Next Article in Journal
Photopolymerization Pattern of New Methacrylate Cellulose Acetate Derivatives
Previous Article in Journal
Effect of Heat Treatment under Different Atmospheres on the Bonding Properties and Mechanism of Ceramiziable Heat-Resistant Adhesive
Previous Article in Special Issue
Amorphous Elastomeric Ultra-High Molar Mass Polypropylene in High Yield by Half-Titanocene Catalysts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Experimental and Theoretical Insights into the Effect of Dioldibenzoate Isomers on the Performance of Polypropylene Catalysts

1
Division of Catalytic Science, SINOPEC (Beijing) Research Institute of Chemical Industry Co., Ltd., Beijing 100013, China
2
Division of Polypropylene Research, SINOPEC (Beijing) Research Institute of Chemical Industry Co., Ltd., Beijing 100013, China
3
Division of Polyethylene Research, SINOPEC (Beijing) Research Institute of Chemical Industry Co., Ltd., Beijing 100013, China
*
Authors to whom correspondence should be addressed.
Polymers 2024, 16(4), 559; https://doi.org/10.3390/polym16040559
Submission received: 11 January 2024 / Revised: 7 February 2024 / Accepted: 12 February 2024 / Published: 19 February 2024
(This article belongs to the Special Issue Catalytic Olefin Polymerization and Polyolefin Materials)

Abstract

:
Experimental investigations and density functional theory (DFT) calculations were carried out to study the comprehensive effect of different 3,5-heptanedioldibenzoate (HDDB) optical isomers as the internal electron donor on the catalytic performance of Ziegler−Natta catalysts. The experimental catalytic activity of HDDB has a positive correlation with the relative content of the mesomer incorporated during catalyst preparation, while the hydrogen response of HDDB displayed a negative correlation with the relative content of the mesomer. In order to apply the DFT calculation results to the macroscopic activity of the catalyst, the content of the active centers of the catalyst was analyzed. Assuming that the content of the active centers is proportional to the internal electron donor content of the catalyst, binary linear regression was carried out, which showed a good linear correlation between experimental activity data and internal electron donor content. Furthermore, the fitted activity of the single active centers aligned well with the calculated activation energies. These results revealed that the catalytic activity of polypropylene (PP) catalysts is dependent on both the active center content and the catalytic activity of an individual active center. Additionally, the lower hydrogen response of HDDB leads to a higher molecular weight of polypropylene obtained from the RS-containing catalyst compared to the SS-containing catalyst. Further study reveals that the hydrogen transfer reactions of 2,4-pentanediol dibenzoate (PDDB)/HDDB are influenced by the orientation of the methyl/ethyl groups in different isomers, which affect the activation energy differences between the hydrogen transfer reaction and the propylene insertion reaction, and finally influence the molecular weight of PP.

