Next Article in Journal
3D FRP Reinforcement Systems for Concrete Beams: Innovation towards High Performance Concrete Structures
Previous Article in Journal
Study of the Rolling Effect on MoS2–Carbon Fiber Density and Its Consequences for the Functionality of Li-Ion Batteries
Previous Article in Special Issue
Changing Template/Al2O3 Ratio in Reaction Gel—An Effective Way to Regulate Nature of Intermediate Phases and Properties of SAPO-11 Molecular Sieves during Crystallization
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Understanding the Structural and Catalytic Properties of Al(IV)-2 Acidic Sites of ZSM-5

1
School of Materials, Sun Yat-sen University, Shenzhen 518107, China
2
Lanzhou Petrochemical Research Center, Petrochemical Research Institute, PetroChina, Lanzhou 730060, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Materials 2024, 17(12), 2824; https://doi.org/10.3390/ma17122824
Submission received: 14 May 2024 / Revised: 5 June 2024 / Accepted: 5 June 2024 / Published: 10 June 2024
(This article belongs to the Special Issue Zeolitic Materials: Structure, Properties, and Applications II)

Abstract

:
It is crucial to identify the structures of active sites to understand how catalysts function and to use that understanding to develop better catalytic materials. ZSM-5 zeolites with dominant Al(IV)-2 sites have been developed in this work. 1H-27Al 2D HMQC and 2D 1H TQ(DQ)-SQ NMR experiments have been performed to investigate the structural properties of this acidic site. The Al(IV)-2 sites have Brønsted and Lewis acid characteristics. The catalytic performance of Al(IV)-2 sites has been tested by n-dodecane cracking reactions. The catalytic results show that the Brønsted acidic strength of the Al(IV)-2 sites is comparable to that of the Al(IV)-1 sites, but the Al(IV)-2 sites’ Lewis acid characteristics provide extra catalytic activity. We have gained valuable insights into the characteristics of Al(IV)-2 acid sites within these materials.

1. Introduction

Due to the large consumption of light olefins and petrochemicals for producing plastics, textiles, drug intermediates, and fine chemicals, there will continue to be a massive demand-supply disparity. Utilizing heavy oil feedstocks to convert them into lighter hydrocarbons to produce refined fuels and petrochemicals is becoming more attractive [1,2]. One of the most used zeolite-based catalysts for producing lighter hydrocarbons from n-dodecane by steam catalytic conversion is ZSM-5 zeolite [3,4]. ZSM-5 was first synthesized in 1965 by Landolt and Argauer [5]. Because of its high selectivity and relatively lengthy cycle, ZSM-5 is one of the most studied zeolites and has gained a great deal of interest in both basic research and industrial applications [6,7]. Interlinked sinusoidal and straight channels comprise the 3D pore structure of the ZSM-5 zeolite, which is greatly wanted in catalysis for providing distinct diffusion paths for reactants and products [8]. The tetrahedrally coordinated framework Al atoms in zeolites create Brønsted acid sites through the charge-balancing function of a proton (Si-OH-Al). These bridging acid sites are located in the cavities with specific shapes and scales and only allow reactants with suitable sizes and geometries to approach. Therefore, Brønsted acid sites in H-form zeolites are essential for their spectacular selectivity. The success of zeolites as heterogeneous catalysts is strongly affected by their acidic characteristics. The structure and distribution of Al species in the catalyst framework are important factors in understanding zeolite structure-activity correlations [9,10,11]. This area lacks knowledge due to the lack of zeolites with controlled Al distribution and precise descriptions of Al distribution. This results from the lack of an experimental method to identify precisely where the aluminum is located within the framework. In zeolite science, characterizing Brønsted acid sites and establishing the relationship between structure and activity are crucial for advancing the utilization of zeolites as catalysts. Davis and his coworkers have conducted extensive research to understand the influence of factors such as the Si/Al ratio, crystallization conditions, and post-synthesis treatments on acid site properties [12,13,14,15,16]. This understanding is essential and a significant step toward enhancing catalyst performance through optimization. Although numerous researches examining the impact of various factors on the catalytic performance of ZSM-5 zeolite are available, studies investigating the structure-property correlation resulting from the acidity are still desired.
Solid-state NMR spectroscopy has developed over the last few decades into a powerful tool for zeolite research as a high-sensitivity characterizing method for the local structure of resonant nuclei. The type, strength, accessibility, and concentration of acid sites can be determined through solid-state NMR spectroscopy [17,18]. Directly investigating hydroxyl protons acting as catalytically active Brønsted acid sites is feasible by 1H MAS NMR spectroscopy. There is currently a lot of research in the literature to determine if single active sites, multiple sites, or distribution of active acid sites exist in this commercially and academically important catalyst [19,20,21]. However, it is unlikely for 1H MAS NMR spectra of bare acidic zeolites to offer precise details regarding the various types and locations of the Brønsted acid protons. Two-dimensional (2D) correlation experiments are necessary to obtain comprehensive structural information about complicated systems, such as the local configurations of atoms concerning one another. Through spy spin-1/2 nuclei, the sequence of HMQC has been proposed as an effective means [22,23,24]. Since there are only two radio frequency pulses on the quadrupolar channel in this sequence, there is less complexity, which improves the efficiency of magnetization transfer. Due to their high gyromagnetic ratio, protons can be used as spy nuclei to increase sensitivity significantly, but this creates a new issue because of the broad 1H line shapes that result from the strong 1H-1H homonuclear dipolar couplings. Modern advancements in fast MAS technology enable the proton-detected D-HMQC (‘D’ means dipolar). In the current work, 1H/{27Al} D-HMQC experiments at very fast MAS were performed to study 1H-27Al heteronuclear correlation in ZSM-5 zeolites.
Through applying two-dimensional 27Al{1H} correlation techniques coupled with a series-connected hybrid (SCH) magnet at 35.2 T, acidic ZSM-5 catalysts can have at least two framework Al(IV) sites with hydroxyl groups [11,25,26]. In addition to the known Al(IV) at the framework bridging acid site (BAS), the second tetrahedrally coordinated Al site (denoted Al(IV)-2) experiences an increased chemical shift relative to the BAS. The second site arises from the partially bonded framework (SiO)4−n-Al(OH)n species, significantly improving catalytic reactivity in benzene hydride-transfer and n-hexane cracking processes. For extending zeolite catalysts to more complex feedstocks, this new structure information is paramount to understanding the function of BAS in HZSM-5 catalysts. They were created either by framework crystallization flaws or insufficient post-synthetic hydrolysis of a framework Al before extra-framework Al was created. The Al siting in the zeolite frameworks was demonstrated to depend on the hydrothermal synthesis conditions rather than being random or regulated by simple rules [27,28]. It is well-known that the strength and density of the acid sites in ZSM-5 can indeed be adjusted or tuned through various synthesis conditions. For example, the Si/Al ratio in the synthesis gel affects the acidity of ZSM-5 and the choice and concentration of mineralizing agents (e.g., sodium hydroxide, tetrapropylammonium hydroxide) in the synthesis gel affect the pH and composition of the gel, which in turn influence the formation of ZSM-5 crystals and the acidity of the resulting material. Therefore, in our work, three different synthesis methods have been used to prepare ZSM-5 zeolites for identifying the acid sites in ZSM-5 [29,30,31]. In the present work, we presented an approach to obtain ZSM-5 with predominate Al(IV)-2 sites and found out that the Al(IV)-2 site is affected by the presence of TPA+ in the synthesis gel. Z-I was synthesized using Na cations and ZSM-5 seeds without TPA+ cations. TPA+ cations were used as OSDAs to synthesize high silica ZSM-5 and nano ZSM-5; neither of these materials’ frameworks contain Al(IV)-2 sites.

