Next Article in Journal
Correlation between Flow Temperature and Average Molar Ionic Potential of Ash during Gasification of Coal and Phosphorus-Rich Biomass
Next Article in Special Issue
Catalytic Ketonization over Oxide Catalysts (Part XIV): The Ketonization and Cross-Ketonization of Anhydrides, Substituted Acids and Esters
Previous Article in Journal
On the Nature of the Rotational Energy Barrier of Atropisomeric Hydrazides
Previous Article in Special Issue
Hydrodeoxygenation of Oxygenates Derived from Biomass Pyrolysis Using Titanium Dioxide-Supported Cobalt Catalysts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Acidity Quantification and Structure Analysis of Amide-AlCl3 Liquid Coordination Complexes for C4 Alkylation Catalysis

School of Chemistry and Chemical Engineering, Inner Mongolia University, Hohhot 010021, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Molecules 2023, 28(23), 7857; https://doi.org/10.3390/molecules28237857
Submission received: 29 October 2023 / Revised: 21 November 2023 / Accepted: 27 November 2023 / Published: 30 November 2023
(This article belongs to the Special Issue Research on Heterogeneous Catalysis)

Abstract

:
Liquid coordination complexes (LCCs), which are formed between metal halides and donor molecules, represent promising catalysts. Six amide-AlCl3 LCCs were successfully synthesized, followed by their characterization through NMR, Raman, and UV-visible spectroscopy. The acidity of these LCCs was quantified by performing computational modelling of fluoride ion affinities (FIA) and experimental Gutmann–Beckett measurements. Spectroscopic analysis indicated bidentate coordination between amide ligands and Al, which induced asymmetric splitting of Al2Cl6 into diverse ions such as [AlCl2L2]+, [AlCl4], [AlCl3L], and [Al2Cl6L]. The computed FIA was found to align well with the experimental acidity trends, thereby confirming the proposed structure of the LCC. In the alkylation tests, the LCC with a high acidity demonstrated an increase in the yields of C5-C7 alkylates. These results provide an in-depth understanding of the tuneable structures of amide-AlCl3 LCCs. The acidity of LCCs can be controlled by tuning the ratio of the organic ligand to AlCl3, which allows bidentate coordination to facilitate asymmetric splitting of Al2Cl6. The LCCs demonstrate a high degree of potential as versatile and sustainable acid catalysts in alkylation reactions. These findings may advance the foundational knowledge of LCCs for the purpose of targeted acid catalyst design.

1. Introduction

From photoelectron catalysis to traditional acid catalysis, liquids containing high concentrations of metals exhibit significant application value. Ideally, these liquids demonstrate exceptional control and tunable metal coordination capabilities, enabling key properties (e.g., viscosity and acidity) to be adjustable and well-regulated. Common types of high-metal-concentration liquids are molten alloys, molten salts, ionic liquids, or deep eutectic solvents [1,2]. However, these compounds exhibit several limitations, including high costs, metal solubility constraints, corrosiveness, and insufficient metal coordination capacity, hindering the realization of ideal systems. Liquid coordination complexes (LCCs) are deemed superior alternatives to ionic liquids and molten salts [3,4,5], exhibiting outstanding performance across various applications. For instance, in the alkylation of isobutane with 2-butene, LCCs prepared by combining metal halides and donor molecules possess more flexible approaches to regulating Lewis acidity and heightened catalytic activity. Additionally, these complexes are cost-effective, easily synthesized, and environmentally friendly characteristics, thus partially mitigating the drawbacks associated with ionic liquids [6,7,8].
Currently, the predominantly encountered LCCs are formed by coordinating metal halides (MClx, x = 2,3) such as ZnCl2, SnCl2, CrCl3, FeCl3, AlCl3, and GaCl3 with donor molecules like urea, alcohol, and amides [1,9]. Previous investigations have demonstrated that many of these ligands induce asymmetric splitting of “M2Cl6” units [10], leading to the formation of ionized compounds like [MCl2L2][MCl4]. However, the intricacy of LCC complexes extends far beyond this. Recent research suggests that LCCs formed by coordinating AlCl3 with amides might accommodate species like [AlCl2(amide)n][AlCl4] (n = 1 or 2). This coordination requires the presence of [AlCl2(amide)2]+, [AlCl4], and [AlCl2(amide)]+ ions [11,12]. The presence of tridentate ions [AlCl2(amide)2]+ and [AlCl2(amide)]+, however, still remains a topic of contention [1,4,9,13]. Existing academic literature underscores that donor molecules possessing high dielectric constants and polarity (e.g., acetonitrile) enhance the formation of ionic materials in AlCl3-based LCCs, while using substances with lower dielectric constants and polarity such as tetrahydrofuran results in the formation of neutral materials [14,15]. Conversely, investigations into coordination sites and modes in amide LCCs remain quite restricted. Campos et al. found that formamide, through coordination with AlCl3 using the O and N atoms, can form ionic species with bidentate structures [16,17]. Hu et al. also indicated that the presence of methyl leads to bidentate coordination structures in N-methylacetamide-AlCl3 and N,N-dimethylacetamide-AlCl3 LCCs for both O and N atoms. In contrast, acetamide only engages monodentate O atoms to coordinate with AlCl3 [11,12]. The characterization of complexes and coordination sites in other amide-AlCl3 LCCs still presents an ambiguity. Do they form [AlCl3L] and [AlCl3L2], or form [AlCl2L2]+ and [AlCl2L]+? Do they represent monodentate or bidentate coordination structures? In light of the economical preparation and simplified synthesis routes of amide LCCs, as well as their increasing relevance in catalytic reactions, a thorough investigation of the indicated issues related to amide LCCs becomes essential.
In addition, both chloroaluminate ionic liquids and AlCl3-LCCs manifest significant advantages in the alkylation reaction of isobutane and 2-butene compared to traditional sulfuric acid catalysts [7,18,19,20,21]. However, accurately quantifying the acid strength of these compounds, which act as Lewis acidic catalysts, remains a challenge [22]. The International Union of Pure and Applied Chemistry (IUPAC) defines Lewis acidity as the thermodynamic propensity to form Lewis acid–base pairs, often measured by the affinity of fluoride ions [23]. The task to ascertain Lewis acidity experimentally can be complex due to the numerous bound molecules used for evaluation [24]. The Gutmann–Beckett (GB) method, a well-established method, hinges on the induced 31P NMR shift of triethylphosphine oxide to determine the Lewis acidity strength of a compound [25,26,27]. In general, the induced 31P NMR shift (δ31P) of triethylphosphine oxide can be measured to obtain the acceptor number (AN), where AN = (δ31P − 41.0) × [100/(86.1 − 41.0)]. A higher AN value (or GB value = Δδ31Pexp = δ31Pexp − 50) indicates a stronger Lewis acidity of the compound. On the other hand, computed fluoride ion affinity (FIA) serves as a valuable metric for evaluating Lewis acidity [28,29], although benchmark studies on the accuracy of FIA computations have been presented only to a limited extent [30]. Recently, Erdmann and Greb employed the DLPNO-CCSD(T) computational model to perform comparative FIA analysis of compounds consisting of various elements from the periodic table [23,24,29,31]. They have also studied the thermodynamic properties, nuclear magnetic resonance characteristics, and GB values of more than 130 Lewis acids. Importantly, this link between theoretical and experimental perspectives enhances our understanding of Lewis acidity in numerous compounds.
In this work, we have synthesized six amide-AlCl3 LCCs using a one-step approach with acetamide, propionamide, and butanamide. Additionally, chloroformamide, chloroacetamide, and 2-chloropropionamide were also used as the donor molecules. A notable contribution of this work is the establishment of a distinct correlation between the molecular structure of Lewis acidic compounds, computed fluoride ion affinity (FIA) values, and experimental Gutmann–Beckett measurements. Utilizing computed FIA and GB values, we employed hypothesis testing methods to investigate the specific molecular structures of the Lewis acidic LCCs. The six amide-LCCs were thoroughly characterized using NMR, Raman, and UV-Vis spectroscopy. These characterizations provided insights into the specific influences of amide structures on the asymmetric splitting of AlCl3. Altogether, this work has expanded the range of LCC catalysts and provided valuable data for the design and synthesis of the C4 alkylation catalysts.