1. Introduction

Ziegler-Natta (ZN) catalysts have achieved great success in the polyolefin industry since their invention [1,2], with an annual production exceeding 150 million tons [3]. Notably, the production of polypropylene (PP) reached 102.8 million tons in 2021, and more than 90% of polypropylene was produced using ZN catalysts. The ZN catalysts consist of TiCl4, MgCl2 support, and internal electron donors (IDs), while external electron donors (EDs) and initiators are added during polymerization [4]. Until now, industrial ZN catalysts have been developed to the fifth generation. Starting from the third generation of ZN catalysts, catalytic performance has been improved predominantly through the development of IDs [5]. The IDs for the fifth generation of industrial ZN catalysts are non-phthalate compounds including 1,3-diethers [6,7], succinates [8,9], and diol diesters [10,11,12]. In addition, new electron donors such as sulfonyl amine [13,14,15] and its derived diesters [16] are also expected to be introduced into industrialization.
Despite the remarkable success and advancements of ZN catalysts in the industry, academic research on the mechanism of ZN catalysts is complicated, due to its multi-component disperse structures and the challenges associated with the microstructure characterizations. Recently, several studies involving the exploration of microscopic active centers have been carried out [17,18]. Fan et al. explored the catalyst active center numbers during polymerization reactions using the quenching method [19,20]. Christophe and co-workers investigated the interactions between the support and the electron donor [21]. Unfortunately, experiments on characterizing the precise structures of active centers are generally expensive and difficult. Since the 21st century, a large number of theoretical studies have been applied to the mechanism research of ZN catalysts [22], including explorations on the interactions among each component [23,24,25,26,27,28,29,30,31,32,33,34,35] and the influences of electron donors on the catalytic performance of ZN catalysts [36,37,38,39,40,41,42]. Although earlier computational studies had inconsistent conclusions with the experiment results due to different models and computational methods [43], the polymerization process of propylene in the MgCl2/TiCl4/ID system has been gradually elucidated in recent years. During this process, TiCl4 is adsorbed on a (110)-facet of MgCl2 to form a stable active center, and the electron donor close to TiCl4 enhances the stereoselectivity of the ZN catalyst. Some studies also revealed that triethylaluminium could affect the stereoselectivity of ZN catalysts [44,45,46].
Compared with the current qualitative research on ZN catalysts that explains the basic mechanism of ZN catalyst systems [47], there is a lack of quantitative simulation for industrial ZN catalysts. Unlike studies on isotacticity, which are easily explored through the comparison of different active energies for a single active center, research on macroscopic properties such as catalytic activity and average molecular weight is rare. Macroscopic property research must consider factors such as the quantity of active centers, various reaction conditions, and changes in the actual reaction environment. Moreover, significant differences in the actual reaction performance would appear when using different preparation methods for the same catalyst composition, even when using the same type of electron donor isomers [43].
Catalysts [11,12,48] containing different isomers of diol diester electron donors have significantly different properties. In our previous study [49], we reported that the ZN catalyst using optical isomers of electron donor PDDB (2,4-pentanedioldibenzoate) has significant differences in its ID content and stereoselectivity. DFT calculations showed that the primary factors contributing to the different content of ID isomers in the catalyst are the distinct conformations observed in the free state and adsorption state of ID molecules, along with different adsorption energies on the MgCl2 support.
In this manuscript, we extend the exploration of diol diester IDs from PDDB to HDDB. Apart from the investigations on stereoselectivity, comprehensive theoretical calculations were carried out to elucidate the molecular level understanding of propylene polymerization catalyzed by ZN catalysts, and to investigate the factors contributing to diverse performances including catalytic activity, molecular weight, and other properties.

2. Experiment and Calculation Sections

2.1. Experiments

2.1.1. Materials

All operations were executed under a nitrogen atmosphere using glove-box and standard Schlenk (Beijing, China) techniques. Propylene and MgCl2 were obtained from Sinopec (Beijing, China); AlEt3, TiCl4, and cyclohexylmethyldimethoxylalkoxysilane (CHMMS) were purchased from J&K Scientific (Beijing, China) without further purification; AlEt3 and CHMMS were diluted in n-hexane to 0.5 mol/L and 0.1 mol/L, respectively. All other solvents were purchased from Sinopharm Chemical Reagent Co. Ltd. (Shanghai, China) and dried with 4Å molecular sieves in advance.

2.1.2. Catalyst Preparation

The stereoisomers of HDDB (characterized using 1H NMR, see Figures S1 and S2) and the MgCl2-supported catalysts [11,12,48] were synthesized or prepared according to the previously reported procedure. A typical synthesis procedure of MgCl2-supported catalysts is as follows: First, 4.8 g magnesium chloride, 95 mL toluene, 4 mL epoxy chloropropane, and 12.5 mL tributyl phosphate were added successively to a nitrogen-purged five-necked reactor. The mixture was heated to 60 °C with stirring and maintained at this temperature for 1.5 h. Then, 1.4 g phthalic anhydride was added to the reaction mixture and the reaction was maintained for 1 h. The solution was then cooled down, and 56 mL TiCl4 was introduced slowly below −20 °C. After the mixture was heated to 80 °C, 5 mmol of IDs (HDDB with different contents of mesomer and racemate, or others) was added and maintained for 1 h. After removing the supernatant, the residue was washed with toluene. The obtained solid precipitate was treated twice with 40 mL TiCl4 and 60 mL toluene at 100 °C for 2 h. After sedimentation and filtration, the residue was washed with 60 mL hexane to obtain the solid catalyst.