2. Experimental Section

2.1. Zeolite Syntheses

ZSM-5 zeolites with dominant Al(IV)-2 acid sites: The preparation of a synthesis gel involved adding Na2SiO4·18H2O and Al2(SO4)2 to an aqueous solution, followed by the addition of ZSM-5 seed crystals. The solution was then stirred at room temperature for one hour. The molar composition of the synthesis gel was Al2O3:45SiO2:4.5Na2O:730H2O. The resulting dense gel was placed in 100 mL Teflon-lined stainless-steel autoclaves and statically heated at 185 °C and under autogenous pressure for 32 h. The solid products were obtained through filtration, washed with water and ethanol, and dried at 100 °C overnight. These ZSM-5 zeolites were labeled as Z-I.
High silica ZSM-5 zeolites: An aqueous solution containing tetrapropylammonium hydroxide was mixed with Na2SiO4 and Al2(SO4)2 in a molar ratio (Si/Al) of 550. Then, 1M H2SO4 was added to attain a pH of 7–8. The synthesis reactions took place in stainless-steel reactors equipped with Teflon cups and were conducted at 180 °C for 32 h. These resulting ZSM-5 zeolites were labeled as Z-II.
Nano ZSM-5 zeolites: The first step is preparing the solution with seeds. The synthesis involved preparing mixtures with a molar composition expressed in terms of oxide ratios (10 (TPA)2O:60SiO2: 0.5Al2O3:936H2O). This was achieved by adding aluminum sulfate, sodium silicate, and tetrapropyl-ammonium hydroxide in water, stirring for 10 min. The clear solutions were poured into 100 mL Teflon-lined stainless-steel autoclaves and heated statically at 120 °C for 24 h, forming 60 mL of gel. In the second step, new synthesis mixtures were created by adding NaAlO2 and Na2SiO4 to an aqueous sodium hydroxide solution. The molar composition of the synthesis gel was Al2O3:45SiO2:4.5Na2O:730H2O. The solution was stirred for one hour at room temperature, after which 5 mL of the gel with seed crystals was added dropwise. The resulting gel was placed in 100 mL Teflon-lined stainless-steel autoclaves and heated statically at 120 °C for 32 h. The solid products were obtained through centrifugation, washed with water and ethanol, and dried at 100 °C overnight. These resulting ZSM-5 zeolites were designated as Z-III.
Z-I and Z-II were ion-exchanged with a 1 M NH4Cl solution 3 times at 80 °C. Then, the solid materials were filtered out, washed with deionized water, and dried at 100 °C. HZSM-5 was obtained by calcination in air at 823 K for 2 h after heating at a 2 K/min rate. However, Z-III is not necessary in these cases, especially when organic TPA acts as the SDAs; in such cases, HZSM-5 can be directly formed after calcination in the air at 823 K.