2. Results and Discussion

2.1. 27Al NMR and Raman Spectroscopic Analysis

Previous studies have attributed the formation of species in amide-AlCl3 based LCC to the asymmetric splitting of Al2Cl6 under the induction of organic ligands (L) [1,32,33]. We sought to gather valuable insights into the structure of LCCs, recognizing that 27Al NMR is an effective tool for identifying aluminum species [34,35,36]. Figure 1a presents 27Al NMR spectra of three amide-AlCl3 based LCCs with a compound of ratio of XAlCl3 = 0.67.
Three distinct peaks have been identified in the 27Al NMR spectra of three LCCs (Figure 1a). For the LCC of acetamide-AlCl3, three peaks emerged at 102, 90, and 75 ppm, associated with the complexes formed between aluminum chloride and organic ligands (L, acetamide). Specifically, the downfield peak at approximately 102 ppm is attributed to the tetrahedral [AlCl4] anion, consistent with previous literature values. Moreover, the upfield-shifted peaks in the 27Al NMR spectrum are attributed to [AlCl3L] (90 ppm) and [AlCl2L2]+ (75 ppm), respectively. Notably, the [AlCl2L2]+ ion was distinctly detected in the NMR spectrum, absent the so-called [AlCl2L]+ ion. When XAlCl3 equals 0.5, a broad band is displayed between δ = 101 − 106 ppm in the acetamide-AlCl3 based LCC (Figure 1b). This broad signal masks the sharp signals of [AlCl4] and [AlCl3L] that were previously observed at the corresponding positions when XAlCl3 = 0.67. The observed broad band is believed to be attributed to the dynamic equilibrium of [AlCl4], [Al2Cl6L], and [AlCl3L].
As illustrated in Figure 2, the 27Al NMR spectra of the chloroamide-AlCl3-based LCCs were examined with XAlCl3 = 0.50. Across all the spectra of the chloroamide-based LCCs, a broad band was observed at approximately δ = 105 − 75 ppm. This observed band could potentially be attributed to the dynamic equilibrium between [AlCl3L], [Al2Cl6L], and [AlCl4] complexes. Moreover, the NMR spectra also identified [AlCl2L2]+ at the distinctive 75 ppm and [AlCl3L2] at the 52 ppm regions. These particular species exhibit intricate equilibrium relationships:
[AlCl2L2]+ + [AlCl4] + [Al2Cl6] + L ⇌ 2[AlCl3L] + [Al2Cl6L]
[AlCl2L2]+ + [AlCl4] + 2L ⇌ 2[AlCl3L2]
[AlCl3L2] + AlCl3 ⇌ [AlCl2L2]+ + [AlCl4]
It is noted that the introduction of an organic ligand L (e.g., amide and chloroamide) or AlCl3 could disrupt the equilibrium of Formulas (1)−(3). This phenomenon is noticeably evident when XAlCl3 = 0.67. As the molar fraction of AlCl3 in acetamide-LCC increases from 0.50 to 0.67 (Figure 2b), a substantial alteration is manifested in the [AlCl2L2]+ signal peak of the acetamide-AlCl3-based LCC, whereas the [AlCl3L2] peak virtually vanishes. This suggests that an increase in AlCl3 in LCC fosters the rightward shift of the reaction equation (3). The NMR spectra of chloroacetamide-AlCl3 provides further evidence supporting this inference. Upon increasing the XAlCl3 proportion from 0.50 to 0.67, the [AlCl3L2] signal peak becomes unobservable. Simultaneously, the signal peak resulting from the equilibrated species [AlCl4], [Al2Cl6L], and [AlCl3L] exhibits a progressive broadening.
The presence of broad bands in the 27Al NMR spectrum obscures the specific structure of Al species in LCCs. In particular, previous studies have consistently suggested that the signal of [Al2Cl7] cannot be identified through 27Al NMR spectroscopy [3,36,37]. Nevertheless, this limitation can be mitigated by employing Raman spectroscopy. The Raman spectra of amide-based LCCs exhibit distinct peaks corresponding to [AlCl4] (350 cm−1) and [Al2Cl7] (315 cm−1) within the range of 400–300 cm−1 (Figure 3a) [5,38,39]. There is a decrease in the intensity of [AlCl4] associated with an increase in alkyl group size. That is, the reaction of 2[AlCl4] ⇌ [Al2Cl7] + Cl is more likely to be influenced towards the right by smaller organic ligands. On the Raman spectrum of chloroacetamide-AlCl3 (Figure 3b), a shoulder related to [Al3Cl10] appears when XAlCl3 = 0.60 (390 cm−1). Upon increasing the XAlCl3 to 0.60 and 0.67, this shoulder develops into a prominent band.

