Next Article in Journal
Exploration of Cucumber Waste as a Potential Biorefinery Feedstock
Previous Article in Journal
Potential Efficacy of Bacillus coagulans BACO-17 to Modulate Gut Microbiota in Rats Fed High-Fat Diet
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Insight into the Acidity and Catalytic Performance on Butane Isomerization of Thermal Stable Sulfated Monoclinic Zirconia

Department of Chemistry and Shanghai Key Laboratory of Molecular Catalysis and Innovative Materials, Fudan University, Shanghai 200438, China
*
Authors to whom correspondence should be addressed.
Processes 2022, 10(12), 2693; https://doi.org/10.3390/pr10122693
Submission received: 15 November 2022 / Revised: 8 December 2022 / Accepted: 8 December 2022 / Published: 13 December 2022
(This article belongs to the Section Catalysis Enhanced Processes)

Abstract

:
Sulfated monoclinic zirconia (M-SZ) with high thermal stability and high catalytic performance on butane isomerization were obtained by hydrothermal method followed with sulfation treatment. The acidity of M-SZ was studied by 31P MAS NMR, with trimethylphosphine (TMP) as the probe molecule, and the catalytic performance of 1-13C-n-butane over M-SZ was monitored by 13C MAS NMR spectroscopy. Both Brønsted and Lewis acids were observed on M-SZ. Only Brønsted acid strength shows close relation to the activation energy of butane isomerization, and the M-SZ catalyst with the strongest Brønsted acid strength shows the lowest activation energy of 46.4 kJ·mol−1. The catalytic stability tests were evaluated at 673 K for 240 h, which shows that sulfated monoclinic zirconia has higher thermal stability than sulfated tetragonal zirconia.

Graphical Abstract

1. Introduction

With the reduction in fossil energy reserves, promoting resource efficiency has become an important topic in current catalysis research. Isomerization of paraffin can convert straight-chain paraffin into higher-value branched-chain paraffin, which is an effective way to improve resource utilization efficiency [1]. As the product of butane isomerization, isobutane is an important feedstock for the additive MTEB used to enhance gasoline octane number of synthetic gasoline. Much progress has been made in the development of industrial catalysts for butane isomerization, such as Pt/Al2O3-Cl, Pt-MOR, and sulfated oxide catalysts [2,3,4]. As an important catalyst with high catalytic performance, sulfated zirconia has been widely studied because of its super acidity, which is of great potential in the butane isomerization reaction [5,6,7,8].
It is generally believed that sulfated tetragonal zirconia has super acidity, while the acidic strength of sulfated monoclinic zirconia is relatively low [9]. Therefore, researchers mainly focus on the in-depth understanding of the structure and catalytic properties of sulfated tetragonal zirconia [9,10] and pay much less attention to monoclinic zirconia sulfation. Stichert et al. compared the monoclinic and tetragonal sulfated zirconia in catalyzing butane isomerization [11]. It was found that both phases of sulfated zirconia are catalytically active in the butane isomerization reaction, indicating that the sulfated monoclinic zirconia can also act as an acid catalyst. They also studied the surface acidity of sulfated monoclinic and tetragonal zirconia by adsorption of CO [12] and found similar acid strength of the two catalysts. However, the catalytic performance for butane isomerization of sulfated monoclinic zirconia is about 25% of that of sulfated tetragonal zirconia. Zhang et al. studied the dehydration cyclization reaction of sorbitol on sulfated zirconia catalysts prepared from zirconium hydroxide, tetragonal zirconia, and monoclinic zirconia [13]. The results showed that the catalytic performance of sulfated tetragonal synthesized from zirconium hydroxide was significantly higher than those prepared from tetragonal and monoclinic zirconia. Meanwhile, the catalytic performance of sulfated monoclinic zirconia obtained from monoclinic zirconia is higher than that of sulfated tetragonal zirconia synthesized from tetragonal zirconia. In addition, the thermodynamic stability of the monoclinic phase (<1448 K) is significantly higher than that of the tetragonal phase [14], so the zirconia could undergo a tetragonal-to-monoclinic phase transformation at high catalytic reaction temperatures. Therefore, the research on sulfated monoclinic zirconia may expand the types of sulfated zirconia superacid catalysts, leading to a deeper exploring of highly thermal stable sulfated zirconia catalysts.
Magic-angle spinning solid-state nuclear magnetic resonance (MAS NMR) is widely used in characterizing the surface acidity of the catalysts, and it is also a useful tool to analyze surface catalytic reactions. On the one hand, the surface acidity of the catalysts may be studied by directly detecting the hydroxy group with 1H NMR [15,16] or characterizing the probe molecules adsorbed on the catalyst, such as 31P, 13C, and 15N NMR spectroscopies with trimethylphosphine (TMP) or trimethylphosphine oxide (TMPO) [17,18], 13C-acetone [19], and 15N-pyridine [20] as probe molecules, respectively. On the other hand, solid-state NMR spectroscopy using isotopically labeled reaction molecules may monitor the evolution of related molecules during the reaction and study the catalytic reaction mechanism [21,22].
In this paper, a series of monoclinic zirconia and their sulfated counterparts were prepared. The catalytic reaction was studied by 13C MAS NMR, with 1-13C-n-butane as the reactant. The surface acidity of the catalyst was characterized by 31P MAS NMR, using TMP as the probe molecule. The effect of surface acidity on the butane isomerization reaction was studied.