2.2. Catalyst Characterization

A Bruker Avance 600 NMR spectrometer (Sun Yat-sen University, Guangzhou, China) was used to record nuclear magnetic resonance (NMR) spectra. A typical Bruker MAS probe head with a rotor diameter of 1.9 mm and a spinning frequency of 25 kHz was employed for the 27Al magic angle spinning (MAS) NMR. Using a single 1 µs excitation pulse and a 1 s interscan delay, the 27Al chemical shift was referenced to a saturated Al(NO3)3 solution. 1H/{27Al} D-HMQC experiments were performed under a MAS frequency of 25 kHz and a recycle delay of 1.5 s. For 1H, the lengths of π/2 and π pulses were 0.9 and 1.8 µs, respectively. SR412 was used as the recoupling sequence to reintroduce 1H-27Al heteronuclear dipolar interaction. The 1H rf field strength during SR412 was adjusted to 50 kHz, two times the spinning frequency of 25 kHz, and the total recoupling duration was optimized. For 27Al, the rf field strength was calibrated using 1 M Al(NO3)3 aqueous solution. The rf field strength for hard pulse was 250 kHz, while it was set to 1.5 kHz for soft pulse.
Further, 2D 1H DQ(TQ)-SQ NMR experiments were carried out using a 1.9 mm MAS probe at 25 kHz spin rates, typical π/2-pulse lengths of 1.5 μs, and a recycle delay of 5 s. The spectra were referenced concerning tetramethylsilane (TMS) using solid tetrakis(trimethylsilyl)silane as a secondary standard (0.27 ppm for 1H). The symmetry pulses were used for DQ(TQ) excitation/reconversion, which has been demonstrated to be an efficient homonuclear DQ recoupling scheme at fast MAS rates.
All NMR spectra are presented in the Supplementary Materials.
X-ray powder diffraction (XRD): XRD patterns were measured with a Rigaku Ultimate VI X-ray diffractometer (40 kV, 40 mA) using Cu-Kα1 (λ = 1.5406 Å) radiation (Wilmington, MA, USA).
Scanning electron microscopy (SEM): SEM measurements were carried out on a Hitachi Regulus SU8230 FE-SEM (Tokyo, Japan).

2.3. Catalytic Activity Measurements

The zeolites were hydrothermally deactivated for 4 h at 800 °C with 100% steam before the test, respectively. Then, the activity of the zeolite was evaluated by using a microactivity test (MAT) unit (MAT-Ⅱ), which was designed according to the ASTM D-3907 method using a dodecane (analytical purity). The tests were carried out at a reaction temperature of 460 °C, a reacting time of 70 s, and a catalyst-to-oil ratio of 3.2:1. The liquid product was analyzed by Varian 3800 (Palo Alto, CA, USA), and the average cracking activity of the zeolite was obtained. The quality of a fuel’s ignition is measured by its octane number. Two methods for determining the octane number are the research octane number (RON) and the motor octane number (MON). In this study, the RON and MON of the liquid product were determined. The composition of the liquid product and its octane number were analyzed using an Agilent 8890 instrument (Santa Clara, CA, USA).

