Next Article in Journal
Setting up In2O3-ZrO2/SAPO-34 Catalyst for Improving Olefin Production via Hydrogenation of CO2/CO Mixtures
Next Article in Special Issue
Isomerization of Hemicellulose Aldoses to Ketoses Catalyzed by Basic Anion Resins: Catalyst Screening and Stability Studies
Previous Article in Journal
Nickel Foam-Supported Hierarchical NiCo2S4 Nanostructures as Efficient Electrocatalysts for the Methanol Oxidation Reaction
Previous Article in Special Issue
Gas-Phase Deoxygenation of Biomass Pyrolysis Tar Catalyzed by Rare Earth Metal Loaded Hβ Zeolite
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Isomerization Properties of Pt/SAPO-11 Catalysts for the Production of Bio-Aviation Kerosene

1
Zhejiang Province Key Laboratory of Biofuel, School of Chemical Engineer, Zhejiang University of Technology, Hangzhou 310014, China
2
Zhejiang Jieda Technology Co., Ltd., Huzhou 313300, China
*
Author to whom correspondence should be addressed.
Catalysts 2023, 13(7), 1100; https://doi.org/10.3390/catal13071100
Submission received: 13 June 2023 / Revised: 11 July 2023 / Accepted: 11 July 2023 / Published: 14 July 2023

Abstract

:
The hydroisomerization of n-alkanes is an important step in the production of high-quality bio-aviation kerosene. A SAPO-11 molecular sieve was synthesized using the hydrothermal synthesis method, and a 0.5 wt% Pt/SAPO-11 catalyst was prepared using the impregnation method. The crystal phase, pore structure, acidity, and morphology of Pt/SAPO-11 were characterized via X-ray diffraction, N2 adsorption-desorption, NH3 temperature-programmed desorption, scanning electron microscopy, and transmission electron microscopy, respectively. The hydroisomerization performance of the catalyst was evaluated with bio-n-hexadecane as the model compound. The results showed that temperature and space velocity had significant effects. Under the conditions of 340 °C, 1.5 MPa, WHSV = 1.0 h−1, V(H2): V(n-hexadecane) = 1000:1, the conversion of n-hexadecane and the selectivity of i-hexadecane were 81.8% and 86.5%, respectively.

1. Introduction

Sustainable aviation fuel (SAF) is recognized as a short-to medium term solution toward the overall reduction of greenhouse gas emissions. It can achieve a 50% reduction in carbon emissions by 2050 compared to 2005. At present, the most important SAF is bio-aviation kerosene (BAK), which is a synthesized paraffinic kerosine produced from hydroprocessed esters and fatty acids. BAK is a synthetic blending component for aviation turbine fuels in civil aircraft and engines. BAK is mainly obtained through the hydrodeoxygenation of esters and fatty acids. The ester and fatty acid hydrogenation route has some advantages, such as being abundant in raw materials, relatively mature in technology, and having a high yield. Therefore, the hydrogenation route is an effective method to prepare bio-aviation fuel [1]. The products obtained by hydrodeoxygenation from ester and fatty acid hydrogenation have a high content of n-alkanes, resulting in poor low-temperature fluidity in bio-aviation kerosene. The hydroisomerization of n-alkanes is one of the most effective methods to improve the low-temperature fluidity of biofuels.
The hydroisomerization reaction of n-alkanes is usually carried out on bifunctional catalysts. Therefore, developing catalysts with high i-alkane yield and reaction stability is crucial for producing high-quality bio-aviation kerosene [2]. Among common catalysts, for the hydroisomerization of strong Bronsted [3], Y zeolite, and MCM-41 catalysts, the conversion of n-decane is about 78% and 30%, respectively [4,5]. For the hydroisomerization performances of ZSM-22 and ZSM-48 catalysts, the conversion of dodecane is about 40% and 52%, respectively [6,7]. Due to the shape-selective isomerization effect, monobranched isomeric alkanes are obtained in higher yields on 10-membered ring zeolite [8]. The typical catalyst is a bifunctional catalyst for alkane hydroisomerization and has been widely researched. The n-alkane first adsorbs to the metal active site and then undergoes a dehydrogenation reaction to form an olefin. The resulting olefin diffuses from the metal active site to the acidic site, where it is protonated to produce an alkyl carbon carbocation. At the acidic site, olefins and alkyl carbocation undergo competitive adsorption-desorption. Subsequently, the alkyl carbocation starts skeletal isomerization or cracks into a new olefin and a new carbocation. Finally, the carbocation deprotonates to an olefin, which diffuses from the acidic site to the metal active site to prepare hydrogenation to an alkane and then desorbs.
In terms of the influence of the carrier, it is mainly the pore distribution of the carrier and the acidity strength that affect the catalyst’s activities. Protonation, isomerization, and cracking reactions take place on the acid carrier. Different molecular sieves have different pore distributions, amounts of acid, and acidity distributions, resulting in different isomeric performances of the catalysts. The carbocation reaction generated from isomerization or cracking is an acid catalytic reaction that is mainly influenced by the acidity of the molecular sieve, the distribution of acid sites, and the structural characteristics. The stronger the acid, the more intense the carbocation reaction is, and the resulting conversion is higher. With the increase in acidity, the isomerization reaction intensifies to a certain extent, but not as much as the cracking reaction. Therefore, to obtain more isomerized products and reduce excessive cracking, molecular sieves with weak and moderately strong acids should be selected. Metals are loaded on molecular sieves to provide mainly hydrogenation or dehydrogenation functions. A stronger hydrogenation function of a metal results in a higher conversion rate. In addition, the strong hydrogenation ability of metals can suppress carbon deposition during the reaction. Thereby, the lifetime of the catalyst can be improved.
The SAPO-11 molecular sieve is one of the most promising acidic carriers due to its one-dimensional 10-membered ring pore structure and suitable acidity [9,10]. Pt metal is often used as a metal site for bifunctional catalysts due to its stronger hydrogenation activity, while SAPO-11 molecular sieves provide acidic centers [11,12]. Thus, the selectivity of isomeric alkanes mainly depends on the metal and acidic sites [13], and the appropriate balance between these variables is closely related to determining the reactivity and selectivity of bifunctional catalysts. Xiaojun Dai et al. synthesized SAPO-11 molecular sieves through a two-step crystallization method. Under the reaction conditions of 360 °C, 2 MPa, 1.5 h−1, and a 600 volume ratio of H2/n-C16, the conversion of n-hexadecane was 90% and the selectivity of i-hexadecane was 80% [14,15]. Yuchao Lyu et al. prepared a Ni/SAPO-11 bifunctional catalyst for the hydroisomerization of n-hexane. Under the reaction conditions of 340 °C, 2 MPa, 1.0 h−1, and a 4.0 volume of H2/n-C6, the conversion of n-hexane was 75% and the selectivity of i-hexane was 90% [16]. Lili Geng et al. studied Pt/SAPO-11 molecular sieves under the reaction conditions of 320 °C, 0.1 MPa, 4.0 h−1 and a 15.0 mole ratio of H2/n-C12; the resulting conversion of n-dodecane was 76% and the selectivity of i-dodecane was 80% [12].
In this study, in order to prepare bio-aviation kerosene, the SAPO-11 molecular sieve was synthesized by the one-step hydrothermal synthesis method. Additionally, 0.5 wt% Pt/SAPO-11 was prepared by loading Pt. The physicochemical properties were characterized using instruments. The catalytic performance was investigated in the n-hexadecane hydroisomerization.