2.1.3. Bulk Polymerization Process

Propylene polymerization was carried out in a 5 L stainless steel reactor with liquefied propylene [11,12,48]. The typical procedure for the bulk polymerization of propylene was as follows: A total of 5 mL AlEt3, 1 mL CHMMS, about 9 mg of the solid catalyst, and 1.2 SL (standard liter) hydrogen were added. Then, 2.3 L liquid propylene was added to the 5 L stainless steel reactor, which had been replaced with propylene gas completely. The reactor was heated to 70 °C and the polymerization was processed at 70 °C for 1 h. Then, polypropylene (PP) resin powder was obtained after reducing both the temperature and pressure.

2.1.4. Characterization

Nuclear magnetic resonance (1H NMR) was characterized using the Bruke dmx300 nuclear magnetic resonance spectrometer (300 MHz, solvent is CDCl3, TMS as internal standard, and measuring temperature is 300 K).
The content of meso-/rac-HDDB was detected using liquid chromatography (LC). The content of meso-/rac-HDDB in compounds was detected directly. The catalyst was first dissolved in an acidic media and then extracted using ethyl acetate, to be examined. Furthermore, separations were performed on the Waters-600E UPLC H-Class (Waters Corporation, 34 Maple Street, Milford, MA, USA) with a column of ACQUITY UPLC BEH C18 (50 mm × 2.1 mm, 1.7 μm). The absolute content of total HDDB (including mesomer and racemate) in the catalyst was calculated by using a standard curve.
The isotactic index (I.I) of the polymer was calculated using the following method: First, 2 g dried polymer sample is extracted with boiling heptane in an extractor for 6 h, then the residual substance is dried to a constant weight, and the ratio of the weight (g) of residual polymer to two is named as the isotactic index.
The melt flow rate (MFR) of the polymer was measured using the test standard GB/T 3682.1–2018 [50] (corresponds to the standard of ASTM D 1238-13). The MFR is the amount of the polymer (g) flowing through the capillary for 10 min, under a pressure of 2.16 kg at 230 °C.
The molecular weight and its distribution (PD) of the polypropylene were determined using a Waters Alliance V2000 gel permeation chromatograph (GPC) equipped with a refractive index detector, using three Polymer Laboratory MIXED-B columns and 1,2,4-trichlorobenzene as the solvent at 150 °C. The number-average and weight-average molecular weight (Mn and Mw, respectively) values were evaluated with reference to a polystyrene standard calibration.

2.2. Computational Details

All DFT calculations were performed with the Gaussian16 program, Revision C.01. All geometry optimizations of intermediates and transition states were carried out using PBE0-D3/6-311G(d,p) [51,52] in combination with the Becke–Johnson (BJ) damping correlation [53]. Frequency calculations were also conducted at the same level of theory to obtain vibrational frequencies to determine the identity of stationary points as intermediates (no imaginary frequencies) or transition states (only one imaginary frequency). To avoid the errors associated with low frequency, Grimme’s quasi-RRHO [54] model was used for Gibbs free energy calculations using the Sherm 2.3 program [55]. Compared with the default RRHO model provided by Gaussian16, the quasi-RRHO model demonstrated a quantitative accuracy improvement. Gibbs free energies were calculated at 343.15 K and 1 atm for the stable structures of the catalyst and transition state structures. The MgCl2 cluster model, derived from δ-MgCl2 crystals, underwent geometry optimization with the Mg atoms constrained at a fixed distance of 3.6363 Å (Mg···Mg) while the other atoms are free.

3. Results and Discussions

3.1. The Polymerization Experiments

The polymerization results of the ZN catalysts with different HDDB stereoisomers as IDs are shown in Table 1 (see Table S1 for the polymerization results of PDDB). With the increase in mesomer content (see Figures S3–S16), the ID content, the activity (AC), and stereoselectivity (characterized using the isotactic index, I.I.) of the catalysts, as well as the molecular weight of the polymer product shows an increasing trend, while the hydrogen response (characterized using the melt flow rate, MFR) and the molecular weight distribution (characterized using the polydispersity, PD) exhibit a decreasing or narrowing trend. These results suggest that the stereoisomer content of HDDB in the catalysts plays a crucial role in the catalytic performance during propylene polymerization.