3. Results and Discussion

The XRD powder patterns and SEM photographs of ZSM-5 samples are shown in Figure 1. The X-ray diffraction pattern shows the crystalline samples with reflections that agree with the pattern of standard ZSM-5 zeolite, and the SEM photographs of Z-I and Z-II are coffin-shaped particles. The composition of the synthesis solution, including the types and concentrations of alkali metals, fluoride ions, organic structure-directing agents (OSDAs), or seeding agents, can significantly impact the morphology and properties of ZSM-5. In this work, three popular hydrothermal synthesis methods have been employed to prepare ZSM-5 zeolite, and XRD patterns and morphologies of ZSM-5 reflect the synthesized conditions. Figure 2 shows the 1H and 27Al MAS NMR spectra of ZSM-5 samples. Modern solid-state NMR spectrometers can distinguish the various types of hydroxyl groups by their chemical shifts due to the resolution of their 1H MAS NMR spectra. Signals between 1.2 and 2.2 ppm point to the presence of SiOH groups, for example, on the outside of silicate and aluminosilicate particles or in zeolite framework defects [32,33], i.e., these surface hydroxyl groups can originate from the hydrolysis of silanol groups (Si-OH) located on the external surface or within the channels of the inner surface of the zeolite. The variation in chemical shifts may arise from their location on the external surface or within the zeolite micropores. Due to bridging OH groups in large cages and tiny structural units, the Si-OH-Al group 1H MAS NMR signals can be observed at 3.6–4.3 and 4.6–5.2 ppm [34]. Therefore, strongly acidic hydroxy groups in zeolites are characterized by peaks in the region of 3.6–6.0 ppm in the 1H MAS NMR. Signals around 4.0–4.5 and 2.5–2.9 ppm in the 1H solid-state NMR spectra of dehydrated H-ZSM-5 have been employed as indicators for the presence of hydroxyl groups on bridging acid sites (BAS) and extra-framework aluminum (EFAl) species, respectively. NMR techniques based on heteronuclear dipolar (through space) interactions between 1H and 27Al nuclei are needed to assign 1H chemical shifts to specific bonding situations. A recently developed NMR pulse sequence called D-HMQC is very well suited for examining the spatial proximity of quadrupolar and spin-1/2 nuclei [24]. Figure 3b demonstrates that the tetrahedral Al site is spatially close to three chemically different protons at 2.68, 1.19, and 0.87 ppm. Based on extensive literature reporting known 1H chemical shift values [18], the chemical shift at 0.87 ppm and 1.19 ppm can be attributed to the silanol groups on the external surface and the lattice defects, respectively. As shown in Figure 3b, these silanol groups are near framework aluminum species [34]. In the literature, the 2.68 ppm 1H signal has been attributed to hydroxyls on extra-framework aluminum species other than tetrahedral Al(IV) species such as Al(III), Al(V), or Al (VI) [18,35]. However, NMR results show that the protons giving rise to the signal at 2.5–2.9 ppm are active Brønsted sites at partially bonded Al(IV)-2 species and Al(IV)-2 sites with a chemical shift at 54 ppm [25]. 27Al MAS NMR and D-HMQC data show that Z-I zeolites mainly have Al(IV)-2 sites associated with protons at 2.68 ppm. The broad signal at 5.70 ppm might be generated by disordered bridging OH groups in the zeolites H-ZSM-5, and these hydroxyl protons interact strongly with nearby oxygen atoms in the framework [18]. Therefore, without organic structure-directing agents such as TPA+, the framework Al sites in the seeded synthesized ZSM-5 comprise Al(IV)-2 sites. The ZSM-5 zeolites with predominate Al(IV)-2 acid sites offer a solid foundation for research into the characteristics of Al(IV)-2 acid sites in their framework. Our synthesis method shows that framework crystallization defects cause the Al(IV)-2 sites. Chen and his coworkers proposed several intermediate structures formed by Al(IV)-2 acid sites [25]. In this work, the structural properties of defects created by Al(IV)-2 are further studied by 2D multiple quantum 1H NMR techniques. Due to the large proton-proton dipolar couplings, fast MAS alone is frequently insufficient to offer high resolution in the crowded regions of 1H MAS spectra. 1H MAS NMR spectrum of ZSM-5 with dominant Al(IV)-2 acid sites shown in Figure 1a is a good example. It is challenging to resolve the signals from 0 to 5 ppm effectively. Homonuclear 1H-1H double quantum (DQ) and triple quantum (TQ) 2D NMR experiments have been performed to assign the other signals. The dipolar coupling is strongly dependent on the inter-nuclear distance. It can be used to determine the geometry of the coupled spin, and then 1H TQ/DQ-SQ MAS NMR correlation spectroscopy has been used to identify the SiOH groups and BAS in ZSM-5 [36,37,38]. In 2D NMR spectra, the cross-correlation signal indicates that species with protons are in mutual proximity within about 5 Å (to a maximum of 8 Å). No signal will occur for isolated 1H species in such spectra due to the DQ filtering. In the 2D 1H DQ(TQ)-SQ spectrum, the spectral span of the DQ (TQ) dimension is twice (or triple) that of the SQ dimension, i.e., the sum of the isotropic chemical shifts of the two (or three) identical or different spins in the SQ dimension yields the chemical shift value of a corresponding contour in the DQ (TQ) dimension. The chemical shift at 0.87 ppm and 1.19 ppm can be attributed to the silanol groups on the external surface and the lattice defects, respectively. According to the 2D 1H TQ-SQ NMR spectrum in Figure 4, the contour signals at (2.68, 2.68 + 2.68 + 2.68) suggest that the defects created by Al(IV)-2 acid sites have three OH terminals with the same chemical shifts, as in the hydroxides in Al-(OH)3 groups (with structural models such as Figure 4d in Chen’s paper [25]). The chemical shift at 7.00 ppm can be attributed to the bridge-bonded H, which is associated with the framework of a partially hydrolyzed but still bonded Brønsted site. A two-dimensional 1H DQ-SQ NMR experiment was performed to reveal the structure formed by Al(IV)-2 acid sites. The correlated contour signals in Figure 5 indicate that the proton with chemical shift at 7.00 ppm is spatially close to the protons with chemical shift at 2.68 ppm. Therefore, the true chemical shifts of Al(IV)-2 acid sites is 7.00 ppm. The signal with 2.68 ppm is from Al-(OH)3 terminals. As suggested by the low 1H chemical shift value of 2.68 ppm, the protons associated with Al-(OH)3 terminals could be weakly acidic. However, the protons associated with Al(IV)-2 acid sites with a high 1H chemical shift value of 7.00 ppm are the frame-bonded Brønsted sites. Figure 3b demonstrates that no correlated signal is from 1H at 7.00 ppm, which indicts that the 1H/27Al D-HMQC technique is not reliable for studying the Al(IV)-2 Brønsted sites for the rapid proton chemical exchange. 