2. Materials and Methods

2.1. Preparation of the Samples

Monoclinic zirconia (M-ZrO2) was prepared by the hydrothermal method. Typically, 9.25 g of zirconyl nitrate hydrate (ZrO(NO3)2·xH2O, 99.5%, Shanghai Aladdin Biochemical Technology Co., Ltd., Shanghai, China) and 4.80~28.81 g of urea (CO(NH2)2, 99%, Sigma-Aldrich, Co., Saint Louis, MO, USA) were dissolved in 100 mL of deionized water. The solution was then poured into a 200 mL Teflon-lined stainless steel autoclave was treated at different temperatures (393, 433, and 473 K) for 12 h. The obtained white solid was washed with deionized water, separated by centrifugation, and dried at 393 K overnight. It was then calcined in static air with a heating rate of 2 K min−1 to 673 K for 4 h in a tube furnace. Sulfated zirconia was prepared by the impregnation method. A total of 1 g of monoclinic zirconia was dispersed in 15 mL of 1 mol·L−1 H2SO4 solution, stirred for 30 min at room temperature, and, finally, filtered and dried at 393 K. The white solid was put into a tube furnace and calcined in static air at 823 K for 3 h with 5 K min−1 rate. Among the synthesized monoclinic zirconia samples, as shown in Figure S1, three samples with Zr:Urea molar ratio of 1:12, 1:6, and 1:2 were selected for the following study and named as M-ZrO2-1, M-ZrO2-2, and M-ZrO2-3, and their sulfated counterparts were named as M-SZ-1, M-SZ-2, and M-SZ-3, respectively. For M-ZrO2-1, 28.81 g of urea was used, and the hydrothermal temperature was 393 K. For M-ZrO2-2, 14.40 g of urea was used, and the hydrothermal temperature was 433 K. For M-ZrO2-3, 4.80 g of urea was used, and the hydrothermal temperature was 473 K. In addition, another sample, M-SZ-4, was prepared by sulfation using M-ZrO2-2 as the precursor, and 2 mol·L−1 H2SO4 solution was used instead of 1 mol·L−1 H2SO4 solution.
Tetragonal zirconia was prepared by the precipitation method [23]. A total of 24.17 g of zirconyl chloride octahydrate (ZrOCl2·8H2O, 99%, Shanghai Aladdin Biochemical Technology Co., Ltd., Shanghai, China) was dissolved into 150 mL of deionized water, and 50 mL of ammonia solution (NH3·H2O 14.8 mol·L−1, Sinopharm Chemical Reagent Co., Ltd., Shanghai, China) was then dropped into the zirconium solution at a rate of 30 mL·h−1. The mixture was stirred at room temperature for 48 h. Using deionized water as the washing reagent, the white sample was washed and filtrated three times. After drying at 393 K overnight, the sample was put into the tube furnace and calcined under air at 673 K for 3 h with 5 K·min−1 rate. A total of 1 g of Zr(OH)4 was sulfated in a similar process to the sulfation of M-ZrO2 at a calcination temperature of 873 K. The obtained white powder was named as T-SZ-general.

2.2. Characterization of the Catalysts

X-ray diffraction (XRD) patterns were obtained on a Bruker D2 PHASER diffractometer. The Cu-Kα radiation was used (λ = 0.15418 nm) at the voltage of 30 kV and the current of 10 mA. The range of 2θ was from 10° to 80°and the step width was 0.02°, while the step time was 0.2 s.
Fourier transform infrared (FT-IR) spectra were recorded on a Nicolet iS 10 infrared spectrometer from Thermo Fisher Scientific (Waltham, MA, USA). The detection range is from 400 to 4000 cm−1. The sample was dispersed in KBr and pressed into a disc before the detection.
X-ray photoelectron spectra were collected using a PHI 5000C ESCA spectrometer from PerkinElmer (Waltham, MA, USA). The source was monochromatic Mg Kα radiation (1253.6 eV) with the voltage of 14 kV and the current of 20 mA. The banding energy of all elements was calibrated by C 1s of the contaminated carbon at 284.8 eV.
The microstructure images of the samples were collected on an FEI Tecnai Ge F20 S-Twin transmission electron microscope (TEM). The samples were dispersed in ethanol by ultrasonic for 600 s, and the suspensions were dripped onto copper grids coated with carbon film.
The N2 adsorption–desorption isotherms were obtained with a Micromeritics Tristar II 3020 porosimetry analyzer at 77 K. Before the measurement, the samples were treated at 573 K for 2 h to remove physically adsorbed gases before the measurement. The specific surface areas of the samples were calculated by Brunauer–Emmett–Teller (BET) equation. The cumulative pore volumes and average pore diameters were calculated from the desorption branch of the isotherms by the Barrett–Joyner–Halenda (BJH) model.
For the 31P MAS NMR experiments, an appropriate amount (0.15 ~ 0.20 g) of catalyst was put into a glass tube, shown in Figure S2b. The tube was connected to a vacuum system and evacuated to about 1~2 Pa. The samples were then heated to 573 K at a heating rate of 10 K·min−1 and kept for 2 h. The TMP probe molecules were adsorbed onto the catalyst through a vacuum system, and the amounts of the probe molecules adsorbed on the samples were controlled by the pressure change. The glass sample tube with the sample was sealed by flame and transferred to the glove box, which is filled with high-purity nitrogen. The sample was then packed into 4 mm zirconia rotors with Kel-F caps in the glove box. Using a Bruker ADVANCE III 400 WB spectrometer, the MAS NMR spectra were recorded. The resonance frequency of 31P was 161.9 MHz to be consistent with the nucleus, and 85 wt% H3PO4 was used as the external standard. 1H high-power decoupling and one-pulse 31P MAS NMR experiments were conducted at 12 kHz magic angle speed, 1.1 μs pulse width (30° pulse), 15 s relaxation delay time, and 120 to 960 scans.