3.2. The Active Center Models for DFT Calculations

HDDB has two unequal stereoisomers: racemate and mesomer. In the free state, the free energy of the mesomer is 1.5 kcal/mol higher than that of the racemate isomers. Two oxygen atoms of carbonyl groups in the RS-isomer are in the same direction, while those of SS-isomers are in the opposite direction (See Figure 1 for the structure of free HDDB).
Based on our previous work [49], where we compared the adsorption energy of PDDB on MgCl2 and the relative energy of various active center models, we assumed that the trans-a model (differing only in left/right order as trans-a/b) represents the active center structure with the lowest energy for both RS-PDDB and SS-PDDB. With this model, the structure for the active center with the lowest energy and stereoselectivity were calculated, respectively, revealing that the configuration of donor HDDB resembles that of PDDB (Table 2). Similarly to PDDB, the key difference in the transition state (TS) structures of insertion reactions with the RS-HDDB or SS-HDDB isomer is the orientation of the ethyl group. For the RS isomer, the ethyl group pointed towards the Ti atom, thereby restricting the space near the Ti center, resulting in a large energy difference in the si- and re-insertion reactions of propylene molecules. Conversely, for the SS isomer, the ethyl group is nearly perpendicular to the (110)-facet of MgCl2, which has less influence on the active center, resulting in less energy difference for the insertion reaction of propylene. So, in this manuscript, we maintain the use of the trans-a model as the active center model for HDDB (See Figure 2 for the active center structure of HDDB).

3.3. DFT Calculations about the Catalytic Activity

Considering that the transition state energy for the si-insertion reaction is lower than that of re-insertion (see Table 1), the si-insertion and propylene–catalyst complex were selected as the reaction channel and energy starting point, respectively to calculate the energy barriers for the active center with a trans-a structure. The structures for the HDDB–catalyst complex and HDDB–catalyst–propylene (si-insertion) complex are listed in Figure 3 and Figure 4, respectively. According to the calculation results on activation energies, the activity of SS-HDDB is slightly higher than that of RS-HDDB (see Table 3). The reason for the difference can be seen in Figure 4. The hydrogen atoms of RS-HDDB are only 1.95 angstroms apart from the hydrogen atoms of the ethyl group representing the polymer chain. This shows a strong steric effect. The same distance for SS-HDDB is 2.49 angstroms, and the interference with the reaction center is almost negligible.

3.4. Analysis of Factors Affecting Catalytic Activity

According to the experimental data from Table 1 and Table S1, it is evident that the catalytic activity of the catalyst increases with a higher RS isomer content, while the overall IDs would also be increased. We suggest that the adsorption modes for different ID molecules on the MgCl2 support are similar during the catalyst preparation processes. Assuming an ideal single-layer adsorption model, where the number of active centers is proportional to the quantity of ID molecules, the model for catalytic activity can be described as
ATmacro = ARSmacro + ASSmacro = ARSCRS + ASSCSS
Here, ATmacro represents the total activity of the catalyst, ARSmacro is the macroscopic activity of the active centers with the RS isomer, and ASSmacro is the macroscopic activity of the active centers with the SS isomer. ARS is proportional to the activity of the single RS isomer active center, CRS is the relative content of the RS isomer in the catalyst, ASS is proportional to the activity of the single SS isomer active center, and CSS is the relative content of the SS isomer in the catalyst. Additionally, we defined the relative activity ratio of the single SS isomer active center to that of the RS isomer as ASS/ARS.
The macroscopic activity data and ID isomer content, as presented in Table 1 and Table S1, underwent binary regression analysis and the findings are summarized in Table 4. The outcomes in Table 4 indicate that the catalytic activity for the active center with the SS isomer is slightly higher than that of the RS isomer for both PDDB and HDDB, which is consistent with the calculated results. Figure 5 shows the changes in macroscopic activity attributed to the active centers when the isomer content increased for both the RS and SS isomers. Notably, the depicted relationship demonstrates a high degree of linearity, with a regression coefficient exceeding 0.999.
The calculation results indicated that the activity for the active center with the RS isomer is lower than that of the SS isomer, while the macroscopic activity for the catalyst exhibits a positive correlation with the RS content in the experiments. These results revealed that the mesomer of dioldibenzoate improves the experimental catalytic activity of polypropylene (PP) catalysts through increasing the active center content, rather than the catalytic activity of individual active centers. This phenomenon can be attributed to the fact that it is easier for the RS isomer to be adsorbed on MgCl2 surfaces (see Table 5), resulting in an increase in the content of active centers. This conclusion is also consistent with our previous work [49], where the mesomer ID molecule is easier to be adsorbed on the (110)-facet of the MgCl2 support with TiCl4, compared with the racemate isomer. For the ratio of TiCl4/ID and the catalyst preparation, the procedure remains unchanged in the series of experiments (Entry 1 to Entry 7 in Table 1), and the active center content is reasonably positively proportional to the ID content. The result for the binary regression analysis in Table 4 and Figure 5 is consistent with the calculated relative activity for the active centers with the SS/RS isomer in Table 3, which reveals that the assumption above is suitable for our catalyst system. These results revealed that the catalytic activity of polypropylene (PP) catalysts is dependent on both the active center content and the catalytic activity of the individual active center.