2.2. Experimental Measurement of LCC Acidity

The purpose of the study on LCC was to catalyze the C4 alkylation reaction between isobutane and 2-butene. However, this reaction requires a high acid strength. According to the Hammett acidity function, only H0 values lower than −10 can initiate the reaction [40]. Conversely, excessive acidity can result in increased side reactions in alkylation, thus lowering the quality of alkylate gasoline. The industrial evaluation of acid strength in alkylation catalysts primarily focuses on Brönsted acidity, with relatively few studies dedicated to the acid strength of Lewis acidic catalysts [18]. The Gutmann method can be used to measure Lewis acid strength by quantifying the acceptor number (AN) of a compound through the measurement of the 31P NMR chemical shift of a basic probe molecule, triethylphosphine oxide (TEPO) [26]. However, the Gutmann method is unwieldy, because it necessitates the extrapolation of data for TEPO-Lewis acid [41]. To simplify the measurement, the Gutmann–Beckett method has been developed [27]. This method enables the quantification of Lewis acidity by measuring the GB value (Δδ31Pexp). The GB value is positively correlated with Lewis acid strength, and it can serve to quantify the acid strength of LCCs and to some extent infer the molecular structure of the Lewis acid.
The 31P NMR spectroscopic analysis revealed that the AN values of LCCs were much higher than those of chloroaluminate ionic liquids with XAlCl3 = 0.50 (e.g., [BMIm]Cl-AlCl3 and [(C2H5)3NH]Cl-AlCl3). Generally, when XAlCl3 = 0.50, the chloroaluminate ionic liquid does not exhibit acidity. Notably, the product of acetamide-AlCl3 contains a significant proportion of C8 olefins, reaching up to 84.1%. This high yield is a result of acid-catalyzed oligomerization of C4 olefins, suggesting that LCCs exhibit Lewis acidity due to the asymmetric splitting of Al2Cl6 when XAlCl3 = 0.50. In addition, the Gutmann–Beckett method was used to determine the AN values (AN = (δ31P − 41.0) × [100/(86.1 − 41.0)]) of the six studied LCCs with XAlCl3 = 0.67, as shown in Table 1. The AN values of the LCCs were higher than those of [(C2H5)3NH]Cl-AlCl3 and [BMIm]Cl-AlCl3. Particularly, the AN values of three chloroamide-AlCl3-based LCCs were approximately 100 (98–102), indicating that the LCCs have significant potential for use in Lewis acid catalysis. Moreover, the results of the alkylation of the six LCC catalysts are shown in Table 1.

2.3. DFT Calculations of FIA Values

NMR and Raman spectroscopy measurements, as well as the determination of AN values, have provided a general direction for the structural determination of LCCs. However, considering the difficulties with experimental determination of the structures of liquid catalysts, it is necessary to obtain additional information from theoretical calculations. The calculation of fluoride ion affinity (FIA) is a valuable descriptor for evaluating the strength of Lewis acids in compounds. Despite its widespread use, a scarcity of data and comparative studies for larger Lewis acids impedes the broad comparability between experimental results and calculated data.
LA + COF3 → [LA-F] + COF2    ΔH1, LA is Lewis Acid.
COF3 → COF2 + F      ΔH2 (experimental 208.8 kJ·mol−1)
LA + F → [LA-F]      FIA = ΔH1 − ΔH2
Herein, we have evaluated the performance of selected methods (e.g., DLPNO-CCSD(T)) for FIA computation based on CCSD(T)/CBS data and Erhmann’s benchmark procedure [31]. The FIA values for the studied LCCs and chloroaluminate ionic liquids have been calculated. The computed FIA values can be compared unbiasedly against each other (Table S2), and these values provide a critical benchmark for evaluating the acid strength of prospective alkylation catalysts. Table 2 presents the computed FIA and GB values for the six studied LCCs.
The results indicate a definitive correlation between the FIA values and GB values among the six LCCs. Specifically, as the computed FIA value increases, the GB value also tends to increase. Butanamide-AlCl3 and chloropropionamide-AlCl3 exhibit relatively high FIA/GB values. The larger GB values imply a more pronounced ability to accept fluoride ions, indicating stronger Lewis acidity. This is consistent with the earlier evaluation results of alkylation experiments, where chloropropionamide-AlCl3 showed higher catalytic activity. Within Table 1, the C5–C7 fractions of these catalysts are significantly greater than those of other LCCs. Additionally, the GB values of amide- or chloroamide-LCCs decrease as the molecular weight of the derivatives increases, suggesting a decrease in overall acidity due to coordination between Lewis acidic species and larger organic ligands.
However, we have noted that the FIA values vary when calculations are conducted on varied structures of the LCC. As depicted in Table 2, calculations of the same LCC employing different species structures yield varying FIA values. The diverse FIA values associated with a compound lead to multiple possibilities for material structure analysis, yet they also introduce some degree of confusion. Therefore, it is necessary to further clarify the relationship between FIA, GB, and species structures.