2. Results and Discussion

2.1. Physicochemical Properties

Figure 1 shows the XRD patterns of the samples SAPO-11 and Pt/SAPO-11. It can be seen from the figure that the characteristic diffraction peaks of the SAPO-11 molecular sieve were observed at 2θ = 8.1°, 9.4°, 13.1°, 15.6°, 20.3°, and 21.2° for the crystalline sample, indicating that the impregnation process of Pt has no significant effect on the crystallinity of SAPO-11, and the loaded Pt/SAPO-11 had good crystallinity and no heterocrystals and still maintained a good AEL structure [17]. Besides the typical characteristic peaks of SAPO-11, Pt/SAPO-11 has a weak peak at 2θ = 59° (marked by ↓) due to loaded Pt.
Figure 2 shows the N2 adsorption-desorption isotherms and pore diameter distribution of the SAPO-11 molecular sieve and Pt/SAPO-11. Table 1 shows the corresponding pore structure parameters. The SAPO-11 molecular sieve is a typical type-I isotherm, indicating that the SAPO-11 sample is a microporous molecular sieve. Pt/SAPO-11 is a type-IV adsorption-desorption isotherm and a type-H4 hysteresis loop. The pore diameter distribution of the SAPO-11 molecular sieve is obvious from 1 nm to 12 nm, and the Pt/SAPO-11 molecular sieve mainly has a pore size distribution in the range of 2–20 nm. As can be seen from Table 1, the specific surface area of the catalyst decreased significantly compared with the SAPO-11 molecular sieve when loaded with 0.5 wt% Pt. The specific surface area decreased from 145.4 m2·g−1 to 103.0 m2·g−1. This is due to the interaction between the active component Pt and the molecular sieve after loading, which occupies a portion of the surface area and results in a decrease in surface area. In addition, agglomeration of grains in the pore channel occurred, which blocked the molecular sieve pore channel, causing a decrease in surface area and microporous pore capacity.
Figure 3 shows the SEM image of SAPO-11 and Pt/SAPO-11. 1–2 shows the SEM image of SAPO-11, and 3–4 shows the SEM image of Pt/SAPO-11. 1 and 3 show the morphology at the scale of 50 μm, and 2 and 4 show the morphology at the scale of 20 μm. It can be seen from the figure that the SAPO-11 molecular sieve has spheroidal particles with relatively rough surfaces. The particle size is about 20–35 μm. The SAPO-11 molecular sieve synthesized by the hydrothermal method has high crystallinity and a relatively uniform particle diameter. The microscale structure of the SAPO-11 is influenced little by Pt loading. Additionally, uniform particle size distribution, high crystallinity, and smooth pore channels facilitate the diffusion of reactants and products in the isomerization reaction.
Figure 4 shows the TEM image of SAPO-11 and Pt/SAPO-11. 1–2 shows the TEM image of SAPO-11, and 3–4 shows the TEM image of Pt/SAPO-11. 1 and 3 show the morphology at the scale of 50 μm, 2 and 4 show the morphology at the scale of 100 μm. It can be seen from the figure that the black spots on the TEM photograph of the catalyst are loaded platinum particles. The particle size is about 10–25 nm. The platinum particles were successfully loaded on the molecular sieve, both dispersed on the outer surface of SAPO-11 and partially dispersed inside the pore channel, and the loaded particles were evenly distributed with good dispersion and no obvious agglomeration of large particles. Similarly, the microscale structure of the SAPO-11 is influenced little by Pt loading. This is because the hydroisomerization performance of the catalyst is affected by the synergistic effect of both the pore structure and the acidic and metal active sites [18]. Platinum is loaded using an equal-volume impregnation method supplemented by ultrasound to evenly distribute platinum particles on the SAPO-11 molecular sieve. Therefore, with the better dispersion of Pt, the catalyst is not easily blocked in the pore.
NH3-TPD is obtained as a TPD detection signal through NH3 adsorption and temperature-programmed desorption. Figure 5 shows the NH3-TPD spectra of the SAPO-11 and Pt/SAPO-11 samples. It can be seen from the figure that the molecular sieve had one desorption peak concentrated at 50–200 °C, which corresponds to the weak acidic site, and another peak concentrated at 200–400 °C, which corresponds to the medium to strong acidic site. Additionally, the area of the low-temperature desorption peak was larger than that of the high-temperature desorption peak. This indicates that the catalyst is dominated by weak acid centers. The loading of platinum had little effect on the acid intensity distribution and the weak and medium-strong acid centers [19]. During the hydrogenation of n-hexadecane, the acidic strength had a significant influence on the performance of the catalyst [20], with medium-to-strong acid centers inducing dimerization and cracking reactions and weak acid centers favoring the isomerization of alkanes [21,22,23]. Table 2 shows the amount of acid in SAPO-11 and Pt/SAPO-11 through integration calculations. The Pt/SAPO-11 molecular sieve’s amount of weak and medium acid decreased slightly compared to the SAPO-11 molecular sieve, possibly due to the metal occupying the acidic site.

2.2. Hydroisomerization of n-Hexadecane

In order to prepare bio-aviation kerosene, the effects of reaction temperature, weight hourly space velocity, the ratio of H2/n-C16, reaction pressure, and metal loadings were considered. Meanwhile, the influence of reaction conditions on conversion and selectivity was studied.

2.2.1. Effect of Reaction Temperature on the Isomerization Performance

The influence of reaction temperature on the hydroisomerization performance of the Pt/SAPO-11 catalyst is shown in Figure 6. Under the reaction conditions of 1.5 MPa pressure, 1.5 h−1 of weight hourly space velocity, and 1000 of volume ratio of H2/n-C16, with the increase in temperature, the conversion of n-hexadecane increased obviously. With the increase in conversion, the selectivity and yield showed a trend of increasing and then decreasing. When the temperature increased from 340 °C to 380 °C, the conversion increased from 81.8% to 85.12%, and the selectivities were 86.5% and 71.2%, respectively. When the reaction temperature exceeded 340 °C, the isomeric olefin intermediates did not desorb in time, and a portion of the carbocations carried out β fracture, and then carbocations occurred with varying degrees of cracking, resulting in a decrease in the selectivity of i-C16 with the increase in conversion [24].

2.2.2. Effect of Weight Hourly Space Velocity on the Isomerization Performance

Figure 7 shows the influence of weight hourly space velocity on the hydroisomerization performance of the Pt/SAPO-11 catalyst. The conversion of n-hexadecane decreased, and the selectivity of i-hexadecane increased with the increase in weight hourly space velocity. When the space velocity increased from 1 h−1 to 3 h−1, the conversion decreased from 81.7% to 17.5%, and the selectivity increased from 86.8% to 98.1%. With the increase in space velocity, the residence time of n-hexadecane in the active center of the catalyst became shorter, the reaction was insufficient, and the conversion decreased. Although the product selectivity was higher at higher space velocity, the conversion was lower and not conducive to improving the quality of the product.

2.2.3. Effect of the Ratio of H2/n–C16 on the Isomerization Performance

Figure 8 shows the effect of the ratio of H2/n–C16 from 600 to 1400 on the isomerization performance. It can be seen that the conversion of n-hexadecane and the selectivity of i-hexadecane on Pt/SAPO-11 almost did not change when increasing the ratio. When the ratio of H2/n–C16 was 600, the conversion of n-hexadecane was 79.5% and the selectivity of i-hexadecane was 85.5%. When the ratio of H2/n–C16 was 1400, the conversion of n-hexadecane was 80.1% and the selectivity of i-hexadecane was 86.2%. A higher H2/n–C16 ratio can increase the hydrogen partial pressure, which is conducive to improving the hydroisomerization of the catalyst.

2.2.4. Effect of Reaction Pressure on the Isomerization Performance

Figure 9 shows the effect of reaction pressure from 1.0 MPa to 3.0 MPa on the isomerization performance. It can be seen that the conversion of n-hexadecane and the selectivity of i-hexadecane changed only slightly, with insignificant effects. When the reaction pressure was 1.0 MPa, the conversion of n-hexadecane was 80.3% and the selectivity of i-hexadecane was 86.9%. When the reaction pressure was 3.0 MPa, the conversion of n-hexadecane was 83.1% and the selectivity of i-hexadecane was 72.5%. Increasing the partial pressure of hydrogen resulted in a slight increase in the conversion of the hydroisomerization reaction and a decrease in selectivity.

2.2.5. Effect of Metal Loadings on the Isomerization Performance

Figure 10 shows the effect of metal loadings from 0.2 wt% to 1.0 wt% on the isomerization performance. It can be seen that the active components of precious metals have a great influence on the hydroisomerization reaction by the Pt/SAPO-11 catalyst. When the active component loading increased from 0.2 wt% to 0.8 wt%, the conversion of n-hexadecane increased to 82.43% and the selectivity of i-hexadecane increased to 87.3%. However, when the active component loading increased from 0.8 wt% to 1.0 wt%, the conversion of n-hexadecane decreased to 63.56% and the selectivity of i-hexadecane decreased to 81.6%. On the one hand, this may be due to excessive Pt occupying partial acidic sites of the catalyst. On the other hand, some Pt may block the pores of the carrier.

2.3. Isomeric Product Distribution

Table 3 shows the product distribution of the Pt/SAPO-11 catalyst. In order to reflect the data more intuitively, Figure 11 is drawn. The influence of reaction temperature on the distribution of mono/multi-branched isomers is shown in Figure 12. The ratio of mono-branched chain isomeric products to multi-branched chain isomeric products decreased gradually with the increase in conversion. At high conversion, the multi-branched isomeric products were greater than the mono-branched isomeric products.
The conversion sequence of n-alkanes on the catalyst is as follows: n-C16  mono-branched chain; i-C16  multi-branched chain; and i-C16  cracking product [24]. At low conversion, the main reaction products were mono-branched chain i-C16. With the increase in reaction temperature, the mono-branched chain i-C16 was converted to the multi-branched chain i-C16 and further cracked into short-chain alkanes. The reaction mechanism of the mono-branched isomer is shown in Figure 13. First, the alkanes at the Pt center are dehydrogenated to form olefin molecules and diffused to the acidic site. Then olefin molecules converted to alkane carbocation at the B-acid center. Subsequently, the carbocation undergoes skeletal isomerization to form a new carbocation. After that, the carbocation deprotonates to olefin. Finally, the mono-branched isomer is formed by hydrogenation on the Pt center.

2.4. Stability of Catalyst

Figure 14 shows the conversion of the Pt/SAPO-11 catalyst and the selectivity of i-hexadecane over time. From the graph, it can be seen that the conversion and selectivity of isomeric hexadecane changed smoothly with the development of reaction time. The conversion of the Pt/SAPO-11 catalyst and the selectivity of isomeric hexadecane remained above 80%. Research has shown that molecular sieves have large specific surface areas and mesoporous volumes and are conducive to increasing the carbon deposition resistance of SAPO-11 catalyst, thus improving the activity of the catalyst and extending its service life. The Pt/SAPO-11 catalyst can run continuously and stably for more than 60 h under the abovementioned conditions.

3. Materials and Methods

3.1. Synthesis of SAPO-11

The SAPO-11 molecular sieve was synthesized by hydrothermal synthesis. The specific synthesis steps were as follows: Pseudo-boehmite (Al2O3·nH2O) and phosphoric acid (H3PO4) were dissolved in deionized water, stirred for 2 h in a constant temperature water bath at 35 °C, and then di-n-propylamine (DPA) and acid silica sol (mSiO2·nH2O) were added to the abovementioned solution and stirred continuously for 2 h. The molar composition was 1.25 DPA:1.0 Al2O3:0.8 P2O5:0.4 SiO2:50 H2O gel. The synthesized sample was transferred to a stainless-steel autoclave lined with polytetrafluorethylene and crystallized in a drying oven at 180 °C for 24 h. The product was cooled to room temperature, and the required sample was obtained after centrifuging, washing, and drying. The sample was placed in a muffle furnace and calcined at 600 °C for 5 h to obtain SAPO-11 to remove the template.

3.2. Preparation of Catalyst

The forming and metal loading of the catalyst were carried out by pressing and dipping methods, respectively. The abovementioned calcined SAPO-11 powder was pressed and formed by a tablet press, screened to 20–40 mesh, and a 10 g sample was taken and placed in a surface dish with a diameter of 10 cm.
Afterward, 0.1 g of H2PtCl6 solid powder was weighed and dissolved in 10 mL deionized water, adding the H2PtCl6 precursor solution dropwise onto the SAPO-11 molecular sieve, ultrasonicating for 30 min, and then left to stand at room temperature for 24 h. Afterwards, it was dried at 105 °C for 12 h and finally calcininated in a Muffle furnace at 550 °C for 3 h.

3.3. Characterization

The determination of the crystal phase structure was performed as follows: we used the X’Pert Pro automatic powder diffractometer from Panaco, Netherlands, with a working voltage of 30 kV and a working current of 30 mA, using a Cu target, a Kα radiation source, a wide angle scanning range of 2θ = 5°–40°, and a scanning rate of 2°·min−1. For N2 adsorption–desorption, we used a 3H-2000PS1 specific surface area and pore diameter analyzer from Beijing Besser Instrument Technology Co., Ltd., Beijing, China. The samples were degassed for 3 h under vacuum conditions at 300 °C, and data were collected under a high-purity N2 atmosphere. The obtained data were analyzed by the BET method, the t-plot method, and the BJH method.
Morphology determination was performed as follows: a SIGMA-type field emission scanning electron microscope from Zeiss (Jane, Germany) was used. The working voltage of the instrument was 2.0 kV, and it was necessary to spray gold before testing to increase the conductivity of the molecular sieve. A transmission electron microscope of the Tecnai G2 F30 S-Twin type from Philips-FEI, The Netherlands, was used for observation. The determination of NH3-TPD was performed with an Auto Chem II 2920 program temperature chemisorption instrument. First, the sample was pretreated with helium at 550 °C for 0.5 h, then adsorbed at 70 °C for 0.5 h. After that, the physically absorbed NH3 was removed in helium for 0.5 h, then the sample temperature was increased to 700 °C at a rate of 10 °C·min−1 to obtain a TPD detection signal.

3.4. Catalytic Performance Tests

Bio-n-hexadecane (n-C16) was obtained by hydrodeoxygenation of fatty acid methyl esters (biodiesel) and distillation, with a purity > 99%. We used bio-n-hexadecane as a model compound to evaluate the isomerization property of the Pt/SAPO-11 catalyst. The reaction was carried out continuously in an XL-RL-type tube fixed-bed reactor from Haian Petroleum Scientific Research Instrument Co., Ltd., Nantong, China. The inner diameter of the reaction tube was 1 cm, and the length was 45 cm. Next, 6 g of catalyst was loaded into the reaction tube, and both ends were filled with quartz sand. The ends of the reaction tubes were blocked with asbestos mesh to prevent the loss of quartz sand. A temperature-measuring tube was inserted in the middle of the bed. After filling, we tightened both ends of the reaction tube and mounted it on the experimental setup. Hydrogen and n-hexadecane were fed through a mass flow controller and a liquid chromatography metering pump, respectively. Hydrogen pressure was controlled through a cylindrical reducing valve and a backpressure valve.
Before running the device, the gas tightness of the device was checked with soapy water to ensure that the experimental device was gas-tight. Prior to the reaction, the loaded Pt/SAPO-11 catalyst was reduced at 380 °C for 4 h in a high-purity H2 atmosphere. Then, the temperature was decreased to the reaction temperature, and the reaction parameters were controlled to carry out the hydroisomerization of n-hexadecane. The hydroisomerization properties of n-hexadecane were evaluated at a reaction temperature of 300 °C to 380 °C, a reaction pressure of 1 MPa to 3 MPa, a weight hourly space velocity of 1 h−1 to 3 h−1, and the ratio of H2/n-C16 from 600 to 1400. After leaving the reactor, the product entered the gas-liquid separator, and the liquid product was separated and collected by two continuously arranged gas-liquid separators in order to minimize the interference caused by collecting and to increase the liquid yield. In order to ensure the comparability of the data, the samples were taken at a stationary time interval. Additionally, the liquid in the gas-liquid separator was emptied after the collection was completed to avoid interference with the next group of data. The experiments were conducted by changing different reaction conditions in order to obtain product data over a large conversion range. After each change in conditions, we waited for the system to stabilize before starting product collection and sampling. However, the minimum stabilization time of the system varied according to the changes in gas and liquid velocities, and in actual operation, a large time interval elapsed before emptying and sampling.
The products were qualitatively and quantitatively analyzed by Agilent GC-MS and Shimadzu GC-2014 gas chromatographs, respectively. The GC-MS ramp-up program was held at 50 °C for 3 min and subsequently increased to a maximum temperature of 300 °C at a rate of 20 °C·min−1. It was held for 2 min, with a total program time of 17.5 min. The GC-2014 was performed on a Rtx-5 capillary column (30 m × 0.32 mm × 0.25 μm) and a FID detector. First, 50 °C was held for 3 min, then increased from 50 °C to 300 °C at a rate of 20 °C·min−1 and held for 2 min.
Formulas (1) and (2) were used to calculate the conversion of n-hexadecane (x) and the selectivity (s) of i-hexadecane, respectively.
  x = m 1     m 2 m 1   ×   100 %
s = m 3 m 1   -   m 2   ×   100 %
In the formulas, m1 and m2 represent the mass of n-hexadecane in the raw materials and products, respectively, g; m3 represents the mass of the isomerized product in the products, g.

4. Conclusions

The catalytic performance of Pt/SAPO-11 was evaluated with bio-aviation kerosene (n-hexadecane) by a hydroisomerization reaction. The SAPO-11 molecular sieve synthesized by the hydrothermal synthesis method exhibited regular columnar crystal aggregation, forming a spherical shape. The load of Pt did not affect the AEL structure of SAPO-11, and the Pt/SAPO-11 molecular sieve exhibited an obvious mesoporous structure and acid content. The reaction temperature and weight hourly space velocity had significant effects on the conversion of n-hexadecane and the selectivity of i-hexadecane, whereas the reaction pressure and the ratio of H2/n-C16 had no obvious influence. Therefore, the Pt/SAPO-11 molecular sieve had a high hydroisomerization rate for n-hexadecane, which is beneficial for improving the low-temperature fluidity of bio-aviation kerosene.

Author Contributions

Conceptualization, X.L.; methodology, X.L.; software, S.Y.; validation, S.Y., Q.M. and X.L.; formal analysis, S.Y.; investigation, X.Z. and N.S.; resources, J.H.; data curation, S.Y.; writing—original draft preparation, S.Y.; writing—review and editing, X.L.; visualization, S.Y.; supervision, W.S.; project administration, M.L.; funding acquisition, X.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Science and Technology Program of Zhejiang Province-Key Research and Development Program: 2021C01063.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sousa-Aguiar, E.F.; Noronha, F.B.; Faro, A., Jr. The Main Catalytic Challenges in GTL (Gas-to-Liquids) Processes. Catal. Sci. Technol. 2011, 1, 698–713. [Google Scholar] [CrossRef]
  2. Dai, X.; Cheng, Y.; Si, M.; Wei, Q.; Chen, D.; Huang, W.; Zhou, Y. SAPO-11 Molecular Sieves Synthesized in Alcohol-Water Concentrated Gel System with Improved Acidity, Mesoporous Volume and Hydroisomerization Performance. Fuel 2022, 314, 123131. [Google Scholar] [CrossRef]
  3. Pagis, C.; Bouchy, C.; Dodin, M.; Franco, R.M.; Farrusseng, D.; Tuel, A. Hollow Y Zeolite Single Crystals: Synthesis, Characterization and Activity in the Hydroisomerization of n-Hexadecane. Oil Gas Sci. Technol. Rev. IFP Energ. Nouv. 2019, 74, 38. [Google Scholar] [CrossRef] [Green Version]
  4. Zhang, Y.; Liu, D.; Lou, B.; Yu, R.; Men, Z.; Li, M.; Li, Z. Hydroisomerization of n-Decane over Micro/Mesoporous Pt-Containing Bifunctional Catalysts: Effects of the MCM-41 Incorporation with Y Zeolite. Fuel 2018, 226, 204–212. [Google Scholar] [CrossRef]
  5. Qin, B.; Zhang, X.; Zhang, Z.; Ling, F.; Sun, W. Synthesis, Characterization and Catalytic Properties of Y-β Zeolite Composites. Pet. Sci. 2011, 8, 224–228. [Google Scholar] [CrossRef] [Green Version]
  6. Park, K.-C.; Ihm, S.-K. Comparison of Pt/Zeolite Catalysts for n-Hexadecane Hydroisomerization. Appl. Catal. A Gen. 2000, 203, 201–209. [Google Scholar] [CrossRef]
  7. Wang, Y.; Liu, W.; Zhang, W.; Sun, J.; Li, S.; Zheng, J.; Fan, B.; Li, R. Comparison of n-Dodecane Hydroisomerization Performance over Pt Supported ZSM-48 and ZSM-22. Catal. Lett. 2021, 151, 3492–3500. [Google Scholar] [CrossRef]
  8. Zschiesche, C.; Himsl, D.; Rakoczy, R.; Reitzmann, A.; Freiding, J.; Wilde, N.; Gläser, R. Hydroisomerization of Long-Chain n-Alkanes over Bifunctional Zeolites with 10-Membered- and 12-Membered-Ring Pores. Chem. Eng. Technol. 2018, 41, 199–204. [Google Scholar] [CrossRef]
  9. Wei, X.; Kikhtyanin, O.V.; Parmon, V.N.; Wu, W.; Bai, X.; Zhang, J.; Xiao, L.; Su, X.; Zhang, Y. Synergetic Effect between the Metal and Acid Sites of Pd/SAPO-41 Bifunctional Catalysts in n-Hexadecane Hydroisomerization. J. Porous Mater. 2018, 25, 235–247. [Google Scholar] [CrossRef]
  10. Lv, G.; Wang, C.; Chi, K.; Liu, H.; Wang, P.; Ma, H.; Qu, W.; Tian, Z. Effects of Pt Site Distributions on the Catalytic Performance of Pt/SAPO-11 for n-Dodecane Hydroisomerization. Catal. Today 2018, 316, 43–50. [Google Scholar] [CrossRef]
  11. Kouzu, M.; Kuwako, T.; Ohto, Y.; Suzuki, K.; Kojima, M. Single Stage Upgrading with the Help of Bifunctional Catalysis of Pt Supported on Solid Acid for Converting Product Oil of Triglyceride Thermal Cracking into Drop-in Fuel. Fuel Process. Technol. 2020, 202, 106364. [Google Scholar] [CrossRef]
  12. Geng, L.; Gong, J.; Qiao, G.; Ye, S.; Zheng, J.; Zhang, N.; Chen, B. Effect of Metal Precursors on the Performance of Pt/SAPO-11 Catalysts for n-Dodecane Hydroisomerization. ACS Omega 2019, 4, 12598–12605. [Google Scholar] [CrossRef] [Green Version]
  13. Said, S.; Zaky, M.T. Pt/SAPO-11 Catalysts: Effect of Platinum Loading Method on the Hydroisomerization of n-Hexadecane. Catal. Lett. 2019, 149, 2119–2131. [Google Scholar] [CrossRef]
  14. Dai, X.; Cheng, Y.; Si, M.; Wei, Q.; Zhao, L.; Wang, X.; Huang, W.; Liu, H.; Zhou, Y. Synthesis of Nickel In Situ Modified SAPO-11 Molecular Sieves and Hydroisomerization Performance of Their NiWS Supported Catalysts. Front. Chem. 2021, 9, 765573. [Google Scholar] [CrossRef]
  15. Dai, X.; Cheng, Y.; Si, M.; Wei, Q.; Zhou, Y. Hydroisomerization of n-Hexadecane Over Nickel-Modified SAPO-11 Molecular Sieve-Supported NiWS Catalysts: Effects of Modification Methods. Front. Chem. 2022, 10, 857473. [Google Scholar] [CrossRef]
  16. Lyu, Y.; Yu, Z.; Yang, Y.; Liu, Y.; Zhao, X.; Liu, X.; Mintova, S.; Yan, Z.; Zhao, G. Metal and Acid Sites Instantaneously Prepared over Ni/SAPO-11 Bifunctional Catalyst. J. Catal. 2019, 374, 208–216. [Google Scholar] [CrossRef]
  17. Bértolo, R.; Silva, J.; Ribeiro, M.; Martins, A.; Fernandes, A. Fernandes Microwave Synthesis of SAPO-11 Materials for Long Chain n-Alkanes Hydroisomerization: Effect of Physical Parameters and Chemical Gel Composition. Appl. Catal. A Gen. 2017, 542, 28–37. [Google Scholar] [CrossRef]
  18. Verma, D.; Kumar, R.; Rana, B.S.; Sinha, A.K. Sinha Aviation Fuel Production from Lipids by a Single-Step Route Using Hierarchical Mesoporous Zeolites. Energy Environ. Sci. 2011, 4, 1667–1671. [Google Scholar] [CrossRef]
  19. Lee, E.; Yun, S.; Park, Y.-K.; Jeong, S.-Y.; Han, J.; Jeon, J.-K. Selective Hydroisomerization of n-Dodecane over Platinum Supported on SAPO-11. J. Ind. Eng. Chem. 2014, 20, 775–780. [Google Scholar] [CrossRef]
  20. Yang, Z.; Li, J.; Liu, Y.; Liu, C. Effect of Silicon Precursor on Silicon Incorporation in SAPO-11 and Their Catalytic Performance for Hydroisomerization of n-Octane on Pt-Based Catalysts. J. Energy Chem. 2017, 26, 688–694. [Google Scholar] [CrossRef] [Green Version]
  21. Tao, S.; Li, X.; Lv, G.; Wang, C.; Xu, R.; Ma, H.; Tian, Z. Highly Mesoporous SAPO-11 Molecular Sieves with Tunable Acidity: Facile Synthesis, Formation Mechanism and Catalytic Performance in Hydroisomerization of n-Dodecane. Catal. Sci. Technol. 2017, 7, 5775–5784. [Google Scholar] [CrossRef]
  22. Yang, L.; Wang, W.; Song, X.; Bai, X.; Feng, Z.; Liu, T.; Wu, W. The Hydroisomerization of n-Decane over Pd/SAPO-11 Bifunctional Catalysts: The Effects of Templates on Characteristics and Catalytic Performances. Fuel Process. Technol. 2019, 190, 13–20. [Google Scholar] [CrossRef]
  23. Wang, W.; Liu, C.-J.; Wu, W. Bifunctional Catalysts for the Hydroisomerization of n-Alkanes: The Effects of Metal–Acid Balance and Textural Structure. Catal. Sci. Technol. 2019, 9, 4162–4187. [Google Scholar] [CrossRef]
  24. Jaroszewska, K.; Fedyna, M.; Trawczyński, J. Hydroisomerization of Long-Chain n-Alkanes over Pt/AlSBA-15+ Zeolite Bimodal Catalysts. Appl. Catal. B Environ. 2019, 255, 117756. [Google Scholar] [CrossRef]
  25. Fan, Y.; Xiao, H.; Shi, G.; Liu, H.; Bao, X. Alkylphosphonic Acid- and Small Amine-Templated Synthesis of Hierarchical Silicoaluminophosphate Molecular Sieves with High Isomerization Selectivity to Di-Branched Paraffins. J. Catal. 2012, 285, 251–259. [Google Scholar] [CrossRef]
  26. Bouchy, C.; Hastoy, G.; Guillon, E.; Martens, J.A. Martens Fischer-Tropsch Waxes Upgrading via Hydrocracking and Selective Hydroisomerization. Oil Gas Sci. Technol. Rev. IFP 2009, 64, 91–112. [Google Scholar] [CrossRef] [Green Version]
Figure 1. XRD patterns of SAPO-11 and Pt/SAPO-11.
Figure 1. XRD patterns of SAPO-11 and Pt/SAPO-11.
Catalysts 13 01100 g001
Figure 2. N2 adsorption–desorption isotherms and BJH pore size distribution curves of SAPO-11 and Pt/SAPO-11.
Figure 2. N2 adsorption–desorption isotherms and BJH pore size distribution curves of SAPO-11 and Pt/SAPO-11.
Catalysts 13 01100 g002
Figure 3. SEM images of SAPO-11 and Pt/SAPO-11. (1,2) are SEM images of SAPO-11, and (3,4) are SEM images of Pt/SAPO-11.
Figure 3. SEM images of SAPO-11 and Pt/SAPO-11. (1,2) are SEM images of SAPO-11, and (3,4) are SEM images of Pt/SAPO-11.
Catalysts 13 01100 g003
Figure 4. TEM diagram of SAPO-11 and Pt/SAPO-11. (1,2) are TEM images of SAPO-11, and (3,4) are TEM images of Pt/SAPO-11.
Figure 4. TEM diagram of SAPO-11 and Pt/SAPO-11. (1,2) are TEM images of SAPO-11, and (3,4) are TEM images of Pt/SAPO-11.
Catalysts 13 01100 g004
Figure 5. NH3-TPD profiles of SAPO-11 and Pt/SAPO-11.
Figure 5. NH3-TPD profiles of SAPO-11 and Pt/SAPO-11.
Catalysts 13 01100 g005
Figure 6. Effect of reaction temperature on isomerization performance. Reaction conditions: P = 1.5 MPa, WHSV = 1.0 h−1, V(H2): V(hexadecane) = 1000.
Figure 6. Effect of reaction temperature on isomerization performance. Reaction conditions: P = 1.5 MPa, WHSV = 1.0 h−1, V(H2): V(hexadecane) = 1000.
Catalysts 13 01100 g006
Figure 7. Effect of weight hourly space velocity on the isomerization performance. Reaction conditions: T = 340 °C, P = 1.5 MPa, V(H2): V(hexadecane) = 1000.
Figure 7. Effect of weight hourly space velocity on the isomerization performance. Reaction conditions: T = 340 °C, P = 1.5 MPa, V(H2): V(hexadecane) = 1000.
Catalysts 13 01100 g007
Figure 8. Effect of the ratio of H2/n–C16 on the isomerization performance. Reaction conditions: T = 340 °C, P = 1.5 MPa, WHSV = 1.0 h−1.
Figure 8. Effect of the ratio of H2/n–C16 on the isomerization performance. Reaction conditions: T = 340 °C, P = 1.5 MPa, WHSV = 1.0 h−1.
Catalysts 13 01100 g008
Figure 9. Effect of reaction pressure on the isomerization performance. Reaction conditions: T = 340 °C, WHSV = 1.0 h−1, V(H2): V(hexadecane) = 1000.
Figure 9. Effect of reaction pressure on the isomerization performance. Reaction conditions: T = 340 °C, WHSV = 1.0 h−1, V(H2): V(hexadecane) = 1000.
Catalysts 13 01100 g009
Figure 10. Effect of metal loadings on the isomerization performance. Reaction conditions: T = 340 °C, P = 1.5 MPa, WHSV = 1.0 h−1, V(H2): V(hexadecane) = 1000.
Figure 10. Effect of metal loadings on the isomerization performance. Reaction conditions: T = 340 °C, P = 1.5 MPa, WHSV = 1.0 h−1, V(H2): V(hexadecane) = 1000.
Catalysts 13 01100 g010
Figure 11. Effect of reaction temperature on hydroisomerization. Green line represents yield.
Figure 11. Effect of reaction temperature on hydroisomerization. Green line represents yield.
Catalysts 13 01100 g011
Figure 12. Effect of reaction temperature on distribution of mono/multi-branched isomers.
Figure 12. Effect of reaction temperature on distribution of mono/multi-branched isomers.
Catalysts 13 01100 g012
Figure 13. Reaction mechanism of mono-branched isomers [25,26].
Figure 13. Reaction mechanism of mono-branched isomers [25,26].
Catalysts 13 01100 g013
Figure 14. Activity stability evaluation of the Pt/SAPO-11 samples. Reaction conditions: T = 340 °C, P = 1.5 MPa, WHSV = 1.0 h−1, V(H2): V(hexadecane) = 1000.
Figure 14. Activity stability evaluation of the Pt/SAPO-11 samples. Reaction conditions: T = 340 °C, P = 1.5 MPa, WHSV = 1.0 h−1, V(H2): V(hexadecane) = 1000.
Catalysts 13 01100 g014
Table 1. Pore structures of SAPO-11 and Pt/SAPO-11.
Table 1. Pore structures of SAPO-11 and Pt/SAPO-11.
SamplesSurface Area (m2·g−1)Pore Volume (cm3·g−1)
BET Surface AreaMicroporeTotalMicropore
SAPO-11145.4130.50.310.069
Pt/SAPO-11103.093.10.290.039
Table 2. The amount of acid distribution of SAPO-11 and Pt/SAPO-11.
Table 2. The amount of acid distribution of SAPO-11 and Pt/SAPO-11.
SamplesAmount of Weak Acid (μmol/g)Amount of Medium Acid (μmol/g)Total Acid (μmol/g)
SAPO-110.7950.3031.098
Pt/SAPO-110.7370.4521.189
Table 3. Product distributions using Pt/SAPO-11.
Table 3. Product distributions using Pt/SAPO-11.
Temperature (°C)Conversion (%)Yield (%)Cracked Product (%)
MoBC16 1MuBC16 2
30043.8131.052.623.2
32063.5741.847.322.7
34081.8056.3614.413.5
36083.6628.0035.024.7
38085.1210.1150.528.8
1 MoBC16, Mono-branched hexadecane isomers. 2 MuBC16, Multi-branched hexadecane isomers.
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

Yang, S.; Liu, X.; Zhang, X.; Sun, W.; Ma, Q.; Song, N.; Lu, M.; Hu, J. Isomerization Properties of Pt/SAPO-11 Catalysts for the Production of Bio-Aviation Kerosene. Catalysts 2023, 13, 1100. https://doi.org/10.3390/catal13071100

AMA Style

Yang S, Liu X, Zhang X, Sun W, Ma Q, Song N, Lu M, Hu J. Isomerization Properties of Pt/SAPO-11 Catalysts for the Production of Bio-Aviation Kerosene. Catalysts. 2023; 13(7):1100. https://doi.org/10.3390/catal13071100

Chicago/Turabian Style

Yang, Sangni, Xuejun Liu, Xin Zhang, Wuji Sun, Qiqi Ma, Nianhua Song, Meizhen Lu, and Jianming Hu. 2023. "Isomerization Properties of Pt/SAPO-11 Catalysts for the Production of Bio-Aviation Kerosene" Catalysts 13, no. 7: 1100. https://doi.org/10.3390/catal13071100

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