3.5. Hydrogen Transfer Reaction

The chain transfer reaction is quite important in the polymerization process, and the relative rate of the hydrogen transfer reaction to the polymerization affects the molecular weight of polypropylene. Previous studies suggest that the chain transfer to the monomer serves as the primary chain transfer reaction, which could be applied in heterogeneous ZN catalysts [56,57], similar to metallocene [58,59] and non-metallocene systems [60,61]. Under conditions where the monomer concentration is low or the alkylaluminium concentration is relatively high, the chain transfer to the metal and the chain transfer to the cocatalyst may become the dominant chain transfer reactions [62]. In the area of industrial ZN catalysts, the use of hydrogen to control molecular weight makes hydrogenolysis (i.e., the hydrogen transfer reaction) the primary chain transfer reaction [56,63]. The hydrogen transfer reaction can be represented by Equation (2):
Ti–CH2–Polymer + H2 → Ti–H + CH3–Polymer
Considering the existence of H2, neglecting other chain transfer reactions (chain transfer to monomer, chain transfer to alkylaluminum, and ethyl β-dehydrogenation) and assuming constant concentrations of each reactant, the relationship between the hydrogen transfer reaction and the molecular weight can be expressed as follows:
Mw α kpCp/(kHCH) α exp[(∆GH − ∆Gp)/RT]
where Mw is molecular weight, kp is the rate of propylene insertion reaction, Cp is the concentration of propylene, kH is the rate of hydrogen transfer reaction, CH is the concentration of hydrogen, and ∆GH and ∆Gp are the activation energies of the hydrogen transfer reaction and the propylene insertion reaction, respectively.
The hydrogen transfer reaction and propylene insertion reaction compete for the same active center. The difference between their activation energies is positively correlated to the molecular weight of polypropylene. Substituting the propylene molecule with an H2 molecule in the model of Figure 4 allowed for the calculation of the hydrogen transfer reaction. The results from the Gibbs energy calculations revealed that there is no stable intermediate complex for H2 and the active center. Instead, the hydrogen transfer reaction proceeded directly from a separated state to a transition state. In contrast to Ref. [33], the transfer of polymer chains (represented by the ethyl group in Figure 6) with H2 was attributed to the interaction between the β-H and Ti center, while the insertion reaction of propylene was attributed to the interaction between the α-H and Ti center (see Figure 6 for the transition state structures). The activation energies are listed in Table 6, which suggests that the transfer activation energy of H2 at the RS active center is higher than that of the SS active center for both PDDB and HDDB. This result indicates that the hydrogen transfer reaction in the RS isomer system is slightly more challenging than in the SS isomer system, consistent with the experimental results where the molecular weight of polypropylene obtained from the RS-containing catalyst is higher than that from the SS-containing catalyst.
The differences in catalytic activity and activation energy showed that the orientation for the methyl/ethyl groups of RS isomers led to the variations in hydrogen transfer reactions, which finally resulted in the distinct hydrogen response and polypropylene molecular weight for the catalysts. Nevertheless, due to the smaller space requirement of the hydrogen transfer reaction than the propylene insertion reaction, the variations in the difficulty of the hydrogen transfer reactions were relatively small. These results show that the variation trend for the activation energy of the hydrogen transfer reaction in different kinds of active centers is consistent with that of the propylene insertion reaction, with the variation range slightly different from each other.