2.4. Computed FIA, Δδ31Pexp, and Structure

Based on the dataset provided by Erdmann and Greb [24], a strong correlation was found between the calculated FIA, Δδ31Pexp, and the molecular structure of Lewis acids containing aluminum (Tables S1 and S2). Accurate FIA values rely on precise molecular structures of Lewis acids. According to the existing GB-FIA relationship, the experimental Δδ31Pexp (GB value) of a new Lewis acid can be used to estimate the FIA value. By comparing the residuals between the predicted FIA and the calculated FIA values with the confidence interval of the GB-FIA model, the accuracy of the molecular structure associated with the calculated FIA can be determined. Step 1: establish a GB-FIA regression model based on previous FIA and experimentalΔδ31Pexp values. Step 2: calculate the residuals between FIA and the predicted FIA usingΔδ31Pexp. Step 3: measure the Δδ31Pexp of the new Lewis acid molecule in advance and calculate the FIA value based on its optimized structure. Step 4: perform a hypothesis test to verify the correctness of FIA and the molecular structure for the new Lewis acid molecule. H0: the new FIA value conforms to the GB-FIA model. H1: the calculated FIA value of the new Lewis acid does not conform to the GB-FIA model, indicating that the molecular structure constructed during the FIA calculation does not match the actual situation. For more details, please see the Supplementary Materials.
Based on the aforementioned hypothesis tests, we focused on analyzing the relationship between the structures, FIA, and GB values of six amide-based LCCs (Table 3). Specifically, in the case of acetamide-LCC at XAlCl3 = 0.5, AlCl3 coordinates with acetamide through a monodentate site via the O atom. Among the various configurations of the AlCl3-acetamide coordination relationship, only the monodentate coordination model exhibits a significant correlation between the computed FIA and the predicted FIA from the experimental GB value (426 vs. 427 in Table 3). As XAlCl3 increases to 0.67, it is more likely that AlCl3 coordinates with the organic ligand through both the O and N atoms of acetamide in a bidentate coordination manner (Figure 4). Clear evidence supporting the aforementioned results is illustrated in the model structures that demonstrate a consistent match between the computed FIA and GB values.
Further investigation shows that regardless of whether XAlCl3 is 0.5 or 0.67, most amide- and chloroamide-based ligands are likely to coordinate with Al species through bidentate coordination. This appears to be the only manner to obtain a match between the calculated FIA values and the experimental GB values. More definitive evidence is provided by the UV-visible spectra of LCCs (Figure 5). In the UV-Vis spectrum of acetamide-AlCl3 (XAlCl3 = 0.50), a peak at 305 nm can be identified. This peak is attributed to ligand-metal charge transfer (LMCT) absorption, induced by coordination between the O atoms and Al atoms. No further significant absorption peaks are observed, indicating that acetamide-AlCl3 (XAlCl3 = 0.50) primarily exists in monodentate coordination. In the UV-Visible spectra of other LCCs, two LMCT absorption peaks are identified at 305 and 355 nm, indicating coordination between N and O atoms with Al atoms. This corresponds to the primary absorption peaks in a bidentate coordination structure. The steric hindrance from the larger alkyl groups or the Cl in the chloroamide leads to an excess of Al species coordinating not only with the O atom, but also the N atom. This provides a possible explanation for the occurrence of monodentate coordination when the XAlCl3 proportion in acetamide-AlCl3 is 0.50. On the other hand, when there is a high concentration of Al species with XAlCl3 equal to 0.67 within acetamide-AlCl3, the predominant form is bidentate coordination. Moreover, these results suggest that bidentate coordination is more favorable for the asymmetric splitting of Al2Cl6 than monodentate coordination. This might be the primary reason for the various forms of Al species in LCCs.

3. Materials and Methods

3.1. Materials

All reagents were used as received without any further purification, unless otherwise stated. Sodium aluminate (99%) and dichloromethane (99%) were purchased from Sigma-Aldrich Company. They were dried over 3 Å molecular sieves prior to use and stored in a glove box. Acetamide (99.5%), propionamide (99.5%), butanamide (99.5%), chloroformamide (99%), chloroacetamide (99%), 2-chloropropionamide (99%), and anhydrous aluminum chloride (99.5%) were obtained from Aladdin Chemistry Company (Shanghai, China).

3.2. Preparation of LCCs

Under N2 protection, LCCs with different molar ratios of amides to aluminum chloride were prepared in a glove box. The synthesis process of acetamide-AlCl3 (XAlCl3 = 0.67) is outlined as follows. A 150 mL three-neck flask equipped with a constant temperature oil bath and a stirrer was used for the synthesis. Anhydrous AlCl3 (13.34 g, 0.10 mol) was first added to the flask, followed by the slow addition of acetyl amide (2.96 g, 0.05 mol) while stirring for 20 min. The mixture was then heated to 90 °C and stirred for 4 h until all solids were completely dissolved, resulting in a liquid phase. The obtained LCCs were stored in the glove box for further use.

3.3. Characterization

The 27Al NMR spectrum was obtained using the Bruker Avance spectrometer, with a resonance frequency of 160.15 MHz at 25 °C. During the measurement, a 1.1 M solution of Al(NO3)3 dissolved in D2O was used as the external reference for the 27Al NMR chemical shift. Neat 85% H3PO4 was used as the reference solution for 31P NMR (to determine the GB value, or AN value). The LCC sample was placed in a standard 10 mm tube, with capillaries containing H3PO4 and Al(NO3)3/D2O inserted at the center. The measurement of AN value was performed as described in Ref. [41]. Raman spectroscopy measurements were carried out using the Perkin-Elmer Frontier 200 spectrometer. To prevent water contamination, the LCC sample was placed in a tightly sealed vial. The excitation light from the argon ion laser at 532 nm had an emission power of approximately 300–750 mW. The optical resolution of the Raman spectrum was 1 cm−1, with a wave number reproducibility of 0.2 cm−1. The ultraviolet-visible spectrum was recorded using a Shimadzu UV-2550 spectrophotometer in a quartz cell with a length of 0.1 cm. The UV-Visible spectrum was measured using a Shimadzu UV-2600 spectrophotometer in a quartz cell with a length of 0.1 cm pathway.