1H/27Al D-HMQC experiments at very fast MAS have also been performed to study ZSM-5 (high silica) and ZSM-5 (nano), but no signals have been obtained. The reason is that the protons associated with the Al(IV)-1 site are delocalized through rapid chemical exchange for the clusters of water molecules around the Al(IV)-1 site. The 1H/27Al D-HMQC technique is not reliable for studying both the Al(IV)-1 Brønsted sites and the Al(IV)-2 Brønsted sites for the rapid proton chemical exchange. Generally, in 27Al magic angle spinning (MAS), NMR spectra of zeolites’ approximate chemical shifts are at 50–60 ppm for tetrahedral oxygen-coordinated aluminum, at 30–50 ppm for tetrahedral non-framework Al species, at 30 ppm for five-fold coordinated species, and at 0 ppm for octahedral aluminum [39]. The chemical shifts of ZSM-5 (high silica) and ZSM-5 (nano) are 55 ppm and 53 ppm, respectively. Therefore, we can suggest that both ZSM-5 (high silica) and ZSM-5 (nano) have only Al(IV)-1 acid sites. These ‘second’ Brønsted sites significantly increase catalyst reactivity in benzene hydride-transfer and n-hexane cracking reactions [25]. Therefore, it is important to note that the protons associated with Al(IV)-2 are not weakly acidic. According to our observation, we can suggest that the strong acidic site is from the proton Brønsted site with 7.00 ppm. The proton’s chemical shift at 2.68 ppm is from aluminol groups, which are weak Lewis acids. This assignment agrees with the literature, in which the 2.68 ppm 1H signal has been attributed to hydroxyls on extra-framework aluminum species [18,35]. The structural property of Al(IV)-2 acidic sites of ZSM-5 is depicted in Scheme 1, which clearly shows that Al(IV)-2 sites have Brønsted and Lewis acid characteristics. The channel intersections where TPA+ cations prefer to be are also the locations of Al(IV)-2 acidic sites. Aluminol groups found it difficult to live alongside TPA+ cations because of their size. Therefore, when TPA+ cations were employed in the synthesis system, Al(IV)-2 acid sites could not be formed.
In this work, the cracking of n-dodecane was performed to determine the acidic strength of two acid sites and evaluate the acid properties of ZSM-5 zeolites. The microactivity test (MAT) unit was employed to gauge the catalytic efficiency of the synthesized zeolites in the steam catalytic cracking of n-dodecane, and the resulting reaction products were meticulously analyzed. During cracking n-dodecane, Brønsted acid sites play the dominant catalytic roles. According to the catalytic cracking performance of n-dodecane shown in Figure 6, the acidic strength of ZSM-5 with Al(IV)-2 acid sites is not less than the acidic strength of ZSM-5 with Al(IV)-1 acid sites. Lewis acidity in protonic zeolite is due to available coordinative unsaturated Al3+ ions. It is often caused by EFAl, which exists in several oxide and hydroxide forms, such as Al(OH)2+, Al(OH)2+, AlOOH, Al2O3, Al(OH)3, or multinuclear clusters [40,41]. Apart from Brønsted acid sites, certain aluminum species in zeolites possess Lewis acidic characteristics, and Lewis acidic aluminum plays a pivotal role in traditional cracking reactions [42,43,44]. Generally, the catalytic cracking of alkanes can initially take place on Brønsted acid sites by forming carbonium ions (i.e., via protolytic cracking) or on Lewis acid sites by forming carbenium ions (i.e., via a classical β-scission mechanism) [45]. Corma et al. conclude that both pathways occur in parallel and protolytic cracking has a lower activation energy than β-scission [46]. The tertiary carbenium ions are more stable in branched compounds produced by isomerization processes. Therefore, Lewis acid is favorable for forming isomerized hydrocarbon. The acidity of ZSM-5 catalysts, primarily due to the presence of Brønsted and Lewis acid sites, plays a crucial role in catalytic reactions. Variations in acidity can impact catalytic activity and selectivity in acid-catalyzed reactions. Brønsted acid sites are characterized by their ability to donate protons to reactant molecules, facilitating acid-catalyzed reactions such as cracking, isomerization, alkylation, and dehydration. The catalytic products suggest that the Brønsted acidic strength of the Al(IV)-2 sites is either a little stronger than or comparable to that of the Al(IV)-1 sites. Lewis acid sites are often associated with aluminum atoms in distorted tetrahedral positions or extra-framework cations. These sites exhibit a strong affinity for electron-rich species. They can catalyze reactions involving Lewis acid-based interactions, such as adsorption, activation, and the transformation of unsaturated molecules (e.g., olefins, aromatics). Lewis acid sites are essential for catalyzing alkene oligomerization, alkyl transfer, and aromatic substitution. Based on the catalytic findings, it is evident that both Z-II and Z-III zeolites possess a greater abundance of Lewis acid sites compared to Z-I zeolites. This abundance enhances their ability to catalyze isomerization reactions, resulting in a higher concentration of iso-paraffins in the liquid product.
Since the 1980s, commercial ZSM-5-based additives have been employed in the FCC process to increase the octane number of gasoline. RON and MON are two characteristics that are widely used in the compression ignition (CI) engine community. It is important to accurately record the octane rating of liquid products to ensure that these studies can further improve the design of fuels. The overall conversion of n-dedicate to produce gasoline reformate of different ZSM-5 zeolites was evaluated by RON and MON and shown in Figure 7. High conversion and steady selectivity were maintained during the FCC reaction, resulting in a product with a highly researched RON and MON. The ZSM-5 with predominant Al(IV)-2 acid sites also have high catalytic activity.