2.3. Solid-State NMR Study on Isomerization of 1-13C-n-Butane

An appropriate amount of catalyst (50~100 mg) was put into a glass tube, which can be well-matched with a 7 mm NMR rotor (Figure S2c). Similar to the pretreatment of the sample in Section 2.2, 400 μmol·g−1 of 1-13C-n-butane was introduced at liquid nitrogen temperature. Then, the glass tube with the catalyst was protected in liquid nitrogen and symmetrically flame-sealed for achieving the desirable MAS speed, as shown in Figure S2. In the NMR study, the sealed glass sample tube was heated in a tube furnace at a desirable reaction temperature for a certain time and then put into the liquid nitrogen to quench the reaction. The 13C MAS NMR spectra were acquired by Bruker ADVANCE III 400 WB spectrometer at 100.7 MHz using tetramethylsilane (TMS) as the external standard. 1H high-power decoupling 13C MAS NMR experiments were conducted at ca. 1 kHz MAS speed, 1.2 μs pulse width (30° pulse), 8 s relaxation delay time, and 120 to 360 scans.
The content of reactant or product (Ci,t) at time t is:
C i , t = I i , t   I i , 0 × 100 %
In this equation, Ii,0 represents the peak area of the reactant and the products in the original NMR spectrum. Ii,t is the peak area of the reactant and the products after reacting for t min. The conversion of 1-13C-n-butane after reacting for t min was calculated as:
Xt = 1 − Ca,t (a represents 1-13C-n-butane)

3. Results and Discussion

3.1. Characterization

The XRD patterns of M-ZrO2-1, M-ZrO2-2, M-ZrO2-3, M-SZ-1, M-SZ-2, and M-SZ-3 are shown in Figure 1a. The diffraction peaks at 28.2° and 31.5° correspond to (-111) and (111) planes of monoclinic zirconia (JCPDS No. 37-1484). The results show that the sulfation only decreases the full width at half maximum (FWHM) of the peaks, indicating the unchanged crystal phase and the increase in particle size after the sulfation. The FT-IR spectra in Figure 1b show that M-SZ-1, M-SZ-2, and M-SZ-3 have distinct characteristic bands at 1232, 1132, 1042, and 993 cm−1, where the first two bands correspond to S=O asymmetry and symmetric stretching vibrations and the latter two bands correspond to the asymmetric and symmetric stretching vibrations of S-O, indicating that the M-ZrO2-1, M-ZrO2-2, and M-ZrO2-3 samples were successfully sulfated.
The XPS spectra of zirconia and their sulfated counterparts are shown in Figure 2. The peaks near 530.8 eV in M-ZrO2 correspond to the O 1s electrons of the oxygen ions in zirconia, while the peaks near 532.2 eV are related to the oxygen ions affected by sulfation [24]. The peaks near 182.6 and 185.0 eV are related to the Zr 3d5/2 and Zr 3d3/2 electrons of the tetravalent-state zirconium ions in zirconia [25]. All XPS peaks of zirconia, as well as sulfated zirconia, indicate no noticeable change in the chemical environment of oxygen ions and zirconium ions.
Figure 3 displays the TEM images of three monoclinic zirconia and their sulfated catalysts. All three samples contain irregular pore structures. Compared with M-ZrO2, no significant changes were observed in the morphology and pore structure.
The N2 adsorption–desorption isotherms and pore size distributions of the samples are shown in Figure 4. The isotherms are type IV with H1 hysteresis loops, indicating that the samples have mesoporous structures. The results related to the texture properties are shown in Table 1. After the calcination in sulfation, all three sulfated zirconia samples decrease significantly in their specific surface area and increase in the pore volume and pore size.
Figure 5a illustrates the 1H-decoupled 31P MAS NMR spectra of three sulfated zirconia after adsorption of TMP, and the relevant results are listed in Table 2. The peaks in the range of 0~−7 ppm are related to TMP adsorbed on Brønsted acid sites (BA), and those between −20 and 60 ppm correspond to TMP adsorbed on Lewis acid sites (LA) [26]. The 31P chemical shifts of the peaks related to BA on M-SZ-1, M-SZ-2, and M-SZ-3 are −3.4, −3.2, and −3.3 ppm, which reflect the BA acid strength being in the order of M-SZ-2 > M-SZ-3 > M-SZ-1. However, it is worth mentioning that a large change in the BA acid strength only corresponds to a relatively narrow change in 31P chemical shifts of adsorbed TMP [27]. Therefore, the 31P chemical shift of adsorbed TMP on M-SZ-1, M-SZ-2, and M-SZ-3 may not be able to clarify their BA acid strength unambiguously. The J-coupling constants (JP-H) of adsorbed TMP on the BA site (TMPH+) may also be used to characterize the BA acid strength. As shown in Figure 5b, the peaks near −3.3 ppm are split into two peaks due to the J-coupling [28]. The JP-H of M-SZ-1, M-SZ-2, and M-SZ-3 was measured as 437, 494, and 474 Hz, respectively. Since higher JP-H reflects stronger acid strength [29], the BA acid strength of the three catalysts is in the order of M-SZ-2 > M-SZ-3 > M-SZ-1, which is consistent with the order based on the 31P chemical shifts of TMP adsorbed on BA acid sites. As for the BA acid amount, M-SZ-2 is 112 μmol·g−1, which is the highest among the three samples.
By deconvolution of the peaks attributed to LAs, the LAs of the three catalysts mainly contain five types of LAs, named LA1~5, as shown in Table 2. Among them, LA4 and LA5 are related to the original LAs on monoclinic zirconia, and LA1, LA2, and LA3 correspond to the sulfate-related LA on the surface of sulfated zirconia [30,31]. By comparing the 31P chemical shifts of TMP adsorbed on LA1, LA2, and LA3 shown in Table 2, the orders of the LA1~3 acid strength of the three catalysts are also M-SZ-2 > M-SZ-3 > M-SZ-1. However, the 31P chemical shifts of TMP related to LA1~3 of the three catalysts are similar, so the differences in LA1~3 acid strength are not significant. The LA amounts of three M-SZ are also recorded in Table 2.