3.4. Computed FIA

The computed FIA values followed the procedure described by Ref. [31]: (1) The geometry optimization of all structures was performed using the PBEh-3c/def2-mSVP method implemented in the ORCA 5.04 software package [42]. It was confirmed that the optimized structures corresponded to energy minima on the potential energy surface by conducting analytical calculations of the PBEh-3c harmonic frequencies. (2) Single-point energy calculations were carried out using the DLPNO-CCSD(T)/aug-cc-pVQZ method [43] to obtain the electronic energy of both Lewis acid (LA) and [LA+F] structures. The total enthalpy of the Lewis acids and their fluorinated adducts was determined by combining the electronic energy obtained in this step with the thermal corrections from the previous geometry optimization. (3) To obtain the absolute value of FIA, the molecular structure of the selected reference system (COF2/COF3) was also optimized using the same method as in the first step. Subsequently, single point energy calculations were performed using the same method as in the second step. (4) The reaction enthalpy for the equation LA + COF3 → [LA-F] + COF2 was calculated. The experimental FIA value [44] of COF2 (208.8 kJ·mol−1) was subtracted from the calculated value. The final computed FIA values can show an accuracy within kJ·mol−1.

3.5. The C4 Alkylation Reaction

The isobutane/2-butene (C4) alkylation reaction was conducted in a 250 mL pressure-resistant reactor with a water bath temperature of (25 ± 0.5 °C) [45]. The stirring speed was maintained at 1000 rpm. Initially, the LCC was loaded into the reactor, followed by the addition of the C4 mixture. The molar ratio of isobutane to olefin in the feed was 5:1. Figure 6 illustrates the schematic diagram of the alkylation reaction. The reaction time for alkylation was 30 min. After completion, the reaction product and catalyst were poured out of the reactor and allowed to settle for approximately 20 min. To remove excess isobutane, the product was subjected to fractional distillation in a distillation column equipped with a reflux section. The product distribution was analyzed using gas chromatography (HP6890, PONA, 50 m capillary column).

4. Conclusions

This work experimentally measures the Lewis acidity of amide-AlCl3 liquid coordination complexes (LCCs), and also demonstrates the method of calculating the acidity of LCCs by constructing molecular models. The LCCs exhibit their potential as versatile and environmentally friendly catalysts for C4 alkylation reactions. The acidity of LCC was precisely quantified by determining the fluoride ion affinity and referencing the Gutmann–Beckett acidity scale observed in 31P NMR spectroscopic analysis, which also facilitated the definitive confirmation of the molecular structures of LCCs. The calculations revealed that increasing the AlCl3 ratio promotes formation of [AlCl2L2]+ and [AlCl4] ions while decreasing [AlCl3L2], correspondingly increasing LCC acidity. When evaluated in the alkylation of isobutane with 2-butene, the LCCs with the highest acid strength produced a higher yield of C5-C7 alkylate hydrocarbons, confirming their potential as adjustable acid catalysts for the reactions. The correlation developed between computational and experimental acidity measurements provides a valuable guide for designing new LCCs with targeted acidity for catalytic applications. In summary, the combination of experimental and computational methods for measuring acidity allows for precise adjustment of LCCs to achieve optimal acid strength and catalytic activity for environmentally friendly alkylation processes.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28237857/s1, Figure S1: The correlation model established by Al-compound GB-FIA database; Table S1: Collection of literature known and measured Gutmann–Beckett values (Δδ31Pexp); Table S2: Computed FIA values of Lewis acids given in kJ/mol−1; Program code: Program for establishing the GB-FIA model and hypothesis test.