4. Conclusions

In this study, the combination of 2D D-HMQC and 2D 1H multi-quantum MAS NMR experiments provides valuable information on Al(IV)-2 acid sites: the low 1H chemical shift value of 2.68 ppm is from hydroxy groups in Al(OH)3 and the high 1H chemical shift value of 7.00 ppm corresponds to frame-bonded Brønsted sites. Two chemically distinct Brønsted sites can be manipulated through appropriate adjustments to the synthesis conditions. The catalytic products suggest that the Brønsted acidic strength of the Al(IV)-2 sites is either a little stronger than or comparable to that of the Al(IV)-1 sites. According to the catalytic results, it is clear that both Z-II and Z-III zeolites exhibit a higher concentration of Lewis acid sites than Z-I zeolites. This capability of controlling aluminum distribution holds great potential for tuning the properties of silicon-rich zeolites utilized as catalysts in various redox and acid reactions and investigating the effects of aluminum distribution on these materials’ properties and catalytic performance.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ma17122824/s1, Figure S1. 1H MAS NMR spectrum of Z-I; Figure S2. 27Al MAS NMR spectrum of Z-I; Figure S3. 1H -27Al D-HMQC 2D MAS NMR spectrum of Z-I; Figure S4. 1H MAS NMR spectrum of Z-II; Figure S5. 27Al MAS NMR spectrum of Z-II; Figure S6. 1H -27Al D-HMQC 2D MAS NMR spectrum of Z-II; Figure S7. 1H MAS NMR spectrum of Z-III; Figure S8. 27Al MAS NMR spectrum of Z-III; Figure S9. 1H -27Al D-HMQC 2D MAS NMR spectrum of Z-III.

Author Contributions

Validation, H.M.; Investigation, Y.T. and L.Z.; Writing—original draft, Y.W.; Writing—review & editing, X.L. All authors have read and agreed to the published version of the manuscript.