3.2. The Isomerization of Butane Catalyzed by M-SZ Catalysts

The isomerization of butane on the three catalysts was monitored by 13C MAS NMR spectroscopy, using 1-13C-n-butane as the reactant. Figure 6 shows the 13C MAS NMR spectra of the catalytic reaction at different times on M-SZ-2. The peak at 13.0 ppm corresponds to the adsorbed 1-13C-n-butane, and the peak at 10.5 ppm is assigned to the gaseous 1-13C-n-butane [32]. As the reaction time increases, peaks at 25.0 and 23.7 ppm appear, corresponding to 2-13C-n-butane and 13C-isobutane (1-13C-isobutane and 2-13C-isobutane), respectively. The by-product 13C-isopentane related to the peak at 21.6 ppm may be observed at 10 min, and a new peak at 15.2 ppm corresponding to 13C-propane appears at 60 min. The results of the catalytic reaction on different M-SZ catalysts at 433 K for 60 min are shown in Table 3, which shows that M-SZ-2 has the highest catalytic performance.
The distributions of reactants and products in the butane isomerization reaction catalyzed by M-SZ-2 at 393, 413, and 433 K are shown in Figure 7. According to the Langmuir Hinshelwood rate equation for the reversible first-order surface reaction [33,34], the rate of butane isomerization is [22,35]:
r = kK nC 4 ( P nC 4 - K iC 4 P iC 4 / K nC 4 K eq ) ( 1 + K nC 4 P nC 4 + K iC 4 P iC 4 )
where k is the reaction rate constant, P nC 4 and P iC 4 are the partial pressure of n-butane and isobutane, K nC 4 and K iC 4 are the adsorption equilibrium constant of n-butane and isobutane, respectively, and Keq is the reaction equilibrium constant.
At relatively low pressure, it is acceptable to suppose KnC4 = KiC4 and KnC4(PnC4 + PiC4) <<1, and Equation (1) can be converted to:
- ln 1 - 1 + 1 K eq P iC 4 P total = kK nC 4 1 + 1 K eq t
By plotting − ln [1 − (1 + 1/Keq)PiC4/Ptotal]/(1 + 1/Keq) against the reaction time (t), Figure 8a with a good linear relationship was obtained, indicating that the catalytic reaction is mainly through a single molecule mechanism at 393~433 K. The rate constant of butane isomerization catalyzed by M-SZ-2 at 393~433 K may be obtained from the slope in Figure 8a and is shown in Table 4. Further plotting ln(kKnC4) against 1000/T to obtain Figure 8b, the activation energy (Ea) of the M-SZ-2 catalyst for catalyzing butane isomerization can be calculated as 52.8 kJ·mol−1. Hsu et al. studied the activation energies of butane isomerization catalyzed by sulfated zirconia, which was 44.8 kJ mol−1 [36]. Ma et al. obtained the Ea of butane isomerization catalyzed by sulfated tetragonal zirconia as 45.6 kJ·mol−1 [34], and Na et al. obtained the Ea of butane isomerization catalyzed by Cs2.5H0.5PW12O40 as 59 kJ·mol−1 [37]. The Ea of M-SZ-2 is close to the literature results above, indicating that M-SZ-2 also has a high acidic catalytic ability.
The results of butane isomerization on M-SZ-1 and M-SZ-3 are shown in Figures S3 and S4. Figures S5 and S6 and Tables S1 and S2 show their corresponding kinetic results. The activation energies of butane isomerization of M-SZ-1 and M-SZ-3 are 148 and 83.8 kJ·mol−1, respectively.
According to the acidity (Table 2) and Ea results of M-SZ-1, M-SZ-2, and M-SZ-3, it can be found that the Ea of M-SZ catalysts in butane isomerization has a weak correlation with BA acid amount, LA acid amount, and LA acid strength. Nevertheless, Ea is strongly related to the BA acid strength, particularly the JP-H of M-SZ-1, M-SZ-2, and M-SZ-3, which means the BA acid strength is one important factor that influences the butane isomerization. To further prove this inference, the M-SZ-4 catalyst was prepared through sulfation of M-ZrO2-2 with 2 M H2SO4 solution. Its characterization results in Figure S7 showed that M-SZ-4 has the same XRD pattern as that of M-SZ-2, and the JP-H value of M-SZ-4 obtained from 31P MAS NMR with TMP probe is 567 Hz, which is higher than the JP-H value of M-SZ-2 (494 Hz). This result shows that M-SZ-4 has stronger BA acid strength than M-SZ-2. The catalytic performance of M-SZ-4 is shown in Figure S8. From the kinetic study in Figure S9 and Table S3, the activation energy of M-SZ-4 is 46.4 kJ·mol−1. Therefore, the catalytic butane isomerization performance and the BA strength of M-SZ-4 further prove the stronger the BA strength, the lower the activation energy of butane isomerization.

3.3. The Stability of M-SZ and T-SZ Catalysts Operated at High Temperature

According to the literature results, the catalytic performance of butane isomerization over most sulfated tetragonal zirconia decreased with the increase in reaction time [38,39,40]. In order to compare the thermal stability of sulfated monoclinic zirconia and tetrahedral zirconia catalysts, M-SZ-2 and T-SZ-general were calcined at 673 K for 240 h and the catalysts before and after calcination were used to catalyze butane isomerization. The results are shown in Table 5. The conversion of 1-13C-n-butane and the yield to isobutane on M-SZ-2 only decrease by 7.8% and 12.9% after calcination, respectively. Although T-SZ-general has higher catalytic activity for catalyzing butane isomerization, after calcination at 673 K for 240 h, the conversion of 1-13C-n-butane and the yield to isobutane decreases by 17.8% and 30.9%, respectively. The results show that M-SZ-2 exhibits higher catalytic stability than T-SZ-general. The XRD patterns of M-SZ-2 and T-SZ-general samples are shown in Figure S10. It can be easily found that calcination at 673 K could result in the phase change of tetragonal zirconia and its sulfated zirconia, while the monoclinic zirconia and its sulfated zirconia remain unchanged.