Author Contributions

Conceptualization, Y.L. and J.B.; methodology, H.L. and Y.L.; formal analysis, Q.W. and H.L.; investigation, Q.W.; resources, J.B.; data curation, H.L.; writing—original draft preparation, Q.W. and Y.L.; writing—review and editing, Y.L.; supervision, Y.L.; funding acquisition, J.B. and Y.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundations of China (No. 22265022 and No. 21766021). This research was also supported by the Natural Science Foundations of Inner Mongolia (No. 2021MS02008 and No. 2020MS02008).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article and Supplementary Materials.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Coleman, F.; Srinivasan, G.; Swadźba-Kwaśny, M. Liquid Coordination Complexes Formed by the Heterolytic Cleavage of Metal Halides. Angew. Chem. Int. Ed. 2013, 52, 12582–12586. [Google Scholar] [CrossRef] [PubMed]
  2. Zhang, Q.; De Oliveira Vigier, K.; Royer, S.; Jérôme, F. Deep Eutectic Solvents: Syntheses, Properties and Applications. Chem. Soc. Rev. 2012, 41, 7108. [Google Scholar] [CrossRef] [PubMed]
  3. Kore, R.; Kelley, S.P.; Sawant, A.D.; Mishra, M.K.; Rogers, R.D. Are Ionic Liquids and Liquid Coordination Complexes Really Different?-Synthesis, Characterization, and Catalytic Activity of AlCl3base Catalysts. Chem. Commun. 2020, 56, 5362–5365. [Google Scholar] [CrossRef] [PubMed]
  4. Hogg, J.M.; Brown, L.C.; Matuszek, K.; Latos, P.; Chrobok, A.; Swadźba-Kwaśny, M. Liquid Coordination Complexes of Lewis Acidic Metal Chlorides: Lewis Acidity and Insights into Speciation. Dalt. Trans. 2017, 46, 11561–11574. [Google Scholar] [CrossRef] [PubMed]
  5. Brown, L.C.; Hogg, J.M.; Swadźba-Kwaśny, M. Lewis Acidic Ionic Liquids. Top. Curr. Chem. 2017, 375, 78. [Google Scholar] [CrossRef]
  6. Kore, R.; Uppara, P.V.; Rogers, R.D. Replacing HF or AlCl3 in the Acylation of Isobutylbenzene with Chloroaluminate Ionic Liquids. ACS Sustain. Chem. Eng. 2020, 8, 10330–10334. [Google Scholar] [CrossRef]
  7. Kore, R.; Berton, P.; Kelley, S.P.; Aduri, P.; Katti, S.S.; Rogers, R.D. Group IIIA Halometallate Ionic Liquids: Speciation and Applications in Catalysis. ACS Catal. 2017, 7, 7014–7028. [Google Scholar] [CrossRef]
  8. Smith, E.L.; Abbott, A.P.; Ryder, K.S. Deep Eutectic Solvents (DESs) and Their Applications. Chem. Rev. 2014, 114, 11060–11082. [Google Scholar] [CrossRef]
  9. Hogg, J.M.; Coleman, F.; Ferrer-Ugalde, A.; Atkins, M.P.; Swadźba-Kwaśny, M. Liquid Coordination Complexes: A New Class of Lewis Acids as Safer Alternatives to BF3 in Synthesis of Polyalphaolefins. Green Chem. 2015, 17, 1831–1841. [Google Scholar] [CrossRef]
  10. Atwood, D.A. Cationic Group 13 Complexes. Coord. Chem. Rev. 1998, 176, 407–430. [Google Scholar] [CrossRef]
  11. Hu, P.; Zhang, R.; Meng, X.; Liu, H.; Xu, C.; Liu, Z. Structural and Spectroscopic Characterizations of Amide–AlCl 3 -Based Ionic Liquid Analogues. Inorg. Chem. 2016, 55, 2374–2380. [Google Scholar] [CrossRef]
  12. Hu, P.; Jiang, W.; Zhong, L.; Zhou, S.-F. Determination of the Lewis Acidity of Amide–AlCl3 Based Ionic Liquid Analogues by Combined in Situ IR Titration and NMR Methods. RSC Adv. 2018, 8, 13248–13252. [Google Scholar] [CrossRef] [PubMed]
  13. Abood, H.M.A.; Abbott, A.P.; Ballantyne, A.D.; Ryder, K.S. Do All Ionic Liquids Need Organic Cations? Characterisation of [AlCl2·nAmide]+ AlCl4− and Comparison with Imidazolium Based Systems. Chem. Commun. 2011, 47, 3523. [Google Scholar] [CrossRef] [PubMed]
  14. Seo, J.-S.; Kim, K.-W.; Cho, H.-G. Solvation of AlCl3 in Deuterated Acetonitrile Studied by Means of Vibrational Spectroscopy. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2003, 59, 477–486. [Google Scholar] [CrossRef]
  15. Dalibart, M.; Derouault, J.; Granger, P.; Chapelle, S. Spectroscopic Investigations of Complexes between Acetonitrile and Aluminum Trichloride. 1. Aluminum Chloride-Acetonitrile Solutions. Inorg. Chem. 1982, 21, 1040–1046. [Google Scholar] [CrossRef]
  16. Campos, T.B.C.; da Silva, E.F.; Alves, W.A. A Raman Study on the Coordination Sites and Stability of the [Al(Formamide)5]Cl3 Complex. Vib. Spectrosc. 2013, 65, 24–27. [Google Scholar] [CrossRef]
  17. Alves, C.C.; Campos, T.B.C.; Alves, W.A. FT-Raman Spectroscopic Analysis of the Most Probable Structures in Aluminum Chloride and Tetrahydrofuran Solutions. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2012, 97, 1085–1088. [Google Scholar] [CrossRef] [PubMed]
  18. Kore, R.; Scurto, A.M.; Shiflett, M.B. Review of Isobutane Alkylation Technology Using Ionic Liquid-Based Catalysts–Where Do We Stand? Ind. Eng. Chem. Res. 2020, 59, 15811–15838. [Google Scholar] [CrossRef]
  19. Liu, Y.; Wu, G.; Hu, R.; Gao, G. Effects of Aromatics on Ionic Liquids for C4 Alkylation Reaction: Insights from Scale-up Experiment and Molecular Dynamics Simulation. Chem. Eng. J. 2020, 402, 126252. [Google Scholar] [CrossRef]
  20. Liu, G.; Wu, G.; Liu, Y.; Hu, R.; Gao, G. Theoretical Study on the C4 Alkylation Mechanism Catalyzed by Cu-Containing Chloroaluminate Ionic Liquids. Fuel 2022, 310, 122379. [Google Scholar] [CrossRef]
  21. Liu, Y.; Hu, R.; Xu, C.; Su, H. Alkylation of Isobutene with 2-Butene Using Composite Ionic Liquid Catalysts. Appl. Catal. A Gen. 2008, 346, 189–193. [Google Scholar] [CrossRef]
