Next Article in Journal
Influence Factors on the Energy Regulation Law of a Coal Seam after Hydraulic Slotting
Previous Article in Journal
The Influence of the Use of Carrot and Apple Pomace on Changes in the Physical Characteristics and Nutritional Quality of Oat Cookies
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Insights into the Effect of a Microwave Field on the Properties of Modified γ-Alumina: A DFT Study

State Key Laboratory of Heavy Oil Processing, China University of Petroleum (Beijing), Beijing 102200, China
*
Author to whom correspondence should be addressed.
Processes 2024, 12(10), 2064; https://doi.org/10.3390/pr12102064
Submission received: 23 August 2024 / Revised: 15 September 2024 / Accepted: 19 September 2024 / Published: 24 September 2024
(This article belongs to the Section Materials Processes)

Abstract

:
γ-Alumina is often used as a support for hydrodesulfurization catalysts due to its excellent performance. During the catalytic reaction, the strong surface acidity of γ-alumina can induce a strong interaction between the active phase and the support. The reaction activity of the catalyst can be affected by changing the present mode of the active phase on the surface of the support. The (110) crystal plane, acting as the strongest acidity plane of γ-alumina, was selected for modification. The supports modified with boron and phosphorus were successfully constructed, and the acid strengths were quantified by simulating the adsorption of the relevant probe molecules: pyridine in correlation with surface electronic properties via density functional theory. The surface adsorption energy calculation shows that the boron-modified surface is able to moderately reduce the adsorption capacity of alumina, while that of the surface modified by phosphorus is found to be enhanced over the sites of a tetrahedral coordination structure; however, at the other unsaturated Al sites, this is obviously reduced. The results of introducing electric fields imply that applying horizontal electric fields changes the surface acidity of alumina under the premise of a stable structure. With the enhancement of the horizontal electric fields, the adsorption capacity of tetra-coordination sites on the original surface gradually decreases, while those of the others gradually increases. However, for the boron-modified surface, introducing horizontal electric fields can reduce the adsorption capacity of all sites. Hence, microwave-electric-field-assisted modification of B further reduces the surface acidity of alumina, making it beneficial for deep hydrodesulfurization reactions.

1. Introduction

The exploitation and utilization of fossil energy have promoted the rapid development of the social economy while aggravating pollution and damaging the environment. The development and promotion of clean fuels are the focus of the petrochemical industry. The catalytic hydrodesulfurization (HDS) process plays an important role in removing sulfur from petroleum fractions as well as petrochemicals. The most widely used catalyst is the alumina-based Co-Mo-S catalyst; among the alumina structures, γ-Al2O3 is broadly used as a support due to its operable pore structure and satisfactory catalytic activity, providing an uphold for the active component to maintain a stable cluster structure on the catalyst surface. There are Lewis acid and Brønsted acid on the surface. The center of the former is aluminum ions with surface unsaturated coordination, and the latter is related to the degree of surface hydroxylation [1]. These acidic and basic sites have been looked upon as the active catalytic centers of γ-alumina [2].
When the active component attaches to γ-Al2O3 with a single or lower layer firmly, it induces lower catalytic activity. On the contrary, high layers of the active component stacks on the support are beneficial to accelerate the reaction process [3,4]. Therefore, weakening the interaction between the active component and support surface and reducing the surface acid strength are critical to maintaining the high activity of the catalyst.
Introducing modified elements, such as boron [5,6,7,8], phosphorus [9,10] and lanthanum [11], is effective to adjust the surface acidity of γ-Al2O3. The additives may improve the acidity of the γ-Al2O3 surface and the dispersion of active components. It also helps prolong the service life of the catalyst and improves the selectivity of HDS [12,13,14].
Microwave irradiation is advantageous in the field of materials due to its selective heating and non-thermal effects [15,16], including HDS technology [17,18]. Several pieces of research reveal that the active components of the catalyst prepared by the microwave method are uniformly distributed, and the load is significantly increased. It is also found that the HDS catalyst prepared by microwave has lower metal–support interaction (MSI) and higher reaction performance. Deep desulfurization requires a thorough understanding of the mechanism of how microwave affects the acidity of pure or modified surfaces and improves HDS efficiency.
The results of previous studies [19] on boron-modified Mo/γ-Al2O3 catalyst prepared by the microwave-assisted method show that the introduction of microwaves effectively weakens the interaction between the active component and the support on the modified catalyst and thus increases the stacking layer of active components. Moreover, the activities of the three- and four-layer active phases significantly increase from 14% and 1%, respectively, to 38% and 22%, and obviously, those of the monolayer and bilayer phases decrease. As the number of stacked layers being improved, the number of adsorbed active sites is effectively enhanced, which contributes to the enhancement of the rate and selectivity of the HDS reaction [20].
The density functional theory (DFT) method has been widely used to investigate solid acids [21]. The change of surface acidity can be effectively explained theoretically by judging the adsorption energy and charge transfer amount of probe molecules on clean and modified surfaces. At present, acid–base sites are usually detected by the adsorption of carbon monoxide [5] and pyridine [22,23]. Carbon monoxide is a weak Lewis base and is preferentially adsorbed. Pyridine is a stronger base, less sensitive to the strength of the site, and is used to detect Lewis or Brønsted acid sites [24].
Therefore, this work screened stable boron- and phosphorus-modified γ-A12O3 surface models via simulation methods, and a systematic and in-depth study of the electronic properties, acidity, and surface hydroxylation characteristics of the γ-A12O3 surface were conducted through DFT methods, so as to reveal its surface structural characteristics; moreover, the relationship between microstructure and macroscopic properties was analyzed. At the same time, the influence of a microwave electric field on the surface properties of alumina and modified alumina was discussed logically.