4. Conclusions

Monoclinic-phase zirconia samples (M-ZrO2-1, M-ZrO2-2, and M-ZrO2-3) were synthesized and further sulfated to obtain sulfated monoclinic zirconia catalysts (M-SZ-1, M-SZ-2, M-SZ-3, and M-SZ-4). The acidity of monoclinic zirconia and M-SZ samples was thoroughly characterized by 31P MAS NMR with TMP probe. Among the acidic properties, including acid type, acid amounts, and acid strength, the BA acid strength plays an important role in catalyzing butane isomerization, i.e., the stronger the BA acid strength, the better the catalytic performance and the lower the activation energy. M-SZ-4 with the strongest Brønsted acidity shows the highest catalytic performance, with its activation energy of 46.4 kJ mol−1, which is close to that of sulfated tetragonal zirconia. The stability tests exhibited that sulfated monoclinic zirconia has better thermal stability than sulfated tetragonal zirconia although sulfated tetragonal zirconia has relatively higher activity. This study may light a way to explore strong solid acid catalysts with high thermal stability operated at relatively high temperatures.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/pr10122693/s1, Figure S1: The XRD patterns of zirconia synthesized at (a) 393 K, (b) 433 K, and (c) 473 K with different Zr:Urea molar ratio. Figure S2: Instruments for the sample treatment and reaction. Figure S3: Butane isomerization catalyzed by M-SZ-1 at (a) 493 K, (b) 513 K and (c) 533 K. Figure S4: Butane isomerization catalyzed by M-SZ-3 at (d) 433 K, (e) 453 K and (f) 473 K. Figure S5: The kinetic study of butane isomerization at different temperatures catalyzed by M-SZ-1. Figure S6: The kinetic study of butane isomerization at different temperatures catalyzed by M-SZ-3. Figure S7: (a) XRD patterns of M-SZ-2 and M-SZ-4, (b) 31P MAS NMR spectrum of M-SZ-4. Figure S8: Butane isomerization catalyzed by M-SZ-4 at (a) 373 K, (b) 413 K, and (c) 433 K. Figure S9: The kinetic study of butane isomerization at different temperatures catalyzed by M-SZ-4. Figure S10: The XRD patterns of (a) T-ZrO2-general, (b) T-SZ-general, (c) M-ZrO2-2,, and (d) M-SZ-2 calcined in air at 673 K for different times. Table S1: The kinetic constants of butane isomerization catalyzed by M-SZ-1 at 493~533 K. Table S2: The kinetic constants of butane isomerization catalyzed by M-SZ-3 at 433~473 K. Table S3: The kinetic constants of butane isomerization catalyzed by M-SZ-4 at 373~433 K.