4. Conclusions

We prepared ZN catalysts for propylene polymerization using the mixtures of racemate and mesomer isomers of HDDB as IDs. It was observed that the meso-isomers preferred to be adsorbed on the catalyst. As the relative content of meso-isomer increased, there was a corresponding rise in the total amount of IDs adsorbed on the catalysts. This increase was concomitant with the improvements of isotacticity, catalytic activity, and average molecular weight, indicating that the mesomer ID enhances the performance of the ZN catalyst, surpassing the benefits derived from racemate isomer.
The combination of DFT calculation results from our previous paper [49] with the current study shows a remarkable alignment between the calculated stereoselectivity of PDDB and HDDB stereoisomers and the experimental findings. Notably, the orientation of the 2,4-methyl or 3,5-ethyl group in the optimized structures of MgCl2-TiCl4-ID varied significantly when employing mesomer or racemate as the ID, which was the critical factor for the catalytic performances, especially for re- and si-insertion reactions.
After carrying out the binary linear regression analysis for the experimental macroscopic activity of the catalyst and the content of ID molecules, it is evident that the apparent activity of the catalyst was proportional to the ID content, indicating that the ID content was also proportional to the number of active centers. The results showed that the activity of the single RS active center was lower than that of the single SS active center, which was consistent with the activation energy obtained from DFT calculations. These results revealed that the catalytic activity of polypropylene (PP) catalysts is dependent on both the active center content and the catalytic activity of the individual active center. Additionally, the calculated activation energy for the hydrogen transfer reaction consistently corresponds to the trend observed for the relative values of molecular weight. These results show that the variation trend for the activation energy of the hydrogen transfer reaction in different kinds of active centers is consistent with that of the propylene insertion reaction, with the variation ranges slightly different from each other.
In summary, our comprehensive analysis, incorporating factors such as the adsorption energy of ID, the activation energy difference for re-/si- insertion reaction at the active center, the activation energy for active centers, and the activation energy for hydrogen transfer reactions calculated through our active center model, aligns remarkably well with the corresponding experimental results including the observed ID content, the relative activity of the catalyst, the stereoselectivity, and the molecular weight. The remarkable consistency supports the appropriateness of our active center model in accurately describing the microstructure of ZN catalysts, which enables us to gain a detailed understanding of the impact derived from electron donors on the performance of ZN catalysts.
Based on the research of this manuscript and our previous paper [49], some insights on designing efficient electron donors have been obtained. Firstly, the electron donors should have enough adsorption energy to be adsorbed on the (110)-facet of the MgCl2 support, and the conformation of the free ID molecule should be the same as that of the adsorbed ID molecule, which can produce more active centers during catalyst preparation. Secondly, the substituent groups for the ID molecule neighboring the active center should be close enough to the Ti atom, which could have an appropriate effect on the re/si-insertion reaction, and can result in higher stereoselectivity. However, the excess steric effect would break down the catalytic activity. The influences for the steric conformation of ID molecules on both steroselectivity (i.e., the re/si-insertion reaction) and catalytic activity should be considered during the design of novel electron donors.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/polym16040559/s1, Figure S1. 1H NMR of meso-HDDB. Figure S2. 1H NMR of rac-HDDB. Figure S3. LC of Entry1 compound. Figure S4. LC of Entry1 catalyst. Figure S5. LC of Entry2 compound. Figure S6. LC of Entry2 catalyst. Figure S7. LC of Entry3 compound. Figure S8. LC of Entry3 catalyst. Figure S9. LC of Entry4 compound. Figure S10. LC of Entry4 catalyst. Figure S11. LC of Entry5 compound. Figure S12. LC of Entry5 catalyst. Figure S13. LC of Entry6 compound. Figure S14. LC of Entry6 catalyst. Figure S15. LC of Entry7 compound. Figure S16. LC of Entry7 catalyst Figure S17. GPC of Entry 1 PP powder. Figure S18. GPC of Entry 2 PP powder. Figure S19. GPC of Entry 3 PP powder Figure S20. GPC of Entry 4 PP powder. Figure S21. GPC of Entry 5 PP powder. Figure S22. GPC of Entry 6 PP powder. Figure S23. GPC of Entry 7 PP powder. Table S1. The polymerization results for the catalysts with different PDDB stereoisomers as ID. NMR, LC, GPC, and DFT calculation data (PDF).