2. Computational Details and Methods

2.1. Details of Molecular Modeling

Numerous studies [25,26,27,28] confirm that the highest proportion of (110) surface is the most active plane in γ-Al2O3. In the present work, the γ-A12O3 model proposed by Digne et al. [24] was adopted, and the γ-A12O3 (110) surface was cut based on the principle of atomic integrity, with a periodic model of the five-layer structure established. The most stable configuration after optimization with the lowest energy was then selected as the research subject of investigation and analysis, as shown in Figure 1.
The pyridine molecule was used as the probe in Figure 2, showing the Mulliken charge of each atom in detail. It is clear that the molecule is electrically neutral. When pyridine adsorbs on the surface, the electrons of pyridine transfer to γ-A12O3, displaying positive electricity, which verifies the presence of acidic sites on the (110) surface.

2.2. Quantum Chemical Calculations

Molecular models were constructed through Visualizer, and all data were carried out using DMol3 via the Materials Studio package (MS).
A plane-wave basis set with medium energy cutoff was set. The generalized gradient approximation (GGA) with the Perde–Burke–Ernzerhof (PBE) exchange-correlation functional and the basis set using double numerical plus polarization (DNP) with an orbital cutoff of 4.4 Å in real space were utilized to produce highly accurate results [22,23]. And the spin state of the electrons was considered, but the structural symmetry was not used. The integration accuracy of Hamiltonian was set as Medium. And the self-consistent field (SCF) tolerance was set to Medium (10−5). The k-point was Medium with 2 × 1 × 2. All electrons in the calculation system were treated. Direct inversion of the iterative subspace (DIIS) size of 6 and thermal smearing of 0.005 Ha were employed for accelerating convergence. Moreover, the mixing charge density was set to 0.2 and the mixing spin density to 0.5. The convergence tolerances of energy, maximum force, and maximum displacement applied in geometry optimization were 210−5 Ha, 410−3 Ha/Å, and 0.005 Å, respectively.
(1)
The adsorption energy
The adsorption energy (Eads) is defined as the energy released by the adsorption of molecules or atoms from the gas phase on a solid surface [3]. The adsorption energy of the probe molecule on the surface was calculated by the following relationships:
r E a d s = E s u r f + p r o b e E s u r f E p r o b e
where E s u r f stands for the total energies of the surface, E p r o b e for the isolated gas-phase molecule, and E s u r f + p r o b e for the adsorbed molecule on the surface. A negative value indicates stable adsorption, corresponding to an exothermic process.
(2)
The displaced energy
The displaced energy determines the difficulty of replacing Al atoms on the surface of γ-Al2O3 with B and P. If the displaced energy is positive, the stability of the displaced crystal model becomes worse. Conversely, negative displace energy means that the reaction is spontaneous. The larger the absolute value, the easier the displacement process. For instance, the calculation method of the energy of B-γ-Al2O3 is as follows:
E d o p e = E B A l 2 O 3 + E A l E A l 2 O 3 E B
where E A l and E B are the energies of single aluminum and boron atoms, and E B A l 2 O 3 is the energy of the optimized B-γ-Al2O3.

3. Results and Discussion

3.1. Electronic Properties of the Modified Surfaces

Aluminum atoms were exposed in the upper two layers, including di-coordination, tetra-coordination, and penta-coordination, numbered Al1~Al8 (Figure S1). The coordination numbers of the aluminum atoms at various positions are given in Table S1. The doping energies of various sites were calculated to determine the best reaction sites. Sites with the lower doping formation energy are considered easy to replace. To investigate the effects of various sites and different coordination numbers, sites with the lowest doping formation energy were selected for research. As shown in Tables S2 and S3, the optimal doping sites for boron on the γ-Al2O3 (110) surface were Al1, Al3, and Al8 sites in turn, and those for phosphorus were Al1, Al2, and Al8. A comparison of the two tables indicates that B is more easily substituted for Al on the surface.
Mulliken population analysis refers to the charge density distributed on the atoms that make up the molecule. Table 1 displays the Mulliken charges of different atoms at substitution sites. The results indicate that electrons transfer from the aluminum, boron, and phosphorus to the adjacent oxygen. Comparison of the same site doped with different atoms demonstrates that Mulliken population charge is the largest for Al on the undoped surface, followed by P and B atoms, respectively, which suggests the high ionic character of Al-O bonds on the (110) surface. B (χ = 2.04) and P (χ = 2.19) presents a stronger electronegativity and ability to attract electrons than Al (χ = 1.61). The electron density at the modified site increased, which is not conducive to the adsorption of foreign molecules on it. Since the electronegativity of B is smaller than that of P, the adsorption capacity of the B modified surface is lower [29].
Moreover, partial density of states (PDOS) [30,31] curves for the γ-A12O3 (110) surface and the surfaces of A1 sites replaced by B and P of γ-A12O3 (110) were analyzed from −22.7 eV to 18.2 eV, as plotted in Figure 3.
For the γ-A12O3 (110) surface, a narrow peak from −21 eV to −15 eV and a wide peak from −8 eV to 0 eV can be observed. The PDOS was further analyzed to define the formation of the peaks. The narrow peak is found to form due to the hybridization of O-2s and Al-3p orbitals. The contribution to the broader peak at 0 eV comes from the O-2p and Al-3s orbitals.
From Figure 3b, on B-γ-A12O3 (110) surface, the narrow peak from −21 eV to −15 eV is composed of O-2s and B-3p orbitals, while the broad peak at −8 eV to 0 eV is composed of O-2p and B-2s orbitals. Between 5 eV and 17.5 eV, there is a relatively wide orbital composition of O-2p, Al-3p, and B-2s.
For the P-γ-A12O3 (110) surface, the narrow peak from −21 eV to −15 eV is composed of O-2s and P-3p orbitals. The broad peak at −8 eV to 0 eV consists of O-2p and Al-3s and P-3s orbits. There is a relatively broad orbital composition of O-2s, Al-3p, and P-3p, ranging from 5 eV to 17.5 eV.
From the above PDOS analysis, B and P are found to have hybridized and bonded with atoms of the γ-A12O3 (110) surface, indicating that the atoms of B and P were successfully displaced.