Author Contributions

Methodology, D.H., B.Y. and H.H.; Conceptualization, B.Y. and H.H.; Data curation, D.H.; Formal analysis, D.H., W.F., B.Y. and H.H.; Investigation, D.H.; Project administration, H.H.; Writing—original draft, D.H.; Supervision, L.Z., B.Y. and H.H.; Funding acquisition, H.H.; Writing—review and editing, D.H., W.F., L.Z., B.Y. and H.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the National Natural Science Foundation of China (22088101), the Ministry of Science and Technology (2022YFA1503504), and SINOPEC (420068-2).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wang, P.Z.; Yue, Y.Y.; Wang, T.H.; Bao, X.J. Alkane isomerization over sulfated zirconia solid acid system. Int. J. Energy Res. 2020, 44, 3270–3294. [Google Scholar] [CrossRef]
  2. Potter, M.E.; Le Brocq, J.J.M.; Oakley, A.E.; McShane, E.B.; Vandegehuchte, B.D.; Raja, R. Butane isomerization as a diagnostic tool in the rational design of solid acid catalysts. Catalysts 2020, 10, 1099. [Google Scholar] [CrossRef]
  3. Echevskii, G.V.; Aksenov, D.G.; Kodenev, E.G.; Ovchinnikova, E.V.; Chumachenko, V.A. Activity of a sulfated zirconia catalyst in isomerization of n-butane fractions. Pet. Chem. 2019, 59, S101–S107. [Google Scholar] [CrossRef]
  4. Villegas, J.I.; Kumar, N.; Heikkila, T.; Smieskova, A.; Hudec, P.; Salmi, A.; Murzin, D.Y. A highly stable and selective Pt-modified mordenite catalyst for the skeletal isomerization of n-butane. Appl. Catal. A Gen. 2005, 284, 223–230. [Google Scholar] [CrossRef]
  5. Yadav, G.D.; Nair, J.J. Sulfated zirconia and its modified versions as promising catalysts for industrial processes. Microporous Mesoporous Mater. 1999, 33, 1–48. [Google Scholar] [CrossRef]
  6. Song, X.M.; Sayari, A. Sulfated zirconia-based strong solid-acid catalysts: Recent progress. Catal. Rev. Sci. Eng. 1996, 38, 329–412. [Google Scholar] [CrossRef]
  7. Adeeva, V.; Liu, H.Y.; Xu, B.Q.; Sachtler, W.M.H. Alkane isomerization over sulfated zirconia and other solid acids. Top. Catal. 1998, 6, 61–76. [Google Scholar] [CrossRef]
  8. Yamaguchi, T. Alkane isomerization and acidity assessment on sulfated ZrO2. Appl. Catal. A Gen. 2001, 222, 237–246. [Google Scholar] [CrossRef]
  9. Morterra, C.; Cerrato, G.; Pinna, F.; Signoretto, M. Crystal phase, spectral features, and catalytic acitivity of sulfate-doped zirconia systems. J. Catal. 1995, 157, 109–123. [Google Scholar] [CrossRef]
  10. Comelli, R.A.; Vera, C.R.; Parera, J.M. Influence of ZrO2 crystalline-structure and sulfate ion concentration on the catalytic activity of SO42−-ZrO2. J. Catal. 1995, 151, 96–101. [Google Scholar] [CrossRef]
  11. Stichert, W.; Schuth, F. Synthesis of catalytically active high surface area monoclinic sulfated zirconia. J. Catal. 1998, 174, 242–245. [Google Scholar] [CrossRef]
  12. Stichert, W.; Schüth, F.; Kuba, S.; Knözinger, H. Monoclinic and tetragonal high surface area sulfated zirconias in butane isomerization: CO adsorption and catalytic results. J. Catal. 2001, 198, 277–285. [Google Scholar] [CrossRef]
  13. Zhang, X.G.; Rabee, A.I.M.; Isaacs, M.; Lee, A.F.; Wilson, K. Sulfated zirconia catalysts for D-sorbitol cascade cyclodehydration to isosorbide: Impact of zirconia phase. ACS Sustain. Chem. Eng. 2018, 6, 14704–14712. [Google Scholar] [CrossRef]
  14. Shukla, S.; Seal, S. Mechanisms of room temperature metastable tetragonal phase stabilisation in zirconia. Int. Mater. Rev. 2005, 50, 45–64. [Google Scholar] [CrossRef]
  15. Morales, M.D.; Infantes-Molina, A.; Lázaro-Martínez, J.M.; Romanelli, G.P.; Pizzio, L.R.; Rodríguez-Castellón, E. Heterogeneous acid catalysts prepared by immobilization of H3PW12O40 on silica through impregnation and inclusion, applied to the synthesis of 3H-1,5-benzodiazepines. Mol. Catal. 2020, 485, 110842. [Google Scholar] [CrossRef]
  16. Yu, H.G.; Fang, H.J.; Zhang, H.L.; Li, B.J.; Deng, F. Acidity of sulfated tin oxide and sulfated zirconia: A view from solid-state NMR spectroscopy. Catal. Commun. 2009, 10, 920–924. [Google Scholar] [CrossRef]
  17. Zheng, A.M.; Liu, S.B.; Deng, F. 31P NMR chemical shifts of phosphorus probes as reliable and practical acidity scales for solid and liquid catalysts. Chem. Rev. 2017, 117, 12475–12531. [Google Scholar] [CrossRef]
  18. Lunsford, J.H.; Sang, H.; Campbell, S.M.; Liang, C.H.; Anthony, R.G. An NMR study of acid sites on sulfated-zirconia catalysts using trimethylphosphine as a probe. Catal. Lett. 1994, 27, 305–314. [Google Scholar] [CrossRef]
  19. Barich, D.H.; Nicholas, J.B.; Xu, T.; Haw, J.F. Theoretical and experimental study of the 13C chemical shift tensors of acetone complexed with Brønsted and Lewis acids. J. Am. Chem. Soc. 1998, 120, 12342–12350. [Google Scholar] [CrossRef]
  20. 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]