Author Contributions

Validation, X.X.; Investigation, H.F. and C.L.; Writing—original draft, H.F.; Writing—review and editing, C.L., S.T. and Z.S.; Supervision, J.Z. and X.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work is supported by the SINOPEC project (#213002 and #219023-1) in the SINOPEC (Beijing) Research Institute of Chemical Industry Co., Ltd. and The APC was funded by SINOPEC project (#219023-1).

Institutional Review Board Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

Author Huasheng Feng was employed by the company SINOPEC (Beijing) Research Institute of Chemical Industry Co. Ltd. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. GB828791A; Method of Selectively Polymerising α-olefins. Montecatini Societa Generale L.: Montecatini, Italy, 1955.
  2. US3438956A; Separation of Macromolecules Having Different Stereoregularities from Polypropylenes Comprising Mixtures of Such Macromolecules. Montedison S.p.A.: Milan, Italy, 1968.
  3. Kumawat, J.; Gupta, V.K. Fundamental Aspects of Heterogeneous Ziegler-Natta Olefin Polymerization Catalysis: An Experimental and Computational Overview. Polym. Chem. 2020, 11, 6107–6128. [Google Scholar] [CrossRef]
  4. Pasquini, N. Polypropylene Handbook, 2nd ed.; Hanser Gardner Publications, Inc.: Munich, Germany, 2005; ISBN 1-56990-385-9. [Google Scholar]
  5. Kaminsky, W. Polyolefins: 50 Years after Ziegler and Natta I—Polyethylene and Polypropylene; Springer: Berlin/Heidelberg, Germany, 2013; ISBN 978-3-642-40807-6. [Google Scholar]
  6. Morini, G.; Albizzati, E.; Balbontin, G.; Mingozzi, I.; Sacchi, M.C.; Forlini, F.; Tritto, I. Microstructure Distribution of Polypropylenes Obtained in the Presence of Traditional Phthalate/Silane and Novel Diether Donors:  A Tool for Understanding the Role of Electron Donors in MgCl2-Supported Ziegler-Natta Catalysts. Macromolecules 1996, 29, 5770–5776. [Google Scholar] [CrossRef]
  7. Sacchi, M.C.; Forlini, F.; Tritto, I.; Locatelli, P.; Morini, G.; Noristi, L.; Albizzati, E. Polymerization Stereochemistry with Ziegler-Natta Catalysts Containing Dialkylpropane Diethers:  A Tool for Understanding Internal/External Donor Relationships. Macromolecules 1996, 29, 3341–3345. [Google Scholar] [CrossRef]
  8. Cecchin, G.; Morini, G.; Pelliconi, A. Polypropene Product Innovation by Reactor Granule Technology. Macromol. Symp. 2001, 173, 195–210. [Google Scholar] [CrossRef]
  9. Zaccaria, F.; Vittoria, A.; Correa, A.; Ehm, C.; Budzelaar, P.H.M.; Busico, V.; Cipullo, R. Internal Donors in Ziegler-Natta Systems: Is Reduction by AlR3 a Requirement for Donor Clean-Up? ChemCatChem 2018, 10, 984–988. [Google Scholar] [CrossRef]
  10. US7388016B2; Solid Catalyst Component for Polymerization of Olefins, Catalyst Comprising the Same and Use Thereof. Beijing Research Institute of Chemical Industry: Beijing, China, 2008.