  22. Sanchez, B.; Campodónico, P.R.; Contreras, R. Gutmann’s Donor and Acceptor Numbers for Ionic Liquids and Deep Eutectic Solvents. Front. Chem. 2022, 10, 861379. [Google Scholar] [CrossRef] [PubMed]
  23. Erdmann, P.; Greb, L. Multidimensional Lewis Acidity: A Consistent Data Set of Chloride, Hydride, Methide, Water and Ammonia Affinities for 183 p-Block Element Lewis Acids. ChemPhysChem 2021, 22, 935–943. [Google Scholar] [CrossRef]
  24. Erdmann, P.; Greb, L. What Distinguishes the Strength and the Effect of a Lewis Acid: Analysis of the Gutmann–Beckett Method. Angew. Chemie 2022, 134, e202114550. [Google Scholar] [CrossRef]
  25. Mayer, U.; Gutmann, V.; Gerger, W. The Acceptor Number—A Quantitative Empirical Parameter for the Electrophilic Properties of Solvents. Monatshefte Für Chem. 1975, 106, 1235–1257. [Google Scholar] [CrossRef]
  26. Gutmann, V. The Donor-Acceptor Approach to Molecular Interactions; Springer: Boston, MA, USA, 1978; pp. 12–29. ISBN 978-1-4615-8827-6. [Google Scholar]
  27. Beckett, M.A.; Strickland, G.C.; Holland, J.R.; Sukumar Varma, K. A Convenient n.m.r. Method for the Measurement of Lewis Acidity at Boron Centres: Correlation of Reaction Rates of Lewis Acid Initiated Epoxide Polymerizations with Lewis Acidity. Polymer 1996, 37, 4629–4631. [Google Scholar] [CrossRef]
  28. Kerridge, D.H. The Chemistry of Molten Acetamide and Acetamide Complexes. Chem. Soc. Rev. 1988, 17, 181–227. [Google Scholar] [CrossRef]
  29. Jupp, A.R.; Johnstone, T.C.; Stephan, D.W. The Global Electrophilicity Index as a Metric for Lewis Acidity. Dalt. Trans. 2018, 47, 7029–7035. [Google Scholar] [CrossRef]
  30. Böhrer, H.; Trapp, N.; Himmel, D.; Schleep, M.; Krossing, I. From Unsuccessful H2-Activation with FLPs Containing B(Ohfip)3 to a Systematic Evaluation of the Lewis Acidity of 33 Lewis Acids Based on Fluoride, Chloride, Hydride and Methyl Ion Affinities. Dalt. Trans. 2015, 44, 7489–7499. [Google Scholar] [CrossRef]
  31. Erdmann, P.; Leitner, J.; Schwarz, J.; Greb, L. An Extensive Set of Accurate Fluoride Ion Affinities for p -Block Element Lewis Acids and Basic Design Principles for Strong Fluoride Ion Acceptors. ChemPhysChem 2020, 21, 987–994. [Google Scholar] [CrossRef]
  32. Meek, D.W.; Drago, R.S. A Coördination Model as an Alternative to the Solvent System Concept in Some Oxychloride Solvents. I. Similarity in the Behavior of Phosphorus Oxychloride and Triethyl Phosphate as Non-Aqueous Solvents. J. Am. Chem. Soc. 1961, 83, 4322–4325. [Google Scholar] [CrossRef]
  33. Abbott, A.P.; Al-Barzinjy, A.A.; Abbott, P.D.; Frisch, G.; Harris, R.C.; Hartley, J.; Ryder, K.S. Speciation, Physical and Electrolytic Properties of Eutectic Mixtures Based on CrCl3·6H2O and Urea. Phys. Chem. Chem. Phys. 2014, 16, 9047–9055. [Google Scholar] [CrossRef] [PubMed]
  34. Pires, E.; Fraile, J.M. Study of Interactions between Brønsted Acids and Triethylphosphine Oxide in Solution by 31P NMR: Evidence for 2:1 Species. Phys. Chem. Chem. Phys. 2020, 22, 24351–24358. [Google Scholar] [CrossRef]
  35. Wilkes, J.S.; Frye, J.S.; Reynolds, G.F. 27A1 and 13C NMR Studies of Aluminum Chloride-Dialkylimidazolium Chloride Molten Salts. Inorg. Chem. 1983, 22, 3870–3872. [Google Scholar] [CrossRef]
  36. Cui, J.; De With, J.; Klusener, P.A.A.; Su, X.; Meng, X.; Zhang, R.; Liu, Z.; Xu, C.; Liu, H. Identification of Acidic Species in Chloroaluminate Ionic Liquid Catalysts. J. Catal. 2014, 320, 26–32. [Google Scholar] [CrossRef]
  37. Zhang, X.; Zhang, R.; Liu, H.; Meng, X.; Xu, C.; Liu, Z.; Klusener, P.A.A. Quantitative Characterization of Lewis Acidity and Activity of Chloroaluminate Ionic Liquids. Ind. Eng. Chem. Res. 2016, 55, 11878–11886. [Google Scholar] [CrossRef]
  38. Gale, R.J.; Gilbert, B.; Osteryoung, R.A. Raman Spectra of Molten Aluminum Chloride: 1-Butylpyridinium Chloride Systems at Ambient Temperatures. Inorg. Chem. 1978, 17, 2728–2729. [Google Scholar] [CrossRef]
  39. Wicelinski, S.P.; Gale, R.J.; Williams, S.D.; Mamantov, G. Raman Spectra of Molten Gallium Chloride: 1-Methyl-3-Ethylimidazolium Chloride at Ambient Temperatures. Spectrochim. Acta Part A Mol. Spectrosc. 1989, 45, 759–762. [Google Scholar] [CrossRef]
  40. Corma, A.; Martínez, A. Chemistry, Catalysts, and Processes for Isoparaffin–Olefin Alkylation: Actual Situation and Future Trends. Catal. Rev. Sci. Eng. 1993, 35, 483–570. [Google Scholar] [CrossRef]
  41. Estager, J.; Oliferenko, A.A.; Seddon, K.R.; Swadźba-Kwaśny, M. Chlorometallate(III) Ionic Liquids as Lewis Acidic Catalysts—A Quantitative Study of Acceptor Properties. Dalt. Trans. 2010, 39, 11375. [Google Scholar] [CrossRef]
  42. Neese, F.; Wennmohs, F.; Becker, U.; Riplinger, C. The ORCA Quantum Chemistry Program Package. J. Chem. Phys. 2020, 152, 224108. [Google Scholar] [CrossRef] [PubMed]
  43. Purvis, G.D.; Bartlett, R.J. A Full Coupled-Cluster Singles and Doubles Model: The Inclusion of Disconnected Triples. J. Chem. Phys. 1982, 76, 1910–1918. [Google Scholar] [CrossRef]
  44. Grant, D.J.; Dixon, D.A.; Camaioni, D.; Potter, R.G.; Christe, K.O. Lewis Acidities and Hydride, Fluoride, and X- Affinities of the BH3-NXn Compounds for (X = F, Cl, Br, I, NH2, OH, and SH) from Coupled Cluster Theory. Inorg. Chem. 2009, 48, 8811–8821. [Google Scholar] [CrossRef] [PubMed]