  21. Ma, Z.Y.; Yang, C.; Wei, W.; Li, W.H.; Sun, Y.H. Catalytic performance of copper supported on zirconia polymorphs for CO hydrogenation. J. Mol. Catal. A Chem. 2005, 231, 75–81. [Google Scholar] [CrossRef]
  22. Zhang, L.; Yue, B.; Ren, Y.H.; Chen, X.Y.; He, H.Y. An aluminum promoted cesium salt of 12-tungstophosphoric acid: A catalyst for butane isomerization. Catal. Sci. Technol. 2013, 3, 2113–2118. [Google Scholar] [CrossRef]
  23. Chuah, G.K.; Jaenicke, S.; Cheong, S.A.; Chan, K.S. The influence of preparation conditions on the surface area of zirconia. Appl. Catal. A Gen. 1996, 145, 267–284. [Google Scholar] [CrossRef]
  24. Bianchi, C.L.; Ardizzone, S.; Cappelletti, G. Surface state of sulfated zirconia: The role of the sol-gel reaction parameters. Surf. Interface Anal. 2004, 36, 745–748. [Google Scholar] [CrossRef]
  25. Bosman, H.J.M.; Pijpers, A.P.; Jaspers, A. An X-ray photoelectron spectroscopy study of the acidity of SiO2-ZrO2 mixed oxides. J. Catal. 1996, 161, 551–559. [Google Scholar] [CrossRef]
  26. Fu, Y.Y.; Zhang, L.; Yue, B.; Chen, X.Y.; He, H.Y. Simultaneous characterization of solid acidity and basicity of metal oxide catalysts via the solid-state NMR technique. J. Phys. Chem. C 2018, 122, 24094–24102. [Google Scholar] [CrossRef]
  27. Chu, Y.Y.; Yu, Z.W.; Zheng, A.M.; Fang, H.J.; Zhang, H.L.; Huang, S.J.; Liu, S.B.; Deng, F. Acidic strengths of Brønsted and Lewis acid sites in solid acids scaled by 31P NMR chemical shifts of adsorbed trimethylphosphine. J. Phys. Chem. C 2011, 115, 7660–7667. [Google Scholar] [CrossRef]
  28. Zhao, B.Y.; Pan, H.J.; Lunsford, J.H. Characterization of [(CH3)3P-H]+ complexes in normal H-Y, dealuminated H-Y, and H-ZSM-5 zeolites using 31P solid-state NMR spectroscopy. Langmuir 1999, 15, 2761–2765. [Google Scholar] [CrossRef]
  29. Lunsford, J.H. Characterization of acidity in zeolites and related oxides using trimethylphosphine as a probe. Top. Catal. 1997, 4, 91–98. [Google Scholar] [CrossRef]
  30. Wang, S.; Fang, Y.; Huang, Z.; Xu, H.L.; Shen, W. The effects of the crystalline phase of zirconia on C-O activation and C-C coupling in converting syngas into aromatics. Catalysts 2020, 10, 262. [Google Scholar] [CrossRef] [Green Version]
  31. Zhang, H.L.; Yu, H.G.; Zheng, A.M.; Li, S.H.; Shen, W.L.; Deng, F. Reactivity enhancement of 2-propanol photocatalysis on SO42-/TiO2: Insights from solid-state NMR spectroscopy. Environ. Sci. Technol. 2008, 42, 5316–5321. [Google Scholar] [CrossRef]
  32. Chen, B.; Munson, E.J. Investigation of the mechanism of n-butane oxidation on vanadium phosphorus oxide catalysts:  Evidence from isotopic labeling studies. J. Am. Chem. Soc. 2002, 124, 1638–1652. [Google Scholar] [CrossRef]
  33. Zarkalis, A.S.; Hsu, C.Y.; Gates, B.C. Solid superacid catalysis: Kinetics of butane isomerization catalyzed by a sulfated oxide containing iron, manganese, and zirconium. Catal. Lett. 1994, 29, 235–239. [Google Scholar] [CrossRef]
  34. Ma, Z.N.; Zou, Y.; Hua, W.M.; He, H.Y.; Gao, Z. In situ 13C MAS NMR study on the mechanism of butane isomerization over catalysts with different acid strength. Top. Catal. 2005, 35, 141–153. [Google Scholar] [CrossRef]
  35. Ma, Z.N.; Hua, W.M.; Ren, Y.; He, H.Y.; Gao, Z. n-Butane isomerization over Cs-salts of H3PW12O40: A mechanistic study by 13C MAS NMR. Appl. Catal. A Gen. 2003, 256, 243–250. [Google Scholar] [CrossRef]
  36. Hsu, C.Y.; Heimbuch, C.R.; Armes, C.T.; Gates, B.C. A highly active solid superacid catalyst for n-butane isomerization: A sulfated oxide containing iron manganese and zirconium. J. Chem. Soc. Chem. Commun. 1992, 22, 1645–1646. [Google Scholar] [CrossRef]
  37. Na, K.; Okuhara, T.; Misono, M. Skeletal isomerization of n-butane over caesium hydrogen salts of 12-tungstophosphoric acid. J. Chem. Soc. Faraday Trans. 1995, 91, 367–373. [Google Scholar] [CrossRef]
  38. Sun, Y.Y.; Yuan, L.N.; Ma, S.Q.; Han, Y.; Zhao, L.; Wang, W.; Chen, C.L.; Xiao, F.S. Improved catalytic activity and stability of mesostructured sulfated zirconia by Al promoter. Appl. Catal. A Gen. 2004, 268, 17–24. [Google Scholar] [CrossRef]
  39. Wang, J.H.; Mou, C.Y. Alumina-promoted mesoporous sulfated zirconia: A catalyst for n-butane isomerization. Appl. Catal. A Gen. 2005, 286, 128–136. [Google Scholar] [CrossRef]
  40. Gao, Z.; Xia, Y.D.; Hua, W.M.; Miao, C.X. New catalyst of SO42-/Al2O3-ZrO2 for n-butane isomerization. Top. Catal. 1998, 6, 101–106. [Google Scholar] [CrossRef]
Figure 1. (a) The XRD patterns and (b) FT-IR spectra of M-ZrO2-1, M-SZ-1, M-ZrO2-2, M-SZ-2, M-ZrO2-3, and M-SZ-3.
Figure 1. (a) The XRD patterns and (b) FT-IR spectra of M-ZrO2-1, M-SZ-1, M-ZrO2-2, M-SZ-2, M-ZrO2-3, and M-SZ-3.
Processes 10 02693 g001
Figure 2. (a) Zr 3d and (b) O 1s XPS spectra of M-ZrO2-1, M-ZrO2-2, M-ZrO2-3, M-SZ-1, M-SZ-2, and M-SZ-3.
Figure 2. (a) Zr 3d and (b) O 1s XPS spectra of M-ZrO2-1, M-ZrO2-2, M-ZrO2-3, M-SZ-1, M-SZ-2, and M-SZ-3.
Processes 10 02693 g002