Funding

The authors gratefully acknowledge financial support from the National Natural Science Foundation of China (NSFC) (22075322).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

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

Conflicts of Interest

Li Zhang and Yi Wang were employed by the PetroChina. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Muraza, O.; Galadima, A. Aquathermolysis of heavy oil: A review and perspective on catalyst development. Fuel 2015, 157, 219–231. [Google Scholar] [CrossRef]
  2. Corma, A.; Mengual, J.; Miguel, P.J. Steam catalytic cracking of naphtha over ZSM-5 zeolite for production of propene and ethene: Micro and macroscopic implications of the presence of steam. Appl. Catal. A Gen. 2012, 417, 220–235. [Google Scholar] [CrossRef]
  3. Diao, Z.; Wang, L.; Zhang, X.; Liu, G. Catalytic cracking of supercritical n-dodecane over meso-HZSM-5@Al-MCM-41 zeolites. Chem. Eng. Sci. 2015, 135, 452–460. [Google Scholar] [CrossRef]
  4. Ji, Y.; Yang, H.; Zhang, Q.; Yan, W. Phosphorus modification increases catalytic activity and stability of ZSM-5 zeolite on supercritical catalytic cracking of n-dodecane. J. Solid State Chem. 2017, 251, 7–13. [Google Scholar] [CrossRef]
  5. Argauer, R.J.; Landolt, G.R. Crystalline Zeolite ZSM-5 and Method of Preparing the Same. US Patent 3,702,886, 1972. [Google Scholar]
  6. Degnan, T.F.; Chitnis, G.K.; Schipper, P.H. History of ZSM-5 fluid catalytic cracking additive development at Mobil. Microporous Mesoporous Mater. 2000, 35, 245–252. [Google Scholar] [CrossRef]
  7. Ono, Y. Transformation of Lower Alkanes into Aromatic Hydrocarbons over ZSM-5 Zeolites. Catal. Rev. 1992, 34, 179–226. [Google Scholar] [CrossRef]
  8. Bensafi, B.; Chouat, N.; Djafri, F. The universal zeolite ZSM-5: Structure and synthesis strategies. A review. Coord. Chem. Rev. 2023, 496, 215397. [Google Scholar] [CrossRef]
  9. Dědeček, J.; Sobalík, Z.; Wichterlová, B. Siting and Distribution of Framework Aluminium Atoms in Silicon-Rich Zeolites and Impact on Catalysis. Catal. Rev. 2012, 54, 135–223. [Google Scholar] [CrossRef]
  10. Jones, A.J.; Carr, R.T.; Zones, S.I.; Iglesia, E. Acid strength and solvation in catalysis by MFI zeolites and effects of the identity, concentration, and location of framework heteroatoms. J. Catal. 2014, 312, 58–68. [Google Scholar] [CrossRef]
  11. Chen, K.; Gan, Z.; Horstmeier, S.; White, J.L. Distribution of Aluminum Species in Zeolite Catalysts: 27Al NMR of Framework, Partially-Coordinated Framework, and Non-Framework Moieties. J. Am. Chem. Soc. 2021, 143, 6669–6680. [Google Scholar] [CrossRef]
  12. Dusselier, M.; Davis, M.E. Small-pore zeolites: Synthesis and catalysis. Chem. Rev. 2018, 118, 5265–5329. [Google Scholar] [CrossRef] [PubMed]
  13. Davis, M.E. Zeolites from a materials chemistry perspective. Chem. Mater. 2014, 26, 239–245. [Google Scholar] [CrossRef]
  14. Davis, M.E.; Lobo, R.F. Zeolite and molecular sieve synthesis. Chem. Mater. 1992, 4, 756–768. [Google Scholar] [CrossRef]
  15. Davis, R.J. New perspectives on basic zeolites as catalysts and catalyst supports. J. Catal. 2003, 216, 396–405. [Google Scholar] [CrossRef]
  16. Caratzoulas, S.; Davis, M.E.; Gorte, R.J.; Gounder, R.; Lobo, R.F.; Nikolakis, V.; Sandler, S.I.; Snyder, M.A.; Tsapatsis, M.; Vlachos, D.G. Challenges of and insights into acid-catalyzed transformations of sugars. J. Phys. Chem. C 2014, 118, 22815–22833. [Google Scholar] [CrossRef]
  17. Haw, J.F.; Nicholas, J.B.; Xu, T.; Beck, L.W.; Ferguson, D.B. Physical Organic Chemistry of Solid Acids:  Lessons from in Situ NMR and Theoretical Chemistry. Acc. Chem. Res. 1996, 29, 259–267. [Google Scholar] [CrossRef]
  18. Hunger, M. Multinuclear solid-state NMR studies of acidic and non-acidic hydroxyl protons in zeolites. Solid State Nucl. Magn. Reson. 1996, 6, 1–29. [Google Scholar] [CrossRef]
  19. Pashkova, V.; Sklenak, S.; Klein, P.; Urbanova, M.; Dědeček, J. Location of Framework Al Atoms in the Channels of ZSM-5: Effect of the (Hydrothermal) Synthesis. Chem. Eur. J. 2016, 22, 3937–3941. [Google Scholar] [CrossRef] [PubMed]
  20. Schallmoser, S.; Ikuno, T.; Wagenhofer, M.F.; Kolvenbach, R.; Haller, G.L.; Sanchez-Sanchez, M.; Lercher, J.A. Impact of the local environment of Brønsted acid sites in ZSM-5 on the catalytic activity in n-pentane cracking. J. Catal. 2014, 316, 93–102. [Google Scholar] [CrossRef]
  21. Chen, K.; Abdolrhamani, M.; Sheets, E.; Freeman, J.; Ward, G.; White, J.L. Direct Detection of Multiple Acidic Proton Sites in Zeolite HZSM-5. J. Am. Chem. Soc. 2017, 139, 18698–18704. [Google Scholar] [CrossRef]
  22. Cavadini, S.; Lupulescu, A.; Antonijevic, S.; Bodenhausen, G. Nitrogen-14 NMR Spectroscopy Using Residual Dipolar Splittings in Solids. J. Am. Chem. Soc. 2006, 128, 7706–7707. [Google Scholar] [CrossRef] [PubMed]
  23. Gan, Z. Measuring Amide Nitrogen Quadrupolar Coupling by High-Resolution 14N/13C NMR Correlation under Magic-Angle Spinning. J. Am. Chem. Soc. 2006, 128, 6040–6041. [Google Scholar] [CrossRef] [PubMed]
  24. Tricot, G.; Trébosc, J.; Pourpoint, F.; Gauvin, R.; Delevoye, L. Chapter Four—The D-HMQC MAS-NMR Technique: An Efficient Tool for the Editing of Through-Space Correlation Spectra Between Quadrupolar and Spin-1/2 (31P, 29Si, 1H, 13C) Nuclei. In Annual Reports on NMR Spectroscopy; Webb, G.A., Ed.; Elsevier: Amsterdam, The Netherlands, 2014; Volume 81, pp. 145–184. [Google Scholar]
  25. Chen, K.; Horstmeier, S.; Nguyen, V.T.; Wang, B.; Crossley, S.P.; Pham, T.; Gan, Z.; Hung, I.; White, J.L. Structure and Catalytic Characterization of a Second Framework Al(IV) Site in Zeolite Catalysts Revealed by NMR at 35.2 T. J. Am. Chem. Soc. 2020, 142, 7514–7523. [Google Scholar] [CrossRef] [PubMed]
  26. Chen, K.; Zornes, A.; Nguyen, V.; Wang, B.; Gan, Z.; Crossley, S.P.; White, J.L. 17O Labeling Reveals Paired Active Sites in Zeolite Catalysts. J. Am. Chem. Soc. 2022, 144, 16916–16929. [Google Scholar] [CrossRef] [PubMed]
  27. van Bokhoven, J.A.; Lee, T.L.; Drakopoulos, M.; Lamberti, C.; Thieß, S.; Zegenhagen, J. Determining the aluminum occupancy on the active T-sites in zeolites using X-ray standing waves. Nat. Mater. 2008, 7, 551–555. [Google Scholar] [CrossRef] [PubMed]
  28. Vjunov, A.; Fulton, J.L.; Huthwelker, T.; Pin, S.; Mei, D.; Schenter, G.K.; Govind, N.; Camaioni, D.M.; Hu, J.Z.; Lercher, J.A. Quantitatively Probing the Al Distribution in Zeolites. J. Am. Chem. Soc. 2014, 136, 8296–8306. [Google Scholar] [CrossRef] [PubMed]