3.2. Effect of Acid Types

3.2.1. Lewis Acid Analysis

Pyridine, the basic probe molecule, was simulated for the adsorption on different γ-alumina acid sites. The adsorption energies were calculated to obtain the extrinsic acid strength at different surface sites. The Mulliken population charge reflects the electric property and quantity of the charge of the atom. The adsorption strength can be determined by analyzing the amount of transferred charge of the probe molecule.
Taking the adsorption of Al1 sites as an example, from Figure 4, the charges of pyridine are 0.075, 0.268, and 0.191, when adsorbed on the surfaces of pure γ-Al2O3, B-γ-Al2O3, and P-γ-Al2O3, respectively. This is because the electrons in pyridine molecules flow into the alumina surface during adsorption, indicating successful adsorption of pyridine on the alumina surfaces. A comparison of the charges at different active sites before and after adsorption confirms that pyridine electrons do flow into alumina, as shown in Table S4.
When pyridine is stably adsorbed on Al1 sites of pure, B-modified, and P-modified γ-Al2O3 (110) surfaces, the adsorption energies are −111.64 kJ/mol, −88.5 kJ/mol, and −165.85 kJ/mol, respectively. That is, pyridine is more easily adsorbed on P-γ-Al2O3 with stronger acidity. The adsorption capacity of the γ-Al2O3 (110) surface is weaker than that of the P-γ-Al2O3 (110) surface and stronger than the B-γ-Al2O3 (110) surface. This demonstrates that the B atom modification is more constructive to weaken the acidity over the Al1 adsorption site, whereas the P modification would enhance the acidity of the Al1 site, which is well in agreement with the reported literature [32,33].
Electronic transfers also occurred during the adsorption of pyridine over penta- and di-coordination sites, which again indicates that pyridine has successfully adsorbed on all three surfaces. Analysis of adsorption energy shows that the energy released by pyridine adsorption on the modified surface reduced, revealing that B and P modification is beneficial to reduce the acidity of the γ-Al2O3 (110) surface.
The adsorptions of pyridine on the surfaces were analyzed by replacing the penta-coordinated sites of Al atoms with B and P. From Table S8, the weakening ability via P modification is too strong, which would affect the adsorption of other substances. However, B can properly weaken the surface acidity without affecting the adsorption of other substances. The adsorption of pyridine at the di-coordination site is similar to the above rule.
Therefore, P modification enhances the acidity of the tetra-coordination sites and significantly reduces the acidity of the di- and penta-coordinated sites, resulting in too strong or too weak adsorption capacities. Conversely, B modification moderately reduces surface acidity.

3.2.2. Brønsted Acid Analysis

Since the aluminum surface is always hydroxylated under normal operating conditions, the presence of the hydroxyl group will change the acid–base properties of γ-alumina [34,35,36]. Hence, the structures of the hydroxyl group displaced on the original and modified γ-Al2O3 (110) surfaces at various adsorption sites were optimized to obtain stable configurations, and the adsorption energy of hydroxyl groups was calculated. The results are shown in Table 2, where the adsorption energies are less than −125.0 kJ/mol, indicating that they are of chemical adsorptions [37]. Additionally, hydroxyl is easy to adsorb on the P-γ-Al2O3 (110) surface, resulting in cover of the Lewis acid, and then the Brønsted acid site is provided. Therefore, only the properties of hydroxylated γ-Al2O3 and B-γ-Al2O3 (110) surfaces are analyzed below.
It can be seen from the data in Table 3 that Mulliken charges of the adsorption site over the hydroxylated γ-Al2O3 surfaces are higher than those on the surface not hydroxylated, and the surface acidity is accordingly enhanced. B-OH had a higher acidity than Al-OH [33]. The main reason is that the electronegativity of the oxygen atom within the hydroxyl group is relatively large, and the adsorption of the hydroxyl group therefore induces the electron flow into the hydroxyl group and increases the number of the positive charges over the adsorption site. However, from Table 4 based on the adsorption energies of pyridine on different surfaces, the opposite conclusion can be drawn: the acidity of hydroxylation surfaces is obviously weakened.
Therefore, the analysis of the surface structure of the hydroxylation species shows that the adsorption performance of the surface is different due to the change in Al-O bond length (In Figure 5). Mayer bond level analysis was performed to understand the Al-O and O-H bonds. It is clear that the shift of the hydroxyl group is the key to the decrease in surface acidity on the hydroxylated γ-Al2O3 surfaces.
It is concluded that the steric hindrance effect of the hydroxyl group affects the adsorption of pyridine, and Brønsted acid on the alumina surface is weaker. B effectively weakens the acidity of a hydroxylation surface and reduces its surface adsorption capacity.
Only when the surface of γ-Al2O3(110) is partially dehydroxylated, the Lewis acid site on the surface can appear, and its adsorption activity will be significantly improved.

