Young-eun
Kim†
ab,
Jaekyum
Kim†
c,
Unho
Jung
a,
Yongha
Park
a,
Min Hye
Youn
d,
Dong Hyun
Chun
d,
Byeong-Seon
An
e,
Jung Kyu
Kim
c,
Byung-Hyun
Kim
*fg,
Ki Bong
Lee
*b and
Kee Young
Koo
*ah
aHydrogen Research Department, Korea Institute of Energy Research (KIER), 152 Gajeong-ro, Yuseong-gu, Daejeon 34129, Republic of Korea. E-mail: kykoo@kier.re.kr; Fax: +82 42 860 3739; Tel: +82 42 860 3192
bDepartment of Chemical and Biological Engineering, Korea University, 145 Anam-ro, Seongbuk-gu, Seoul 02841, Republic of Korea. E-mail: kibonglee@korea.ac.kr; Tel: +82 2 3290 4851
cSchool of Chemical Engineering, Sungkyunkwan University (SKKU), 2066, Seobu-ro, Jangan-gu, Suwon, 16419, Republic of Korea
dCarbon Conversion Research Laboratory, Korea Institute of Energy Research (KIER), 152 Gajeong-ro, Yuseong-gu, Daejeon 34129, Republic of Korea
eAnalysis Center for Energy Research, Korea Institute of Energy Research (KIER), 152 Gajeong-ro, Yuseong-gu, Daejeon 34129, Republic of Korea
fDepartment of Energy and Bio Sciences, Hanyang University ERICA, 55 Hanyangdaehak-ro, Sangnok-gu, Ansan, Gyeonggi-do 15588, Republic of Korea. E-mail: bhkim00@hanyang.ac.kr; Tel: +82 31 400 5495
gDepartment of Applied Chemistry, Center for Bionano Intelligence Education and Research, Hanyang University ERICA, 55 Hanyangdaehak-ro, Sangnok-gu, Ansan, Gyeonggi-do 15588, Republic of Korea
hHydrogen Energy Engineering, University of Science and Technology (UST), 217 Gajeong-ro, Yuseong-gu, Daejeon, 34113, Republic of Korea
First published on 10th October 2025
Linear alpha-olefins (LAOs), containing a terminal double bond at their carbon chain, are widely used as raw materials in the production of polyolefins, lubricants, alcohol-based detergents, and alpha-olefin sulfonates. One common method for synthesizing LAOs is alcohol dehydration, for example, conversion of 1-octanol to 1-octene. However, this reaction often produces double-bond isomers, increasing production costs due to the need for further purification. Therefore, catalysts that enhance alcohol dehydration while minimizing isomer formation are needed. This study proposes alkaline-earth metal (AEM)-impregnated Al2O3 catalysts to improve 1-octene purity. Our approach focuses on understanding how AEM impregnation alters the electron density of Al2O3 and, consequently, affects the catalytic activity. Among the catalysts tested, Ba-impregnated Al2O3 demonstrated the highest 1-octene purity, which is attributed to the increased electron density at Al3+ sites, facilitated by active electron transfer from the AEM to Al3+. Density functional theory calculations further reveal that this electron density increase reduces the energy of 1-octene re-adsorption, limiting its subsequent isomerization. These results highlight the relationship between the electronic modulation of the Al3+ active site and reaction mechanism, which can contribute to the development of more efficient catalysts in the petrochemical industry.
In alcohol dehydration, water is removed from alcohols to form olefins and ethers. Acidic metal oxides like zeolites,19,20 TiO2,21 WOx,22 and Al2O3 have been investigated as dehydration catalysts. The acid/base properties of the catalyst directly influence performance, as acidic sites act as active sites for alcohol dehydration. Al2O3 is the most widely used due to its high activity, mild acidity, and high 1-olefin selectivity.23–27 For light alcohols such as ethanol and propanol, the position of the double bond is generally unaltered. However, linear alpha-olefins (LAOs) and linear internal olefins (LIOs) with different positions of the double bond can be obtained during the dehydration of long-chain alcohols (>C4). LAOs, which have a double bond at the terminal position, serve as essential feedstocks in the production of polyolefins, lubricants, alcohol-based detergents, and alpha-olefin sulfonates.28 LAOs can be synthesized via alcohol dehydration using Sasol's 1-heptene upgrading29 and biomass upgrading.30 However, these methods often generate LIOs with similar boiling points and molecular geometries, complicating separation and purification. Thus, suppressing LIO formation is essential for improving alcohol dehydration efficiency.31–33
Dehydration of 1-octanol yields 1-octene, a valuable LAO used extensively in the petrochemical industry as a co-monomer of linear low-density polyethylene, combined with 1-butene and 1-hexene. Among the various 1-octene production methods, Al2O3-catalyzed 1-octanol dehydration has gained attention due to its high C8 selectivity and oxygenate upgrading.30,34 The mechanism underlying alpha-olefin formation during alcohol dehydration remains not fully understood. Previous studies have proposed E1-type mechanisms involving the existence of alkoxy species as intermediates and E2-type elimination mechanisms consisting of the participation of β-H in the reaction.35–38 For alumina catalysts, the E2 mechanism is generally favored over the E1 mechanism. In the E1-type elimination, the intermediate carbocation may undergo double-bond isomerization, leading to high isomer selectivity. Conversely, in the E2 mechanisms, β-H positioned adjacent to the hydroxyl group favors the formation of alpha-olefin during primary alcohol dehydration. Thus, Al2O3, which promotes E2 elimination, typically results in higher alpha-olefin selectivity. In this E2 pathway, the hydroxyl group adsorbs onto the acid site of Al3+ on Al2O3, and then, β-H is attracted to the O atom of the metal oxide (Scheme 1). Subsequently, the β-H and hydroxyl group form water, which is eliminated to yield the olefin product. This mechanism is especially favorable for producing LAOs when using weakly acidic catalysts like Al2O3.39
![]() | ||
| Scheme 1 1-Octanol dehydration catalyzed by Al2O3via an E2-type mechanism.40 | ||
However, the LAO formed can be re-adsorbed onto the acidic sites of the catalyst and rehydrated to form a secondary alcohol.25 Further dehydration of this secondary alcohol results in a mixture of LAO and LIO products, such as 2-olefins. During 1-octanol dehydration, multiple LIOs such as cis-2-octene, trans-2-octene, trans-3-octene, and trans-4-octene are simultaneously formed due to the long carbon chain of 1-octanol. The prevalence of LIOs is further driven by the thermodynamic stability of internal double bonds. Various studies have sought to improve LAO selectivity using various metal oxide catalysts. Davis et al.40 reported a relationship between the basicity of metal oxide catalysts such as Al2O3, ZrO, In2O3, SiO2, MgO, and CaO, and the distribution of dehydration products, specifically, the formation of the Hofmann product (1-alkenes). Kawk et al.41 explored the regioselectivity of 2-butanol dehydration and showed that the stabilization energy of the transition state depends on van der Waals interactions between the adsorbed alcohol and the catalyst surface.
Several previous studies have employed density functional theory (DFT) calculations to investigate the mechanisms and kinetics of alcohol dehydration, aiming to establish a correlation between the acid/base properties and catalytic activity. Roy et al.42 reported an E2-type mechanism with a lower energy barrier (134.7 or 153.0 kJ mol−1) than the corresponding E1-type elimination (290.3 kJ mol−1). The Al3+ site of Al2O3 facilitates OH group removal, while surface O ions interact with the β-H. The higher acidity of Al2O3, compared to TiO2 and ZrO2, led to a lower overall energy barrier, as the adsorption energy of 1-alcohol is influenced by the acid–base properties of the catalyst.43 Furthermore, double-bond isomer selectivity is affected by the surface properties of the catalyst. During alcohol dehydration over Al2O3, isomer formation occurs via a “dehydration–hydration–dehydration” pathway.25 The iso-alcohol formed through 1-olefin hydration features its hydroxyl group within the carbon chain; thus, the location of the resulting double bond depends on the position of the β-H atom involved in the elimination. Dabbagh et al.44,45 conducted a DFT study on 2-butanol dehydration and discovered that the double bond position of the olefin product is governed by the position of the β-H participating in the reaction. The activation was shown to depend on the distance between the β-H atom and the basic surface sites in the transition state.46,47 Consequently, the anti-Saytzeff effect was observed, where the Hofmann product (LAO or the cis-isomer) was favored over the Saytzeff product (trans-isomer). Most existing studies have focused on short-chain alcohols, such as ethanol, isopropanol, and butanol. Although many DFT studies have been performed on the reaction mechanism and activation energies (Eas), the role of olefin–catalyst interactions in determining product selectivity remains unexplored. In particular, the impact of olefin adsorption sites on double-bond isomer formation has been investigated mainly in the context of n-paraffin/olefin separation48–50 using zeolite-based catalysts with Brønsted acidic sites. However, the surface properties of the catalyst and the produced olefins have not been determined.
This study examines the effect of catalyst basicity on the activity and selectivity of 1-octanol dehydration by impregnating a γ-Al2O3 catalyst with various alkaline-earth metals (AEMs), including Mg, Ca, Sr, and Ba. We further computed changes in surface electron density of γ-Al2O3 following AEM incorporation and correlated the 1-octene adsorption energy with product selectivity using DFT calculations.
| ΔE1-octene* = E1-octene* − (E* + E1-octene) |
Charge density difference (Δρ) was computed using VASPKIT58 using the following equation:
| Δρ = ρ1-octene* − (ρ* + ρ1-octene) |
| Catalyst | Metal contents | NH3 desorptionb (μmol g−1) | CO2 desorptionb (μmol g−1) | Surface areac (m2 g−1) | Mean pore sizec (nm) | |
|---|---|---|---|---|---|---|
| (wt%)a | (mol%) | |||||
| a Confirmed by ICP-MS. b Calculated by integrating the results of TPD. c Estimated from N2 adsorption–desorption experiments performed at −196 °C. | ||||||
| Al2O3 | — | — | 249 | 18 | 162.24 | 10.98 |
| Mg/Al2O3 | 0.23 | 0.96 | 232 | 19 | 162.70 | 11.02 |
| Ca/Al2O3 | 0.46 | 1.17 | 221 | 25 | 160.79 | 11.03 |
| Sr/Al2O3 | 0.87 | 1.01 | 206 | 28 | 162.80 | 10.85 |
| Ba/Al2O3 | 1.36 | 1.01 | 195 | 33 | 163.94 | 10.82 |
To understand the electronic influence of AEM incorporation, we analyzed the surface chemistry of the catalysts using XPS. On the Al2O3 surface, Al3+ and lattice O ions serve as acidic and basic sites, respectively.61–63 All XPS spectra were aligned to the C–C bond peak at 284.8 eV in the C 1s spectra (Fig. S4a). A single dominant Al 2p peak was observed at ∼74.3 eV for all samples, with a slight decrease in BE as the ionic radius of the AEM increased. This negative shift is attributed to the electron transfer from the incorporated AEM ions to Al centers, indicating a modified electronic environment and the Lewis basicity of the Al2O3 surface, which can significantly affect catalytic activity.
The electronic modifications observed via XPS were further correlated with acid–base properties, as revealed by probe molecule studies. The Py-DRIFT spectrum (Fig. 1(d) and Table S1) of Al2O3 exhibited peaks at 1621, 1612, and 1596 cm−1. These peaks were classified according to the acid strength of the Lewis acid site (LAS), which was composed of several unsaturated Al3+ sites.64 The 1621 cm−1 peak arises from Py coordination to tetrahedral Al3+ in Al2O3, forming a strong LAS through interaction with the Py 8a orbital.65 The peaks at 1612 and 1596 cm−1 arising from Py and unsaturated Al3+ were also observed, representing medium and weak LASs, respectively.66 Among these, the strong LAS showed the highest sensitivity to AEM impregnation.67 Notably, the intensity of this peak corresponding to the strong LAS significantly decreased with increasing atomic radius of the AEM.
Complementing the acid site characterization, CO2-DRIFT spectroscopy was performed to investigate the basic characteristics of the catalysts (Fig. 1(e)). The CO2-DRIFT spectrum of Al2O3 displayed two peaks at 1661 and 1435 cm−1, attributed to bidentate CO2 groups.65,68 Upon AEM impregnation, the intensity of the unidentate CO2 characteristic peak (1586 cm−1) increased with a direct correlation to the period of the AEM impregnation. When the CO2 interacts with the strong basic O sites of the metal oxides, monodentate CO2 species appear, which serve as an indicator of strong basic sites.69 The increase in the basicities of the impregnated Al2O3 catalysts aligns well with previously reported basicities of corresponding metal oxides.70,71
Quantitative validation of these qualitative trends observed in DRIFT spectroscopy was attained via TPD measurements using NH3 and CO2 (Table 1 and Fig. S6). The Al2O3 exhibited the highest acidity, with an NH3 desorption value of 249 μmol g−1, while AEM-impregnated catalysts showed progressively lower NH3 desorption values as the atomic radius of the AEM increased: Mg/Al2O3 (232 μmol g−1) > Ca/Al2O3 (221 μmol g−1) > Sr/Al2O3 (206 μmol g−1) > Ba/Al2O3 (195 μmol g−1), representing a 22% reduction from pure Al2O3 to Ba/Al2O3. Conversely, CO2 desorption increased across the series as follows: Al2O3 (18 μmol g−1) < Mg/Al2O3 (19 μmol g−1) < Ca/Al2O3 (25 μmol g−1) < Sr/Al2O3 (28 μmol g−1) < Ba/Al2O3 (33 μmol g−1), with Ba/Al2O3 showing an 83% increase compared to pure Al2O3. This inverse relationship between acidity and basicity supports an electron transfer mechanism. According to the hard and soft acid–base theory, the electron density increases with the atomic radius of the AEM,68 facilitating electron transfer from the AEM to Al ions and thus enhancing the basicity of the catalyst surface while diminishing its acidity.
These trends at the atomic level are further elucidated by DFT calculations. The γ-Al2O3 structure proposed by Digne et al. was adopted to model the AEM/Al2O3 surfaces. Each AEM (Mg, Ca, Sr, and Ba) was individually introduced onto the γ-Al2O3 (110) surface, since the (110) surface is the experimentally dominant and most stable facet.64,72–74 To focus on the electronic structure changes when AEMs are impregnated and present on the Al2O3 surface in the form of AEM oxides, we introduced a single AEM atom on top of the Al2O3 surface. Five potential surface sites for AEM incorporation were considered, classifying Al atoms into tri- and tetra-coordinated species (Al3c and Al4c) and O into two top layers (O3c and O2c) and one subsurface layer (O3c_sub) (Fig. S7(a)). Among these, the O3c_sub site was thermodynamically favorable for all AEMs, as indicated by formation energy calculations (Fig. S7(b)). Geometry optimization of the AEM-doped surfaces revealed that AEMs pushed and penetrated the lattice Al ions of Al2O3, forming the AEM/oxide configuration (Fig. S8). Bader charge analysis confirmed that the number of electrons in the surrounding Al atoms increased with the atomic radius of the AEM (Table S2). This trend was attributed to the lower electronegativities of heavier AEMs, which readily lose electrons compared to Al with a Pauling electronegativity of 1.61, thereby donating charge to surrounding Al atoms after lattice penetration (Fig. 1(f)). The visualized charge density difference (Fig. 1(g–j)) for Mg/Al2O3, Ca/Al2O3, Sr/Al2O3, and Ba/Al2O3 clearly illustrates this electron donation. Furthermore, the average Al 2p core-level shifts (CLS) were calculated (Fig. S9), and the average CLS systematically increases in magnitude from Mg to Ca, Sr, and Ba, tracking the degree of charge transfer to adjacent Al sites. The average CLS values exhibit excellent quantitative agreement with the experimental XPS peak positions (R2 = 0.937). As the atomic radius of AEMs increases from Mg to Ba, it is easier for the AEM to lose electrons and donate them to the surrounding Al. This confirms that the AEM acts as an electron donor, increasing the Lewis basicity of the electron-rich AEM/Al2O3 surface as compared to that of the bare Al2O3 surface.
:
octene ratio, was affected by the basicity of the catalyst. Catalysts with lower acidities (Al2O3 > Mg/Al2O3 > Ca/Al2O3 > Sr/Al2O3 > Ba/Al2O3) exhibited lower 1-octanol conversion. DOE, formed by the reaction of two 1-octanol molecules, was obtained with high selectivity because of the high fraction of unreacted 1-octanol in the reactor, owing to the lower alcohol conversion.25 Ethers (including diethyl ether) analogous to DOE decompose into alcohols and olefins at 300–400 °C, similar to alcohol dehydration.
According to Phung et al.,75 ethers adsorb onto acidic sites of the catalyst, like alcohols, and the conversion increases with the density of acidic sites. Consequently, AEM-impregnated catalysts with low acidity showed high DOE and low 1-octene selectivity at 300 °C. Nevertheless, at 400 °C, all catalysts exhibited >2% DOE selectivity and <90% 1-octanol conversion. 1-Octene purity, which affects downstream separation, decreased with increasing 1-octanol conversion. Notably, the highly acidic Al2O3 catalyst produced a lower 1-octene purity (80.9%) than the Mg/Al2O3 catalyst (88.2%) at a similar conversion rate at 300 °C. High catalyst acidity influenced the formation of the double-bond isomer of 1-octene and 1-octanol conversion, leading to lower 1-octene purity. Comparison of the cis/trans ratio, which indicates the ratio of cis- to trans-isomers among the double-bond isomer products, showed that catalysts with AEM additives exhibited higher cis/trans ratios than bare Al2O3. Specifically, catalysts impregnated with highly basic AEMs, such as Ba/Al2O3, showed negligible trans-isomer formation even at 350 °C. These findings suggest that AEM addition enhances the anti-Saytzeff effect, thereby suppressing trans-isomer production. Fig. 3 illustrates the product distributions across various catalysts at different reaction temperatures. Al2O3 showed the highest isomer fraction (∼20%) at 300 °C and an even greater fraction (67.8%) at 400 °C, which minimized the difference between 1-octanol conversions of AEM-impregnated catalysts. Isomer formation proceeded via the re-adsorption of 1-octene by Al3+, rather than olefin isomerization. Al2O3, mainly composed of LASs, produced double-bond isomers via a “dehydration–hydration–dehydration” sequence involving 1-olefin re-adsorption.25,36 Akizuki et al.76 investigated the relationships between catalyst acid type and selectivity in the reactions of 1-octene and 2-octanol and revealed that cis/trans ratios were affected by the catalyst properties. The Ea of 1-octene hydration at LASs (263 ± 17 kJ mol−1) was lower than that of 1-octene isomerization (304 ± 38 kJ mol−1).
![]() | ||
| Fig. 3 Product distribution during 1-octanol dehydration over (a) Al2O3, (b) Mg/Al2O3, (c) Ca/Al2O3, (d) Sr/Al2O3, and (e) Ba/Al2O3 at various reaction temperatures. | ||
The mechanism of 1-octanol dehydration proceeds through the following steps (Scheme 1). First, 1-octanol is adsorbed onto the Al3+ LASs of the catalyst. This adsorption is primarily governed by strong LASs. Then, the β-H on the terminal carbon is adsorbed on the O ions of the catalyst, enabling a concerted β-H elimination mechanism.36 During this step, the β-H atom combines with the hydroxyl group of 1-alcohol, forming water and an olefin product. However, the generated 1-olefin is re-adsorbed onto the Al3+ sites. Upon re-adsorption, activation of the π bond of the 1-olefin allows the hydroxyl group to recombine with the C chain.77 According to Markovnikov's rule,78,79 hydration preferentially produces secondary or tertiary alcohol products, primarily producing 2-octanol.
Subsequently, 2-octanol undergoes a dehydration step similar to that of 1-octanol via adsorption onto the acidic sites of the catalyst, forming water and an olefin product. As the hydroxyl group is located on the second carbon, elimination involving β-H atoms on adjacent carbons leads to the formation of either 1-octene (via removal of the H on C1) or 2-octene (via removal of the H on C3).36 At elevated temperatures, the proportion of internal isomers increases due to higher thermodynamic stability relative to 1-octene. Nevertheless, isomer formation was reduced when AEM-impregnated catalysts were used, even at high temperatures. This reduction can be attributed to the weakened acidity of Al3+ sites, which reduces the re-adsorption of 1-octene.
Conversely, cis-isomer ratios ranged between 10.7 and 20.7% across all catalysts, and small variations were observed between the cis-isomer ratios of the catalysts when compared with the case of the trans-isomer ratios. This was attributed to the mechanism of isomerization. Unlike 1-octanol dehydration, 1-octene isomerization mainly occurs in the absence of water. Since the trans-isomer exhibits a more linear and geometrically stable structure than the cis-isomer, 1-octene preferentially isomerizes to the trans-isomer, with a cis/trans ratio of <1 for all catalysts. However, our experimental results for 1-octanol dehydration revealed cis/trans ratios >1 under all conditions except for Al2O3 at 400 °C. This fundamental difference confirms that the isomers were generated via “dehydration–hydration–dehydration” instead of direct 1-octene isomerization.
The experimental results demonstrated that AEM impregnation significantly reduced 1-octene re-adsorption, thereby improving selectivity (Fig. 4). To explore the relationship between the reduced 1-octene re-adsorption and the increasing atomic radius of AEMs on the Al2O3 surface, DFT calculations were conducted on 1-octene adsorption using various AEM-impregnated surface models. A comprehensive COHP analysis was performed to investigate the interaction between surface Al atoms and the C atoms of 1-octene, facilitating a comparative assessment of 1-octene adsorption on diverse AEM-modified models (Fig. 5(a)). In each projected COHP (−pCOHP) plot, the upper and lower parts represent antibonding and bonding states, respectively. As the ionic radius of the AEM increased, the high-energy window considerably shifted toward the Fermi level. This shift suggests a weakened interaction between 1-octene and the Al2O3 surface, attributable to the reduced positive charge of the Al2O3 surface, and is consistent with prior research highlighting enhanced olefin adsorption on surfaces with pronounced positive charges.80,81 Al2O3 surfaces impregnated with Ca, Sr, and Ba exhibited significant antibonding states below the Fermi level, suggesting weaker 1-octene adsorption, which is contrary to pristine Al2O3. The presence of occupied antibonding states below the Fermi level hinders adsorption by lowering the adsorption energy, thereby promoting desorption of 1-octene. Thus, the upward shift of bonding states and occupancy of antibonding states in AEM-impregnated Al2O3 surfaces led to weaker adsorptions. In agreement with the COHP analysis, the calculated adsorption energies of 1-octene decreased with increasing AEM radius. The adsorption energy of 1-octene on bare Al2O3 was −1.22 eV, whereas that on Ba-impregnated Al2O3 was −0.68 eV. This weaker binding on Ba-impregnated Al2O3 facilitates the easier release of 1-octene compared to the pristine Al2O3 surface. Additionally, the lower adsorption energy correlated with an elongated Al–C bond, while the C1–C2 bond in 1-octene remained at 1.34 Å, consistent with typical C
C bond lengths (Table S5). A correlation was observed between the integrated pCOHP (IpCOHP) and 1-octene adsorption energy across various AEM-impregnated surfaces (Fig. 5(c)). Since higher IpCOHP indicates stronger binding of 1-octene to the catalyst surface, the reduced bond strength in AEM-impregnated Al2O3 further supports the change in 1-octene conversion. Therefore, in the absence of AEMs, Al2O3 should undergo 1-octene re-adsorption, resulting in high 1-octene conversion. Conversely, Ba-impregnated Al2O3, with the lowest adsorption energy (−0.68 eV), exhibited the weakest re-adsorption tendency. In summary, modulating the electronic properties of Al sites via AEM impregnation reduces 1-octene re-adsorption, thereby enhancing selectivity by directly altering its conversion dynamics.
AEM incorporation altered the electronic properties of Al2O3 surfaces. Py-FTIR spectra showed a decrease in the peak intensity of LASs with the increasing AEM ionic radius (for Mg, Ca, Sr, and Ba), while CO2-FTIR spectra revealed a corresponding increase in strongly basic sites, exhibiting characteristic unidentate CO2 adsorption peaks. These acid/base property changes were attributed to the electron density redistribution at Al3+ sites. DFT calculations confirmed electron transfer from AEMs to Al3+ upon AEM incorporation, with the extent of electron transfer increasing with the AEM ionic radius, as confirmed by XPS analysis. The primary cause of low 1-octene purity was identified as a “dehydration–hydration–dehydration” mechanism. This mechanism was experimentally validated using 1-octene as a probe molecule and direct 2-octanol dehydration studies, revealing distinct cis/trans selectivity patterns. The electronic structure modifications directly translated to improved catalytic performance. While pure Al2O3, despite its high acidity, produced low-purity 1-octene due to extensive isomer formation, AEM addition enhanced catalyst basicity and 1-octene purity. AEMs with larger ionic radii reduced the positive charge density of Al3+ sites more effectively, leading to lower 1-octene adsorption energy and minimizing isomer formation. Ba/Al2O3 achieved the optimal combination of electronic properties, demonstrating the highest 1-octene purity through effective re-adsorption inhibition.
Overall, this study provides valuable insights into the development of efficient catalysts for the production of high-purity 1-octene, a valuable material used in the petrochemical industry. However, it only examined the effects of individual model compounds on the formation of high-purity 1-octene. Hence, further studies are required to evaluate the influences of co-existing water generated in the reaction on the production of high-purity 1-octene.
Footnote |
| † These authors equally contributed to this manuscript. |
| This journal is © The Royal Society of Chemistry 2026 |