  45. Wu, G.; Liu, Y.; Liu, G.; Hu, R.; Gao, G. Role of Aromatics in Isobutane Alkylation of Chloroaluminate Ionic Liquids: Insights from Aromatic—Ion Interaction. J. Catal. 2021, 396, 54–64. [Google Scholar] [CrossRef]
Figure 1. 27Al NMR spectra of amide-AlCl3 systems: (a) butanamide-, propionamide-, and acetamide-based LCCs; (b) Acetamide-AlCl3 based on various ligands at XAlCl3 = 0.50, 0.67.
Figure 1. 27Al NMR spectra of amide-AlCl3 systems: (a) butanamide-, propionamide-, and acetamide-based LCCs; (b) Acetamide-AlCl3 based on various ligands at XAlCl3 = 0.50, 0.67.
Molecules 28 07857 g001
Figure 2. 27Al NMR spectra of LCCs: (a) chloroformamide-, chloroacetamide-, and chloropropionamide-AlCl3 at XAlCl3 = 0.50; (b) chloroacetamide-AlCl3 based on various ligands at XAlCl3 = 0.50, 0.67.
Figure 2. 27Al NMR spectra of LCCs: (a) chloroformamide-, chloroacetamide-, and chloropropionamide-AlCl3 at XAlCl3 = 0.50; (b) chloroacetamide-AlCl3 based on various ligands at XAlCl3 = 0.50, 0.67.
Molecules 28 07857 g002
Figure 3. Raman spectra: (a) amide-AlCl3-based LCCs with XAlCl3 = 0.67; (b) chloroacetamide-AlCl3-based LCC with dominant bands for Al-Cl stretching frequencies assigned to chloroaluminate anions.
Figure 3. Raman spectra: (a) amide-AlCl3-based LCCs with XAlCl3 = 0.67; (b) chloroacetamide-AlCl3-based LCC with dominant bands for Al-Cl stretching frequencies assigned to chloroaluminate anions.
Molecules 28 07857 g003
Figure 4. (a) Monodentate coordination; (b) bidentate coordination of acetamide−AlCl3. Element color: C gray, Cl green, Al orange, O red, N blue, and H light gray.
Figure 4. (a) Monodentate coordination; (b) bidentate coordination of acetamide−AlCl3. Element color: C gray, Cl green, Al orange, O red, N blue, and H light gray.
Molecules 28 07857 g004
Figure 5. UV—vis spectra of LCCs (1 wt.%) in the dichloromethane solution.
Figure 5. UV—vis spectra of LCCs (1 wt.%) in the dichloromethane solution.
Molecules 28 07857 g005
Figure 6. Scheme of the C4 alkylation reaction.
Figure 6. Scheme of the C4 alkylation reaction.
Molecules 28 07857 g006
Table 1. AN values and C4 alkylation performances of Lewis acid systems.
Table 1. AN values and C4 alkylation performances of Lewis acid systems.
No.CatalystsANXAlCl3Composition of Alkylate, wt%
C5–7C8 1C9+
1[(C2H5)3NH]Cl−AlCl370.70.50---
2[(C2H5)3NH]Cl−AlCl392.20.6730.234.235.6
3[BMIm]Cl−AlCl391.70.6723.542.034.5
4acetamide−AlCl382.20.502.284.1 213.7
5butanamide−AlCl396.40.6720.445.634.0
6propionamide−AlCl397.00.6720.248.331.5
7acetamide−AlCl397.60.6720.143.436.5
8chloropropionamide−AlCl398.20.6725.448.226.4
9chloroacetamide−AlCl399.10.6724.249.626.2
10chloroformamide−AlCl3102.00.6737.544.717.8
1 C8 include trimethyl-pentane and dimethyl-hexane; 2 Herein, C8 are mainly the mixture of C8 olefins.
Table 2. GB and computed FIA values.
Table 2. GB and computed FIA values.
Lewis AcidsGBConstructed Species Model, FIA
IIIIII
[(C2H5)3NH]Cl−AlCl3 122.9(C2H5)3NH+/AlCl4, 345
[(C2H5)3NH]Cl−AlCl332.6(C2H5)3NH+/Al2Cl7, 500
[BMIm]Cl−AlCl332.4BMIm+/Al2Cl7, 490
acetamide−AlCl3 123.1AlCl2L2+/AlCl4, 426AlCl3L2/AlCl3L/Al2Cl6L, 517
butanamide−AlCl334.5AlCl2L2+/Al2Cl7, 505AlCl3L/Al2Cl6L, 531AlCl3L2/AlCl3L/Al2Cl6L, 589
propionamide−AlCl334.8AlCl2L2+/Al2Cl7, 515AlCl3L/Al2Cl6L, 539AlCl3L2/AlCl3L/Al2Cl6L, 594
acetamide−AlCl335.0AlCl2L2+/Al2Cl7, 520AlCl3L/Al2Cl6L, 550AlCl3L2/AlCl3L/Al2Cl6L, 603
chloropropionamide−AlCl335.3AlCl2L2+/Al2Cl7, 530AlCl3L/Al2Cl6L, 538AlCl3L2/AlCl3L/Al2Cl6L, 600
chloroacetamide−AlCl335.7AlCl2L2+/Al2Cl7, 547AlCl3L/Al2Cl6L, 580AlCl3L2/AlCl3L/Al2Cl6L, 602
chloroformamide−AlCl337.0AlCl2L2+/Al2Cl7, 575AlCl3L/Al2Cl6L, 603AlCl3L2/AlCl3L/Al2Cl6L, 616
1 XAlCl3 = 0.50, and XAlCl3 = 0.67 for the other Lewis acids.
Table 3. Relationships of GB, computed FIA, and predicted FIA values.
Table 3. Relationships of GB, computed FIA, and predicted FIA values.
AcidsFIA 1GBLCCGBPredicted FIA 2Computed FIA 3
AlMe336816.7acetamide−AlCl3 423.1427426
AlEt337915.9butanamide−AlCl334.5593589
[AlCalixMe]40218.5propionamide−AlCl334.8595594
AlN(OC7H4)341221.2acetamide−AlCl335.0597603
AlCl356532.2chloropropionamide−AlCl335.3600599
AlBr357733.1chloroacetamide−AlCl335.7604602
AlI358234.4chloroformamide−AlCl337.0616615
Al(C6F5)359636
1 Ref. [24]; 2 FIA values were predicted by the regression model. 3 FIA values were calculated by the DFT method. 4 XAlCl3 = 0.50, and XAlCl3 = 0.67 for the other LCCs.
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

Li, H.; Wu, Q.; Liu, Y.; Bao, J. Acidity Quantification and Structure Analysis of Amide-AlCl3 Liquid Coordination Complexes for C4 Alkylation Catalysis. Molecules 2023, 28, 7857. https://doi.org/10.3390/molecules28237857

AMA Style

Li H, Wu Q, Liu Y, Bao J. Acidity Quantification and Structure Analysis of Amide-AlCl3 Liquid Coordination Complexes for C4 Alkylation Catalysis. Molecules. 2023; 28(23):7857. https://doi.org/10.3390/molecules28237857

Chicago/Turabian Style

Li, Hao, Qiong Wu, Ying Liu, and Jinrong Bao. 2023. "Acidity Quantification and Structure Analysis of Amide-AlCl3 Liquid Coordination Complexes for C4 Alkylation Catalysis" Molecules 28, no. 23: 7857. https://doi.org/10.3390/molecules28237857

Article Metrics

Back to TopTop