3.3. Effects of the Electric Field

Compared to P, B-modified alumina has a better abatement in surface adsorption. Hereby, the effect of electric field on surface adsorption performance was explored.
The addition of the electric fields can change the internal charge arrangement of the alumina crystal system, which in turn would induce a change in the crystal structure. The electric fields were applied in different directions, including X, Y, Z, positive, and negative directions, as shown in Figure 6.

3.3.1. Pure γ-Al2O3 (110)

The calculations show that the surface of the alumina crystal is largely deformed by the perpendicular electric fields, which is not favorable for adsorption. Horizontal electric fields of different directions and intensities were applied to the γ-Al2O3 (110) surface, and Mulliken charge analysis was performed on the Al1, Al3, and Al8 sites of the surface. From Figure 7, the maximum field strengths for the stable pure crystal structure of alumina in the X and Y directions are 0.02 a.u. and 0.015 a.u. respectively.
When an electric field parallel to the X+ was applied, the charge of the Al1 remained almost constant, while those of the Al3 and Al8 sites changed in N-typed curves with increasing intensities of the electric field. The electric field in the Y- direction has an irregular effect on the sites’ charge, e.g., the charge at the Al1 site did not change significantly, while that over the Al8 site decreases and the Al3 site increases.
When the electric fields in the X- and Y+ directions are applied, it can be determined that the number of charges at different sites varies regularly with the electric field intensities. That is, the charge at the Al1 site decreases linearly, while the charges at the Al3 and Al8 sites increases linearly.
The decreases in the positive charge at the adsorption site is beneficial for reducing the adsorption capacity and vice versa for increasing it. Correspondingly, it is conjectured that the adsorption capacity of the Al1 site decreases and those of the Al3 and Al8 sites increases after applying the X- and Y+ horizontal electric fields. Figure 8 shows the structural model and Mulliken charge for the adsorption sites at the maximum field strength.
The adsorption energies of pyridine at Al1, Al3, and Al8 sites on the γ-Al2O3 surface were calculated under the electric fields of X- and Y+, as shown in Table 5 and Table 6.
As the electric field is enhanced, the absolute value of the adsorption energy at Al1 site gradually decreases, indicating that the adsorption capacity is gradually weakened. However, the absolute values of adsorption energy for Al3 and Al8 are enhanced, and the adsorption capacities also clearly increase. The variation of the adsorption energy of the pyridine molecule at different sites as a function of the field strength in the Y+ direction followed the same pattern as above.
The change in adsorption energy corresponds to the change in the amount of charge at the adsorption site before adsorption, suggesting that the charge at the adsorption site affects the surface adsorption capacity. Thus, under the assumption of a stable structure, an increase in the strength of electric field weakens the adsorption capacity at the Al1 site of the pure γ-Al2O3 surface and simultaneously enhances it at the Al3 and Al8 sites.

3.3.2. B-γ-Al2O3 (110)

The maximum field strengths that B-γ-Al2O3 withstood in the X and Y directions are 0.020 a.u. and 0.015 a.u., respectively, which coincides with those for the non-modification surface.
Calculation of PDOS of B-γ-Al2O3 analyzes whether the B atom still exists in the alumina structure after introducing the electric field and whether it has a good bonding state with the surrounding atoms.
Taking the PDOS of Al1 site on the B-γ-Al2O3 under the electric field of EX− = 0.020 a.u. and EY+ = 0.015 a.u. as examples, are compared with that of the Al1 site without an electric field.
As can be seen from Figure 9a,b, for instance, the relatively narrow peak appearing at around −20 eV to −15 eV in the Y+ under the applied electric field is a result of the hybrid bonding of the O-2s orbit and B-3p orbit. The dominant contributions to the broader peak at about −7.8 eV to 0 eV comes from the O-2p and B-2s orbitals. At about 0 eV to 17 eV, there is a very gentle broad peak consisting of B-3p and O-2s orbitals. Comparing the PDOS curves for the EX− and EY+ electric fields, the peak positions are found to be essentially the same, indicating that applying different directions and intensities of electric fields does not change the bonding states of the boron and oxygen atoms.
According to Figure 9c, there is a peak range from 0 eV to 17 eV at the PDOS curve of boron without the electric field. However, after the introduction of the electric field, it is found that there are two peaks at this position, with the new peaks locating at about 13 eV to 17 eV. The corresponding curves are relatively flat, implying that the electron delocalization become stronger. The introduction of the electric field disrupts the charge balance in the system and favors the electrons movement. The PDOS curve changes no significantly for oxygen atoms, which is likely due to the high electronegativity of the oxygen atoms and the low degree of delocalization of the surrounding electrons. To sum up, the addition of electric field does not cause significant changes in the structure of B-γ-Al2O3.
The Mulliken charges at the adsorption sites for different field intensities and directions are shown in Figure S7. It is not difficult to judge that the electric fields in the X+ and Y- directions have no advantage in weakening the adsorption capacity of the B-γ-Al2O3 (110) surface, while that in X- and Y+ may be beneficial to weakening.
The adsorption energies of the adsorption sites are obtained and are listed in Table 7 and Table 8. At increasingly strong electric fields, the energy released by pyridine adsorption become lower and the Al1 site changes dramatically. So, the X+ electric field weakens the adsorption ability of the B-γ-Al2O3 (110) surface, especially the Al1 site. A consistent conclusion is reached by analyzing the action of the Y- electric field.
The effects of X- and Y+ electric fields on pyridine adsorption on the γ-Al2O3 (110) and B-γ-Al2O3 (110) surfaces are compared in Figure 10. The positive half axis of the horizontal represents the electric field strength in the Y+ direction, and the negative half axis represents that in the X- direction.
It is easy to find that the electric fields in X- and Y+ directions weaken the adsorption capacity of the Al1 site on the γ-Al2O3 surface and enhance those of the Al3 and Al8 sites. Meanwhile, they effectively weaken the adsorption capacity of the adsorption site on the B-γ-Al2O3 surface, so as to weaken the surface acidity.
Because the direction of the microwave electric field is variable, the surface acidity can be effectively weakened when the horizontal electric field is applied to the B-γ-Al2O3 surface. For the pure γ-Al2O3 surface, the horizontal electric field has different effects on various active sites. Therefore, microwave-assisted B modification is beneficial to reduce the surface acidity of γ-Al2O3.