Figure 3. The TEM image of (a) M-ZrO2-1, (b) M-ZrO2-2, (c) M-ZrO2-3, (d) M-SZ-1, (e) M-SZ-2, and (f) M-SZ-3.
Figure 3. The TEM image of (a) M-ZrO2-1, (b) M-ZrO2-2, (c) M-ZrO2-3, (d) M-SZ-1, (e) M-SZ-2, and (f) M-SZ-3.
Processes 10 02693 g003
Figure 4. (a) N2 adsorption–desorption isotherms and (b) pore distribution of M-ZrO2-1, M-ZrO2-2, M-ZrO2-3, M-SZ-1, M-SZ-2, and M-SZ-3.
Figure 4. (a) N2 adsorption–desorption isotherms and (b) pore distribution of M-ZrO2-1, M-ZrO2-2, M-ZrO2-3, M-SZ-1, M-SZ-2, and M-SZ-3.
Processes 10 02693 g004
Figure 5. (a) 1H-decoupled 31P MAS NMR spectra and (b) single-pulse 31P MAS NMR spectra of M-SZ-1, M-SZ-2, and M-SZ-3.
Figure 5. (a) 1H-decoupled 31P MAS NMR spectra and (b) single-pulse 31P MAS NMR spectra of M-SZ-1, M-SZ-2, and M-SZ-3.
Processes 10 02693 g005
Figure 6. 13C MAS NMR spectra of 1-13C-n-butane reacted at 433 K with M-SZ-2 as the catalyst at different times.
Figure 6. 13C MAS NMR spectra of 1-13C-n-butane reacted at 433 K with M-SZ-2 as the catalyst at different times.
Processes 10 02693 g006
Figure 7. Butane isomerization catalyzed by M-SZ-2 at (a) 393 K, (b) 413 K, and (c) 433 K. (■) 1-13C-n-butane, (◆) 2-13C-n-butane, (●) 1-13C-isobutane and 2-13C-isobutane, (▲) 13C-isopentane, (▼) 13C-propane, and (★) the total amount of 13C-labeled species.
Figure 7. Butane isomerization catalyzed by M-SZ-2 at (a) 393 K, (b) 413 K, and (c) 433 K. (■) 1-13C-n-butane, (◆) 2-13C-n-butane, (●) 1-13C-isobutane and 2-13C-isobutane, (▲) 13C-isopentane, (▼) 13C-propane, and (★) the total amount of 13C-labeled species.
Processes 10 02693 g007
Figure 8. The kinetic study of butane isomerization at different temperatures catalyzed by M-SZ-2. (a) − ln[1 − (1 + 1/Keq)PiC4/Ptotal]/(1 + 1/Keq) vs. reaction time at (■) 393 K, (●) 413 K, and (▲) 433 K, and (b) ln(kKnC4) vs. 1000/T.
Figure 8. The kinetic study of butane isomerization at different temperatures catalyzed by M-SZ-2. (a) − ln[1 − (1 + 1/Keq)PiC4/Ptotal]/(1 + 1/Keq) vs. reaction time at (■) 393 K, (●) 413 K, and (▲) 433 K, and (b) ln(kKnC4) vs. 1000/T.
Processes 10 02693 g008
Table 1. The specific surface area, pore volume, and pore diameter of the samples.
Table 1. The specific surface area, pore volume, and pore diameter of the samples.
SamplesBET Surface Area
(m2·g−1)
Pore Volume
(cm3·g−1)
Average Pore
Diameter (nm)
M-ZrO2-11190.1295.4
M-SZ-1760.1957.4
M-ZrO2-21690.21311.4
M-SZ-2670.26515.5
M-ZrO2-31360.24613.1
M-SZ-3570.26318.8
Table 2. The acidity of M-SZ-1, M-SZ-2, and M-SZ-3 acquired from 31P MAS NMR spectra.
Table 2. The acidity of M-SZ-1, M-SZ-2, and M-SZ-3 acquired from 31P MAS NMR spectra.
SampleAcid TypeChemical Shift
(ppm)
Acid Amount
(μmol·g−1)
JP-H
(Hz)
M-SZ-1BA−3.468437
LA1, LA2, LA3, LA4, LA5−32.9, −36.7, −40.8, −45.2, −49.74, 27, 26, 12, 89-
M-SZ-2BA−3.2112494
LA1, LA2, LA3, LA4, LA5−31.8, −35.6, −40.5, -, −54.29, 19, 12, -, 187-
M-SZ-3BA−3.369474
LA1, LA2, LA3, LA4, LA5−32.7, −36.2, −40.6, −46.2, −53.59, 22, 22, 4, 157-
Table 3. The conversion and product distribution of the butane isomerization at 433 K for 60 min.
Table 3. The conversion and product distribution of the butane isomerization at 433 K for 60 min.
CatalystsX60 (%)2-13C-n-Butane (%)13C-Isobutane (%)
M-SZ-1000
M-SZ-269.228.729.5
M-SZ-335.021.410.1
Table 4. The kinetic constants of butane isomerization catalyzed by M-SZ-2 at 393 ~ 433 K.
Table 4. The kinetic constants of butane isomerization catalyzed by M-SZ-2 at 393 ~ 433 K.
Reaction Temperature (K)kKnC4 × 105 (mol·g−1·s−1)R
3934.410.990
4138.980.986
43319.660.998
Table 5. Comparison of the butane isomerization catalyzed by M-SZ-2 and T-SZ-general before and after calcination at 673 K for 240 h at different reaction temperatures for 60 min.
Table 5. Comparison of the butane isomerization catalyzed by M-SZ-2 and T-SZ-general before and after calcination at 673 K for 240 h at different reaction temperatures for 60 min.
CatalystsReaction
Temperature (K)
UncalcinedCalcined at 673 K for 240 h
X60 (%)Isobutane (%)X60 (%)Isobutane (%)
M-SZ-243369.229.563.825.7
T-SZ-general33861.132.750.222.6
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Huang, D.; Feng, W.; Zhang, L.; Yue, B.; He, H. Insight into the Acidity and Catalytic Performance on Butane Isomerization of Thermal Stable Sulfated Monoclinic Zirconia. Processes 2022, 10, 2693. https://doi.org/10.3390/pr10122693

AMA Style

Huang D, Feng W, Zhang L, Yue B, He H. Insight into the Acidity and Catalytic Performance on Butane Isomerization of Thermal Stable Sulfated Monoclinic Zirconia. Processes. 2022; 10(12):2693. https://doi.org/10.3390/pr10122693

Chicago/Turabian Style

Huang, Daofeng, Wenhua Feng, Li Zhang, Bin Yue, and Heyong He. 2022. "Insight into the Acidity and Catalytic Performance on Butane Isomerization of Thermal Stable Sulfated Monoclinic Zirconia" Processes 10, no. 12: 2693. https://doi.org/10.3390/pr10122693

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