  29. Yokoi, T.; Mochizuki, H.; Namba, S.; Kondo, J.N.; Tatsumi, T. Control of the Al Distribution in the Framework of ZSM-5 Zeolite and Its Evaluation by Solid-State NMR Technique and Catalytic Properties. J. Phys. Chem. C 2015, 119, 15303–15315. [Google Scholar] [CrossRef]
  30. Park, S.; Biligetu, T.; Wang, Y.; Nishitoba, T.; Kondo, J.N.; Yokoi, T. Acidic and catalytic properties of ZSM-5 zeolites with different Al distributions. Catal. Today 2018, 303, 64–70. [Google Scholar] [CrossRef]
  31. Biligetu, T.; Wang, Y.; Nishitoba, T.; Otomo, R.; Park, S.; Mochizuki, H.; Kondo, J.N.; Tatsumi, T.; Yokoi, T. Al distribution and catalytic performance of ZSM-5 zeolites synthesized with various alcohols. J. Catal. 2017, 353, 1–10. [Google Scholar] [CrossRef]
  32. Hunger, M.; Anderson, M.W.; Ojo, A.; Pfeifer, H. Study of the geometry and location of the bridging OH groups in aluminosilicate and silicoaluminophosphate type zeolites using 1H MAS NMR sideband analysis and CP/MAS NMR. Microporous Mater. 1993, 1, 17–32. [Google Scholar] [CrossRef]
  33. Hunger, M.; Ernst, S.; Steuernagel, S.; Weitkamp, J.P.D. High-field 1H MAS NMR investigations of acidic and non-acidic hydroxyl groups in zeolites H-Beta, H-ZSM-5, H-ZSM-58, and H-MCM-22. Microporous Mater. 1996, 6, 349–353. [Google Scholar] [CrossRef]
  34. Jiang, Y.; Huang, J.; Dai, W.; Hunger, M. Solid-state nuclear magnetic resonance investigations of the nature, property, and activity of acid sites on solid catalysts. Solid State Nucl. Magn. Reson. 2011, 39, 116–141. [Google Scholar] [CrossRef] [PubMed]
  35. Ernst, H.; Freude, D.; Wolf, I. Multinuclear solid-state NMR studies of Brønsted sites in zeolites. Chem. Phys. Lett. 1993, 212, 588–596. [Google Scholar] [CrossRef]
  36. Schroeder, C.; Siozios, V.; Mück-Lichtenfeld, C.; Hunger, M.; Hansen, M.R.; Koller, H. Hydrogen Bond Formation of Brønsted Acid Sites in Zeolites. Chem. Mater. 2020, 32, 1564–1574. [Google Scholar] [CrossRef]
  37. Schroeder, C.; Mück-Lichtenfeld, C.; Xu, L.; Grosso-Giordano, N.A.; Okrut, A.; Chen, C.Y.; Zones, S.I.; Katz, A.; Hansen, M.R.; Koller, H. A Stable Silanol Triad in the Zeolite Catalyst SSZ-70. Angew. Chem. Inter. Ed. 2020, 59, 10939–10943. [Google Scholar] [CrossRef]
  38. Shantz, D.F.; Schmedt auf der Günne, J.; Koller, H.; Lobo, R.F. Multiple-Quantum 1H MAS NMR Studies of Defect Sites in As-Made All-Silica ZSM-12 Zeolite. J. Am. Chem. Soc. 2000, 122, 6659–6663. [Google Scholar] [CrossRef]
  39. Haouas, M.; Taulelle, F.; Martineau, C. Recent advances in the application of 27Al NMR spectroscopy to materials science. Prog. Nucl. Magn. Reson. Spectrosc. 2016, 94, 11–36. [Google Scholar] [CrossRef]
  40. Bhering, D.L.; Ramírez-Solís, A.; Mota, C.J.A. A Density Functional Theory Based Approach to Extraframework Aluminum Species in Zeolites. J. Phys. Chem. B 2003, 107, 4342–4347. [Google Scholar] [CrossRef]
  41. Catana, G.; Baetens, D.; Mommaerts, T.; Schoonheydt, R.A.; Weckhuysen, B.M. Relating Structure and Chemical Composition with Lewis Acidity in Zeolites:  A Spectroscopic Study with Probe Molecules. J. Phys. Chem. B 2001, 105, 4904–4911. [Google Scholar] [CrossRef]
  42. Busca, G. Acidity and basicity of zeolites: A fundamental approach. Microporous Mesoporous Mater. 2017, 254, 3–16. [Google Scholar] [CrossRef]
  43. Wang, Q.L.; Giannetto, G.; Guisnet, M. Dealumination of zeolites III. Effect of extra-framework aluminum species on the activity, selectivity, and stability of Y zeolites in n-heptane cracking. J. Catal. 1991, 130, 471–482. [Google Scholar] [CrossRef]
  44. Tarach, K.A.; Góra-Marek, K.; Martinez-Triguero, J.; Melián-Cabrera, I. Acidity and accessibility studies of desilicated ZSM-5 zeolites in terms of their effectiveness as catalysts in acid-catalyzed cracking processes. Catal. Sci. Technol. 2017, 7, 858–873. [Google Scholar] [CrossRef]
  45. Vogt, E.T.C.; Weckhuysen, B.M. Fluid catalytic cracking: Recent developments on the grand old lady of zeolite catalysis. Chem. Soc. Rev. 2015, 44, 7342–7370. [Google Scholar] [CrossRef] [PubMed]
  46. Corma, A.; Planelles, J.; Sanchez-Marin, J.; Tomas, F. The role of different types of acid site in the cracking of alkanes on zeolite catalysts. J. Catal. 1985, 93, 30–37. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of (a) Z-I, (b) Z-II, and (c) Z-III; SEM pictures of (d) Z-I, (e) Z-II, and (f) Z-III.
Figure 1. XRD patterns of (a) Z-I, (b) Z-II, and (c) Z-III; SEM pictures of (d) Z-I, (e) Z-II, and (f) Z-III.
Materials 17 02824 g001
Figure 2. 1H MAS NMR and 27Al MAS NMR spectra of (a,a’) Z-I, (b,b’)Z-II, and (c,c’) Z-III.
Figure 2. 1H MAS NMR and 27Al MAS NMR spectra of (a,a’) Z-I, (b,b’)Z-II, and (c,c’) Z-III.
Materials 17 02824 g002
Figure 3. (a) 1H MAS NMR spectrum and (b) 1H/{27Al} D-HMQC NMR spectrum of Z-I.
Figure 3. (a) 1H MAS NMR spectrum and (b) 1H/{27Al} D-HMQC NMR spectrum of Z-I.
Materials 17 02824 g003
Figure 4. 1H TQ-SQ MAS NMR spectrum of Z-I.
Figure 4. 1H TQ-SQ MAS NMR spectrum of Z-I.
Materials 17 02824 g004
Figure 5. 1H DQ-SQ MAS NMR spectrum of Z-I.
Figure 5. 1H DQ-SQ MAS NMR spectrum of Z-I.
Materials 17 02824 g005
Scheme 1. Structures of Al(IV)-2 acid sites of ZSM-5.
Scheme 1. Structures of Al(IV)-2 acid sites of ZSM-5.
Materials 17 02824 sch001
Figure 6. Hydrocarbons distribution in gasoline components over three ZSM-5 catalysts.
Figure 6. Hydrocarbons distribution in gasoline components over three ZSM-5 catalysts.
Materials 17 02824 g006
Figure 7. The octane number of products created by three ZSM-5 catalysts.
Figure 7. The octane number of products created by three ZSM-5 catalysts.
Materials 17 02824 g007
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

Tong, Y.; Zhang, L.; Ma, H.; Wang, Y.; Liu, X. Understanding the Structural and Catalytic Properties of Al(IV)-2 Acidic Sites of ZSM-5. Materials 2024, 17, 2824. https://doi.org/10.3390/ma17122824

AMA Style

Tong Y, Zhang L, Ma H, Wang Y, Liu X. Understanding the Structural and Catalytic Properties of Al(IV)-2 Acidic Sites of ZSM-5. Materials. 2024; 17(12):2824. https://doi.org/10.3390/ma17122824

Chicago/Turabian Style

Tong, Yan, Li Zhang, Hong Ma, Yi Wang, and Xiaolong Liu. 2024. "Understanding the Structural and Catalytic Properties of Al(IV)-2 Acidic Sites of ZSM-5" Materials 17, no. 12: 2824. https://doi.org/10.3390/ma17122824

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