4. Conclusions

The DFT results show that the surface acidity modified by B has moderately better weakened acidity than that modified by P. The acidity of the hydroxylated surface is weaker, and Brønsted acid is much weaker than Lewis acid. The steric hindrance of hydroxyl affects the adsorption of pyridine. Similarly, the acidity of the hydroxylated surface is obviously reduced via B modification.
In addition, the introduction of electric fields effectively weakens the adsorption capacity of the B-γ-Al2O3 surface and has no advantage in the pure γ-Al2O3 surface. Through comprehensive comparison, it is found that microwave-assisted B modification is beneficial to reduce the surface acidity of γ-Al2O3.
In the future, molecular simulation and other means will be used to search for suitable and effective atoms in a wide range for the modification of γ-Al2O3. Finally, efficient HDS catalyst support will be prepared by experimental means and with microwave assistance.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/pr12102064/s1, Figure S1: γ-Al2O3 (110) surface structure model. Figure S2: Structural model of B doped at various positions. Figure S3: Structural model of P doped at various positions. Figure S4: The configuration and Mulliken charge number before and after adsorption. Figure S5: Alumina structure at field strength EZ = 0.002 a.u. Figure S6: Boron-modified alumina structure at field strength EZ = 0.002 a.u. Figure S7: Variation of Mulliken charge at adsorption sites in horizontal electric field. Figure S8: Energy screening of γ-Al2O3 (110) crystal face model. Table S1: Coordination numbers of Al at different positions. Table S2: Doping formation energy of B at various positions. Table S3: Doping formation energy of P at various positions. Table S4: Charges of pyridine adsorption at Al1 site. Table S5: Adsorption energy of pyridine adsorption at Al1 site. Table S6: Charges of various sites atoms before and after adsorption. Table S7: Mulliken charge of pyridine at four adsorption sites. Table S8: Charges at modified sites before and after adsorption. Table S9: Adsorption energy before and after three surface modifications (kJ/mol). Table S10: Charges of pyridine before and after modification. Table S11: Charges at modified sites before and after adsorption. Table S12: Adsorption energy before and after modification(kJ/mol). Table S13: Hydroxyl adsorption energy on pure surface. Table S14: Mulliken charges of adsorption sites. Table S15: Hydroxylated surface structure parameters before and after pyridine adsorption. Table S16: Structural parameters of boron modified hydroxylated surfaces before and after pyridine adsorption. Table S17: Structural model screening energy.

Author Contributions

X.F.: Conceptualization, formal analysis, investigation, writing—original draft preparation, writing—review and editing; T.L.: Conceptualization, methodology, formal analysis, software, visualization, validation; H.S.: Methodology, supervision, writing—reviewing and editing; Z.X.: Validation, resources, data curation; J.Y.: Investigation, formal analysis; A.D.: Methodology, correction. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (No. 21878330) and CNPC DQZX-KY-21-007.

Data Availability Statement

The original contributions presented in the study are included in the article/Supplementary Material, further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Chupas, P.J.; Chapman, K.W.; Halder, G.J. Elucidating the structure of surface acid sites on γ-Al2O3. J. Am. Chem. Soc. 2011, 133, 8522–8524. [Google Scholar] [CrossRef]
  2. Sanchez Escribano, V.; Garbarino, G.; Finocchio, E.; Busca, G. γ-Alumina and amorphous silica–alumina: Structural features, acid sites and the role of adsorbed water. Top. Catal. 2017, 60, 1554–1564. [Google Scholar] [CrossRef]
  3. Zhang, Q.; Shang, H.; Xue, Z.; Duan, A. The effect of microwave electric field on sulfur vacancies formation over the edge sites of Co/Ni-promoted and unpromoted MoS2 catalysts through DFT investigations. Fuel 2022, 318, 123553. [Google Scholar] [CrossRef]
  4. Baeza, P.; Villarroel, M.; Villanueva, R.; Gil-Llambías, F.J.; Avila, P.R.T. Synergism of extruded and monolithic Co–Mo/γ-AL2O3–sepiolite catalytic systems in the hydrodesulphurization reaction. Top. Catal. 2016, 59, 252–258. [Google Scholar] [CrossRef]
  5. Fu, R. The preparation of Boron-doped activated Alumina by impregnating boron in the period of boehmite. MATEC Web Conf. 2021, 353, 01026. [Google Scholar] [CrossRef]
  6. Chen, Z.; Wang, J.; Yang, Z.; Tang, Z.; Jiang, H. Effect of boron promoter on the performance of NiMo/γ-Al2O3 hydrotreating catalyst. Acta Pet. Sin. 2016, 32, 56–63. [Google Scholar] [CrossRef]
  7. Maity, S.; Lemus, M.; Ancheyta, J. Effect of preparation methods and content of boron on hydrotreating catalytic activity. Energy Fuels 2011, 25, 3100–3107. [Google Scholar] [CrossRef]
  8. Vatutina, Y.V.; Klimov, O.V.; Nadeina, K.A.; Danilova, I.G.; Gerasimov, E.Y.; Prosvirin, I.P.; Noskov, A.S. Influence of boron addition to alumina support by kneading on morphology and activity of HDS catalysts. Appl. Catal. B Environ. 2016, 199, 23–32. [Google Scholar] [CrossRef]
  9. Xia, B.; Cao, L.; Luo, K.; Zhao, L.; Wang, X.; Gao, J.; Xu, C.J.E. Effects of the active phase of CoMo/γ-Al2O3 catalysts modified using cerium and phosphorus on the HDS performance for FCC gasoline. Energy Fuels 2019, 33, 4462–4473. [Google Scholar] [CrossRef]
  10. Shan, S.; Liu, H.; Shi, G.; Bao, X. Tuning of the active phase structure and hydrofining performance of alumina-supported tri-metallic WMoNi catalysts via phosphorus incorporation. Front. Chem. Sci. Eng. 2018, 12, 59–69. [Google Scholar] [CrossRef]
  11. Yue, Y. Effect on the Surface Acidity of the Lanthanum Modified γ-Al2O3; China University of Petroleum-Beijing: Beijing, China, 2018. [Google Scholar]
  12. Abbasi, A.; Karimi, A.; Aghabozorg, H.; Sadighi, S.; Aval, M.A. Cobalt-promoted MoS2 nanosheets: A promising novel diesel hydrodesulfurization catalyst. Int. J. Chem. Kinet. 2020, 52, 159–166. [Google Scholar] [CrossRef]
  13. Wang, Z.; Wu, P.; Lan, L.; Liu, K.; Hu, Y.; Ji, S. Preparation, characterization and hydrodesulfurization performances of Co-Ni2P/SBA-15 catalysts. J. Energy Chem. 2015, 24, 185–192. [Google Scholar] [CrossRef]
  14. Yan, X.; Zhang, Z.; Li, Z.; Ji, H. Effect of B, P, K elements modification on properties of alumina support and Ni-Mo catalyst. Chem. React. Eng. Technol. 2020, 36, 230–238. [Google Scholar]
  15. Peng, R.; Li, Y.; Peng, Y.; Su, H.; Lu, Y.; Chen, M.; Huang, F.; Zhang, H. Sintering and microwave dielectric properties of LiCaPO4 ceramic: Experiment and first principle calculation. J. Mater. Res. Technol. 2020, 9, 13988–13993. [Google Scholar] [CrossRef]
  16. Peng, R.; Li, Y.; Su, H.; Lu, Y.; Shi, L.; Yun, Y.; Liao, B. The modification of sintering and microwave dielectric properties of Mn2+ doped LiZnPO4 ceramic. J. Mater. Res. Technol. 2020, 9, 4994–5006. [Google Scholar] [CrossRef]
  17. Shang, H.; Zhang, H.; Du, W.; Liu, Z. Development of microwave assisted oxidative desulfurization of petroleum oils: A review. J. Ind. Eng. Chem. 2013, 19, 1426–1432. [Google Scholar] [CrossRef]
  18. Zhang, X.; Hayward, D.O. Applications of microwave dielectric heating in environment-related heterogeneous gas-phase catalytic systems. Inorganica Chim. Acta 2006, 359, 3421–3433. [Google Scholar] [CrossRef]
  19. Shang, H.; Guo, C.; Ye, P.; Zhang, W. Synthesis of boron modified CoMo/Al2O3 catalyst under different heating methods and its gasoline hydrodesulfurization performance. Front. Chem. Sci. Eng. 2021, 15, 1088–1098. [Google Scholar] [CrossRef]
  20. Shang, H.; Ye, P.; Yan, H.; Zhang, W. Microwave enhanceing preparation hydrodesulfurization performance study on γ-Al2O3 supporter. Vac. Electron. 2018, 1, 33–38. [Google Scholar] [CrossRef]
  21. Salam, M.A.; Abdullah, B. Catalysis mechanism of Pd-promoted γ-alumina in the thermal decomposition of methane to hydrogen: A density functional theory study. Mater. Chem. Phys. 2017, 188, 18–23. [Google Scholar] [CrossRef]
