M. (2025). A Review on Production of Ethylene Oxide from Epoxidation of Ethylene: Catalysis, Mechanism and Kinetics. Catalysts, 15(6), 560. https://doi.org/10.3390/catal15060560

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop