Home / Publications / Acetylene semi-hydrogenation: recent advances

Acetylene semi-hydrogenation: recent advances

Share
Cite this
GOST
 | 
Cite this
GOST Copy
Gyrdymova Y. V., Rodygin K. S. [{"id":"WAyKaSjYOt","type":"paragraph","data":{"text":"Acetylene semi-hydrogenation: recent advances"}}] // Russian Chemical Reviews. 2025. Vol. 94. No. 8. RCR5182
GOST all authors (up to 50) Copy
Gyrdymova Y. V., Rodygin K. S. [{"id":"WAyKaSjYOt","type":"paragraph","data":{"text":"Acetylene semi-hydrogenation: recent advances"}}] // Russian Chemical Reviews. 2025. Vol. 94. No. 8. RCR5182
RIS
 | 
Cite this
RIS Copy
TY - JOUR
DO - 10.59761/RCR5182
UR - https://rcr.colab.ws/publications/10.59761/RCR5182
TI - [{"id":"WAyKaSjYOt","type":"paragraph","data":{"text":"Acetylene semi-hydrogenation: recent advances"}}]
T2 - Russian Chemical Reviews
AU - Gyrdymova, Yulia V.
AU - Rodygin, Konstantin S.
PY - 2025
DA - 2025/08/25
PB - ANO Editorial Board of the journal Uspekhi Khimii
SP - RCR5182
IS - 8
VL - 94
ER -
BibTex
 | 
Cite this
BibTex (up to 50 authors) Copy
@article{2025_Gyrdymova,
author = {Yulia V. Gyrdymova and Konstantin S. Rodygin},
title = {[{"id":"WAyKaSjYOt","type":"paragraph","data":{"text":"Acetylene semi-hydrogenation: recent advances"}}]},
journal = {Russian Chemical Reviews},
year = {2025},
volume = {94},
publisher = {ANO Editorial Board of the journal Uspekhi Khimii},
month = {Aug},
url = {https://rcr.colab.ws/publications/10.59761/RCR5182},
number = {8},
doi = {10.59761/RCR5182}
}
MLA
Cite this
MLA Copy
Gyrdymova, Yulia V., and Konstantin S. Rodygin. “[{"id":"WAyKaSjYOt","type":"paragraph","data":{"text":"Acetylene semi-hydrogenation: recent advances"}}].” Russian Chemical Reviews, vol. 94, no. 8, Aug. 2025, p. RCR5182. https://rcr.colab.ws/publications/10.59761/RCR5182.
Publication views
21

Keywords

Acetylene
catalyst cocktail
selective hydrogenation
single-atom catalysts

Abstract

The development of new catalysts with improved properties for organic reactions is relevant both for solving fundamental chemical challenges and for their implementation into production. In industry, acetylene hydrogenation is a catalytic process. The role of a catalyst is to increase the selectivity of hydrogenation to ethylene and prevent overhydrogenation to ethane. Research in the field of selective hydrogenation of acetylene, in addition to the applied relevance, is also of academic interest, since it discovers the nature of catalysis in details. Amazingly, acetylene hydrogenation has become, in a sense, a criterion for assessing the properties of synthesized catalysts, demonstrating the "structure–properties" relationship. Changes in the structure of a catalyst immediately result in changes in its catalytic characteristics, which, in turn, led to a more complete understanding of the processes occurring on the catalyst surface. A sharp increase in the number of researches on this subject that occurred in the period from 2021 to 2024 was due to the possibilities of obtaining and describing single-atom catalysts with the identification of the key dependence "particle size–activity". This review not only complements previous summarizing works on selective hydrogenation of acetylene, but also differs significantly in the arrangement of the material. Various approaches to the design of catalysts for this reaction are discussed based on the insight into its pathway. Mono- and bimetallic catalysts based on active metals such as palladium, nickel, copper, etc. are considered. Particular attention is paid to the application of single-atom catalysts. At the conclusion of this review, we summarized the catalytic characteristics of various systems in a table.

The bibliography includes 267 references.

1. Introduction

Catalytic processes have been the subject of research throughout their creation, both from the academic and industrial perspective. The first industrial catalytic processes of nitrogen fixation to ammonia and the development of catalysts for ethylene polymerization marked a new era in the development of chemistry. Many important chemical transformations are catalytic today, e.g., cross-coupling,[1-4] hydrogenation,[5-9] alkylation, and many others.[10-16] The use of transition metal compounds were for these purposes a common case. The variable oxidation state and d-orbital in the atomic structure of these metals provide many opportunities for tuning the electronic structure of the metal centre and increase the number of possible compounds that it can form.

Pd, Pt, Ru, Ni, and Cu-based catalysts are extremely active even in small amounts. Typically, catalytic systems including noble metals are characterized by low selectivity and promote the reaction in all possible directions. At the same time, they are prone to poisoning by impurities that can block their active centres. Yet another issue is sintering (coking) due to high temperatures. All these factors adversely affect the selectivity and performance of catalysts, reducing their durability. The said drawbacks stimulate research to improve the properties of available catalysts and develop new more efficient catalytic systems.

The study of the properties of catalysts is usually carried out on model reactions. Ideally, the most appropriate reactions should proceed completely and selectively, and the product should be clearly identified in the mixture. Hydrogenation is one of the key catalytic reactions in chemical synthesis,[17-27] that meets the listed requirements. In particular, acetylene hydrogenation can be considered as a convenient model for studying the properties of metal catalysts.

Why acetylene hydrogenation? Acetylene is the simplest alkyne consisting of triple-bonded two carbon atoms. Unsaturated character of acetylene is widely used in the synthesis of vinyl monomers, heterocycles, and a wide range of valuable chemicals.[28-33] Hydrogenation of acetylene can proceed with the formation of ethylene or ethane. Semi-hydrogenation of acetylene is a critical process providing selective conversion of acetylene to ethylene, while preventing further hydrogenation to ethane. Ethylene is one of the key raw materials of the chemical industry, representing the main supplier of the C2 monomer unit. In 2023, global polyethylene production amounted to 141 million tons.[34] Ethylene obtained by high-temperature cracking contains up to 2% of acetylene as an impurity.[35-39] Since acetylene is a catalytic poison in further ethylene polymerization, ethylene must be carefully purified from acetylene, and the residual amount of acetylene should not exceed 5 ppm.[40] Separating ethylene and acetylene is a complicated, labour-intensive task, that requires special approaches.[41-44] Currently, the removal of acetylene from ethylene prior to polymerization is usually realized via selective hydrogenation of acetylene to ethylene.

Acetylene hydrogenation is a tandem reaction including two competing parallel reaction pathways:* sequential hydrogenation of acetylene to produce ethylene (reaction (1)) and its further hydrogenation to ethane (reactions (2) and (3)).[45]

Prolonged hydrogenation of acetylene in the reducing medium in presence of a catalyst results in the formation of oligomerization products of a linear structure with an even number of carbon atoms (C4+, reaction (4)). The liquid fraction of the hydrocarbons is called ‘green oil’.[46] The oil is formed on the surface of catalysts only in the combined presence of acetylene and hydrogen,[47][48] and blocks active centres, deactivating a catalyst.[49][50] Most traditionally employed noble metal-based hydrogenation catalysts [51-60] promote all possible reactions, including the formation of the ‘green oil’. Therefore, the best or ideal catalyst, in addition to high selectivity to acetylene semi-hydrogenation and anti-coking properties, should inhibit the formation of oligomerization products and be resistant to their effects (if green oil is formed). The research of many scientific groups is focused on solving precisely these problems.

According to the SciFinder database, the next trend in publication activity in the field of selective catalysts development and their application in hydrogenation of acetylene occurred over the period 2021 – 2024 (Fig. 1). This increase in the number of works was largely determined by the wide access to the synthesis and study of single-atom catalysts, which significantly changed the ideas about the nature of catalysis per se in general.

Fig. 1
Analysis of publication activity on selective hydrogenation of acetylene in the period 1990 – 2025, carried out using the CAS SciFinder search system

The previous reviews have contributed greatly to the summarization of methods for the selective hydrogenation of acetylene,[61] the systematization of approaches to bimetallic[62][63] and single-atom catalysts,[64] and also revealed the role of solid solutions of insertion in the palladium-catalyzed hydrogenation of acetylene to ethylene.[65][66]

In this review, we summarized information on the achievements in the field of metal-catalyzed selective hydrogenation of acetylene over the last five years. The available data on the nature of the element used as a catalytically active component of the reaction have been systematized. The work is not limited to palladium and its compounds, but includes studies on the activity of nickel, copper, and bimetallic systems. The authors do not limit themselves to considering only heterophase processes, but also jointly consider the hydrogenation of an ethylene mixture with an admixture of acetylene and pure acetylene into ethylene.

* After the reaction equation, the values of their standard enthalpies (DH°) are given.

2. Study of the mechanisms of catalytic processes and development of new approaches to design of hydrogenation catalysts

An effective catalytic system involves obtaining the desired compounds by the chemical reaction in high yields in the absence of by-products. The challenge of catalytic transformations in chemistry is significantly complicated by additional needs, viz., the possibility of catalyst regeneration,[67] the use of minimal catalyst loadings, the absence of stabilizers and expensive ligands,[68] the use of cheap metals.[69] Homogeneous and heterogeneous catalysts are successfully used to solve many problems in organic synthesis.[70-75] Homogeneous catalysts are effective and selective, mainly due to the fact that they generate certain rather stable molecular species in the reaction medium. However, it is usually not possible to reuse such catalysts. Heterogeneous catalysts may be used several times. However, with an increase in the number of cycles, their activity and selectivity may significantly decrease, and the detection of metal-containing species in the solution indicates the transition of part of the catalyst into the solution. This calls into question the heterogeneous nature of the processes and also leads to a decrease in the amount of catalyst and the impossibility of its reuse.

In recent years, there has been a significant accumulation of new knowledge and an evolution of our understanding of catalysis and catalysts. An important conceptual achievement is the development of ideas that expand our understanding of homogeneous and heterogeneous systems and the discovery of dynamic catalyst cocktail systems.[76][77] Catalyst cocktail systems were first described by Ananikov and co-workers.[78] In this Section, mechanistic studies of the dynamic nature of catalysis using various reactions as examples are described, which is necessary for forming ideas about the methods for studying dynamic phenomena in catalysis. At the end of this Section, the effects of ‘cocktail’ systems on the mechanisms of hydro­genation reactions are considered. In the following Sections, specific examples of hydrogenation reactions are given, many of which highlight various aspects of the dynamic behaviour of catalytic systems.

The type (or nature) of catalysis is largely determined by the particle size.[79][80] Nanoparticles and their assemblies [81][82] have fundamentally different properties compared to microparticles. Further reduction in particle size leads to the formation of clusters consisting of several atoms, as well as individual atoms.[83-86] An even greater reduction in the concentration of active sites on the surface of a heterogeneous catalyst is achieved by isolating the active metal with another metal, which can lead to the formation of alloys or intermetallides. All of these particles exhibit different behaviour patterns, and the variability of the processes is astounding. Catalytically active species can stick together, pass into solution as complexes and precipitate back as particles, migrate along the surface of the substrate, be immersed in the substrate, and become coated with a layer of other compounds. All this significantly changes the catalytic activity and selectivity [87][88] and makes the nature of catalysis dynamic.[89][90]

Thus, it was shown that in a model cross-coupling reaction [91][92] the used Pd/NHC (NHC is N-heterocyclic carbene) complex catalysts are transformed in the course of the reaction into metal nanoparticles and clusters due to the rupture of Pd-NHC coordination bonds and the transformation of the ligand, as evidenced by electrospray ionization high resolution mass-spectrometry (ESI-HRMS).[93] It is important to note that the catalyst transformations did not reduce the catalytic activity of the system. This was probably due to the formation of new catalytically active species via the leaching of Pd particles from the surface of the metal catalyst. The reversibility of the transition of species into solution (leaching) and back was calculated using density functional theory (DFT) modelling methods on model reactions of C – H-arylation, Heck, and amination.[94] As was found out, Pd leached into solution from the catalyst was the progenitor of a ‘cocktail’ of compounds that were classified into three main groups (Fig. 2):

1) catalytically active Pd nanoparticles with a high concentration of surface defects,

2) monomeric and oligomeric L[ArPdX]nL (L is a ligand) particles,

3) irreversibly deactivated Pd species.

Fig. 2
Possible processes of leaching, trapping and oligomerization in Pd catalytic systems. Pd nanoparticles are shown in blue, Pd species that have passed into solution are shown in green. Reproduced from Ref. 94 under a CC-BY-NC-ND license.

The study of complex systems requires a special approach. The traditionally used mercury test, despite its rapidity and versatility for identifying homogeneous and heterogeneous catalysis, demonstrated low analytical accuracy and low reproducibility for Pd and Pt complexes.[95] A low-temperature solid-state NMR spectroscopy technique was proposed to monitor hybrid catalytically active species and study the mechanisms of their reactions.[96] It was shown that for certain nanoparticle sizes, only six NHC ligands are sufficient to cover 80% of the surface of Pd particles measuring ~ 1 nm. Later, the concept of 4D catalysis was proposed, combining the achievements of instrumental analysis methods and the potential of machine learning for studying the mechanisms of dynamic catalysis.[97]

Leaching of a metal catalyst is uneven and depends on the type of facet of the metal catalyst surface. In bimetallic nanocatalysts, predominantly one metal is etched. Surface particles can migrate to the surface, stick together or separate from the surface and settle on it again. The presence of oxygen in the system can contribute to the formation of defects in the crystal lattice of metal particles and clusters, which sometimes plays a decisive role in some catalytic processes. All possible processes are characteristic of the formed catalyst particles: growth and dissolution of particles, movement along the surface and changes in the internal structure at the level of individual atoms, clusters and nanoparticles (NPs). Atoms are also detached from the edges and corners of NPs. Atoms can precipitate from the solution on the catalyst surface both on free surface areas and on NPs. In the first case, atoms can re-enter the catalytic cycle, in the second, they are poisoned by nanoparticles, changing the morphology of the particles and causing their growth.

Not only palladium but also nickel form catalytic systems of the ‘cocktail’ type. In a study,[98] a comparative characteristics and properties of two widely used nickel and palladium catalysts are given. Due to several oxidation states, nickel provides higher activity and larger variety of chemical transformations than palladium. This becomes possible despite the higher activation barrier of nickel catalysts than palladium ones. The authors note that Pd catalytic systems can easily form a ‘cocktail of species’ from various palladium compounds, but with similar oxidation states (Pd1, Pdn, Pd NPs), while nickel can behave as a ‘cocktail of species’ in different oxidation states with a smaller number of compounds.

Dynamic transformations of the catalyst in chemical processes significantly complicate the study of reaction mechanisms and catalytic cycles. Nevertheless, these data are necessary for understanding not only the course of reactions, but also the essence and nature of catalysis per se. Hydrogenation of alkynes can be considered as one of the representative examples for studying the nature of catalysis.[99][100]

Metal-catalyzed hydrogenation with hydrogen is carried out in a homogeneous or heterogeneous mode using catalysts based on palladium, nickel and some other transition metals. According to the Horiuti – Polanyi principle (dissociative mechanism), the heterogeneous hydrogenation pathway involves the sorption of the substrate and hydrogen on the catalyst surface. In this case, two variants of the reaction are possible: hydrogenation on the surface of metal particles occurs selectively to an alkene. Sorption of reagents in the entire volume of the catalyst promotes the hydrogenation of an alkyne to an alkane. Adsorption and activation of hydrogen are important steps in the hydrogenation process.

On metal catalysts, the bond in the hydrogen molecule can undergo both homolytic and heterolytic cleavage. This is because transition metals have unoccupied d-orbitals to accommodate hydrogen electrons. At the same time, the heterolytic cleavage of the H – H bond creates a vacant s-orbital for the transfer of electrons from the metal.

In the Langmuir – Hinshelwood mechanism, which remains the most common for heterogeneous catalytic systems, two pathways can be distinguished (Fig. 3). The Horiuti – Polanyi mechanism, or the so-called dissociative mechanism, proposed in 1934, involves the homolytic dissociation of hydrogen on the metal surface, followed by the sequential addition of H atoms to the adsorbed alkynes.[100] Hydrogenation of an alkyne with molecular hydrogen begins with the sorption of the alkyne and hydrogen on the catalyst by bonding the Pd atom with the electrons of the π-bond and forming the [alkyne – Pd] complex (Fig. 3, route AB).[101] The heat of adsorption of ethyne on palladium through the coordination of both π-bonds of the alkyne molecule with the metal is 112 kJ mol–1. The next step (route BC) comprises the migration and addition of the adsorbed H2 molecule to the [alkyne – Pd] complex. In this case, a σ-bond is formed between the adjacent carbon and Pd, yielding an intermediate (Z)-vinylalkyl compound C or σ-vinyl[Pd]. Upon addition of an additional H atom, σ-vinyl[Pd] is converted to π-alkene[Pd]. This process is thermodynamically favourable (ΔH = −136 kJ mol–1).

Fig. 3
Mechanisms of hydrogenation of alkynes and alkenes. Red balls are H atoms in molecular hydrogen, green balls are H or D atoms attached to the unsaturated substrate. Reproduced from Ref. 101 with permission from the American Chemical Society.

Further, the hydrogenation process can proceed in two directions. The first consists of the decomposition of the π-alkene[Pd] complex to give free alkene D* (semi-hydrogenation). The second direction is further hydrogenation of the resulting alkene. π-Alkene[Pd] can add adsorbed hydrogen to form a σ-bonded alkyl complex E. This process is energetically less favourable than the formation of a free alkene (route D – E, ΔH = −35 kJ mol–1). At the final stage E – F, the σ-bond Pd – C is broken with the complete hydrogenation product F.

The mechanism of alkyne hydrogenation over alloy catalysts is somewhat more complex, since it is necessary to study and take into account the effect of each metal in the alloy separately. The mechanism of acetylene hydrogenation on the Pd – Mn/Al2O3 alloy catalyst is similar to that on the monometallic Pd catalyst.[102] The authors noted that the initial stage of formation of the π-complex of acetylene with the active centre of the catalyst is fast. The subsequent stages of hydrogen addition there to give ethylene and the formation of the [alkene-metal] complex occur much more slowly. In a study,[103] the mechanism of acetylene hydrogenation over the Pd – Ag/α-Al2O3 alloy catalyst was studied using the theory of quantum mechanics and the Monte Carlo method. The doping of the Pd catalyst with Ag reduced the adsorption energies of C2H2 and C2H4 on the catalyst surface (Fig. 4) and facilitated the desorption of ethylene from the catalyst surface, thereby improving the selectivity of hydrogenation. It is noted that the adsorption energy of acetylene and ethylene decreased with a decrease in the number of Pd atoms or an increase in the number of Ag atoms. Therefore, the activity of the catalysts was proportional to the number of Pd atoms and inversely proportional to the number of Ag atoms, which was consistent with the experimental results.

Fig. 4
Effect of Ag doping on the activation energy of acetylene hydrogenation. Reproduced from Ref. 103 with permission from Taylor & Francis.

During the hydrogenation, the dispersed Pd – Ag/α-Al2O3 catalytic particles undergo agglomeration, and their average size increases almost 2-fold.[104] Analysis of images obtained by scanning electron microscopy (SEM) (Fig. 5) showed that in the course of hydrogenation, polymeric compounds and coke accumulate on the catalyst surface, which in turn leads to its deactivation.

Fig. 5
SEM images of fresh (a) and spent (b) catalysts. Darker areas correspond to polymer formations on the catalyst surface. Reproduced from Ref. 104 under a Creative Commons Attribution (CC BY) license.

Homogeneous hydrogenation can involve activation of the alkyne within the metal complex. Pd/NHC complexes are widely used as catalysts in homogeneous hydrogenation reactions. Denisova et al.[105] described the experience of successful use of Pd/NHC complexes in homogeneous-heterogeneous (transfer) catalytic hydrogenation. It was found that using a 13C-labeled NHC ligand, the Pd/NHC catalysts are converted to Pd NPs (1 – 9 nm) in the first minutes of the diphenylacetylene transfer hydrogenation, indicating a mixed type of catalysis. The authors proposed a mechanism of mixed hydrogenation, in which oxidative addition occurs on the surface of Pd particles, and the stages of syn-addition of alkynes and reductive elimination occur in the zone of Pd nanoparticles modified with NHC ligands. Using DFT calculations on the Pd3 catalyst model, a mechanism of three-stage hydrogenation was proposed that includes the reaction of oxidative addition of molecular hydrogen to the Pd cluster, syn-addition of H2 to the triple CºC bond of acetylene and reductive elimination to give an ethylene molecule. The presence of a ligand on the metal surface complicates the addition of hydrogen.

Similar studies were carried out for the Pd/NHC-catalyzed semi-hydrogenation of 1,2-diphenylacetylene.[106] Using modern NMR, high resolution mass spectrometry with electrospray ionization (ESI-HRMS) and SEM/TEM methods, it was shown that during the hydrogenation process, the metal complexes are transformed into Pd NPs with the formation of a heterogeneous phase. Modification of the surface of these NPs with organic ligands under the reaction conditions significantly affected their catalytic activity.

It is worth noting that studies in the field of homogeneous catalysis confidently demonstrate the presence of dynamic processes. The composition of particles is constantly changing, which means that their properties and the very catalytic activity change. Under conditions of heterogeneous catalysis, the number of possible processes seems more limited. However, even in the solvent-free heterogeneous systems, it has been shown that metal atoms can migrate along the substrate surface and undergo various changes, e.g., combine into larger clusters. This suggests that the heterogeneous process is also an evolutionary system. This raises a question that bridge across the present Section with the following ones: is it possible to discuss the dynamic nature of catalysis in relation to heterogeneous systems? In the authors opinion, the reaction of acetylene hydrogenation is a very convenient tool for proving this fact.

3. Monometallic Pd as catalyst for acetylene semi-hydrogenation

Palladium is a heterogeneous effective hydrogenation catalyst for a wide range of compounds.[107-115] The high activity of Pd catalysts is both an advantage and disadvantage. High-pressure hydrogenation of acetylene in the presence of Pd compounds produces mainly ethane rather than ethylene. Suppression of excess Pd activity to increase selectivity of acetylene semi-hydrogenation reactions is one of the main areas of catalysis. Common tools for varying the properties of a catalyst include reducing the size of catalytic species, the use of bimetallic catalysts, and varying the support (carrier) for the metal particles. Below are the main advances in the field of increasing the selectivity of Pd hydrogenation catalysts.

Single-atom (SA) catalysts have attracted particular attention due to their high reactivity and excellent catalytic efficiency,[116-120] which the small particle size and the resulting weak adsorption of ethylene on its surface. selectivity of acetylene semi-hydrogenation was using the DFT method. The dependence of the increase in selectivity of acetylene semi-hydrogenation on the decrease in the catalyst particle size was calculated using DFT method.[121] Pd/C catalyst obtained by impregnation with subsequent hydrogenation in a hydrogen stream at 500°C was used as a model catalyst. S-containing silane agents served as nanoparticle size regulators. Metal NPs were firmly fixed on the support due to the formation of Pd – S bonds with thiol groups, and were uniformly distributed over the surface. Interestingly, the Pd particle size of 2 nm was crucial, since particles of this size contributed to the formation of C4 hydrocarbons and ‘green oil’.

For example, 26 nm Pd NPs with a close-packed (Pd(111)) surface demonstrated 16 times greater ethylene selectivity than commercially available Pd/C.[122] At 100% acetylene conversion, the selectivity to acetylene was 76.5% and 4.5% for Pd(111) and Pd/C, respectively (Fig. 6, TEM is transmission electron microscopy). The authors suggested that the increase in the selectivity of nanostructured Pd stemmed from lower hydrogen adsorption on the metal catalyst because of the formation of palladium hydride sites Pd4H3 in the hydrogen stream.

Fig. 6
TEM images of a commercially available Pd/C catalyst (a) and octahedral Pd/C (b). Acetylene conversion and selectivity to ethylene in acetylene hydrogenation over commercial Pd/C (c) and octahedral Pd/C (d) catalysts at 293 K, p(C2H2) = 0.02 atm, gas hourly space velocity (GHSV) is 120 000 h−1. The insets in the upper corners show enlarged high resolution TEM images. Reproduced Ref. 122 with permission from the American Chemical Society.

In fact, the selective hydrogenation of acetylene requires the reaction to occur on the surface of the catalyst, not in its bulk. The same idea on the need to separate the surface and bulk reactions of acetylene hydrogenation to increase selectivity to ethylene was highlighted in the work[123] (Fig. 7). According to in situ X-ray photoelectron spectroscopic analysis of solid Pd catalysts, non-selective hydrogenation occurred at a low carbon content in the catalyst, and the process itself occurred on hydrogen-saturated β-hydride in the catalyst bulk.

Fig. 7
Schematic view of the processes of non-selective and selective hydrogenation of acetylene. Grey balls represent the metal catalyst, red balls represent hydrogen atoms in molecular H2 as, blue balls represent hydrogen atoms in acetylene, and black balls represent carbon atoms.

The high selectivity to ethylene in the semi-hydrogenation of acetylene at high temperatures on the Pd(111) catalyst was in good agreement with the calculated value (82% at 300 K), obtained by ab initio molecular dynamics (AIMD) methods. The calculated values of the free energy barriers for ethylene desorption on Pd(111) were 0.51 eV at 300 K and 0.39 eV at 500 K.[124] The surface of the close-packed monoatomic Pd(211) catalyst was more active than that of Pd(111). At the same time, differential chemisorption energy of Pd(211) catalyst was higher with a lower barrier to further hydrogenation of ethylene (Table 1).[125] DFT calculations and coverage-dependent microkinetic simulations showed that ethylene was produced on the close-packed regions of the catalyst, while the defective surface regions were significantly more active in selective ethane formation.

Table 1
\[ \]
Adsorption energies (Eads, eV) and free adsorption energies (Gads, eV) of acetylene and ethylene on palladium surfaces.125
(1)

According to DFT calculations, hydrogenation of ethylene to ethane on the SA – Pd catalyst required overcoming two barriers of 0.93 and 1.22 eV.[126] The chemical adsorption energy of ethylene is 0.76 eV, while the hydrogenation barriers are only 0.62 and 0.35 eV, located below the ethylene level.

Template method was used for the Pd catalyst design. Isolated Pd atoms (Pd-ISA) were anchored on the inner walls of mesoporous nitrogen-doped carbon foam nanospheres (MPNC).[127] Such catalyst provided complete conversion of acetylene hydrogenation at only 60°C! However, the selectivity of the catalyst was lower than 20%. Increasing the temperature improved the hydrogenation selectivity. The ISA-Pd/MPNC and ISA-Pd/non-MPNC (non-MPNC is non-mesoporous analogue of MPNC) catalysts were used at 120 and 140°C, respectively, and were more selective at lower loadings (200 vs. 300 mg): acetylene conversion was 83% and 32%, respectively, and ethylene selectivity was 82% and 17%, respectively.

Atomically dispersed Pd1/C3N4 catalyst and Pd NP catalyst supported on graphite carbon nitride (gC3N4), Al2O3 and SiO2 were synthesized by wet-impregnation of the support with a solution of Pd(acac)2 (acac is acetyl acetonate) followed by calcination in a flow of 10% O2/He and H2.[128] Surprisingly, when hydrogenating acetylene in excess ethylene, the Pd NPs/gC3N4 catalysts demonstrated higher selectivity to ethylene than the common Pd/Al2O3 and Pd/SiO2 catalysts. The high catalytic performance was, probably, due to the significant charge transfer from Pd NPs to the gC3N4 support. A single-atom Pd1/C3N4 catalyst demonstrated superior catalytic performance compared to Pd NPs/gC3N4 catalysts with higher selectivity to ethylene and higher coking stability in both reducing and oxidizing media at high temperatures. Atomically dispersed palladium species Pd-ISAS/CN immobilized on N-doped carbon nanospheres obtained from covalent organic frameworks (COFs) (demonstrated 80% selectivity to ethylene and 92% acetylene conversion at 100°C, with the catalyst activity of 712 molC2H2 (molPd h)–1.[129]

Ultrasmall Pd NPs prepared using H2 instead of NaBH4 to reduce NaBH4 and immobilized in the micropores of layered covalent triazine frameworks (CTFs) demonstrated 100% acetylene conversion with 99.9% ethylene selectivity, as well as excellent stability at 250°C and a space velocity of 110 000 h–1.[130] The nanocatalyst demonstrated high recycling potential with full conversion and selectivity maintained for 5 cycles. The authors explained the improved selectivity rates by the layering CTFs into thinner and smaller sheets during the reaction, which provides greater access of reagents to the catalytically active Pd sites and increased mass transfer.

Varying the support for immobilization of metals is a convenient tool for controlling the catalyst properties. MgO-supported metal Pd particles (Pd/MgO) showed high selectivity (82%) in the hydrogenation of acetylene to ethylene at 200°C with ultra-low catalyst loading (0.05 wt.%).[131] Such catalyst with a nominal Pd loading of 0.001% demonstrated excellent performance, providing high acetylene conversion (97%) and good selectivity to ethylene (89%).[132] Amazingly, but under the same conditions, the ethylene selectivity of 1% Pd/MgO catalyst was lower than 4%. Kinetic isotope effect and IR spectroscopy data suggested heterolytic H2 dissociation on isolated Pd atoms to form the O – H bond. This probably determines the high selectivity in the selective hydrogenation of acetylene. Moreover, dispersed Pd atoms were inactive for ethylene hydrogenation, which prevented the formation of ethane.

Single-atom Pd1/MgO acetylene semi-hydrogenation catalyst was obtained mechanochemically in a ball mill by grinding palladium phthalocyanine and magnesium phthalocyanine, followed by calcination at 600°C.[133] The nature of single-atom Pd species after hydrogenation at 100°C was confirmed by aberration-corrected high-angle annular dark-field scanning transmission electron microscopy (AC-HAADF-STEM) (Fig. 8). The strong interaction between Pd atoms and MgO support not only inhibited the aggregation of Pd atoms, but also resulted in high ethylene desorption ability of Pd, thereby providing high selectivity of acetylene semi-hydrogenation: 100% C2H2 conversion with 60% selectivity to acetylene at 140°C and TOF = 0.074 s–1.

Fig. 8
AC-HAADF-STEM images of the surface of Pd1/MgO (a) and (b) Pd1/MgO-H100, which is parent Pd1/MgO particles reduced at 100°C with 10 vol.% H2 (b). A part of isolated Pd atoms are designated as yellow circles. Reproduced from Ref. 133 with permission from Wiley.

Pd1/CeO2 and Pd1/α-Fe2O3 catalysts were prepared by reducing conventional supported Pd/CeO2 and Pd/α-Fe2O3 obtained by impregnating the supports with Pd(NO3)2 solution, followed by reduction with 10 vol.% H2/He at 200, 500, and 600°C and oxidation with 20 vol.% O2/He at 500°C to selectively encapsulate Pd nanoparticle clusters.[134] Single atoms exhibited strong metal – support interactions only at higher reduction temperatures than for nanoparticles/clusters, thus contributed to increased selectivity and stability in acetylene semi-hydrogenation. Pd/CeO2 catalyst prepared by impregnation with the initial humidity and calcined at 600°C and 800°C demonstrated excellent ethylene selectivity (~ 100%) in acetylene hydrogenation in large excess of ethylene at 50 – 200°C.[135] This was probably due to the formation of PdxCe1 – xO2 – y or Pd – O – Ce phases on the catalyst surface, as confirmed by the results of X-ray diffraction (XRD) and X-ray photoelectron spectroscopy and indicated a strong interaction of Pd and Ce.

SA catalyst coordinated by 1,10-phenanthroline-5,6-dione (PDO) was successfully used for the semi-hydrogenation of acetylene under mild conditions.[136] The hydrogenation selectivity was highly dependent on the metal : ligand molar ratio. Surprisingly, the increase in ligand loading promoted the hydrogenation of acetylene to ethylene. In the absence of an organic ligand, 100% conversion of acetylene to ethane was observed. At a PDO : Pd ratio of 1 : 1, the ethylene selectivity was 30%. The best selectivity (80%) was obtained at a ligand : Pd ratio of 6 : 1, which remained constant after a further increase in the ligand amount (Fig. 9). The authors suggested, that the metal atoms are maximally coordinated and are unable to form bonds with other ligand molecules at a ratio of 6 : 1. According to calculations, the binding of hydrogen to the ligand switches the reaction towards ethylene formation. The measured activation energies for the PDO – Pd(1 : 1)/CeO2 and PDO – Pd(6 : 1)/CeO2 catalysts were 36 kJ mol–1 in both cases. It was later shown that the organic ligand stabilized the single metal atoms through the strong Pd – O or Pd – N interaction.[137] Ethylene molecules were stronger adsorbed on the Pd1 – PDO/CeO2(111) catalyst than C2H2 molecules, which resulted in higher selectivity in the semi-hydrogenation.

Fig. 9
Conversion (a) and ethylene selectivity (b) of Pd – PDO catalysts as a function of the PDO : Pd ratio in acetylene hydrogenation, T = 120°C. Reproduced from Ref. 136 with permission from Elsevier.

Boron nitride (BN) used as a Pd support acted as a regulator of the electron density of Pd and promoted the desorption of ethylene in acetylene semi-hydrogenation.[138] The Pd/BN catalyst provided 100% acetylene conversion and 93% ethylene selectivity with good stability.

Single-atom Pd species coordinated by dicyandiamide (Ndicy) and supported on hierarchical carbon nanocells (hNCNC) demonstrated an 18-fold increase in activity compared to similar catalysts in the absence of organic ligand (Pd1/hNCNC) and provided 99% C2H2 conversion and 87% selectivity at 160°C due to C2H4 desorption on isolated single Pd sites in Pd1 – Ndicy/hNCNC.[139] N-doped well-organized carbonaceous cells were prepared in situ by the template method at 800°C from MgO and pyridine or benzene as a carbon source. The Pd1 – Ndicy/hNCNC catalyst was prepared by pyrolysis of a mixture of the support, PdCl2 and dicyanamide at 700°C.

Single-atom Pd1/TiO2 catalyst was prepared by mechanochemical mixing of palladium and titanium phthalocyanines and TiOPc in a ball mill followed by calcination and reduction with hydrogen. For comparison, the same catalyst was obtained by the classical impregnation method. Such catalysts showed low yields in the hydrogenation of acetylene. It should be noted that light irradiation increased the conversion of acetylene hydrogenation from 20 to 80% with a simultaneous decrease in ethylene selectivity. DFT calculations demonstrated, that photoinduced electrons transferred from TiO2 to adjacent Pd atoms facilitated the activation of acetylene. The selectivity of Pd1/TiO2 particles in the form of a small number of encapsulated clusters was significantly higher than that of individual catalytic particles in photothermocatalytic hydro­genation of acetylene at 70°C.[140]

The supported Pd/TiO2 catalysts containing 0.10 wt.% Pd prepared by precipitation — reduction with NaBH4 at 300 – 700°C were used in the hydrogenation of acetylene to ethylene.[141] The Pd/TiO2-HT300* sample calcined at 300°C performed best, providing 100% conversion of acetylene and 80% selectivity for ethylene at 55°C (TOF = 0.12 s–1). The authors attributed the catalytic properties to the high content of oxygen vacancies in the sample (71.8%), as confirmed by in situ X-ray photoelectron spectroscopy, in situ Raman spectroscopy and H2 desorption.

Pd NPs encapsulated in a thin SrTiO3 shell were synthesized via a hydrothermal method from Sr(OH)2· 8 H2O and Na2PdCl4 followed by calcination in air and H2/Ar at 200°C for 1 h in a tubular furnace. These catalysts featured lower ethylene adsorption and provided 98% conversion and 92% selectivity to ethylene with a specific activity of 5552 molC2H2 (molPd h)–1 at 100°C, significantly exceeding the most catalysts comprising non-encapsulated Pd NPs.[142]

The Pd/Al2O3 nanocatalyst with a palladium content of 5% was synthesized by flame spray pyrolysis method and provided 97% conversion of acetylene and 62% selectivity to ethylene, TOF = 5 s–1.[143] In this case, the catalytic properties depended on the type of Al2O3 used. The use of Pd supported on pentacoordinated Al2O3 (Pd/meso-Al2O3) in the hydrogenation of a mixture of 0.33% C2H2 , 0.66% H2, 33% C2H4 demonstrated higher selectivity than Pd/com-Al2O3 (commercially available Al2O3) at various reaction temperatures (Fig. 10).[144] At acetylene conversion > 99%, the selectivity of Pd/meso-Al2O3 catalyst was 83%, whereas with Pd/com-Al2O3 it was 25%. According to energy dispersive X-ray spectroscopy (EDS) data, the dispersion of active Pd sites in the Pd/meso-Al2O3 catalyst was high and uniform, which contributed to the high catalyst productivity. The dispersion of Pd in the catalyst on the commercial support was low, and the metal formed conglomerates with an average diameter of 5.55 nm according to HR-TEM (HR is high resolution) data.

Fig. 10
The effect of temperature on acetylene conversion and ethylene selectivity to 0.02% Pd/meso-Al2O3 and 0.02% Pd/com-Al2O3 catalysts (a); and their time dependence under the following reaction conditions: 130°C, gas mixture of 0.33% C2H2, 0.66% H2, 33% C2H4 , GHSV = 6000 h–1). Reproduced from Ref. 144 with permission from Elsevier.

An important role in the manifestation of the properties of a metal catalyst is played by its precursor. The electronic structure of the metal in the starting compound determines the state of the metal in the catalyst. For example, an increase in the electron density on the metal was reported [145] to reduce the adsorption of ethylene on the negatively charged sites of Pd thereby improving the selectivity of the catalyst. This effect was clearly demonstrated on the example of the catalysts obtained from Na2PdCl4, PdSO4 and Pd(acac)2 by the wetness impregnation method followed by calcination in 10% H2/N2 at 400°C for 4 h. The Pd/α-Al2O3 catalyst synthesized from Pd(acac)2 had a higher electron density of the active Pd sites and demonstrated higher selectivity to ethylene than catalysts obtained from Na2PdCl4 and PdSO4 salts (Fig. 11). In addition, Pd/α-Al2O3 prepared from Pd(acac)2 was more stable than similar catalysts obtained from Na2PdCl4 and PdSO4 due to the formation of PdC phase during the prolonged reaction.

Fig. 11
Plots of Acetylene conversion vs. temperature (a), ethylene selectivity vs. acetylene conversion (b), ln(TOF) vs. corrected temperature (c); Arrhenius plots for acetylene conversion and selectivity to ethylene vs. time (d) for Pd/α-Al2O3 catalyst obtained from Na2PdCl4 (1), PdSO4 (2), and Pd(acac)2 (3). Fig. 11,c shows apparent activation energy values (Ea), kJ mol–1. Reproduced from Ref. 145 with permission from Springer Nature.

Highly dispersed stable PdC phase supported on α-Al2O3 was prepared using N-containing silane coupling agent as a modifier.[146] The resulting catalyst demonstrated ethylene selectivity 15% higher than the conventional PdAg alloy with total acetylene conversion at 240°C and greater time-on-stream stability for 200 h. The excellent ethylene selectivity was mainly due to the special structure of the PdC phase NPs with a distorted lattice, which contributed to weaker ethylene adsorption and led to strong interaction of Pd NPs with the support surface.

The Pd/MgAl2O4 catalyst, intended for the selective hydrogenation of acetylene, was prepared from Na2PdCl4 by impregnation and then supported on mixed oxides followed by calcination at 600°C and reduction with H2.[147] Partially positively charged Pd atoms effectively reduced the adsorption of intermediates during hydrogenation, providing 96% acetylene conversion with 87% ethylene selectivity and excellent stability for a 50-hours test at 120°C.

Using carbon nanotubes (NTs) as a support for Pd instead of α-Al2O3 suppressed the formation of palladium hydrides, thereby increasing the yield of ethylene.[148] X-ray absorption spectroscopy, SEM and temperature programmed desorption (TPD) study of C2H4 in combination with DFT calculations demonstrated the presence of a unique local environment of Pd containing subsurface carbon atoms, which increased the acetylene conversion. According to kinetic study of the liquid-phase hydrogenation of acetylene over a commercial 0.5 wt% Pd/Al2O3 spherical catalyst in N-methylpyrrolidone (NMP) in a reactor at 60 – 100°C and 15 – 250 psi (1 – 17 bar), the reaction occurs on the surface and is limited by minor diffusion resistances between and within the catalyst species.[149]

Pd0.87/MoS2 catalyst was synthesized by deposition of Pd particles on MoS2 nanosheets obtained by ultrasonic exfoliation in formamide.[150] The catalyst provided 75% selectivity to ethylene and 100% conversion at a space velocity of acetylene of 40 000 mL h–1 g–1. According to SEM, Pd particles were dispersed on the glossy MoS2 surface with a low nominal Pd loading. Strong Pd–MoS2 interactions arising from the charge transfer from Pd to MoS2 were observed by XPS.

The effective interaction between Pd and the support resulted in high selectivity acetylene hydrogenation to ethylene (100% acetylene conversion at 240°C) of the Fe5C2-supported Pd/α-Al2O3 catalyst.[151] The catalyst was prepared by reduction of FeCl3 with NaBH4 in isopropanol. The resulting Fe NPs reduced Pd2+ to Pd0 when adding Pd(OAc)2 to the solution. The addition of oleylamine and heating under reflux at 350°C afforded the desired catalyst. DFT calculations confirmed that the energy barrier for the conversion of ethylene to ethane on this catalyst was very high, making the complete hydrogenation unlikely.

Natural minerals such as complex silicates and alumosilicates have also found application as supports for hydrogenation catalysts. For example, pre-calcined halloysite nanotubes (HNT-c) were used to support Pd NPs from PdCl2 and Pd(OAc)2 obtained in solutions of ammonia (NH3), sodium chloride (NaCl), hydrochloride (HCl) and in acetone (Ac).[152] The highest (~ 65%) ethylene selectivity at acetylene conversion < 70% was achieved using Pd-HNT-c-Ac. The other samples demonstrated close selectivity (55 – 60%), probably due to palladium poisoning with chlorine and amines. Catalysts based on silanized HNTs (HNT-as) provided the selectivity in the range of 45 – 59% with conversion below 80%. The highest selectivity was demonstrated by Pd-Hal-as-HCl. It should be noted that the performance of Pd-HNT catalysts was comparable to that of palladium catalysts supported on Al2O3, SiO2 and TiO2 . However, their ethylene selectivity is still insufficient for commercial use.

Palygorskite (Pal) is a naturally abundant aluminosilicate mineral that was used as a support for ultralow content Pd (0.8%) hydrogenation catalysts.[153] This mineral was treated ammonia (Pd/N-Pal) or air (A-Pal). The treatment with ammonia to produce Pd/n-Pal ensured easy and strong fixation of metal particles on the support and increased their dispersion due to a decrease in the crystal size and adsorption of H2 at a temperature of 90 – 110°C, which ultimately contributed to an increase in the selectivity to C2H4. On the contrary, air-treated Pd/A-Pal catalysts demonstrated a weak interaction of Pd particles with the support and low dispersion, therefore, the selectivity to ethylene for these catalysts was lower. The TOF values for Pd/A-Pal and Pd/N-Pal in acetylene hydrogenation at 90°C and the reaction time of 360 min were 0.345 and 0.05 s–1, respectively. It is worth noting that the increase in temperature promoted PdHx hydride phase formation on the Pd/A-Pal catalyst, providing the complete hydrogenation of acetylene to ethane.

N-Doped carbon-supported Pd/NC catalysts were prepared from H2PdCl4, activated carbon and 1,3-diethylimidazole acetate/N-methylpyridinone dihydrogen phosphate ionic liquids by wetness impregnation technique followed by calcination at 650°C and hydrogenation at 400°C for 4 h in a tube furnace.[154] Nitrogen doping altered the electronic structure of Pd NPs, allowing rapid desorption of ethylene and preventing ethane formation during overhydrogenation. The ethylene selectivity to this catalyst was 91.3% over 40 h with 94.6% acetylene conversion.

The single-atom Pd1-N8/CNT catalyst on multi-walled carbon nanotubes (CNT) was studied by cyclic voltammetry (CVA).[155] Using voltammograms, the azide ion oxidation into azide radicals with subsequent formation of well-ordered polynitrogen species, as well as the palladium reduction from its oxide were identified. PdO/CNT was used as the working electrode. The N8 fragment was formed from NaN3 in a buffer solution under electrolysis and attached to the single-atom Pd, stabilizing it. Calculations using acetylene temperature programmed desorption (C2H2 – TPD) and DFT confirmed that C2H2 favours π-bonding on Pd-SA, and H2 dissociated on the N atom attached to Pd. This Pd – N synergistic effect favoured the formation of C2H4 during acetylene hydrogenation and prevented its complete hydrogenation to C2H6 .

Modification of the Pd/SiO2 catalyst with histidine followed by calcination at 240°C provided 1.0 nm Pd NPs compared to the untreated catalyst.[156] According to the XPS data, Pd NPs of the Pd – His/SiO2 catalyst were highly dispersed and positively charged, which provided a high semi-hydrogenation selectivity of 81.5% (100% acetylene conversion at 160°C) and high stability at 160°C for 50 hours. The desorption temperature of C2H2 decreased from 123°C for Pd/SiO2 to 112°C for Pd – His/SiO2, while ethylene was easily desorbed from the latter catalyst during the reaction, which prevented the formation of ethane.

Encapsulated palladium nanoparticles or nanoclusters exhibit high selectivity in selective hydrogenation due to the fact that the thin shell of the support or capsule protects the surface of the active metal. Pd nanoclusters coated with a thin SrTiO3 shell demonstrated selectivity to ethylene of 92% with acetylene conversion of 98%.[142] The specific activity of such a catalyst was 5552 molC2H2 (molPd h)–1 at 100°C, which significantly exceeds the similar parameter of most palladium-based catalysts. Selectivity was achieved due to the sorption of C2H2 on Ti3+ capsule sites instead of Pd, and H2 was sorbed on palladium and transferred to the catalyst shell, where reduction occurred.

An unusual type IL@Pd@NMC palladium catalysts of the core-shell type on N-doped mesoporous carbon (NMC) with a metal content of 20 – 50% was synthesized by wet impregnation and subsequent treatment with MeCN, ionic liquids (ILs) and drying. These catalysts provided ethylene selectivity > 93% with complete acetylene conversion in the temperature range of 100 – 200°C.[157] Importantly, they demonstrated excellent stability, providing 100 h on stream for the hydrogenation of acetylene to ethylene without any obvious decrease in productivity. The efficiency of IL@Pd@NMC was comparable to Pd@NMC, making it a potential candidate for industrial use.

Similar Pd catalysts Pd@S-1@IL obtained by impregnation of zeolite (S-1) with solutions of tetrapropylammonium hydroxide (TPAOH), tetraethyl orthosilicate (TEOS), hexadecyltrimethoxysilane (HTS) and followed by incapsulation with [Prmim][Cl] ionic liquid (Prmim is 1-methyl-3-n-propylimidazolium) exhibited excellent ethylene selectivity (> 90%) regardless of the level of substrate conversion and excellent stability over 800 h on stream even under relatively harsh conditions.[158] The high selectivity of the catalyst was due to the structure of the catalyst, which resembled a kind of ‘filter’ (Fig. 12). Encapsulation of Pd NPs inside the support prevented metal migration, and the IL layer regulated gas sorption-desorption on the catalyst surface, suppressing the process of overhydrogenation to ethane. In addition, the IL layer provided heat removal from the metal catalyst and prevented the formation of hot spots, thereby hampering the formation of polymerization products. The high stability of the atomically dispersed Pd hydrogenation catalysts supported on TiO2 and coated with 1-n-butyl-3-methylimidazolium tetrafluoroborate ([BMIM][BF4]) was due to the electrostatic interaction of the IL and Pd/TiO2.[159] Moreover, the synergism of IL and Pd/TiO2 suppressed coke deposition on the catalyst, ensuring its higher performance and stability. And the ionic liquid shell prevented the formation of ‘hot spots’, preventing the formation of ‘green oil’, oligomerization products.

Fig. 12
Schematic representation of Pd@S-1@ILs catalysts. Reproduced from Ref. 158 with permission from Elsevier.

An unusual single-atom catalyst was synthesized by in situ confinement of Pd(acac)2 between two-dimensional layers formed by Cu, 4,4'-bipyridine (bipy) ligand and SiW12O404– (SiW) anions in the outer sphere.[160] The small distance (6.25 Å) between two SiW anions restricted the movement of Pd(acac)2, and the SA-Pd can be directly fixed by oxygen atoms on the SiW surface during the reduction without agglomeration. The resulting Pd1@Cu – SiW catalyst demonstrated 92.6% ethylene selectivity and complete acetylene conversion at 120°C. Surprisingly, the ethylene selectivity remained > 90% with increasing reaction temperature, whereas it usually drops sharply in the presence of common catalysts.

A useful approach to catalytic hydrogenation using parahydrogen (p-H2) was shown in a series of works by I.V.Koptyug. Parahydrogen is a hydrogen molecule in a singlet state, in which two nuclear spins are antiparallel. The p-H2 spin order was transferred to product molecules in hydrogenation, provided that the process involved the simultaneous transfer of two hydrogen atoms of the parent p-H2 molecule to the substrate. This phenomenon of parahydrogen-induced polarization is used in NMR spectroscopy for mechanistic studies of reactions involving molecular hydrogen, e.g., hydrogenation. Strongly enhanced absorption/emission signals were recorded in the NMR spectra of the hydrogenated products. The observed signal enhancement can theoretically reach 105 times and even more than when using ordinary hydrogen.[161] Thus, catalytic hydrogenation with parahydrogen of propyne,[162-166] propene[167][168] and some other compounds[169] was studied.

Selective hydrogenation of acetylene is also of fundamental importance, since it is practically the only way to enrich the nuclear spin isomers of ethylene using p-H2 . For example, in the hydrogenation of acetylene with p-H2 (C2H2 : p-H2 = 1 : 4) over Pdo37/SiO2 with octahedral Pd nanoparticles at 150°C (Scheme 1), a strong enhancement of the signals of the oligomerization products was observed in the NMR spectra against the background of the hydrogenation products of the C≡C triple bond (Fig. 13a).[170] Although ethylene was the predominant reaction product, no enhancement of the signals of the protons at the C=C bond was observed. The reason was in blocking of the nuclear spin polarization by the magnetic equivalence of all protons in the ethylene molecule. The NMR signals of the oligomerization products, on the contrary, were significantly enhanced (nuclear spin polarization > 1.7%), despite the low concentration of these compounds in the mixture. For comparison, the spectrum obtained after the nuclear spins returned to thermal equilibrium is shown in Fig. 13b. The authors[170] found, that the signal amplification was higher the larger the catalyst NPs used in hydrogenation. The fact that oligomers accumulated higher polarization than alkenes was additionally confirmed by experiments on propyne hydrogenation.

Scheme 1
Fig. 13
1H NMR spectra obtained by hyperpolarization of the reaction mixture (a) and after thermal equilibrium has been reached in the stopped-flow experiment (b). Reproduced from Ref. 170 with permission from the American Chemical Society.

Hyperpolarized ethylene was obtained by continuously passing a mixture of acetylene and p-H2 over a Pd/TiO2 catalyst.[171] A similar experiment was carried out using [D2]acetylene, which formed DHC=CHD when hydrogenated with p-H2. This proved that the addition of parahydrogen occurred in pairs and preserved the correlation of individual nuclear spins of the substrate molecule. It was shown that ethylene obtained with p-H2 exhibited hyperpolarized signals in an anisotropic medium.

Selective hydrogenation of acetylene with parahydrogen over Pd catalysts and iridium complexes was studied.[172] NMR spectroscopy studies showed that isomers of individual nuclear spin complexes of ethylene formed by the addition of p-H2 to acetylene equilibrated within 3 – 6 s for spin systems with the same symmetry (syn-addition of hydrogen) and within 1700 – 2200 s for spin systems with different symmetry (anti-addition).

Acetylene molecules contain the isotope 13C, the number of which is small and amounts to 2.2% of the total number, according to the natural abundance of the isotope 13C. To achieve specific objectives, the content of the 13C isotope can be artificially increased by 13C-labelling the desired molecule.[173][174] Such molecules, when hydrogenated with parahydrogen, catalyzed by Rh and Ir complexes, would give increased intensity signals of 1H and 13C in the corresponding NMR spectra.[175] The enhancement of the 1H NMR signal of 13C-ethylene protons was 140! And the enhancement of the signals of carbon nuclei was enough to observe 13C-ethylene even at extremely low concentrations due to the polarization caused by parahydrogen. The shape of the line of polarized 13C-ethylene (Fig. 14) unambiguously indicated that the addition of p-H2 to the acetylene triple bond on immobilized iridium complexes proceeds stereoselectively as syn-addition. In fact, this method is an efficient approach for enrichment of nuclear spin isomers of ethylene.[175]

Fig. 14
1H NMR spectra of ethylene formed by hydrogenation of acetylene with the SiO2-supported catalyst [Ir2(COD)2(μ-Cl)2] (flow rate of the gas mixture 4.6 mL s–1) before (a) and after relaxation of hyperpolarization to thermal equilibrium (d); (COD is cycloocts-1,5-diene); simulated spectra of 13C-ethylene formed by syn- (b) or anti- (c) addition of p-H2 (c). Reproduced from Ref. 175 with permission from Springer Nature.

* The figures in the designation of the support indicate the reduction temperature - 300°C.

4. Bimetallic palladium catalysts

The introduction of another metal into the catalyst to control the geometry and electronic structure of Pd NPs is widely used strategy to effectively improve its overall catalytic activity.[176][177] The comparison, analysis and summary of the synthetic methods of Pd-based bimetallic compounds can provide useful guidelines for the further development of such catalysts.

A mixed PdAg catalyst supported on the mesoporous carbonaceous material Sibunit was obtained by impregnation followed by reduction. Interestingly, the catalysts synthesized by alternating impregnation of the support with solutions of Pd(NO3)2 and AgNO3 in a H2 medium demonstrated lower ethylene selectivity (68 – 73%) due to the formation of species that were non-uniform in both composition and size (about 4 – 60 nm). This was probably due to the removal of O-containing functional groups from the carbon surface during the hydrogenation process, which contributed to the formation of larger catalyst particles. Impregnation with a mixture of salts followed by reduction provided the formation of more uniform particles.[178]

A series of bimetallic Pd and Ag catalysts were tested in the selective semihydrogenation of acetylene to identify the leading catalysts.[179] Based on the calculations and a series of experiments, the authors were able to find out that the nature of the catalyst surface directly determined its ability to adsorb acetylene and ethylene. It was also obvious that the nature of the surface can be changed depending on the ratio of metals in an alloy. The screening strategy described in this research, allowed for rational engineering, i.e., theoretical determination of those forms of the catalyst that demonstrated maximum activity and selectivity. Based on this strategy, the catalysts Ag7@Pd, Pd1@Ag, and Pd1Ag3 were proposed. Further, by means of microkinetic analysis, it was found that the Pd1Ag3 alloy demonstrated the superior cheracteristics of the tested samples. The calculation methods also confirmed that the presence of silver in the alloy catalyst facilitated the desorption of ethylene, thereby increasing the hydrogenation selectivity.[103] Using the calculations and experimental research, interesting patterns were found for alloy binary PdAg catalysts.[180] The formation energy of palladium alloys was 0.06 eV atom–1. The role of silver is to limit the dissolution of hydrogen in palladium. The more silver atoms per palladium atom, the more isolated the palladium active sites are on the surface. Silver atoms are combined on the surface, but this segregation is reversible under the action of hydrogen, so the catalyst surface is constantly changing over time. The nature of the palladium species is also changed: they also tend to aggregate, which led to a deterioration of the catalytic properties of the catalyst as a whole. The highest catalytic activity was found in samples of the Pd1Ag3 composition, which is also confirmed by other studies.[179]

An alternative approach to the bimetallic PdAg catalyst was based on mechanochemical action on solid starting compounds.[181] Compared to a similar catalyst obtained by the standard wet impregnation procedure, this catalyst was more resistant to deactivation and withstood 15 h on stream rather that 5 – 10 h. Therefore, the mechanochemical approach can be considered as an alternative to the synthesis of complex catalytic systems with increased time-on-stream. A number of bimetallic catalysts with different Pd : Ag ratios was obtained by the original procedure from the corresponding nitrates, a stabilizer and sodium cyanide as an auxiliary reagent.[182] The catalysts were then placed on an alumina support and tested in acetylene hydrogenation. Compared to standard hydrogenation catalysts (Pd Nanoselect LF200 and Pd/Al2O3), the bimetallic catalyst demonstrated 80% selectivity to ethylene even after 48 h.

As noted above, non-selective hydrogenation of acetylene can produce ethane and C4+ oligomeric compounds, the so-called ‘green oil’, which is a catalyst poison. A study [183] has been conducted to assess the contribution of buta-1,3-diene formed during hydrogenation to the course of the subsequent process. For this purpose, 1,3-butadiene was introduced into the experimental setup for selective hydrogenation of acetylene using PdAg/Al2O3 as the model catalyst. It was found that C4H5 species were intermediates in the formation of C6 species. It was also proved that butadiene does not react with acetylene to give longer-chain hydrocarbons, but is actively actively hydrogenated to 1-butene, cis- and trans-2-butenes and butane. The amount of resulting C4 hydrocarbons directly depends on the reaction time and the catalyst type. Although direct reaction of acetylene and 1,3-butadiene was not observed, two acetylene molecules can react with each other to form C4H4 species, which can then react with acetylene to give benzene. In the case of the mixed PdAg catalyst, the treatment of the catalyst with carbon monoxide led to segregation of Pd atoms at room temperature and results in the additional enrichment of the surface with palladium (up to 31%).[184] The IR spectrum of CO on the monometallic Pd catalyst (Fig. 15, the black curve) showed a maximum at 2088 cm–1, which corresponds to linearly adsorbed CO. The maximum at 1990 cm–1 and a broad shoulder at 1944 cm–1 corresponds to multicoordinated CO. Based on the signal intensities, multicoordination clearly dominated, i.e. adsorption on two and three palladium atoms. For the mixed PdAg2/Al2O3 catalyst, linearly adsorbed CO dominated (Fig. 15, red curve), which indicated the disappearance of polyatomic Pd centres and the formation of the single-atom Pd catalytic sites, which are separated by Ag atoms. Thus, pre-treatment of the catalyst with CO increases its in activity at 100% selectivity in the acetylene hydrogenation.

Fig. 15
FT-IR diffuse reflectance spectrum of CO adsorbed by Pd/Al2O3 (top); and by PdAg2/Al2O3 (bottom). Reproduced from Ref. 184 with permission from Elsevier.

A series of core-shell catalysts was prepared using a standard impregnation technique followed by deposition on alumina.[185][186] The resulting catalysts consisted of monoatomic alloys of Pd1Ag6 and Pd1Ag and showed selectivity to ethylene 90% at conversion 80%. However, with an increase in conversion to 100%, the selectivity was less than 60%.

A comparative study of the properties of bimetallic Pd catalysts and their analogues modified with S or Si atoms in acetylene semi-hydrogenation using experimental and computational methods has been reported.[187] According to the calculations, the catalytic efficiency of C2H2 semi-hydrogenation largely depends on the spatial extension of the active site and the electronic effect, which consists of a change in the density of the d-zone near the Fermi level and has a significant influence on the sorption characteristics of the catalyst. The electronic effect of the catalyst caused by S atoms was more significant than the effect of Si atoms. At the same time, the geometry of the active centre for Si-analogues changed more than in S-analogues. Thus, for bimetallic PdAg catalysts, the synergistic geometric and electronic effects caused by S atoms on S/Pd6Ag, enhanced the activity and selectivity to C2H4 formation and significantly prevented the formation of a ‘green oil’. Modification of the catalysts with silicon was slightly less effective than with sulfur, and they both increased the selectivity to C2H4 and prevented the formation of ‘green oil’ in the hydrogenation of acetylene on Si/Pd1Au1 and Si/Pd1Ag1 catalysts.

Bimetallic supported Ag@Pd/SiO2 catalysts coated with Ga2O3 demonstrated high selectivity to ethylene (up to 95%) with complete conversion of acetylene in ethylene excess.[188] The Ga2O3-coated monometallic Pd/SiO2 catalyst showed moderate activity. At the same time, the activity and high stability of the Ga2O3-coated Ag@Pd/SiO2 catalyst were nearly 50 times higher than those of the reference Pd1Ag/SiO2.

Alumina is often used as a support for acetylene hydrogenation catalysts. However, its small surface area limits the dispersion of active metals. To address this issue, alumina was prepared by a special in situ growth method (Fig. 16).[189] Then, the alloy bimetallic catalyst PdAg was applied to the obtained support, resulting in 90% acetylene conversion and 89% selectivity. It was due to even greater isolation of the active Pd sites, a more degradation-resistant substrate structure and a decrease in the rate of heat release. Decreased heat release led to a decrease in the number of hot spots on the catalyst surface, and, consequently, to a decrease in the rate of the endothermic coking reaction. As a result, the catalyst pores remained open for a longer time, thereby increasing the catalyst stability.

Fig. 16
SEM images of the surface of standard spherical aluminum oxide (a) and aluminum oxide with a modified flower-shaped array (b). Reproduced from Ref. 189 with permission from the American Chemical Society.

Liquid-phase hydrogenation of a N2-stabilized gas mixture consisting of C2H2 (8%) CO (20%) and H2 (50%) was carried out in NMP in the presence of PdAg/SiO2 showed 92% selectivity to ethylene.[190] The similar characteristics for gas-phase hydrogenation of the same mixture was only 71%.

A mixed SiO2-supported palladium – nickel catalyst was prepared in several steps.[191] First, the Ni/SiO2 nanocatalyst was synthesized by impregnation with subsequent reduction in a hydrogen stream. Then, single palladium atoms were deposited on the surface. Acetylene hydrogenation over this catalyst at a temperature of 80°C proceeded with a conversion of 92% and a selectivity for ethylene of 88%. The authors were able to prove the presence of SA-Pd on the catalyst surface by scanning transmission electron microscopy – energy dispersive X-ray spectroscopy (STEM-EDS) (Fig. 17).

Fig. 17
STEM-EDS image of the mixed PdNi/SiO2 catalyst. Yellow dots indicate Ni atoms; green dots indicate Pd atoms. Reproduced from Ref. 191 with permission from Elsevier.

Metal single-atom alloy (SAA) catalysts based on Pd and Ni were prepared by galvanic replacement (GR) method.[192] For this purpose, nickel nanocrystals were first obtained by the colloidal method. Then, palladium atoms were deposited on nickel via galvanic replacement. And the finished nanocrystalline Pd1Ni alloy was placed on SiO2 support. The resulting alloy was tested in the selective hydrogenation of acetylene in comparison with a similar catalyst obtained by common impregnation. It was found that the catalyst obtained by the galvanic replacement showed higher selectivity in the range of 30 – 180°C. According to DFT calculations, electrons were transferred from nickel to palladium atoms in the alloy catalyst.

The efficiency of mixed PdNi catalysts was improved by replacing Pd with monoatomic Pd. Metallic single-atom alloys Pd1Ni/SiO2-GR obtained by the galvanic replacement method for selective deposition of colloidal Ni nanocrystals on the SiO2 support with subsequent deposition of Pd atoms on Ni atoms.[192] The Pd atoms remained atomically dispersed and are located in proximity to Ni atoms (Fig. 18), which improved the ethylene selectivity in the acetylene semi-hydrogenation reaction compared to the reference PdNi/SiO2 catalyst prepared using the impregnation method. The Pd1Ni/SiO2-GR catalyst showed higher C2H4 selectivity over a wide temperature range (30 – 180°C), being superior to the Pd1Ni/SiO2, Pd/NiO, and Pd/SiO2 SAA catalysts prepared by the impregnation method. According to DFT calculations, the enhanced selectivity stems from the electron transfer from a Ni nanocrystal to single Pd atoms due to the close proximity of these species.

Fig. 18
STEM-EDS mapping showing the close contact of Pd and Ni in the Pd1Ni nanocrystal (a); line scan profile of Ni and Pd in Pd0.1Ni2/SiO2-GR-500R (b); AC-HAADF-STEM images of Pd0.1Ni2/SiO2-GR-500R showing the presence of single Pd and Ni atoms in different scale (c, d). Reproduced from Ref. 192 with permission from Elsevier.

The performance of palladium catalyst was significantly improved by inserting tin atoms onto the surface.[193] The resulting Pd2Sn/C catalyst demonstrated ~ 95% selectivity in hydrogenation of acetylene with almost complete conversion of the substrate. According to experimental studies and DFT calculations, the addition of tin led to a redistribution of Pd atoms over the catalyst surface, which in turn affected the ethylene sorption. Ethylene molecules adsorbed at 300 – 600°C were weakly bound to the surface and easily left it being non-dissociated. At the same time, ethylene molecules that were adsorbed at 439 and 500°C were bound to the surface more strongly and could further react with dissociated hydrogen to form ethane.

Mixed Al2O3-supported PdMn catalyst was prepared in hydrogen reducing medium.[194] According to analytical studies, the surface of thus modified catalyst lacks the weakly bound hydrogen, which was explained by the presence of a uniform distribution of Mn atoms over the surface. In tests for selective hydrogenation of acetylene, the catalyst demonstrated increased activity and a 20% increase in selectivity compared to the unmodified palladium catalyst. The complex Pd/In2O3@SiO2 catalyst was prepared using a standard deposition–precipitation procedure.[195] Doping the catalytic system with indium suppressed palladium activity and also facilitated the adsorption of di-σ-bound ethylene, which significantly improved its desorption. As a result, the selectivity to ethylene was 80% with 100% conversion of acetylene at 140°C for 24 h.

The bimetallic PdZn catalyst demonstrated 85% ethylene selectivity in acetylene hydrogenation even after 70 hours on stream.[196] Comparing to other alloy catalysts, that was not the best result; however, it was very promising because of the room temperature reaction conditions. Based on the calculated data, the authors explained the role of zinc by donating electrons from the occupied zinc d-orbitals to the Pd d-orbitals, which facilitated the interaction of acetylene with the surface. Moderate results were obtained for the bimetallic Sibunit-supported PdZn catalyst.[197] The addition of zinc to Pd : Zn in a ratio of 1 : 1 increased the catalyst selectivity to 67% compared to 42% for monometallic palladium. Further increase in zinc content decreased the selectivity to 62%. According to analytical studies, zinc was incorporated into the palladium crystal lattice during the reduction of metals. With an excess of palladium, the interaction occurred more completely with the formation of an intermetallic phase. With an excess of zinc, lesser incorporation was observed, while zinc was retained in the form of large oxide particles. Further studies demonstrated that an increase in the reduction time of Pd – Zn/C samples in H2 at 500°C from 1 to 20 h was accompanied by an increase in the number of bimetallic sites and an increase in the coordination number of PdZn from 2.4 to 4.3, as well as a decrease in the Pd0/PdZn ratio from 0.63 to 0.18.[198]An increase in the number of modified active sites led to higher selectivity to ethylene in the process of liquid-phase hydrogenation of acetylene.

Spinel monolithic Al–LDH catalysts with porous woodpile architecture were fabricated by extrusion three-dimensional (3D) printing of layered aluminate-intercalated double hydroxide (Al-LDH) followed by low-temperature calcination.[199] Porous monolithic catalysts with a high specific surface area facilitated mass and heat transfer and exposed active catalyst sites (Fig. 19). Thus, the 3D-printed spinel-based catalyst loaded with Pd (3D-AI-Pd/MMO (MMO is MgAl-mixed metal oxide) provided 100% conversion in acetylene hydrogenation with selectivity to ethylene more than 84% at 60°C.

Fig. 19
Schematic representation of the selective hydrogenation of acetylene to ethylene over a 3D-printed 3D-AI-Pd/MMO catalyst with a hierarchical porous structure (a) and a comparison diagram of different catalysts for this process (b). CNF is carbon nanofibers. Reproduced from Ref. 193 with permission from Springer Nature.

Pd1 – Cun/Al2O3 catalysts with single palladium atoms (Pd1) and copper clusters (Cun) were deposited on different supports. The synergistic effect of one palladium atom and a copper cluster provided the isolation effect of catalytic sites and effectively prevented their agglomeration. The activity of the Pd1 – Cun/Al2O3 catalyst increased by 60% compared to the Pd1/Al2O3 catalyst, and its stability increased to more than 100 h with the same conversion values for both catalysts with ethylene selectivity of 83%.[200] The PdCu/siloxene catalyst was prepared by direct reduction in the liquid phase and was characterized by a low loading of Pd distributed in ultrafine Cu species.[201] This catalyst provided 91% acetylene conversion and 93% ethylene selectivity at 200°C in the acetylene semi-hydrogenation reaction. Moreover, PdCu/siloxene also demonstrated high activity and stability under ethylene-rich industrial gas conditions. In addition, the catalyst was resistant to coking at high operating temperatures for 39 hours.

The Pd1Au1 dimeric catalyst supported on a metal-organic framework demonstrated an outstanding (99.99%) acetylene conversion with a selectivity of over 90% under industrial-scale conditions (1% acetylene, 89% ethylene, 10% H2 mixture).[202] The catalyst worked efficiently over a wide temperature range (35 – 150°C), with only ethane as a by-product. The reaction proceeded with an apparent activation energy of ∼ 1 kcal mol–1 even at 35°C and with operating windows (> 100°C) and mass hourly space velocities (66 000 mL g–1 cat h–1) within industrial specifications. Experimental and theoretical studies demonstrated that palladium atoms performed the main catalytic function, while gold atoms and thioether moieties of the metal-organic framework participated in hydrogen dissociation.

An interesting technique for regulating the ratio of metals in a bimetallic catalyst was implemented on the example of a PdCo supported on the carbon material Sibunit.[203] The catalysts were prepared by wet impregnation followed by hydrogenation in the stream of hydrogen at various temperatures. At temperatures below 500°C, palladium was reduced to a metallic state, and cobalt was only partially reduced. As the temperature raised to 700°C, larger amount of cobalt was reduced to metal. In this way, catalysts of compositions from Pd0.6Co0.4 to Pd0.45Co0.55 were obtained. With an increase in the catalyst treatment temperature, the yield of ethylene in the acetylene hydrogenation also increased from 56% to 68%. For comparison, the yield of ethylene in hydrogenation in the presence of monometallic palladium on carbon was only 52%. According to the data obtained, the improvement in the catalyst characteristics was associated with a change in the electronic state of palladium due to its surrounding by cobalt atoms.

The bimetallic Pd-Ga/γ-Al2O3 catalyst obtained by impregnation Al2O3 with solutions of the appropriate nitrates demonstrated selectivity to ethylene about 66 – 70%.[204] Outstanding results were obtained for the calcite-supported PdBi catalyst (PdBi/Calcite) prepared by multistep synthesis that included of metal salt reduction with NaBH4 and deposition on the support.[205] This catalyst showed 99% selectivity to ethylene at 100% conversion over the entire temperature range (50 – 300°C)! Thus, the working range with acceptable characteristics was 150°C. The authors carried out a series of experiments to clarify the nature of such high activity. An assumption was made about the influence of bismuth oxide on the activity of Pd metal. However, it was found that palladium and bismuth formed intermetallic compounds that significantly contribute to the reaction mechanism, and the key process parameters depend on their formation. At the same time, the contribution of oxides was insignificant. The intermetallic catalysts contained isolated and electron-rich Pd atoms as active catalytic sites and had an extremely weak ability to absorb ethylene and prevent the formation of palladium hydrides (β-PdHx) on the catalyst surface.

A bimetallic CuPd catalyst was synthesized especially for the selective hydrogenation of acetylene to ethylene.[201]The undoubted novelty of the research was the use of siloxene as a support. Catalytically active sites consisted of isolated palladium species surrounded by ultra-small copper species, and all the species were uniformly distributed in siloxene, which has high porosity and surface area. The catalyst demonstrated high resistance to sintering for 39 h and provided 91% conversion of acetylene with 93% selectivity to ethylene at 200°C.

The synthesis of mixed hydrogenation catalysts obtained by deposition of bismuth on cubic palladium nanoparticles (Pd – NC) was described.[206] Dotted applicated bismuth poisoned palladium atoms, adjusting the catalyst activity. The PdNCs@Bi0.06/Al2O3 catalyst obtained by applying bismuth to the corners of Pd cubes, provided a selectivity to C2H4 of 10.1% with 100% conversion of C2H2 , which exceeded the catalytic characteristics of the PdNCs/Al2O3 catalyst. With an increase in the degree of bismuth inclusion (i.e., with an increase of poisoning), the C2H2 conversion gradually decreased, and the C2H4 selectivity increased (Fig. 20), which made it possible to vary the conditions depending on the tasks at hand. For example, the Pd-NC@Bi catalyst with bismuth deposited on the corners and faces of Pd nanocubes had C2H2 conversion of 99.7% and selectivity to C2H4 94.3% at 170°C in acetylene semi-hydrogenation in excess of ethylene.

Fig. 20
Temperature dependence of conversion (a) and selectivity (b) of acetylene hydrogenation on Pd-NCs@Bi0.5/Al2O3 catalyst. Reproduced from Ref. 206 under a Creative Commons Attribution (CC BY) license

A mixed Pd – La nanocatalyst was obtained by ultrasonic impregnation using nitrogen-doped biochar derived from Pennisetum giganteum as a support (N-pgBC). PdLa/N-pgBC provided 94.2% selectivity to ethylene with full acetylene conversion in the acetylene hydrogenation.[207] The synergistic effect of the inherent high specific surface area of the carbon support together with the separation of the lanthanum-doped catalytic sites promoted adsorption of acetylene and the rapid desorption of ethylene, blocking overhydrogenation of the latter. In addition, the PdLa/N-pgBC catalyst showed a strong dependence on the component composition, which might be due to electronic or geometric modification of metal particles.

5. Pd-Free catalysts for acetylene generation. Ni-based catalysts

Nickel hydrogenation catalysts as completely exposed small metal nanoclusters (Fig. 21), were obtained on a nanodiamond (ND or graphene (G) support.[208] Calculation methods and H2/D2 exchange experiments demonstrated that acetylene and hydrogen are completely activated on the catalytic sites. Moreover, the desorption of ethylene on exposed nickel clusters was more energetically favourable than further dissociation of an H2 molecule. Therefore, exposed nickel clusters were able to completely convert acetylene to ethylene with 85% selectivity at 190°C. When replacing the carbon substrate with a mesoporous material MCM-41 based on silicates and aluminosilicates, the nickel nanocatalyst provided 96% conversion of acetylene and selectivity to ethylene of 87%.[209]

Fig. 21
High-resolution transmission electron microscope images of nanodiamond and graphene (a), and nickel atoms (Ni1), nickel NPs (Nin) and nickel nanoclusters (Nip) supported thereon. Nickel atoms and nickel clusters are circled in white. Reproduced from Ref 208 with permission from the American Chemical Society.

A very interesting idea was realized in the hydrogenation of acetylene to ethylene in the presence of nickel catalysts with a nitrogen-doped carbon support (gC3N4).[210] The authors used acetylene obtained by hydrolysis of calcium carbide as a source of ethylene, i.e., ethylene was produced not from the standard cracking of petroleum products. It should be noted, that the use of acetylene from calcium carbide in synthetic transformation is a common practice.[211-218] Nevertheless, the authors managed to obtain impressive results: the catalyst provided 96% conversion of acetylene for 10 hours at 200°C, while the selectivity to ethylene reached 46%.

A similar catalyst was described in a study [219] but it was additionally doped with sulfur. Sulfur-doped nickel supported on gC3N4 – T (the T and M locants denote the sulfur-doped support and the sulfur-free support, respectively) showed ethylene selectivity > 63% with complete acetylene conversion at 175 – 300°C. XRD, HRTEM, EDS and XPS revealed that the catalyst represents Ni cations enclosed in gC3N4-T cavities or a Ni – S solid solution with surface segregation of S different from the Ni species for Ni/gC3N4-M. The Ni2+ sites above the gC3N4-T support promote the acetylene activation via electrostatic interaction.

Single-atom nickel-dispersed catalysts were obtained by introducing zinc additive.[220] A nitrogen-doped carbon matrix was used as a support. As a result, the presence of the Ni – N4 structure in the carbon matrix was detected and confirmed. The conversion of acetylene in a real mixture with ethylene was > 97%, which was explained by weak π-adsorption of ethylene on single Ni atoms.

Ni catalysts on various supports (SiO2, γ-Al2O3 and zeolite ZSM-12) were investigated for the hydrogenation of acetylene.[221] The Ni/ZSM-12 catalyst showed low selectivity to ethylene and low stability. However, this catalyst performed better than the Ni/SiO2 and Ni/γ-Al2O3 catalysts. due to the synergistic action of Lewis and Brønsted acid sites. Doping with Sn afforded a more effective Ni7Sn/ZSM-12 catalyst, which showed 92.51% selectivity to ethylene with complete acetylene conversion at 250°C. These values were 19.4% higher than those in the absence of Sn. According to DFT calculations, the bimetallic catalyst had a rather high energy barrier to the formation of the intermediate compound C2H5* in acetylene hydrogenation, which determined its high selectivity.

Outstanding results necessary to understand the effect of additives on nickel catalysts were demonstrated using Au, Ag and Cu additives as an example.[222] It was found, that the energy of initial adsorption of acetylene changed in the series AuNi0.5/SiO2 > AgNi0.5/SiO2 > CuNi0.125/SiO2. Moreover, deactivation of the catalysts was mainly due to coking rather than to a particle size changes. The properties of the NiAg catalyst depended on the depth of penetration of metal species into the matrix, which was definitely a very interesting and unusual marker.[223] Thus, an Al2O3-supported complex Ni – Ag@C/Al2O3 catalyst on, consisting of carbon-encapsulated Ag NPs and non-encapsulated Ni NPs, was synthesized. Both types of metal NPs were located on the Al2O3 support. This was possible due to the catalyst synthesis strategy, which involved the co-precipitation of AgNO3 and Ni(NO3)2 · 6 H2O onto Al2O3, followed by the transformation of salts into carbonates and calcination in an acetylene atmosphere at 120°C to obtain encapsulated silver NPs (Fig. 22). The last step of reduction of NiO – Ag@C/Al2O3 with hydrogen at 500°C provided the formation of Ni NPs on the support. After testing, it was found that the Ni – Ag@C/Al2O3 catalyst provided ethylene selectivity of up to 81% with complete acetylene conversion within 30 h (Fig. 23). Interestingly, the use of the monometallic catalyst (Ni/Al2O3), gave virtually no target product (yield 6%), and the non-encapsulated Ni – Ag/Al2O3 catalyst showed only 54% conversion of acetylene. The improved selectivity of the catalyst was achieved due to the ability of carbon particles to provide hydrogen migration across the surface of the Al support of the catalyst. At the same time, the carbon shell blocked the formation of oligomers (i.e. the chain growth) on Ag NPs.

Fig. 22
Schematic illustration of the synthesis of Ni – Ag@C/Al2O3. Reproduced from Ref. 223 with permission from Elsevier.
Fig. 23
Temperature dependence of acetylene conversion (a) and ethylene selectivity (b) in acetylene hydrogenation over Ag/Al2O3, Ni/Al2O3, Ni-Ag/Al2O3 and Ni-Ag@C/Al2O3 catalysts. Reproduced from Ref. 223 with permission from Elsevier.

Intermetallic NiSb catalyst was used for hydrogenation of acetylene in the presence of ethylene.[224] This catalyst was prepared using the strategy of entrapping molten antimony with nickel (Fig. 24). The authors attributed the impressive results (selectivity up to 93.2% with complete acetylene conversion) not only to the weak adsorption of ethylene and strong adsorption of acetylene on the catalyst surface. The electronegative properties of the ‘guest’ metal (Sb) were used to control the properties of the ‘host’ metal (Ni), so that Ni1Sb2 trimers enclosed in NiSb intermetallic particles were formed on the surface. According to theoretical calculations, these regions provided the selective adsorption of acetylene.

Fig. 24
Schematic illustration of the synthesis of a complicated trimeric catalyst Ni1Sb2 in the intermetallic compound NiSb. Reproduced from Ref. 224 under a Creative Commons CC BY license.

The carbide-type Ni3ZnC0.7/C catalyst was obtained from the NiZn-ZIF-8/LDH precursor (ZIF is zeolite imidazolate framework) and showed 85% selectivity to ethylene at 100% conversion.[225] The Bader atomic charge distribution analysis demonstrated that the subsurface carbon divided the nickel atoms on the surface into two types: electron-depleted with the subsurface carbon atom and electron-rich with the subsurface zinc atom. According to calculations, the nickel atoms arranged into Ni3 assemblies promoted ethylene desorption and prevented hydrocarbon chain growth, which ultimately led to an increase in the process selectivity.

The bimetallic NiZn/ZnO catalyst performed better in acetylene semi-hydrogenation in comparison with the intermetallic catalyst Ni3Zn/ZnO of the ligand type. According to research and DFT calculations, the high selectivity of NiZn/ZnO was due to the non-competitive adsorption of acetylene by the adjacent isolated metal atoms.[226]

Features of the influence of the substrate shape on the properties of metal catalyst particles in acetylene hydrogenation were discovered in the Al2O3-supported NiCu NPs.[227] It was found that the support atoms govern the arrangement of metal atoms. The spindle-shaped support promoted the formation of NiCu alloys with a predominance of copper on the surface. And the flaky shape led to a switch to monometallic copper and nickel-rich bimetallic systems. The rod-shaped support promoted the appearance of Ni1Cu1 alloys, and the coordinatively unsaturated Al3+ sites played a key role in the formation of metal NPs, in which nickel atoms were orderly distributed among copper atoms. Uniformly distributed Ni – Cu alloy species with individual nickel sites provided 86% selectivity to ethylene with full conversion at 130°C.

According to calculations, in selective acetylene hydrogenation on intermetallic NiIn catalysts of various compositions, oligomerization products originate exactly from acetylene, not from ethylene.[228] This important conclusion provides deeper insight into the nature of the process and once again emphasizes the need to convert acetylene into ethylene. The calculations also showed that the indium additive to the catalyst improves the selectivity of the process.

NiGa intermetallics were prepared with the aim to elucidate the role of completely isolated nickel sites in the semi-hydrogenation of acetylene.[229] [230] It was found out, that isolated nickel sites preferentially π-adsorbed acetylene and promoted ethylene desorption. Three prepared catalysts based on Ni, Ni5Ga3 and NiGa had similar particle sizes, but different atomic arrangements, as well as different environment of nickel sites. As can be seen from Fig. 25a, Ni and Ni5Ga3 demonstrated higher activity and lower selectivity to ethylene especially at temperatures above 100°C. This suggests that the complete isolation of nickel particles on the surface led to an increase in selectivity. Temperature desorption profiles (Fig. 25b) clearly indicated three peaks for Ni and Ni5Ga3 catalysts. The first two peaks might be attributed to weakly adsorbed ethylene, and the third peak at a higher temperature was a signal of strongly adsorbed acetylene. The absence of a high-temperature peak was observed for the NiGa catalyst with completely isolated nickel sites. Therefore, it can be concluded that isolated nickel sites promoted acetylene desorption. Analysis of acetylene conversion shows that in excess ethylene (Fig. 25c), the catalyst with isolated nickel sites provided higher selectivity to ethylene. It is also evident from Fig. 25d, that the addition of gallium not only increased stability of the catalyst, but also inhibited the formation of ‘green oil’, i.e. hydrocarbons with longer chains (Fig. 25e). In another study,[231] on the properties of multi-intermetallic NiGa with isolated nickel particles, a whole series of additives (Fe, Cu, Ge) were proposed to stabilize the nickel particles on the surface.

Fig. 25
Temperature dependence of acetylene conversion and ethylene selectivity in the absence of ethylene (a); TPD-profiles for acetylene (top) and ethylene (bottom) (b); temperature dependence of acetylene conversion and ethylene selectivity in the presence of ethylene (c); time dependence of acetylene conversion on stream at 190°C (d); thermogravimetric curves of Ni catalysts used: Ni, Ni5Ga3 and NiGa (e). Reproduced from Ref. 229 with permission from Wiley.

A series of nickel-based catalysts was synthesized to elucidate the effect of various additives on the selective hydrogenation of acetylene.[232] It was found that the introduction of Mg results in a complete hydrogenation of acetylene to ethane, while the addition of tin improves selectivity, making it possible to stop hydrogenation at the step of ethylene formation. The authors associated these patterns with the facilitation of ethylene desorption from the surface of Ni7Sn/SiO2, which performed best: 65% yield of ethylene at 250°C. Mg – Al-supported Ni/MgAl2O4 catalyst provided 99% acetylene conversion with an selectivity to ethylene of more than 80%.[233] The authors attributed such outstanding results to the large surface area (309 m2 g–1) at the «tuned’ acidity.

Thus, the conversion and selectivity to ethylene in the acetylene hydrogenation are directly dependent on the size of the nickel species on the substrate surface. Single-atom catalysts demonstrate higher selectivity compared to nano- or microparticles. The catalytic properties of nickel catalysts are significantly affected by the nature of the support, which can either donate electrons to the metal atom or deplete it of electrons. If the support contains more electronegative atoms relative to nickel, the selectivity of such catalysts is superior to those with a support with less electronegative atoms. Lower electron density on the metal atom leads to easier desorption of ethylene and higher hydrogenation selectivity. The presence of a ‘neighbour’ element also has a significant effect on the reaction parameters. Firstly, the second component of the nickel alloy allows for a better distribution of particles over the surface due to greater dilution. Secondly, the nickel atom can transfer part of its electron density to ‘neighbour’ atoms (the number of which may be greater), which prevents the appearance of ‘hot spots’ and ultimately increases the selectivity of the process as a whole.

6. Cu-based catalysts for selective acetylene hydrogenation

Similar to Pd- and Ni-based catalysts, the catalytic activity of Cu catalysts depends on the particle size. It was found that the copper particle size played a crucial role in many catalytic transformations.[234] In order to find out how the properties of a copper catalyst are changed in the selective hydrogenation of acetylene, the single-atom catalyst and copper NPs of different sizes (3.4, 7.3, 9.3 nm) were obtained with their subsequent deposition onto aluminium oxide surface. It was found, that with a decrease in the particle size, the catalyst activity decreased. At the same time, selectivity and stability are significantly increased. In particular, single-atom copper catalysts showed 91% selectivity to ethylene with complete conversion and stability for at least 40 hours on stream. The difference in the properties of catalysts with varying particle sizes can be explained by different types of sorption — desorption of ethylene/acetylene (Fig. 26). According to the temperature profiles, acetylene desorption is observed at a lower temperature for the catalyst with a smaller particle size (89°C vs. 105°C). Therefore, the weak binding of acetylene on catalysts with a smaller particle size leads to lower sorption and lower concentration and hence to lower activity. A similar trend is observed for ethylene. In the case of ethylene, this leads to the less sorption of ethylene by the catalyst and, as a result, to less hydrogenation. Therefore, the selectivity of the process as a whole increases.

Fig. 26
TPD profiles of C2H2 (top) and C2H4 (bottom) on 0.5 Cu/Al2O3 and 2.0 Cu/Al2O3 (The figures indicate the mass weight of metal in the catalysts). Reproduced from Ref. 234 with permission from the American Chemical Society.

The bimetallic catalyst, Zn – CuxC was obtained by co-precipitation of the appropriate starting reagents. The addition of Zn contributed to a decrease in the size and an increase in the number of CuxC crystals.[235] In turn, it caused an increase in the activity of the catalyst in transformation of acetylene to ethylene, and the catalyst per se had excellent stability (> 140 h on stream) at 100°C even in the presence of a large excess of ethylene.

A composite catalyst based on copper and iron was prepared to study the mutual influence of the Fe support atoms on copper atoms.[236] Varying the ratio of elements in the catalyst, the authors managed to achieve 95% selectivity to ethylene with full conversion using the CuFe0.16MgOx catalyst. Based on the analytical data and calculations, the formation of the Cuδ–Feδ+0.16MgOx complex was observed, which played a crucial role in the activation of acetylene and hydrogen. This complex of a certain composition was bonded to both other copper and other iron atoms (with those that were not part of the complex). According to the data obtained, the desorption of ethylene on this complex was more preferable than its further hydrogenation. As the amount of iron in the catalyst increased, its catalytic activity expectedly increased, since the number of complexes and, accordingly, the routes for hydrogen dissociation increased. However, at the atomic ratio Cu/Fe > 0.16, i.e. increasing the proportion of iron, the activity of the catalyst decreased, since the iron particles covered the copper particles. Thus, the activation of acetylene and hydrogen occurred on the copper sites of the catalyst, while the formation of the C – C bond occurred on the Cu/Fe complex, with one carbon atom being bound to a copper atom and the other to an iron atom (Fig. 27).

Fig. 27
The proposed mechanism of acetylene semi-hydrogenation on the Cu/Fe0.16MgOx catalyst. Reproduced from Ref. 236 with permission from the American Chemical Society.

Copper atoms were immobilized on a defective nanodiamond–graphene and N-doped nanodiamond graphene to reveal the influence of the coordinated environment (Cu – C or Cu – N bonds) on the catalytic activity in acetylene hydrogenation.[237] It was found, that more oxidized copper atoms with Cu – C bonds showed high selectivity (up to 93%) with quantitative conversion and high stability (> 100 h on stream). According to the DFT calculations and isotope exchange experiments, the presence of a copper atom bound to carbon promoted the binding of hydrogen for subsequent dissociation, which was the rate-limiting step of the reaction. The coordination of three carbon atoms near the copper atom led to rapid charge consumption on the copper atoms, which in turn boosted the catalyst activity and, at the same time, increased its resistance to metal leaching from the surface.

Significant progress has been achieved in reduction of long-chain hydrocarbons during the selective hydrogenation of acetylene.[238] For this purpose, a copper catalyst was modified with a hydrophobic C16 chain, which was grafted to the metal surface using hexadecyltrimethoxysilane. The similar polarity of the alkyl chain of the catalyst and acetylene facilitated the sorption of the substrate on the catalyst surface, thus ultimately increasing the activity of the catalyst. In addition, the presence of a long hydrocarbon chain on the surface led to steric hindrance on the surface and challenges in the formation of hydrocarbons with a longer chain (C4+ hydrocarbons), which, in turn, significantly increased the catalyst lifetime. This modification provided an access to a catalyst capable of semi-hydrogenating acetylene with a selectivity of up to 84% at 140°C for 120 h.

The selective hydrogenation of acetylene on copper catalysts produces copper carbide CuC2, as well as a number of the other carbides (including mixed ones).[239] The catalyst was prepared from copper hydroxide with subsequent exposing to acetylene and reduction with hydrogen (Fig. 28). The resulting catalyst demonstrated moderate 15% selectivity and lower; however, the very fact of the formation of copper carbides is very interesting. Since the presence of copper acetylide in the solid phase was unambiguously proved, this compound turned out to be a precursor of the true catalyst or an intermediate in the hydrogenation reaction. On the one hand, the formation of copper carbide is a reversible process, and, probably, it is through the carbide that acetylene is hydrogenated. On the other hand, copper carbide transforms easily into more complicated carbides when heated, which was an irreversible process and could actually be considered as coking of the metal. Copper carbide immersed in a carbon matrix was also found in the catalysts comprising basic copper carbonate.[240] It was found that copper carbides acted as catalytic sites for the dissociation of hydrogen molecules, and the carbon matrix created steric hindrances for further growth of the hydrocarbon chain, which, accordingly, prevented undesirable oligomerization of acetylene. Copper carbides can also be formed by adding the other metals to the catalyst.[235] For example, the addition of zinc promoted the formation of copper carbides, and at the same time the size of the carbide particles was significantly smaller than in the synthesis of copper carbide without the addition of salts of other metals. In the same work, it was established that the addition of zinc does not affect the electronic structure of the carbides.

Fig. 28
The synthesis of a carbide CuCx catalyst. Reproduced from Ref. 239 under a CC-BY-NC-ND license.

According to experimental data and DFT calculations, the activation barrier for complete hydrogenation of acetylene on a single-atom copper catalyst was higher than on a copper nanocatalyst.[241] In addition, it was found that ethylene sorption was significantly more complicated on a single-atom catalyst, which led to inhibition of over-hydrogenation.

To prevent the ‘hot spot’ formation and over-hydrogenation, an electrochemical approach carried out at room temperature was proposed.[242] Hydrogen generated in situ during water electrolysis was used as the hydrogenating agent. Copper dendrites, produced electrochemically (electrodeposition from a Cu(NO3)2 solution provided an extremely high selectivity of 97% with an acetylene content of only 4 ppm in the final mixture, which corresponds to industrial standards for ethylene purity for further polymerization. These results were achieved using a flow electrochemical cell with a large electrode surface area (Fig. 29).

Fig. 29
Scheme of electrocatalytic purification of ethylene from acetylene. Reproduced from Ref. 242 with permission from Springer Nature.

Another approach for the electrochemical hydrogenation of acetylene to ethylene assumed the diffusion of acetylene through a carbon-based gas diffusion layer (GDL) and then through a microporous layer (MPL) directly to the surface of the catalyst.[243] Thus, the reaction interface was a gas/electrolyte/catalyst system (Fig. 30). The process occurred at room temperature and was completed with 99.9% conversion of acetylene with 90.1% ethylene selectivity. These data allow considering electro­chemical approach involving the use of hydrogen generated in situ by water electrolysis as an alternative to the standard thermal method. Yet another example of electrochemical hydrogenation of acetylene was also demonstrated using a Cu catalyst, which was copper nanodots.[244] In this case, the acetylene content in the final mixture was less than 1 ppm.

Fig. 30
Schematic representation of cathode for acetylene hydrogenation. Reproduced from Ref. 243 with permission from Springer Nature.

A mixed catalyst based on copper and nickel with an aluminosilicate zeolite ZSM-12 support was prepared.[245] During catalytic tests, it was shown that the catalyst can completely convert acetylene with 82.5% selectivity to ethylene. Among the CuNix/ZSM-12 catalysts (x = 5, 7, 9, 11), CuNi7/ZSM-12 performed best. According to the XPS data, the Ni2p peaks of the Ni/ZSM-12 and CuNi7/ZSM-12 catalysts correspond to three components: the satellite peak of Ni, Ni2+, and Ni0 (Fig. 31a). The signal intensity in the spectra of Ni/ZSM-12 was much higher than that of CuNi7/ZSM-12. All the three peaks have a slight shift towards an increase in the binding energy after the formation of CuNi7 particles, which indicates a strong interaction of copper and nickel particles, expressed in the transfer of electrons. a red shift after the formation of CuNi7 particles, which indicated a strong interaction of copper and nickel particles, expressed in electron transfer. The Cu2p signals in Cu/ZSM-12 and CuNi7/ZSM-12 demonstrated binding energies of 935 and 955 eV, which corresponded to the Cu2p3/2 and Cu2p1/2 states, respectively (see Fig. 31b). When the support was replaced by ZrO2, the hydrogenation selectivity increased to 93% also at full conversion of the substrate.[246] The authors have proved the formation of CuNi alloy particles and, using computational methods and analytical studies, explained the improved results by easier adsorption of acetylene and key intermediates.

Fig. 31
XPS spectra of Ni2p (a) and Cu2p (b) in Cu/ZSM-12 and CuNi7/ZSM-12 catalysts. Reproduced from Ref. 245 under a Creative Commons Attribution (CC BY) license.

A series of copper and iron-based kaolin-supported catalysts were synthesized for testing in the semi-hydrogenation of acetylene.[247] The Cu3Fe0.5/kaolin mixture performed best, providing 95% selectivity at 93% conversion. The authors explained this result by the shift of the characteristic Cu peak in the XPS spectra to a lower energy level caused by changes in the electronic structure of the Cu3Fe0.5/kaolin catalyst, as well as the interaction of copper with Fe and support atoms (Al – Si). In addition, electrons from the partially occupied 4s orbital of copper were able to transfer to the partially occupied 3d orbital of iron atoms, which led to hybridization of orbitals in the alloy and a change in the electronic structure of active sites. Under hybridization, the Cu2p3/2 signals in the alloy spectra were slightly shifted to lower energies.

Thus, single-atom copper catalysts show increased selectivity in comparison with copper NPs, which was shown by the authors of the original works using particles of different sizes as an example. The interaction of copper with support atoms and adjacent atoms in the case of alloy catalysts can also be noted. Separately, it is worth mentioning the identified phases of copper carbides formed during the immersion of metal atoms into a carbonaceous support Apparently, the formation of carbides is characteristic of most metals. The role of carbides in the hydrogenation of acetylene has not yet been reliably identified, and it remains to be established how the transition of active forms to carbide (or vice versa) occurs. Using a copper catalyst as an example, it was shown that grafting a long-chain aliphatic substituent to the metal decreases the amount of acetylene oligomerization products, which can most likely be used in the case of other metals.

7. Other metal catalysts in semi-hydrogenation of acetylene

The common route for removing acetylene from ethylene involves preliminary hydrogen production. An original approach based on the simultaneous production of hydrogen and semi-hydrogenation of acetylene on a Au/α-MoC catalyst was proposed.[55] Hydrogen was obtained by the reaction of carbon monoxide with water. The excess oxygen was removed with CO, which was further oxidized to CO2, and the released hydrogen was immediately consumed in the hydrogenation reaction. To clarify the mechanism of the process, reagent mixtures with labelled atoms were used: CO + D2O and 13CO + H2O. Analysis of the kinetic isotope effect of the H/D exchange processes and 12C/13C in the formation of ethylene revealed the ratios of the formation rates (r):

r(H)/r(D) = 2.59,

r(CO)/r(13CO) = 1.14.

In fact, this meant that both the transfer of surface hydrogen to acetylene and the reaction of adsorbed CO with surface oxygen can contribute to the reaction rate. Based on these data, a reaction model was constructed, involving dissociation of water to give hydroxyls, the transfer of hydrogen to acetylene, and the reaction of CO with oxygen with subsequent desorption of CO2. According to this model, the equilibrium constants of adsorption were determined, which demonstrated a strong competition of acetylene adsorption with the other reagents on the catalyst surface. The authors concluded, that the reaction of the three compounds (CO + H2O + C2H2) occurred by removing of boundary oxygen thereby exposing molybdenum sites for acetylene adsorption and water dissociation to hydroxy anions. Then, hydrogen was transferred from hydroxyls to acetylene (formation of ethylene) with the removal of residual oxygen by carbon monoxide. The best results that were achieved in this process were high acetylene conversion (99%) and very promising ethylene selectivity (83%) at low temperature (80°C). The authors indicated, that the proposed path might be an alternative to traditional energy-intensive hydrogenation processes involving molecular hydrogen.

Alumina-supported gold NPs were used as a catalyst in the selective semi-hydrogenation of acetylene to ethylene.[248] The main idea of the work was to decorate the support with carbon atoms (CA) or carbon and nitrogen atoms (CNA), which were obtained by carbonization of the appropriate precursors on the support surface. The catalysts with doped support demonstrated 5 times higher catalytic activity in the catalytic tests (FIg. 32). According to XPS data, the electron transfer between the Au atom and the C and N atoms can play a crucial role in the activity of gold catalytic sites.

FIg. 32
Acetylene conversion and ethylene selectivity as a function of hydrogenation temperature (a); TOF values for the Au catalysts on various supports at 250°C (b). Reproduced from Ref. 248 with permission from Springer Nature.

Etching the bimetallic AuCu catalyst with acetic acid generates Cu vacancies on the abrupt catalyst surface, which are suitable for alkyne hydrogenation and, in combination with electron-deficient Au atoms, contribute to the increased catalytic properties.[249] The activity of etched Au3Cu1/TiO2 catalysts was 4.6 times higher than that of untreated catalysts, and the ethylene selectivity was 94.0%.

ZnO with the disc-like (ZnO – D) and flower-like (ZnO – F) morphology were prepared hydrothermally from Zn(NO3)2 · 6 H2O and used as a support for Au/ZnO nanocatalysts.[250] Subsequent treatment of the catalysts with a gas mixture of 10 vol.% O2/Ar furnished Au/ZnO – F – O, Au/ZnO – D – O and their reduced derivatives, Au/ZnO – F – H and Au/ZnO – D – H, when treating the starting catalysts with a mixture of 10 vol.% H2/Ar (Fig. 33). The catalytic activity in the semi-hydrogenation of acetylene was directly dependent on the Au3+ content in the catalyst and decreased in the series Au/ZnO – F > Au/ZnO – F – H > Au/ZnO – D > Au/ZnO – D – H. The authors explained this effect by the ability of Au3+ to take electrons from acetylene into vacant 5d orbitals, providing stronger acetylene adsorption than Au+ and Au0. The best acetylene conversion (~ 58%) and ethylene selectivity (76%) were obtained in the presence of Au/ZnO – F (Fig. 34).

Fig. 33
SEM images of Au/ZnO – F – O (a1, a2), Au/ZnO – F – H (b1, b2), Au/ZnO – D – O (c1, c2), Au/ZnO – D – H (d1, d2). Reproduced from Ref. 250 with permission from the American Chemical Society.
Fig. 34
Dependence of conversion/selectivity of acetylene semi-hydrogenation on the chemical state of Au species. Reproduced from Ref. 250 with permission from the American Chemical Society

The core-shell type composite Au@Pt NPs were synthesized by impregnation and then investigated as catalysts for selective semi-hydrogenation of acetylene.[251] An imidazole framework was used as a catalyst support. The synthetic procedure involved growing a platinum shell on gold nanorods, followed by oxidative etching of the resulting hollow nanocomposite. At the second step, the resulting composite particles were encapsulated into the substrate. The Au@Pt-NTs@ZIF-67 catalyst was the most active in the process and provided the 49.3% yield of ethylene at the acetylene conversion of 69.1%. Probably, high activity of such catalysts was due to their tubular structure with a high content of active sites on the catalyst surface. The existence of highly active platinum particles around gold particles caused a synergistic effect, which provided conversion of acetylene into ethylene. After the catalytic process, the composite nanotubes retained their hollow tubular structure, i.e. the morphology of the particles remained unchanged, opening up abilities for the multiple use of the catalysts.

The bimetallic Pt3Fe nanocatalyst, individual particles of which were encapsulated in a silicon shell, were also studied in the hydrogenation of acetylene to ethylene. High-temperature (773 – 973 K) hydrogen treatment of the Pt3Fe surface was carried out, which changed the structure of the catalytic species: disordered Pt species (A1 plane) was formed by ordered Pt3 assemblies surrounded by Fe atoms (L12 plane) (Fig. 35).[252] The structure of this ensemble structure provided charge transfer from Fe atoms to neighbouring Pt atoms in the catalytic species, facilitating ethylene desorption from it and, accordingly, increasing and maintaining the selectivity of acetylene semi-hydrogenation at 83%.

Fig. 35
Geometry of the arrangement of Pt and Fe atoms in Pt3Fe nanoparticles. Configurations of atoms in the dominant planes A1 (a) and L12 (b). Reproduced from Ref. 252 with permission from Elsevier.

A systematic and detailed study was carried out for a pair of platinum catalysts PtCu and PtCo.[253] The first important observation concerned the formation of alloy particles. The authors showed that the alloys are formed not only in the course of hydrogenation of metals from their oxides during synthesis of the catalysts, but also in the course of hydrogenation of acetylene. It was also found that larger platinum particles in Co catalysts promoted a higher hydrogenation reaction with increasing temperature thereby improving the yield of ethane. In contrast, in copper catalysts, a decrease in the platinum particle size led to not only to an improved yield of ethylene, but also to losses in the reaction rate. According to the X-ray absorption spectra of the near-threshold fine structure, the threshold energies for PtCu catalysts were 11 566.8 eV, and 11 563.8 eV for PtCo. Accordingly, higher energy implied higher density of unoccupied d-orbitals of Pt, which was a key indicator of improved activity in the semi-hydrogenation of acetylene (87.6% on Cu catalysts vs. 62.9% on Co catalysts). DFT calculations also showed that the redistribution of electron density towards platinum in the PtCu catalyst was more pronounced than in the PtCo catalyst. platinum atoms paired with copper atoms contained more electrons than those paired with cobalt atoms. This redistribution results in weaker π-bonding of ethylene on the surface and, accordingly, inhibition of complete hydrogenation. The authors also pointed out, that in addition to obvious Pt – O and Pt – Cu/Co bonds, Pt – O – Cu/Co bonds were formed.[254] In fact, it’s about unreduced metal oxides that were used as precursors of metal particles on the catalyst surface. On the one hand, the presence of such particles is obvious. On the other hand, their role and their contribution to the reaction are rarely assessed. It was found that oxygen atoms repel platinum atoms from each other, which furnish Pt – O – Cu and Pt – O – Co atomic structures. Under hydrogenation conditions, the coordination number of the Pt – Co alloy was 4.35, and that of the Pt – Cu alloy was 2.21. At the same time, the coordination numbers of Pt – O – CuOa(a < 1) bonds was 7.45, which is 6.48 times greater than that of the Pt – CoO bond (1.15). Another issue was in the absence of the Cu – Cu bond in the Cu catalyst. All these issues provided weak binding of ethylene to the catalyst surface, which in turn decreases the rate of complete hydrogenation.

A series of complicated catalysts containing copper, platinum and cerium simultaneously were prepared for the selective hydrogenation of acetylene.[59] The initially prepared, PtCe alloy was deposited on a support consisting of various zeolites (Y, ZY), as well as on a copper-modified zeolite support to obtain PtCe/Cu–ZY. In addition, the PtCe alloy was deposited on a support of copper-modified oxalate and borohydride ligands (PtCe/CuX – ZY and PtCe/CuB–ZY, respectively). According to experimental data, the ethylene yield was 55.3% using PtCe/ZY catalysts at 120°C; and 71.4% in the presence of PtCe/Cu – ZY at 180°C. At the same time, in the presence of PtCe/CuX – ZY and PtCe/CuB – ZY catalysts, the yields were increased to 81.9 and 92.1% at 180 and 160°C, respectively. Both results demonstrated the positive effect of copper on the catalyst properties.

A silver carbide-like (AgxC) catalyst was synthesized by reduction of silver acetylide to AgxC under H2 atmosphere.[255] This catalyst demonstrated promising catalytic activity in the conversion of acetylene to ethylene. The acetylene conversion was 100%, the ethylene selectivity was 88.7%, and the catalyst demonstrated good stability under continuous operation for at least 30 h. High performance was due to the ability of the AgxC phase to significantly reduce the activation barrier of H2 dissociation and easy desorption of ethylene, which prevented overhydrogenation.

It is known, that supported rhodium catalysts are inefficient in semi-hydrogenation reactions. The addition of electron-donating acetylene ligands to atomically dispersed rhodium supported on zeolite provided high selectivity to ethylene (> 90%) in the acetylene semi-hydrogenation at 373 K in the presence of a large excess of ethylene.[256]

8. Conclusion

Selective hydrogenation of acetylene continues to be a relevant scientific issue. Obviously, the industrial process of ethylene purification from acetylene impurities stimulates research in this area, but it can be noted that the fundamental interest in this reaction is also extremely high. This is largely due to the need to gain deeper insight into the nature of catalysis, and the reaction of selective hydrogenation of acetylene is a kind of ‘textbook’ illustrating the selectivity patterns. Another boost for further research in this area was the development of complex catalysts and the discovery of new possibilities for their synthesis. The state-of-the-art experimental methods allow not only to reduce the size of metal particles to the nanoscale, but also provide uniform distribution of particles over the surface with atomic precision. Tuning the structure of complex species opens an access to single-atom catalysts, as well as ensembles of two or more elements of a certain composition and structure. For example, single atoms of one element can be placed in the corners of nanocubic particles of a second. Much of this precision is due not to the procedures per se, but rather to improvements in analytical techniques and the associated ability to detect the smallest details of the atomic world.

The highest efficiency in semi-hydrogenation of acetylene is demonstrated by catalysts based on palladium, copper, nickel, silver and their alloys. The general patterns of their action consist of an increase in the selectivity index for ethylene and a simultaneous decrease in the degree of acetylene conversion as the size of the metal particles decreases. Due to the small size of the metal particles and their distance from each other, the addition of hydrogen to acetylene occurs on the surface of the catalyst, and not in its bulk. ‘Dilution’ of the active metal with a ‘guest’ element also improves selectivity by transferring excess heat from the catalytic centre to the ‘satellite’ element. Another method for increasing selectivity is to reduce the electron density on the active metal. This is achieved by introducing one (or more) components to form alloy particles, or by appropriately designing the support, the atoms of which can deplete the active metal with electron density. A decrease in the electron density at the catalytic centre leads to easier desorption of ethylene formed during hydrogenation, a shorter time of its coordination with the catalyst surface and, accordingly, complicates further conversion into ethane.

As the types and numbers of active species on the catalyst surface continue to increase, and due to the difficulties in interpreting the catalytic pathway, it is likely that new barriers in the analysis of big catalytic data sets will emerge in the future. From this point of view, research in the field of machine learning and the use of artificial intelligence, which allows for the automatic analysis of hundreds and thousands of spectra with automatic processing of the results obtained, seem promising. Today, machine learning is already actively used in the development of approaches to the selective hydrogenation of acetylene [257] and in the design of catalysts for this process,[179][180] in the development of nanocluster [258] and single-atom [259] catalysts, in the search for active catalytic centres,[260] as well as for the identifying of intermediate structures on the surface of catalysts.[261]

In conclusion of the review, the comparison of the catalytic characteristics of the above-discussed single-atom catalysts with other types of catalytic systems used in the selective hydrogenation reaction of acetylene is given in Table 2.[262-267]

Table 2
\[ \]
Comparison of single-atom catalysts with other types of catalytic systems used in semi-hydrogenation of acetylene.
(2)
Table 3
\[ \]
Table 2 (continued)
(3)

9. List of abbreviations and designations

Eads — adsorption energy,

Gads — adsorption free energy,

ΔH° — standard enthalpy (of reaction),

r — rate of formation,

Ac — acetone,

acac — acetylacetonate,

AC-HAADF-STEM — aberration-corrected high high-angle annular dark-field scanning transmission electron microscopy,

AIMD — initio ab initio molecular dynamics (method),

CN — nitrogen-doped carbon nanospheres obtained from covalent organic frameworks,

CNT — multi-walled carbon nanotubes,

COD — cycloocta-1,5-diene,

COF — covalent organic frameworks,

CTF — layered covalent triazine frameworks,

ESI-HRMS — high resolution mass spectrometry with electrospray ionization,

EDS — energy dispersive X-ray spectroscopy,

G — graphene,

GHSV — gas hourly space velocity,

GR — galvanic replacement (method),

gC3N4 — graphene-like carbon nitride,

GDL — carbon-based gas diffusion layer,

hNCNC — hierarchical carbon nanocells,

HR — high resolution,

HTS — hexadecyltrimethoxysilane,

HNT — halloysite nanotubes,

IL — ionic liquid,

ISA — isolated single-atom centres,

L — ligand,

LDH — layered double hydroxides,

MPL — microporous layer,

MPNC — mesoroporous N-doped carbon foam nanospheres,

NC — nanocubes,

ND — nanodiamonds,

NHC — N-heterocyclic carbene,

NMC — N-doped mesoroporous carbon,

NMP — N-methyl-2-pyrrolidon,

NP — nanoparticles,

NT — nanotubes,

Ndicy — dicyandiamide,

Pal — palygorskite,

PDO — 1,10-phenanthroline-5,6-dione,

Prmim — 1-methyl-3-n-propylimidazolium,

SA — single-atom (catalyst),

SAA — single-atom alloy,

SEM — scanning electron microscopy,

TEM — transmission electron microscopy,

TEOS — tetraethyl orthosilicate,

TOF — catalyst turnover frequency, which is a measure of catalytic activity and shows the number of substrate molecules transformed into product on one active centre of the catalyst per unit of time,

TPAOH — tetrapropylammonium hydrozide,

TPD — temperature programmed desorption,

XPS — X-ray photoelectron spectroscopy,

XRD — X-ray (powder) diffraction,

ZIF — zeolite imidazolate framework.

References

1.
Bond Formations between Two Nucleophiles: Transition Metal Catalyzed Oxidative Cross-Coupling Reactions
Liu C., Zhang H., Shi W., Lei A.
Chemical Reviews, American Chemical Society (ACS), 2011
2.
More than bystanders: the effect of olefins on transition-metal-catalyzed cross-coupling reactions.
4.
Atom-Efficient Metal-Catalyzed Cross-Coupling Reaction of Indium Organometallics with Organic Electrophiles
Pérez I., Sestelo J.P., Sarandeses L.A.
Journal of the American Chemical Society, American Chemical Society (ACS), 2001
5.
Chiral Tridentate Ligands in Transition Metal-Catalyzed Asymmetric Hydrogenation
Wang H., Wen J., Zhang X.
Chemical Reviews, American Chemical Society (ACS), 2021
6.
Transition Metal-Catalyzed Enantioselective Hydrogenation of Enamines and Imines
Xie J., Zhu S., Zhou Q.
Chemical Reviews, American Chemical Society (ACS), 2010
7.
Highlights of Transition Metal-Catalyzed Asymmetric Hydrogenation of Imines
Fleury-Brégeot N., de la Fuente V., Castillón S., Claver C.
ChemCatChem, Wiley, 2010
9.
Single-site catalyst promoters accelerate metal-catalyzed nitroarene hydrogenation
Wang L., Guan E., Zhang J., Yang J., Zhu Y., Han Y., Yang M., Cen C., Fu G., Gates B.C., Xiao F.
Nature Communications, Springer Nature, 2018
10.
Pd Metal Catalysts for Cross-Couplings and Related Reactions in the 21st Century: A Critical Review
Biffis A., Centomo P., Del Zotto A., Zecca M.
Chemical Reviews, American Chemical Society (ACS), 2018
12.
Recent advances in selective acetylene hydrogenation using palladium containing catalysts
McCue A.J., Anderson J.A.
Frontiers of Chemical Science and Engineering, Higher Education Press, 2015
13.
3d-Metal Catalyzed N- and C-Alkylation Reactions via Borrowing Hydrogen or Hydrogen Autotransfer
14.
Alcohols as Substrates in Transition-Metal-Catalyzed Arylation, Alkylation, and Related Reactions
15.
Transition-Metal-Catalyzed C–H Alkylation Using Alkenes
Dong Z., Ren Z., Thompson S.J., Xu Y., Dong G.
Chemical Reviews, American Chemical Society (ACS), 2017
16.
Transition-metal-catalyzed C–H bond alkylation using olefins: recent advances and mechanistic aspects
Mandal D., Roychowdhury S., Biswas J.P., Maiti S., Maiti D.
Chemical Society Reviews, Royal Society of Chemistry (RSC), 2022
18.
Technology Trends of Catalysts in Hydrogenation Reactions: A Patent Landscape Analysis
Stoffels M.A., Klauck F.J., Hamadi T., Glorius F., Leker J.
Advanced Synthesis and Catalysis, Wiley, 2020
19.
Metal–organic-framework-based catalysts for hydrogenation reactions
Chen Z., Chen J., Li Y.
Chinese Journal of Catalysis, Elsevier, 2017
20.
Application of supported metallic catalysts in catalytic hydrogenation of arenes
Qi S., Wei X., Zong Z., Wang Y.
RSC Advances, Royal Society of Chemistry (RSC), 2013
21.
The development of gold catalysts for use in hydrogenation reactions
Cárdenas-Lizana F., Keane M.A.
Journal of Materials Science, Springer Nature, 2012
22.
Model studies of heterogeneous catalytic hydrogenation reactions with gold
Pan M., Brush A.J., Pozun Z.D., Ham H.C., Yu W., Henkelman G., Hwang G.S., Mullins C.B.
Chemical Society Reviews, Royal Society of Chemistry (RSC), 2013
23.
Ceria-Based Catalysts for Selective Hydrogenation Reactions: A Critical Review
Razmgar K., Altarawneh M., Oluwoye I., Senanayake G.
Catalysis Surveys from Asia, Springer Nature, 2021
24.
Iron-Catalyzed Hydrogenation Reactions
Guo N., Zhu S.
Chinese Journal of Organic Chemistry, Shanghai Institute of Organic Chemistry, 2015
25.
Highly Selective Hydrogenation of C═C Bonds Catalyzed by a Rhodium Hydride
Gu Y., Norton J.R., Salahi F., Lisnyak V.G., Zhou Z., Snyder S.A.
Journal of the American Chemical Society, American Chemical Society (ACS), 2021
26.
The H2-Treated TiO2 Supported Pt Catalysts Prepared by Strong Electrostatic Adsorption for Liquid-Phase Selective Hydrogenation
Kuhaudomlap S., Mekasuwandumrong O., Praserthdam P., Fujita S., Arai M., Panpranot J.
Catalysts, MDPI, 2018
27.
Selective hydrogenation of biomass derived substrates using ionic liquid-stabilized ruthenium nanoparticles
Julis J., Hölscher M., Leitner W.
Green Chemistry, Royal Society of Chemistry (RSC), 2010
28.
Recent Advances of Calcium Carbide in Organic Reactions
Zeng F., Lv Q., Chen X., Yu B.
Current Chinese Chemistry, Bentham Science Publishers Ltd., 2020
30.
Budaev A.B., Ivanov A.V., Petrova O.V., Samsonov V.A., Ushakov I.A., Tikhonov A.Y., Sobenina L.N., Trofimov B.A.
Mendeleev Communications, OOO Zhurnal "Mendeleevskie Soobshcheniya", 2019
31.
Selective synthesis of 1-vinylpyrroles directly from ketones and acetylene: Modification of trofimov reaction
Mikhaleva A.I., Shmidt E.Y., Ivanov A.V., Vasil’tsov A.M., Senotrusova E.Y., Protsuk N.I.
Russian Journal of Organic Chemistry, Pleiades Publishing, 2007
32.
The (3+2)- and formal (3+3)-cycloadditions of N-vinylpyrroles with cyclic nitrones and C,N-cyclic azomethine imines
Afanaseva K.K., Efremova M.M., Kuznetsova S.V., Ivanov A.V., Starova G.L., Molchanov A.P.
Tetrahedron, Elsevier, 2018
33.
A Highly Efficient and Stereoselective Cycloaddition of Nitrones to N-Vinylpyrroles
Molchanov A., Savinkov R., Stepakov A., Starova G., Kostikov R., Barnakova V., Ivanov A.
Synthesis, Georg Thieme Verlag KG, 2014
34.
Mirovye moshchnosty po proizvodstvu polietilena znachitel’no vyrastut k 2028 godu. Sait Kompanii MRS (Market Report) [Global Polyethylene Production Capacity to Grow Significantly by 2028. MRS (Market Report)]; website; https://www.mrc.ru/news/413825-mirovie-moschnosti-po-proizvodstvu-polietilena-znachitelno-virastut-k-2028-godu(Last access 19.07.2025)
35.
Technologies for the synthesis of ethylene and propylene from natural gas
Treger Y.A., Rozanov V.N.
Review Journal of Chemistry, Pleiades Publishing, 2016
37.
An Overview of Light Olefins Production via Steam Enhanced Catalytic Cracking
Akah A., Williams J., Ghrami M.
Catalysis Surveys from Asia, Springer Nature, 2019
39.
Production of Acetylene from Viable Feedstock: Promising Recent Approaches
Gyrdymova Y., Lebedev A., Du Y., Rodygin K.
ChemPlusChem, Wiley, 2024
40.
How to Control the Selectivity of Palladium-based Catalysts in Hydrogenation Reactions: The Role of Subsurface Chemistry
Armbrüster M., Behrens M., Cinquini F., Föttinger K., Grin Y., Haghofer A., Klötzer B., Knop-Gericke A., Lorenz H., Ota A., Penner S., Prinz J., Rameshan C., Révay Z., Rosenthal D., et. al.
ChemCatChem, Wiley, 2012
41.
The Adsorptive Separation of Ethylene from C2 Hydrocarbons by Metal–Organic Frameworks
Xie W., Yang L., Zhang J., Zhao X.
Chemistry - A European Journal, Wiley, 2023
42.
New Paradigms in Porous Framework Materials for Acetylene Storage and Separation
Verma G., Ren J., Kumar S., Ma S.
European Journal of Inorganic Chemistry, Wiley, 2021
43.
Ionic liquids for acetylene and ethylene separation: Material selection and solubility investigation
Palgunadi J., Kim H.S., Lee J.M., Jung S.
Chemical Engineering and Processing: Process Intensification, Elsevier, 2010
44.
Efficient separation of ethylene from acetylene/ethylene mixtures by a flexible-robust metal–organic framework
Li L., Lin R., Krishna R., Wang X., Li B., Wu H., Li J., Zhou W., Chen B.
Journal of Materials Chemistry A, Royal Society of Chemistry (RSC), 2017
45.
Acetylene selective hydrogenation: a technical review on catalytic aspects
Takht Ravanchi M., Sahebdelfar S., Komeili S.
Reviews in Chemical Engineering, Walter de Gruyter, 2017
46.
Composition of the Green Oil in Hydrogenation of Acetylene over a Commercial Pd-Ag/Al2O3Catalyst
Zhang J., Sui Z., Zhu Y., Chen D., Zhou X., Yuan W.
Chemical Engineering and Technology, Wiley, 2016
47.
Hydrogenation of acetylene over supported metal catalysts. Part 1.—Adsorption of [14C]acetylene and [14C]ethylene on silica supported rhodium, iridium and palladium and alumina supported palladium
Al-Ammar A.S., Webb G.
Journal of the Chemical Society Faraday Transactions 1 Physical Chemistry in Condensed Phases, Royal Society of Chemistry (RSC), 1978
48.
Hydrogenation of acetylene over supported metal catalysts. Part 2.—[14C]tracer study of deactivation phenomena
Al-Ammar A.S., Webb G.
Journal of the Chemical Society Faraday Transactions 1 Physical Chemistry in Condensed Phases, Royal Society of Chemistry (RSC), 1978
49.
Green oil poisoning of a Pd/a1203 acetylene hydrogenation catalyst
Sárkány A., Weiss A.H., Szilágyi T., Sándor P., Guczi L.
Applied Catalysis, Elsevier, 1984
51.
Pd-Ag/Al2O3 catalyst: Stages of deactivation in tail-end acetylene selective hydrogenation
Takht Ravanchi M., Sahebdelfar S.
Applied Catalysis A: General, Elsevier, 2016
52.
A novel configuration for Pd/Ag/α-Al2O3 catalyst regeneration in the acetylene hydrogenation reactor of a multi feed cracker
Rahimpour M.R., Dehghani O., Gholipour M.R., Shokrollahi Yancheshmeh M.S., Seifzadeh Haghighi S., Shariati A.
Chemical Engineering Journal, Elsevier, 2012
53.
Metal single-atom catalysts for selective hydrogenation of unsaturated bonds
Sun Z., Wang S., Chen W.
Journal of Materials Chemistry A, Royal Society of Chemistry (RSC), 2021
54.
Single-atom catalysis for organic reactions
Hu H., Xi J.
Chinese Chemical Letters, Elsevier, 2023
55.
Acetylene hydrogenation to ethylene by water at low temperature on a Au/α-MoC catalyst
Huang R., Xia M., Zhang Y., Guan C., Wei Y., Jiang Z., Li M., Zhao B., Hou X., Wei Y., Chen Q., Hu J., Cui X., Yu L., Deng D., et. al.
Nature Catalysis, Springer Nature, 2023
56.
Step Site-Specific Semihydrogenation of Acetylene on the Au Surface
Wu Z., Wang F., Sun G., Xiong F., Teng B., Huang W.
Journal of Physical Chemistry Letters, American Chemical Society (ACS), 2022
57.
Enhancement of Alkyne Semi-Hydrogenation Selectivity by Electronic Modification of Platinum
Wang Z., Garg A., Wang L., He H., Dasgupta A., Zanchet D., Janik M.J., Rioux R.M., Román-Leshkov Y.
ACS Catalysis, American Chemical Society (ACS), 2020
58.
Pt-O4 moiety induced electron localization toward In2O-Triggered acetylene Semi-Hydrogenation
Li Y., Cao Y., Ge X., Zhang H., Yan K., Zhang J., Qian G., Jiang Z., Gong X., Li A., Zhou X., Yuan W., Duan X.
Journal of Catalysis, Elsevier, 2022
60.
Challenges and Opportunities for Exploiting the Role of Zeolite Confinements for the Selective Hydrogenation of Acetylene
Vito J., Shetty M.
ACS applied materials & interfaces, American Chemical Society (ACS), 2023
61.
Catalysts for selective hydrogenation of acetylene: A review
Xie K., Xu K., Liu M., Song X., Xu S., Si H.
Materials Today Catalysis, Elsevier, 2023
64.
Progress in single-atom methodology in modern catalysis
Mashkovsky Igor S., Markov Pavel V., Rassolov Alexander V., Patil Ekaterina D., Stakheev Alexander Yu.
Russian Chemical Reviews, Autonomous Non-profit Organization Editorial Board of the journal Uspekhi Khimii, 2023
66.
67.
Nanoparticles as Recyclable Catalysts: The Frontier between Homogeneous and Heterogeneous Catalysis
Astruc D., Lu F., Aranzaes J.R.
Angewandte Chemie - International Edition, Wiley, 2005
68.
Palladium-Catalyzed Ligand-Directed C−H Functionalization Reactions
Lyons T.W., Sanford M.S.
Chemical Reviews, American Chemical Society (ACS), 2010
69.
Identification of Non-Precious Metal Alloy Catalysts for Selective Hydrogenation of Acetylene
Studt F., Abild-Pedersen F., Bligaard T., Sørensen R.Z., Christensen C.H., Nørskov J.K.
Science, American Association for the Advancement of Science (AAAS), 2008
71.
Visible Light Photoredox Catalysis with Transition Metal Complexes: Applications in Organic Synthesis
Prier C.K., Rankic D.A., MacMillan D.W.
Chemical Reviews, American Chemical Society (ACS), 2013
72.
Heck reaction as a sharpening stone of palladium catalysis
Beletskaya I.P., Cheprakov A.V.
Chemical Reviews, American Chemical Society (ACS), 2000
73.
Palladium(II)-Catalyzed CH Activation/CC Cross-Coupling Reactions: Versatility and Practicality
Chen X., Engle K., Wang D., Yu J.
Angewandte Chemie - International Edition, Wiley, 2009
74.
Gold-Catalyzed Organic Reactions
Hashmi A.S.
Chemical Reviews, American Chemical Society (ACS), 2007
75.
Palladium-Catalyzed Cross-Coupling Reactions in Total Synthesis
Nicolaou K.C., Bulger P.G., Sarlah D.
Angewandte Chemie - International Edition, Wiley, 2005
77.
Reaching For Meta Substitution
CARMEN DRAHL
Chemical & Engineering News, American Chemical Society (ACS), 2012
82.
Cu and Cu-Based Nanoparticles: Synthesis and Applications in Catalysis
Gawande M.B., Goswami A., Felpin F., Asefa T., Huang X., Silva R., Zou X., Zboril R., Varma R.S.
Chemical Reviews, American Chemical Society (ACS), 2016
83.
Isolated Metal Atom Geometries as a Strategy for Selective Heterogeneous Hydrogenations
Kyriakou G., Boucher M.B., Jewell A.D., Lewis E.A., Lawton T.J., Baber A.E., Tierney H.L., Flytzani-Stephanopoulos M., Sykes E.C.
Science, American Association for the Advancement of Science (AAAS), 2012
84.
Single-atom catalysts: a new frontier in heterogeneous catalysis.
Yang X., Wang A., Qiao B., Li J., Liu J., Zhang T.
Accounts of Chemical Research, American Chemical Society (ACS), 2013
85.
A stable single-site palladium catalyst for hydrogenations.
Vilé G., Albani D., Nachtegaal M., Chen Z., Dontsova D., Antonietti M., López N., Pérez-Ramírez J.
Angewandte Chemie - International Edition, Wiley, 2015
86.
Heterogeneous single-atom catalysis
Wang A., Li J., Zhang T.
Nature Reviews Chemistry, Springer Nature, 2018
87.
Leaching Mechanism of Different Palladium Surface Species in Heck Reactions of Aryl Bromides and Chlorides
Gnad C., Abram A., Urstöger A., Weigl F., Schuster M., Köhler K.
ACS Catalysis, American Chemical Society (ACS), 2020
89.
Insight into the Dynamic Evolution of Supported Metal Catalysts by In Situ/Operando Techniques and Theoretical Simulations
Nian Y., Huang X., Liu M., Zhang J., Han Y.
ACS Catalysis, American Chemical Society (ACS), 2023
91.
Transition metal “cocktail”-type catalysis
Prima D.O., Kulikovskaya N.S., Galushko A.S., Mironenko R.M., Ananikov V.P.
Current Opinion in Green and Sustainable Chemistry, Elsevier, 2021
93.
Evidence for “cocktail”-type catalysis in Buchwald–Hartwig reaction. A mechanistic study
Prima D.O., Madiyeva M., Burykina J.V., Minyaev M.E., Boiko D.A., Ananikov V.P.
Catalysis Science and Technology, Royal Society of Chemistry (RSC), 2021
95.
Pd and Pt Catalyst Poisoning in the Study of Reaction Mechanisms: What Does the Mercury Test Mean for Catalysis?
Chernyshev V.M., Astakhov A.V., Chikunov I.E., Tyurin R.V., Eremin D.B., Ranny G.S., Khrustalev V.N., Ananikov V.P.
ACS Catalysis, American Chemical Society (ACS), 2019
96.
Revealing the mechanism of combining best properties of homogeneous and heterogeneous catalysis in hybrid Pd/NHC systems
Prima D., Kulikovskaya N., Novikov R., Kostyukovich A., Burykina J., Chernyshev V., Ananikov V.P.
Angewandte Chemie - International Edition, Wiley, 2024
97.
4D Catalysis Concept Enabled by Multilevel Data Collection and Machine Learning Analysis
Galushko A.S., Ananikov V.P.
ACS Catalysis, American Chemical Society (ACS), 2023
98.
Nickel and Palladium Catalysis: Stronger Demand than Ever
Chernyshev V.M., Ananikov V.P.
ACS Catalysis, American Chemical Society (ACS), 2022
99.
Selective Hydrogenation over Supported Metal Catalysts: From Nanoparticles to Single Atoms
Zhang L., Zhou M., Wang A., Zhang T.
Chemical Reviews, American Chemical Society (ACS), 2019
100.
Catalysts and mechanisms for the selective heterogeneous hydrogenation of carbon-carbon triple bonds
101.
Investigating the Mechanism of Alkyne Hydrogenation through an Open-Ended, Inquiry-Based Undergraduate Research Project Exploring Heterogeneous Catalysis
Mirich A., Enmeier M., Cunningham K., Grossman K., Recker G., Jarman S., Weinmaster T., Mehaffey R., Huldin G., Bacchin G., Kallepalli S., Cogua L., Johnson L., Mattson B.
Journal of Chemical Education, American Chemical Society (ACS), 2020
102.
Kinetic regularities and mechanism of acetylene hydrogenation over PdMn catalyst
Stytsenko V.D., Melnikov D.P., Glotov A.P., Vinokurov V.A.
Molecular Catalysis, Elsevier, 2022
103.
Study on the reaction mechanism of acetylene selective hydrogenation catalysts Pd-Ag/Al2O3
Che C., Gou G., Wen H., Liang Y., Han W., Zhang F., Cai X.
Inorganic and Nano-Metal Chemistry, Taylor & Francis, 2020
105.
“Hidden” Nanoscale Catalysis in Alkyne Hydrogenation with Well-Defined Molecular Pd/NHC Complexes
Denisova E.A., Kostyukovich A.Y., Fakhrutdinov A.N., Korabelnikova V.A., Galushko A.S., Ananikov V.P.
ACS Catalysis, American Chemical Society (ACS), 2022
106.
Mechanistic analysis of transformative Pd/NHC catalyst evolution in the 1,2-diphenylacetylene semihydrogenation using molecular hydrogen
Saybulina E.R., Mironenko R.M., Galushko A.S., Ilyushenkova V.V., Izmailov R.R., Ananikov V.P.
Journal of Catalysis, Elsevier, 2024
107.
Recent Progress in Pd-Based Nanocatalysts for Selective Hydrogenation
Zhao X., Chang Y., Chen W., Wu Q., Pan X., Chen K., Weng B.
ACS Omega, American Chemical Society (ACS), 2021
108.
Recent research advances on catalysts for selective hydrogenation of ethyne
Guo J., Lei Y., Liu H., Li Y., Li D., He D.
Catalysis Science and Technology, Royal Society of Chemistry (RSC), 2023
110.
Intermetallic compounds in heterogeneous catalysis—a quickly developing field
Armbrüster M., Schlögl R., Grin Y.
Science and Technology of Advanced Materials, Taylor & Francis, 2014
112.
Reaction‐Induced Metal‐Metal Oxide Interactions in Pd‑In2O3/ZrO2 Catalysts Drive Selective and Stable CO2 Hydrogenation to Methanol
Pinheiro Araujo T., Morales-Vidal J., Giannakakis G., Mondelli C., Eliasson H., Erni R., Stewart J., Mitchell S., Lopez N., Pérez-Ramírez J.
Angewandte Chemie - International Edition, Wiley, 2023
113.
Selective hydrogenation of unsaturated vinyl ethers in two‑chamber gauge‐reactor
Gyrdymova Y., Saybulina E., Mironenko R., Rodygin K.
Asian Journal of Organic Chemistry, Wiley, 2024
114.
Pd-Catalyzed Hydrogenation of Unsaturated Terpenoids in Two-Chamber Reactors
Gyrdymova Y.V.
Russian Journal of General Chemistry, Pleiades Publishing, 2024
115.
Metal-Catalyzed Chemical Activation of Calcium Carbide: New Way to Hierarchical Metal/Alloy-on-Carbon Catalysts
Lebedev A.N., Rodygin K.S., Mironenko R.M., Saybulina E.R., Ananikov V.P.
Journal of Catalysis, Elsevier, 2022
116.
Dopamine polymer derived isolated single-atom site metals/N-doped porous carbon for benzene oxidation
Wu K., Zhan F., Tu R., Cheong W., Cheng Y., Zheng L., Yan W., Zhang Q., Chen Z., Chen C.
Chemical Communications, Royal Society of Chemistry (RSC), 2020
117.
Well-Defined Materials for Heterogeneous Catalysis: From Nanoparticles to Isolated Single-Atom Sites
Li Z., Ji S., Liu Y., Cao X., Tian S., Chen Y., Niu Z., Li Y.
Chemical Reviews, American Chemical Society (ACS), 2019
118.
A General Strategy for Fabricating Isolated Single Metal Atomic Site Catalysts in Y Zeolite
Liu Y., Li Z., Yu Q., Chen Y., Chai Z., Zhao G., Liu S., Cheong W., Pan Y., Zhang Q., Gu L., Zheng L., Wang Y., Lu Y., Wang D., et. al.
Journal of the American Chemical Society, American Chemical Society (ACS), 2019
119.
Selective acetylene hydrogenation over single metal atoms supported on Fe3O4(001): A first-principle study
Yuk S.F., Collinge G., Nguyen M., Lee M., Glezakou V., Rousseau R.
Journal of Chemical Physics, AIP Publishing, 2020
120.
The Progress and Outlook of Metal Single-Atom-Site Catalysis
Liang X., Fu N., Yao S., Li Z., Li Y.
Journal of the American Chemical Society, American Chemical Society (ACS), 2022
123.
The Roles of Subsurface Carbon and Hydrogen in Palladium-Catalyzed Alkyne Hydrogenation
Teschner D., Borsodi J., Wootsch A., Révay Z., Hävecker M., Knop-Gericke A., Jackson S.D., Schlögl R.
Science, American Association for the Advancement of Science (AAAS), 2008
126.
Direct observation of noble metal nanoparticles transforming to thermally stable single atoms
Wei S., Li A., Liu J., Li Z., Chen W., Gong Y., Zhang Q., Cheong W., Wang Y., Zheng L., Xiao H., Chen C., Wang D., Peng Q., Gu L., et. al.
Nature Nanotechnology, Springer Nature, 2018
127.
Mesoporous Nitrogen‐Doped Carbon‐Nanosphere‐Supported Isolated Single‐Atom Pd Catalyst for Highly Efficient Semihydrogenation of Acetylene
Feng Q., Zhao S., Xu Q., Chen W., Tian S., Wang Y., Yan W., Luo J., Wang D., Li Y.
Advanced Materials, Wiley, 2019
128.
Enhancing both selectivity and coking-resistance of a single-atom Pd1/C3N4 catalyst for acetylene hydrogenation
Huang X., Xia Y., Cao Y., Zheng X., Pan H., Zhu J., Ma C., Wang H., Li J., You R., Wei S., Huang W., Lu J.
Nano Research, Springer Nature, 2017
129.
Atomically Dispersed Pd-N1C3 Sites on a Nitrogen-Doped Carbon Nanosphere for Semi-hydrogenation of Acetylene
Wei S., Liu X., Wang C., Liu X., Zhang Q., Li Z.
ACS Nano, American Chemical Society (ACS), 2023
130.
Confinement of ultrasmall Pd nanoparticles by layered covalent triazine frameworks for semihydrogenation of acetylene
Zhang J., Zhang G., He L., Shi Y., Miao R., Zhu Y., Guan Q.
Applied Surface Science, Elsevier, 2021
131.
Highly Active Isolated Single-Atom Pd Catalyst Supported on Layered MgO for Semihydrogenation of Acetylene
Tao X., Nan B., Li Y., Du M., Guo L., Tian C., Jiang L., Shen L., Sun N., Li L.
ACS Applied Energy Materials, American Chemical Society (ACS), 2022
134.
Pd single-atom catalysts derived from strong metal-support interaction for selective hydrogenation of acetylene
135.
Selective hydrogenation of acetylene over Pd/CeO2
Li K., Lyu T., He J., Jang B.W.
Frontiers of Chemical Science and Engineering, Higher Education Press, 2020
136.
Ligand-coordination effects on the selective hydrogenation of acetylene in single-site Pd-ligand supported catalysts
Wasim E., Din N.U., Le D., Zhou X., Sterbinsky G.E., Pape M.S., Rahman T.S., Tait S.L.
Journal of Catalysis, Elsevier, 2022
139.
Impact of Pd single-site coordination structure on catalytic performance for semihydrogenation of acetylene
Zeng Y., Xia M., Gao F., Zhou C., Cheng X., Liu L., Jiao L., Wu Q., Wang X., Yang L., Fan Y., Hu Z.
Nano Research, Springer Nature, 2024
140.
Photo-thermo semi-hydrogenation of acetylene on Pd1/TiO2 single-atom catalyst
Guo Y., Huang Y., Zeng B., Han B., AKRI M., Shi M., Zhao Y., Li Q., Su Y., Li L., Jiang Q., Cui Y., Li L., Li R., Qiao B., et. al.
Nature Communications, Springer Nature, 2022
141.
Dual effect of  TiO 2  vacancy on Pd catalyst in acetylene hydrogenation: Boosting performance and capturing reaction heat
Yang Y., Yang J., Shi S., Luo Y., Xu X., Liu Y., Li D., Feng J.
AICHE Journal, Wiley, 2023
142.
Decoupling Active Sites Enables Low-Temperature Semihydrogenation of Acetylene
Li Z., Zhang J., Tian J., Feng K., Chen Y., Li X., Zhang Z., Qian S., Yang B., Su D., Luo K.H., Yan B.
ACS Catalysis, American Chemical Society (ACS), 2024
143.
Combined experimental and theoretical study of acetylene semi-hydrogenation over Pd/Al2O3
Gonçalves L.P., Wang J., Vinati S., Barborini E., Wei X., Heggen M., Franco M., Sousa J.P., Petrovykh D.Y., Soares O.S., Kovnir K., Akola J., Kolen'ko Y.V.
International Journal of Hydrogen Energy, Elsevier, 2020
144.
Anchoring Pd species over defective alumina to achieve high atomic utilization and tunable electronic structure for semi-hydrogenation of acetylene
Xu L., Hua S., Zhou J., Xu Y., Lu C., Feng F., Zhao J., Xu X., Wang Q., Zhang Q., Li X.
Applied Catalysis A: General, Elsevier, 2022
147.
Regulating metal-support interactions of Pd/MgAl2O4 for efficient selective hydrogenation of acetylene
148.
Structural and Kinetics Understanding of Support Effects in Pd-Catalyzed Semi-Hydrogenation of Acetylene
Cao Y., Ge X., Li Y., Si R., Sui Z., Zhou J., Duan X., Zhou X.
Engineering, Elsevier, 2021
149.
Kinetic Studies of Liquid Phase Hydrogenation of Acetylene for Ethylene Production Using a Selective Solvent over a Commercial Palladium/Alumina Catalyst
150.
Preparation of MoS2 nanosheets to support Pd species for selective steerable hydrogenation of acetylene
Guo Y., Xie J., Jia L., Shi Y., Zhang J., Chen Q., Guan Q.
Journal of Materials Science, Springer Nature, 2020
151.
Effective Inhibition of Ethane Generation on Fe5C2 Nanoparticles Doped with ppm Level of Pd for Selective Hydrogenation of Acetylene
Sun M., Wang F., Lv G., Zhang X.
Industrial & Engineering Chemistry Research, American Chemical Society (ACS), 2022
152.
Strategies for palladium nanoparticles formation on halloysite nanotubes and their performance in acetylene semi‑hydrogenation
Melnikov D., Reshetina M., Novikov A., Cherednichenko K., Stavitskaya A., Stytsenko V., Vinokurov V., Huang W., Glotov A.
Applied Clay Science, Elsevier, 2023
153.
Different catalytic behavior of Pd/Palygorskite catalysts for semi-hydrogenation of acetylene
Dai H., Xiao X., Huang L., Zhou C., Deng J.
Applied Clay Science, Elsevier, 2021
154.
Adsorption Behavior and Electron Structure Engineering of Pd-IL Catalysts for Selective Hydrogenation of Acetylene
Chen X., Xu Q., Zhao B., Ren S., Wu Z., Wu J., Yue Y., Han D., Li R.
Catalysis Letters, Springer Nature, 2021
155.
N8 stabilized single-atom Pd for highly selective hydrogenation of acetylene
Hu M., Wu Z., Yao Z., Young J., Luo L., Du Y., Wang C., Iqbal Z., Wang X.
Journal of Catalysis, Elsevier, 2021
157.
Interfacial Effect-Induced Electron Localization toward Pd Sites for Efficient Acetylene Semihydrogenation
Wang Q., Li Z., Zhou Y., Wen Y., Zhai J., Xu L., Zhang Q., Qin Y., Lu C., Zhang Q., Li X.
Industrial & Engineering Chemistry Research, American Chemical Society (ACS), 2024
158.
Wang Q., Xu Y., Zhou J., Xu L., Yu L., Jiang D., Lu C., Pan Z., Zhang Q., Li X.
Journal of Industrial and Engineering Chemistry, Korean Society of Industrial Engineering Chemistry, 2021
160.
Polyoxometalate‐Based Metal–Organic Framework as Molecular Sieve for Highly Selective Semi‐Hydrogenation of Acetylene on Isolated Single Pd Atom Sites
Liu Y., Wang B., Fu Q., Liu W., Wang Y., Gu L., Wang D., Li Y.
Angewandte Chemie - International Edition, Wiley, 2021
161.
Current concepts on hyperpolarized molecules in MRI
Viale A., Aime S.
Current Opinion in Chemical Biology, Elsevier, 2010
162.
Selective Single-Site Pd−In Hydrogenation Catalyst for Production of Enhanced Magnetic Resonance Signals using Parahydrogen
Burueva D.B., Kovtunov K.V., Bukhtiyarov A.V., Barskiy D.A., Prosvirin I.P., Mashkovsky I.S., Baeva G.N., Bukhtiyarov V.I., Stakheev A.Y., Koptyug I.V.
Chemistry - A European Journal, Wiley, 2018
163.
Heterogeneous addition of H2 to double and triple bonds over supported Pd catalysts: a parahydrogen-induced polarization technique study.
Kovtunov K.V., Beck I.E., Zhivonitko V.V., Barskiy D.A., Bukhtiyarov V.I., Koptyug I.V.
Physical Chemistry Chemical Physics, Royal Society of Chemistry (RSC), 2012
164.
Evaluation of Activation Energies for Pairwise and Non-Pairwise Hydrogen Addition to Propyne Over Pd/Aluminosilicate Fiberglass Catalyst by Parahydrogen-Induced Polarization (PHIP)
Salnikov O.G., Barskiy D.A., Burueva D.B., Gulyaeva Y.K., Balzhinimaev B.S., Kovtunov K.V., Koptyug I.V.
Applied Magnetic Resonance, Springer Nature, 2014
165.
Parahydrogen-induced polarization in alkyne hydrogenation catalyzed by Pd nanoparticles embedded in a supported ionic liquid phase.
Kovtunov K.V., Zhivonitko V.V., Kiwi-Minsker L., Koptyug I.V.
Chemical Communications, Royal Society of Chemistry (RSC), 2010
166.
Pd on Nanodiamond/Graphene in Hydrogenation of Propyne with Parahydrogen
Burueva D.B., Sviyazov S.V., Huang F., Prosvirin I.P., Bukhtiyarov A.V., Bukhtiyarov V.I., Liu H., Koptyug I.V.
Journal of Physical Chemistry C, American Chemical Society (ACS), 2021
167.
Contrasting Metallic (Rh0) and Carbidic (2D-Mo2C MXene) Surfaces in Olefin Hydrogenation Provides Insights on the Origin of the Pairwise Hydrogen Addition
Meng L., Pokochueva E.V., Chen Z., Fedorov A., Viñes F., Illas F., Koptyug I.V.
ACS Catalysis, American Chemical Society (ACS), 2024
168.
Spatially Resolved NMR Spectroscopy for Operando Studies of Heterogeneous Hydrogenation with Parahydrogen
Skovpin I.V., Trepakova A.I., Kovtunova L.M., Koptyug I.V.
Applied Magnetic Resonance, Springer Nature, 2023
169.
Heterogeneous Catalysis and Parahydrogen‐Induced Polarization
Pokochueva E.V., Burueva D.B., Salnikov O.G., Koptyug I.V.
ChemPhysChem, Wiley, 2021
170.
Acetylene Oligomerization over Pd Nanoparticles with Controlled Shape: A Parahydrogen-Induced Polarization Study
Zhivonitko V.V., Skovpin I.V., Crespo-Quesada M., Kiwi-Minsker L., Koptyug I.V.
Journal of Physical Chemistry C, American Chemical Society (ACS), 2016
171.
Nuclear spin isomers of ethylene: enrichment by chemical synthesis and application for NMR signal enhancement.
Zhivonitko V.V., Kovtunov K.V., Chapovsky P.L., Koptyug I.V.
Angewandte Chemie - International Edition, Wiley, 2013
172.
Manipulating Stereoselectivity of Parahydrogen Addition to Acetylene to Unravel Interconversion of Ethylene Nuclear Spin Isomers
Sviyazov S.V., Babenko S.V., Skovpin I.V., Kovtunova L., Chukanov N., Stakheev A.Y., Burueva D.B., Koptyug I.V.
Physical Chemistry Chemical Physics, Royal Society of Chemistry (RSC), 2024
173.
[13C+D] Double labeling with calcium carbide: incorporation of two labels in one step
Gyrdymova Y.V., Samoylenko D.E., Rodygin K.S.
Chemistry - An Asian Journal, Wiley, 2022
174.
13C‐Labeling as a Method in Organic Synthesis, Catalysis and Biochemical Applications
Rodygin K.S., Bogachenkov A.S., Gyrdymova Y.V., Potorochenko A.N.
Chemistry - Methods, Wiley, 2024
175.
Nonequilibrium Nuclear Spin States of Ethylene during Acetylene Hydrogenation with Parahydrogen over Immobilized Iridium Complexes
Skovpin I.V., Sviyazov S.V., Burueva D.B., Kovtunova L.M., Nartova A.V., Kvon R.I., Bukhtiyarov V.I., Koptyug I.V.
Doklady Physical Chemistry, Pleiades Publishing, 2023
176.
X.Cao, B.W.L.Jang, J.Hu, L.Wang, S.Zhang
177.
Catalysis of semihydrogenation of acetylene to ethylene: current trends, challenges, and outlook
Shittu T.D., Ayodele O.B.
Frontiers of Chemical Science and Engineering, Higher Education Press, 2022
178.
Hydrogenation of Acetylene over Pd–Ag/Sibunit Catalysts: Effect of the Deposition Sequence of Active Component Precursors
Yurpalova D.V., Afonasenko T.N., Trenikhin M.V., Leont’eva N.N., Arbuzov A.B., Temerev V.L., Shlyapin D.A.
Petroleum Chemistry, Pleiades Publishing, 2023
179.
Rational Design of PdAg Catalysts for Acetylene Selective Hydrogenation via Structural Descriptor-based Screening Strategy
180.
In Situ Surface Structures of PdAg Catalyst and Their Influence on Acetylene Semihydrogenation Revealed by Machine Learning and Experiment
Li X., Chen L., Shang C., Liu Z.
Journal of the American Chemical Society, American Chemical Society (ACS), 2021
181.
Selective hydrogenation of highly concentrated acetylene streams over mechanochemically synthesized PdAg supported catalysts
Kley K.S., De Bellis J., Schüth F.
Catalysis Science and Technology, Royal Society of Chemistry (RSC), 2023
182.
Controlled One‐pot Synthesis of PdAg Nanoparticles and Their Application in the Semi‐hydrogenation of Acetylene in Ethylene‐rich Mixtures
Delgado J.A., Benkirane O., Lachaux S., Claver C., Ferré J., Curulla‐Ferré D., Godard C.
ChemNanoMat, Wiley, 2022
184.
Boosting the activity of PdAg2/Al2O3 supported catalysts towards the selective acetylene hydrogenation by means of CO-induced segregation: A combined NAP XPS and mass-spectrometry study
Bukhtiyarov A.V., Panafidin M.A., Prosvirin I.P., Mashkovsky I.S., Markov P.V., Rassolov A.V., Smirnova N.S., Baeva G.N., Rameshan C., Rameshan R., Zubavichus Y.V., Bukhtiyarov V.I., Stakheev A.Y.
Applied Surface Science, Elsevier, 2022
185.
Single-Atom Alloy Pd1Ag6/Al2O3 Egg-Shell Catalyst for Selective Acetylene Hydrogenation
Mashkovsky I.S., Melnikov D.P., Markov P.V., Baeva G.N., Smirnova N.S., Bragina G.O., Stakheev A.Y.
Doklady Chemistry, Pleiades Publishing, 2023
186.
Markov P.V., Mashkovsky I.S., Baeva G.N., Melnikov D.P., Stakheev A.Y.
Mendeleev Communications, OOO Zhurnal "Mendeleevskie Soobshcheniya", 2023
188.
Integration of Bimetallic Electronic Synergy with Oxide Site Isolation Improves the Selective Hydrogenation of Acetylene
Liu F., Xia Y., Xu W., Cao L., Guan Q., Gu Q., Yang B., Lu J.
Angewandte Chemie - International Edition, Wiley, 2021
189.
Array Modified Molded Alumina Supported PdAg Catalyst for Selective Acetylene Hydrogenation: Intrinsic Kinetics Enhancement and Thermal Effect Optimization
Miao C., Cai L., Wang Y., Xu X., Yang J., He Y., Li D., Feng J.
Industrial & Engineering Chemistry Research, American Chemical Society (ACS), 2021
190.
Experimental study and modeling of the liquid phase hydrogenation of acetylene
Kang Z., Wang Y., Liu B., Huang Z., Lan X., Wang T.
Chemical Engineering Journal, Elsevier, 2024
191.
Depositing Pd on the outmost surface of Pd1Ni/SiO2 single-atom alloy via atomic layer deposition for selective hydrogenation of acetylene
Zhong M., Zhao J., Fang Y., Wu D., Zhang L., Du C., Liu S., Yang S., Wan S., Jiang Y., Huang J., Xiong H.
Applied Catalysis A: General, Elsevier, 2023
192.
Atomic intimacy in Pd1Ni/SiO2 single-atom alloy prepared by galvanic replacement for enhanced semi-hydrogenation of acetylene
Liu S., Wu D., Yang F., Chen K., Luo Z., Li J., Zhang Z., Zhao J., Zhang L., Zhang Y., Zhang H., Wan S., Peng Y., Zhang K.H., Xiong H., et. al.
Chemical Engineering Journal, Elsevier, 2024
193.
Selective hydrogenation of acetylene over Pd-Sn catalyst: Identification of Pd2Sn intermetallic alloy and crystal plane-dependent performance
Li R., Yue Y., Chen Z., Chen X., Wang S., Jiang Z., Wang B., Xu Q., Han D., Zhao J.
Applied Catalysis B: Environmental, Elsevier, 2020
194.
Selective Hydrogenation of Acetylene over Pd-Mn/Al2O3 Catalysts
Melnikov D., Stytsenko V., Saveleva E., Kotelev M., Lyubimenko V., Ivanov E., Glotov A., Vinokurov V.
Catalysts, MDPI, 2020
195.
Selective hydrogenation of acetylene over Pd/In2O3@SiO2: The effect of the catalyst reduction temperature
Wu Q., Shen C., Sun K., Liu M., Liu C.
Chemical Engineering Journal, Elsevier, 2024
196.
Acetylene Semi‐hydrogenation at Room Temperature over Pd‐Zn Nanocatalyst
Tiwari G., Sharma G., Verma R., Gakhad P., Singh A.K., Polshettiwar V., Jagirdar B.R.
Chemistry - A European Journal, Wiley, 2023
197.
Zinc Addition Influence on the Properties of Pd/Sibunit Catalyst in Selective Acetylene Hydrogenation
Glyzdova D.V., Afonasenko T.N., Khramov E.V., Leont’eva N.N., Trenikhin M.V., Prosvirin I.P., Bukhtiyarov A.V., Shlyapin D.A.
Topics in Catalysis, Springer Nature, 2020
198.
Effect of Reductive Treatment Duration on the Structure and Properties of Pd‐Zn/C Catalysts for Liquid‐Phase Acetylene Hydrogenation
Glyzdova D.V., Afonasenko T.N., Khramov E.V., Trenikhin M.V., Prosvirin I.P., Shlyapin D.A.
ChemCatChem, Wiley, 2022
199.
3D printed hierarchical spinel monolithic catalysts for highly efficient semi-hydrogenation of acetylene
Yuan Z., Liu L., Ru W., Zhou D., Kuang Y., Feng J., Liu B., Sun X.
Nano Research, Springer Nature, 2022
200.
Highly efficient acetylene semi-hydrogenation over Cun cluster stabilized Pd1 single-atom catalysts
Xu L., Qin Y., Zhang Q., Zhou J., Zhao J., Feng F., Sun T., Xu X., Zhu Y., Lu C., Zhang Q., Wang Q., Li X.
Chemical Engineering Journal, Elsevier, 2024
201.
Cu–Pd bimetal-decorated siloxene nanosheets for semi-hydrogenation of acetylene
Pei X., Zhang D., Tang R., Wang S., Zhang C., Yuan W., Sun W.
Nanoscale, Royal Society of Chemistry (RSC), 2024
202.
A MOF-supported Pd1–Au1 dimer catalyses the semihydrogenation reaction of acetylene in ethylene with a nearly barrierless activation energy
Ballesteros-Soberanas J., Martín N., Bacic M., Tiburcio E., Mon M., Hernández-Garrido J.C., Marini C., Boronat M., Ferrando-Soria J., Armentano D., Pardo E., Leyva-Pérez A.
Nature Catalysis, Springer Nature, 2024
203.
Design of efficient supported Pd-Co catalysts for selective hydrogenation of acetylene
Yurpalova D.V., Afonasenko T.N., Prosvirin I.P., Bukhtiyarov A.V., Kovtunova L.M., Vinokurov Z.S., Trenikhin M.V., Gerasimov E.Y., Khramov E.V., Shlyapin D.A.
Journal of Catalysis, Elsevier, 2024
204.
The nature of modifying effect of gallium on Pd-Ga/Al2O3 catalyst for liquid-phase selective acetylene hydrogenation
Afonasenko T.N., Temerev V.L., Glyzdova D.V., Leont'eva N.N., Trenikhin M.V., Prosvirin I.P., Shlyapin D.A.
Materials Letters, Elsevier, 2021
205.
Highly Selective Acetylene Semihydrogenation Catalyst with an Operation Window Exceeding 150 °C
Lou B., Kang H., Yuan W., Ma L., Huang W., Wang Y., Jiang Z., Du Y., Zou S., Fan J.
ACS Catalysis, American Chemical Society (ACS), 2021
206.
Controllable Deposition of Bi onto Pd for Selective Hydrogenation of Acetylene
Kang H., Wu J., Lou B., Wang Y., Zhao Y., Liu J., Zou S., Fan J.
Molecules, MDPI, 2023
207.
Selective hydrogenation of acetylene on the PdLa@N-doped biochar catalyst surface: the evolution of active sites, catalytic performance, and mechanism
Chen Y., Ning P., Miao R., Shi Y., He L., Guan Q.
New Journal of Chemistry, Royal Society of Chemistry (RSC), 2020
208.
Fully Exposed Nickel Clusters for Semihydrogenation of Acetylene
Sui C., Ma H., Huang F., Wang M., Cai X., Diao J., Ren P., Wen X., Jin L., Wang G., Ma D., Liu H.
ACS Catalysis, American Chemical Society (ACS), 2024
209.
Preparation of MCM-41 supported nickel NPs for the high-efficiency semi-hydrogenation of acetylene
Zhao J., He L., Yu J., Shi Y., Miao R., Guan Q., Ning P.
New Journal of Chemistry, Royal Society of Chemistry (RSC), 2021
212.
The Vinyl Group: Small but Mighty ‐ Transition Metal Catalyzed and Non‐Catalyzed Vinylation Reactions
213.
Direct Synthesis of 3-Coumaranones with Calcium Carbide as an Acetylene Source.
Fu R., Lu Y., Yue G., Wu D., Xu L., Song H., Cao C., Yu X., Zong Y.
Organic Letters, American Chemical Society (ACS), 2021
216.
Atom-economical synthesis of 1,2-bis(phosphine oxide)ethanes from calcium carbide with straightforward access to deuterium- and 13C-labeled bidentate phosphorus ligands and metal complexes
Lotsman K.A., Rodygin K.S., Skvortsova I., Kutskaya A.M., Minyaev M.E., Ananikov V.P.
Organic Chemistry Frontiers, Royal Society of Chemistry (RSC), 2023
217.
Towards Sustainable Carbon Return from Waste to Industry via C2-Type Molecular Unit
Rodygin K.S., Lotsman K.A., Samoylenko D.E., Kuznetsov V.M., Ananikov V.P.
International Journal of Molecular Sciences, MDPI, 2022
218.
Calcium carbide residue – promising hidden source of hydrogen
Lotsman K., Rodygin K.
Green Chemistry, Royal Society of Chemistry (RSC), 2023
219.
Sulfur-Doped g-C3N4-Supported Ni Species with a Wide Temperature Window for Acetylene Semihydrogenation
Zhou H., Li B., Fu H., Zhao X., Zhang M., Wang X., Liu Y., Yang Z., Lou X.
ACS Sustainable Chemistry and Engineering, American Chemical Society (ACS), 2022
220.
π-Adsorption promoted electrocatalytic acetylene semihydrogenation on single-atom Ni dispersed N-doped carbon
Ma W., Chen Z., Bu J., Liu Z., Li J., Yan C., Cheng L., Zhang L., Zhang H., Zhang J., Wang T., Zhang J.
Journal of Materials Chemistry A, Royal Society of Chemistry (RSC), 2022
221.
Robust NixSn/ZSM-12 Catalysts with Zeolite as the Support and Sn as the Promoter for Acetylene Semi-hydrogenation
Wang D., Ye R., Zhang C., Jin C., Lu Z., Shakouri M., Han B., Wang T., Zhang Y., Zhang R., Hu Y., Zhou J., Feng G.
Energy & Fuels, American Chemical Society (ACS), 2023
222.
Effect of IB-metal on Ni/SiO2 catalyst for selective hydrogenation of acetylene
Liu H., Chai M., Pei G., Liu X., Li L., Kang L., Wang A., Zhang T.
Chinese Journal of Catalysis, Elsevier, 2020
224.
Mechanism driven design of trimer Ni1Sb2 site delivering superior hydrogenation selectivity to ethylene
Ge X., Dou M., Cao Y., Liu X., Yuwen Q., Zhang J., Qian G., Gong X., Zhou X., Chen L., Yuan W., Duan X.
Nature Communications, Springer Nature, 2022
225.
Subsurface Carbon as a Selectivity Promotor to Enhance Catalytic Performance in Acetylene Semihydrogenation
Wang Y., Liu B., Lan X., Wang T.
ACS Catalysis, American Chemical Society (ACS), 2021
226.
Homogeneous Ni Single-Atom Sites in the NiZn Intermetallic Nanostructure for Efficient Semihydrogenation of Acetylene
Zhou H., Fu H., Li B., Yan X., Su Y., Pan X., Tang Q., Wang X., Zhao X., Liu Y., Yang Z., Lu Z., Lou X., Li L.
ACS Sustainable Chemistry and Engineering, American Chemical Society (ACS), 2023
227.
Understanding the Role of Coordinatively Unsaturated Al3+ Sites on Nanoshaped Al2O3 for Creating Uniform Ni–Cu Alloys for Selective Hydrogenation of Acetylene
Song Y., Weng S., Xue F., McCue A.J., Zheng L., He Y., Feng J., Liu Y., Li D.
ACS Catalysis, American Chemical Society (ACS), 2023
228.
Acetylene Semi-Hydrogenation on Intermetallic Ni–In Catalysts: Ni Ensemble and Acetylene Coverage Effects from a Theoretical Analysis
Almisbaa Z., Aljama H.A., Almajnouni K., Cavallo L., Sautet P.
ACS Catalysis, American Chemical Society (ACS), 2023
229.
Adsorption Site Regulation to Guide Atomic Design of Ni–Ga Catalysts for Acetylene Semi‐Hydrogenation
Cao Y., Zhang H., Ji S., Sui Z., Jiang Z., Wang D., Zaera F., Zhou X., Duan X., Li Y.
Angewandte Chemie - International Edition, Wiley, 2020
230.
Enhanced acetylene semi-hydrogenation on a subsurface carbon tailored Ni–Ga intermetallic catalyst
Ge X., Ren Z., Cao Y., Liu X., Zhang J., Qian G., Gong X., Chen L., Zhou X., Yuan W., Duan X.
Journal of Materials Chemistry A, Royal Society of Chemistry (RSC), 2022
231.
Nickel‐Based High‐Entropy Intermetallic as a Highly Active and Selective Catalyst for Acetylene Semihydrogenation
Ma J., Xing F., Nakaya Y., Shimizu K., Furukawa S.
Angewandte Chemie - International Edition, Wiley, 2022
232.
Promoter Effect on Ni/SiO2 Catalysts for Acetylene Semi-hydrogenation to Ethylene
Zhang C., Wu L., Ye R., Feng G., Zhang R.
Catalysis Letters, Springer Nature, 2024
234.
Copper Catalysts in Semihydrogenation of Acetylene: From Single Atoms to Nanoparticles
Shi X., Lin Y., Huang L., Sun Z., Yang Y., Zhou X., Vovk E., Liu X., Huang X., Sun M., Wei S., Lu J.
ACS Catalysis, American Chemical Society (ACS), 2020
235.
Enhanced Hydrogenation Activity over a Zn-Modified Cu-Based Catalyst in Acetylene Hydrogenation
Lu C., Zeng A., Wang Y., Wang A.
Industrial & Engineering Chemistry Research, American Chemical Society (ACS), 2022
236.
Interfacial Bifunctional Effect Promoted Non-Noble Cu/FeyMgOx Catalysts for Selective Hydrogenation of Acetylene
Fu F., Liu Y., Li Y., Fu B., Zheng L., Feng J., Li D.
ACS Catalysis, American Chemical Society (ACS), 2021
237.
Insight into the Activity of Atomically Dispersed Cu Catalysts for Semihydrogenation of Acetylene: Impact of Coordination Environments
Huang F., Peng M., Chen Y., Gao Z., Cai X., Xie J., Xiao D., Jin L., Wang G., Wen X., Wang N., Zhou W., Liu H., Ma D.
ACS Catalysis, American Chemical Society (ACS), 2021
238.
Hydrophobic Surface Modification of Cu-Based Catalysts for Enhanced Semihydrogenation of Acetylene in Excess Ethylene
Liu T., Xiong J., Luo Q., Mao S., Wang Y.
ACS Catalysis, American Chemical Society (ACS), 2024
239.
Copper-Based Catalysts for Selective Hydrogenation of Acetylene Derived from Cu(OH)2
Lu C., Zeng A., Wang Y., Wang A.
ACS Omega, American Chemical Society (ACS), 2021
241.
Cu Single‐Atom Catalysts for High‐Selectivity Electrocatalytic Acetylene Semihydrogenation
Jiang X., Tang L., Dong L., Sheng X., Zhang W., Liu Z., Shen J., Jiang H., Li C.
Angewandte Chemie - International Edition, Wiley, 2023
242.
Selective electrocatalytic semihydrogenation of acetylene impurities for the production of polymer-grade ethylene
Bu J., Liu Z., Ma W., Zhang L., Wang T., Zhang H., Zhang Q., Feng X., Zhang J.
Nature Catalysis, Springer Nature, 2021
243.
Room-temperature electrochemical acetylene reduction to ethylene with high conversion and selectivity
Shi R., Wang Z., Zhao Y., Waterhouse G.I., Li Z., Zhang B., Sun Z., Xia C., Wang H., Zhang T.
Nature Catalysis, Springer Nature, 2021
244.
Electrosynthesis of polymer-grade ethylene via acetylene semihydrogenation over undercoordinated Cu nanodots
Xue W., Liu X., Liu C., Zhang X., Li J., Yang Z., Cui P., Peng H., Jiang Q., Li H., Xu P., Zheng T., Xia C., Zeng J.
Nature Communications, Springer Nature, 2023
245.
Semi-Hydrogenation of Acetylene to Ethylene Catalyzed by Bimetallic CuNi/ZSM-12 Catalysts
Hu S., Zhang C., Wu M., Ye R., Shi D., Li M., Zhao P., Zhang R., Feng G.
Catalysts, MDPI, 2022
246.
Unveiling the origin of enhanced catalytic performance of NiCu alloy for semi-hydrogenation of acetylene
Li Z., Zhang J., Tian J., Feng K., Jiang Z., Yan B.
Chemical Engineering Journal, Elsevier, 2022
248.
C, N Co-Decorated Alumina-Supported Au Nanoparticles: Enhanced Catalytic Performance for Selective Hydrogenation of Acetylene
249.
Construction of Enriched Metal Vacancies to Enhance Catalytic Behavior for Selective Hydrogenation Over Au–Cu Nanoalloys
Zheng T., Zhu Y., Ma R., Bai L., Yu H., Zhao S., He Y., Liu Y., Li D.
Industrial & Engineering Chemistry Research, American Chemical Society (ACS), 2024
250.
Au3+ Species Boost the Catalytic Performance of Au/ZnO for the Semi-hydrogenation of Acetylene
Zhou H., Li B., Zhang Y., Yan X., Lv W., Wang X., Yuan B., Liu Y., Yang Z., Lou X.
ACS applied materials & interfaces, American Chemical Society (ACS), 2021
251.
Au@Pt Nanotubes within CoZn-Based Metal-Organic Framework for Highly Efficient Semi-hydrogenation of Acetylene
Wang J., Xu H., Ao C., Pan X., Luo X., Wei S., Li Z., Zhang L., Xu Z., Li Y.
iScience, Elsevier, 2020
252.
Enriching Pt3 ensemble with isolated 3-fold hollow site by crystal-phase engineering of Pt3Fe single-nanoparticle for acetylene hydrogenation
253.
XAS and DFT investigation of atomically dispersed Cu/Co alloyed Pt local structures under selective hydrogenation of acetylene to ethylene
Bolarinwa Ayodele O., Daniel Shittu T., Olayinka Togunwa S., Yu D., Tian Z.
Chemical Engineering Journal, Elsevier, 2024
255.
Carburization in the Crystalline Structure of the Metal Ag for Efficient Selective Hydrogenation of Acetylene
Zhou S., Zhou W., Lu C., Chen Y., Zhou C., Zeng A., Wang A., Zhang Y., Tan L., Dong L.
Industrial & Engineering Chemistry Research, American Chemical Society (ACS), 2024
256.
Atomically dispersed zeolite-supported rhodium complex: Selective and stable catalyst for acetylene semi-hydrogenation
Zhao Y., Bozkurt Ö.D., Kurtoğlu-Öztulum S.F., Su Yordanli M., Hoffman A.S., Hong J., Perez-Aguilar J.E., Saltuk A., Akgül D., Demircan O., Ateşin T.A., Aviyente V., Gates B.C., Bare S.R., Uzun A., et. al.
Journal of Catalysis, Elsevier, 2024
257.
Highly Selective Low-Temperature Acetylene Semihydrogenation Guided by Multiscale Machine Learning
Chen L., Li X., Ma S., Hu Y., Shang C., Liu Z.
ACS Catalysis, American Chemical Society (ACS), 2022
258.
Harnessing Physics-inspired Machine Learning to Design Nanocluster Catalysts for Dehydrogenating Liquid Organic Hydrogen Carriers
Lin C., Lee B.C., Anjum U., Prabhu A.M., Chaudhary N., Xu R., Choksi T.S.
Applied Catalysis B: Environmental, Elsevier, 2025
259.
Machine learning for design principles for single atom catalysts towards electrochemical reactions
Tamtaji M., Gao H., Hossain M.D., Galligan P.R., Wong H., Liu Z., Liu H., Cai Y., Goddard W.A., Luo Z.
Journal of Materials Chemistry A, Royal Society of Chemistry (RSC), 2022
260.
Leveraging Machine Learning Potentials for In-Situ Searching of Active sites in Heterogeneous Catalysis
Cheng X., Wu C., Xu J., Han Y., Xie W., Hu P.
Precision Chemistry, American Chemical Society (ACS), 2024
261.
Data‐driven Machine Learning for Understanding Surface Structures of Heterogeneous Catalysts
Li H., Jiao Y., Davey K., Qiao S.
Angewandte Chemie - International Edition, Wiley, 2023
262.
Thermal effect optimization endows a selective and stable PdCu single atom alloy catalyst for acetylene hydrogenation
263.
Atomically Dispersed Pd on Nanodiamond/Graphene Hybrid for Selective Hydrogenation of Acetylene.
Huang F., Deng Y., Chen Y., Cai X., Peng M., Jia Z., Ren P., Xiao D., Wen X., Wang N., Liu H., Ma D.
Journal of the American Chemical Society, American Chemical Society (ACS), 2018
264.
Atomic Layers of B2 CuPd on Cu Nanocubes as Catalysts for Selective Hydrogenation
Gao Q., Yan Z., Zhang W., Pillai H.S., Yao B., Zang W., Liu Y., Han X., Min B., Zhou H., Ma L., Anaclet B., Zhang S., Xin H., He Q., et. al.
Journal of the American Chemical Society, American Chemical Society (ACS), 2023
265.
Grafting nanometer metal/oxide interface towards enhanced low-temperature acetylene semi-hydrogenation
Zou S., Lou B., Yang K., Yuan W., Zhu C., Zhu Y., Du Y., Lu L., Liu J., Huang W., Yang B., Gong Z., Cui Y., Wang Y., Ma L., et. al.
Nature Communications, Springer Nature, 2021
266.
Anchoring Cu1 species over nanodiamond-graphene for semi-hydrogenation of acetylene
Huang F., Deng Y., Chen Y., Cai X., Peng M., Jia Z., Xie J., Xiao D., Wen X., Wang N., Jiang Z., Liu H., Ma D.
Nature Communications, Springer Nature, 2019
267.
Synergizing metal–support interactions and spatial confinement boosts dynamics of atomic nickel for hydrogenations
Gu J., Jian M., Huang L., Sun Z., Li A., Pan Y., Yang J., Wen W., Zhou W., Lin Y., Wang H., Liu X., Wang L., Shi X., Huang X., et. al.
Nature Nanotechnology, Springer Nature, 2021