  22. Zhang, P.; Zhao, Q.; Liu, J.; Yang, B. Research on inhibitors and hindered groups in ultra-deep hydrodesulfurization based on density functional theory. Catal. Today 2018, 314, 170–178. [Google Scholar] [CrossRef]
  23. Morterra, C.; Magnacca, G. A case study: Surface chemistry and surface structure of catalytic aluminas, as studied by vibrational spectroscopy of adsorbed species. Catal. Today 1996, 27, 497–532. [Google Scholar] [CrossRef]
  24. Digne, M.; Sautet, P.; Raybaud, P.; Euzen, P.; Toulhoat, H. Use of DFT to achieve a rational understanding of acid–basic properties of γ-alumina surfaces. J. Catal. 2004, 226, 54–68. [Google Scholar] [CrossRef]
  25. Stuart, N.M.; Sohlberg, K. The microstructure of γ-Alumina. Energies 2021, 14, 6472. [Google Scholar] [CrossRef]
  26. Svarovskaya, N.; Glazkova, E.; Bakina, O.; Kazantsev, S.; Lozhkomoev, A.; Lerner, M. Hierarchical γ-alumina: From pure phase to nanocomposites. Recent Pat. Nanotechnol. 2020, 14, 92–101. [Google Scholar] [CrossRef]
  27. Drecun, O.; Striolo, A.; Bernardini, C.; Sarwar, M. Hydration structures on γ-Alumina surfaces with and without electrolytes probed by atomistic molecular dynamics simulations. J. Phys. Chem. B 2022, 126, 9105–9122. [Google Scholar] [CrossRef]
  28. Li, H.; Xue, H.; Chu, B.; Ma, Q.; He, H. Promoted activity of surface hydroxyls on γ-Al2O3 mineral dust with the coexistence of SO2)and NH3. J. Phys. Chem. Lett. 2022, 13, 10335–10341. [Google Scholar] [CrossRef]
  29. Yu, Z.; Huang, W.; Wang, X.; Deng, K.; Wei, Q.; Zhou, Y. B-doped Al2O3@C support for CoMo hydrodesulfurization catalyst and their hydrodesulfurization performance. Chem. Ind. Eng. Prog. 2022, 42, 3550. [Google Scholar] [CrossRef]
  30. Cao, F.; Su, S.; Xiang, J.; Sun, L.; Hu, S.; Zhao, Q.; Wang, P.; Lei, S. Density functional study of adsorption properties of NO and NH3 over CuO/γ-Al2O3 catalyst. Appl. Surf. Sci. 2012, 261, 659–664. [Google Scholar] [CrossRef]
  31. Gu, J.; Wang, J.; Leszczynski, J. Co and Ni single sites on the (111)n surface of γ-Al2O3—a periodic boundary DFT study. Ind. Chem. Mater. 2023, 1, 117–128. [Google Scholar] [CrossRef]
  32. Kang, H.; Wang, D.; Zhang, L. Dispersion of cobalt and molybdenum on phosphorus or potassium modified γ-Al2O3 support. Chin. J. Appl. Chem. 1997, 14, 40–43. [Google Scholar] [CrossRef]
  33. Zhao, Y. Acdity of alumina, modified alumina and alumina silicate. Ind. Catal. 2002, 10, 54–58. [Google Scholar]
  34. Suganuma, Y.; Elliott, J.A. Effect of varying stiffness and functionalization on the interfacial failure behavior of isotactic polypropylene on hydroxylated γ-Al2O3 by MD simulation. ACS Appl. Mater. Interfaces 2023, 15, 6133–6141. [Google Scholar] [CrossRef] [PubMed]
  35. Bai, B.; Bai, H.; Zhang, L.; Cao, H.-J.; Gao, Z.-H.; Huang, W. Insight into the formation mechanism of CC chain in ethanol synthesis at the interface of partially hydroxylated γ-Al2O3 (110D) surface and polyethylene glycol solvent. Mol. Catal. 2018, 455, 164–178. [Google Scholar] [CrossRef]
  36. Wang, J.; Yu, X.; Zhang, H.; Li, J.; Sun, K.; Guo, H.; Chen, Q.; Long, M. Hydroxylation and fluorination on γ-Al2O3(110) surfaces: First-principles and transition state calculations. J. Fluor. Chem. 2021, 249, 109840. [Google Scholar] [CrossRef]
  37. Yu, Y. Molecular Simulation and Experimental Study on Ethanol Dehydration to Ethylene on γ-Al2O3. Tianjin University: Tianjin, China, 2012. [Google Scholar]
Figure 1. Optimized structure of the γ-Al2O3 (110) surface.
Figure 1. Optimized structure of the γ-Al2O3 (110) surface.
Processes 12 02064 g001
Figure 2. Model and the Mulliken charge of pyridine.
Figure 2. Model and the Mulliken charge of pyridine.
Processes 12 02064 g002
Figure 3. PDOS plots for (a) γ-Al2O3, (b) B-γ-Al2O3, and (c) P-γ-Al2O3.
Figure 3. PDOS plots for (a) γ-Al2O3, (b) B-γ-Al2O3, and (c) P-γ-Al2O3.
Processes 12 02064 g003
Figure 4. Configurations and Mulliken charge of pyridine at Al1 sites of different surfaces before and after adsorption. (a) γ-Al2O3 (110) surface; (b) pyridine adsorption on the γ-Al2O3 (110) surface; (c) B-γ-Al2O3 (110) surface; (d) pyridine adsorption on the B-γ-Al2O3 (110) surface; (e) P-γ-Al2O3 (110) surface; (f) pyridine adsorption on the P-γ-Al2O3 (110) surface.
Figure 4. Configurations and Mulliken charge of pyridine at Al1 sites of different surfaces before and after adsorption. (a) γ-Al2O3 (110) surface; (b) pyridine adsorption on the γ-Al2O3 (110) surface; (c) B-γ-Al2O3 (110) surface; (d) pyridine adsorption on the B-γ-Al2O3 (110) surface; (e) P-γ-Al2O3 (110) surface; (f) pyridine adsorption on the P-γ-Al2O3 (110) surface.
Processes 12 02064 g004aProcesses 12 02064 g004b
Figure 5. Adsorption of pyridine on hydroxylated B-γ-Al2O3. (a) Al1 site before adsorption; (b) Al1 site after adsorption; (c) Al3 site before adsorption; (d) Al3 site after adsorption; (e) Al8 site before adsorption; (f) Al8 site after adsorption.
Figure 5. Adsorption of pyridine on hydroxylated B-γ-Al2O3. (a) Al1 site before adsorption; (b) Al1 site after adsorption; (c) Al3 site before adsorption; (d) Al3 site after adsorption; (e) Al8 site before adsorption; (f) Al8 site after adsorption.
Processes 12 02064 g005aProcesses 12 02064 g005b
Figure 6. Schematic diagrams of the direction of electric field application. (a) Horizontal electric field; (b) vertical electric field.
Figure 6. Schematic diagrams of the direction of electric field application. (a) Horizontal electric field; (b) vertical electric field.
Processes 12 02064 g006
Figure 7. Mulliken charge at different adsorption sites in the horizontal electric field.
Figure 7. Mulliken charge at different adsorption sites in the horizontal electric field.
Processes 12 02064 g007
Figure 8. Structural model and Mulliken charge of adsorption sites under the horizontal electric field. (a) EX+ = 0.020 structural model; (b) EX− = 0.020 structural model; (c) EY+ = 0.015 structural model; (d) EY− = 0.015 structural model.
Figure 8. Structural model and Mulliken charge of adsorption sites under the horizontal electric field. (a) EX+ = 0.020 structural model; (b) EX− = 0.020 structural model; (c) EY+ = 0.015 structural model; (d) EY− = 0.015 structural model.
Processes 12 02064 g008
Figure 9. PDOS of B-γ-Al2O3 at Al1 under various electric fields. (a) EY+ = 0.015 a.u. (b) EX− = 0.020 a.u. (c) without electric fields.
Figure 9. PDOS of B-γ-Al2O3 at Al1 under various electric fields. (a) EY+ = 0.015 a.u. (b) EX− = 0.020 a.u. (c) without electric fields.
Processes 12 02064 g009aProcesses 12 02064 g009b
Figure 10. Adsorption energy of various adsorption sites under the electric field.
Figure 10. Adsorption energy of various adsorption sites under the electric field.
Processes 12 02064 g010
Table 1. Mulliken charges of different atoms at substitution sites.
Table 1. Mulliken charges of different atoms at substitution sites.
Substitution SitesMulliken Charge
AlBP
Al11.6150.7741.164
Al21.6991.292
Al31.7121.022
Al81.7271.0141.419
Table 2. Hydroxyl adsorption energy on the modified surfaces.
Table 2. Hydroxyl adsorption energy on the modified surfaces.
Absorption SitesAdsorption Energy (kJ/mol)
γ-Al2O3B-γ-Al2O3P-γ-Al2O3
1−152.26−134.30−404.67
2−185.42--−246.22
3−267.59−156.77--
8−265.58−208.86−338.98
Table 3. Mulliken charges for different adsorption sites on various surfaces.
Table 3. Mulliken charges for different adsorption sites on various surfaces.
Absorption Sitesγ-Al2O3B-γ-Al2O3
PureHydroxylatedPureHydroxylated
11.7071.7830.7740.854
31.8382.0361.0231.804
81.8532.0181.0131.806
Table 4. Adsorption energy of pyridine on various surfaces (kJ/mol).
Table 4. Adsorption energy of pyridine on various surfaces (kJ/mol).
Absorption Sitesγ-Al2O3B-γ-Al2O3
PureHydroxylatedPureHydroxylated
1−111.69−107.40−88.54−25.58
3−214.05−60.60−163.97−125.12
8−209.47−15.76−140.77−96.50
Table 5. Adsorption energy in the X- electric field.
Table 5. Adsorption energy in the X- electric field.
Absorption SitesAdsorption Energy (kJ/mol)
EX− = 0EX− = 0.005EX− = 0.010EX− = 0.015EX− = 0.020
1−111.64−109.43−107.14−104.85−103.26
3−214.05−216.97−220.57−223.92−226.80
8−209.47−211.28−214.76−216.52−218.95
Table 6. Adsorption energy in the Y+ electric field.
Table 6. Adsorption energy in the Y+ electric field.
Absorption SitesAdsorption Energy (kJ/mol)
EY+ = 0EY+ = 0.005EY+ = 0.010EY+ = 0.015
1−111.64−107.15−101.36−96.34
3−214.05−217.65−219.95−223.25
8−209.47−212.11−214.83−216.49
Table 7. Adsorption energy at the X- electric field.
Table 7. Adsorption energy at the X- electric field.
Absorption SitesAdsorption Energy (kJ/mol)
EX− = 0EX− = 0.005EX− = 0.01EX− = 0.015EX− = 0.02
1−88.50−83.97−74.63−68.14−64.72
3−163.97−161.34−157.75−154.47−153.29
8−140.77−137.92−135.46−133.15−131.82
Table 8. Adsorption energy at the Y+ electric field.
Table 8. Adsorption energy at the Y+ electric field.
Absorption SitesAdsorption Energy (kJ/mol)
EY+ = 0EY+ = 0.005EY+ = 0.010EY+ = 0.015
1−88.50−76.72−63.23−54.20
3−163.97−161.47−158.69−156.53
8−140.77−137.15−133.61−129.39
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Fan, X.; Li, T.; Shang, H.; Xue, Z.; Yang, J.; Duan, A. Insights into the Effect of a Microwave Field on the Properties of Modified γ-Alumina: A DFT Study. Processes 2024, 12, 2064. https://doi.org/10.3390/pr12102064

AMA Style

Fan X, Li T, Shang H, Xue Z, Yang J, Duan A. Insights into the Effect of a Microwave Field on the Properties of Modified γ-Alumina: A DFT Study. Processes. 2024; 12(10):2064. https://doi.org/10.3390/pr12102064

Chicago/Turabian Style

Fan, Xiayu, Tong Li, Hui Shang, Zonghao Xue, Jie Yang, and Aijun Duan. 2024. "Insights into the Effect of a Microwave Field on the Properties of Modified γ-Alumina: A DFT Study" Processes 12, no. 10: 2064. https://doi.org/10.3390/pr12102064

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

Article Metrics

Back to TopTop