  11. Gao, M.; Liu, H.; Wang, J.; Li, C.; Ma, J.; Wei, G. Novel MgCl2-Supported Catalyst Containing Diol Dibenzoate Donor for Propylene Polymerization. Polymer 2004, 45, 2175–2180. [Google Scholar] [CrossRef]
  12. Sun, Z.; Liu, H.; Gao, M. Performance of Catalyst with Different 2,4-Pentanediol Diester as Internal Donor. Petrochem. Technol. 2010, 39, 1236–1240. [Google Scholar]
  13. US8404789B2; Olefin Polymerization Catalyst and Preparation Method and Use Thereof. Petrochina Company Limited: Beijing, China, 2013.
  14. Li, H.S.; Yi, J.J.; Cui, C.M. Bis(Trifluoromethylsulfonyl)-Phenylamines as Internal Donors for Ziegler-Natta Polymerization Catalysts. China Pet. Process. Petrochem. Technol. 2008, 10, 51–54. [Google Scholar]
  15. Wang, L.; Yin, B.Z.; Yi, J.J.; Cui, C.M. Propylene Polymerization Catalysts with Sulfonyl Amines as Internal Electron Donors. China Pet. Process. Petrochem. Technol. 2013, 15, 19–23. [Google Scholar]
  16. Matta, A.; Chammingkwan, P.; Singh, B.K.; Terano, M.; Kaneko, T.; Taniike, T. Truxillic and Truxinic Acid-based, Bio-derived Diesters as Potent Internal Donor in Ziegler-Natta Catalyst for Propylene Polymerization. Appl. Catal. A-Genet. 2018, 554, 80–87. [Google Scholar] [CrossRef]
  17. Patil, H.; Karthikeyan, S.; Kote, V.; Sengupta, P.; Samanta, P.; Kadam, P.; Natarajan Venkateswaran, N.; Gupta, V. An Insight into Ziegler-Natta Catalyst Active Site Distribution for Polyolefins: Application of Jitter Differential Evolution. Polym. Bull. 2023, 80, 1425–1445. [Google Scholar] [CrossRef]
  18. Piovano, A.; Wada, T.; Amodio, A.; Takasao, G.; Ikeda, T.; Zhu, D.; Terano, M.; Chammingkwan, P.; Groppo, E.; Taniike, T. Formation of Highly Active Ziegler-Natta Catalysts Clarified by a Multifaceted Characterization Approach. ACS Catal. 2021, 11, 13782–13796. [Google Scholar] [CrossRef]
  19. Shen, X.; Hu, J.; Fu, Z.; Lou, J.; Fan, Z. Counting the Number of Active Centers in MgCl2-Supported Ziegler-Natta Catalysts by Quenching with 2-Thiophenecarbonyl Chloride and Study on the Initial Kinetics of Propylene Polymerization. Catal. Commun. 2013, 30, 66–69. [Google Scholar] [CrossRef]
  20. Weng, Y.; Jiang, B.; Fu, Z.; Fan, Z. Mechanism of Internal and External Electron Donor Effects on Propylene Polymerization with MgCl2-Supported Ziegler-Natta Catalyst: New Evidences Based on Active Center Counting. J. Appl. Polym. Sci. 2018, 135, 46605. [Google Scholar] [CrossRef]
  21. Yakimov, A.; Xu, J.; Searles, K.; Liao, W.; Antinucci, G.; Friederichs, N.; Busico, V.; Copéret, C. DNP-SENS Formulation Protocols to Study Surface Sites in Ziegler-Natta Catalyst MgCl2 Supports Modified with Internal Donors. J. Phys. Chem. C 2021, 125, 15994–16003. [Google Scholar] [CrossRef]
  22. Bahri-Laleh, N.; Hanifpour, A.; Mirmohammadi, S.A.; Poater, A.; Nekoomanesh-Haghighi, M.; Talarico, G.; Cavallo, L. Computational Modeling of Heterogeneous Ziegler-Natta Catalysts for Olefins Polymerization. Prog. Polym. Sci. 2018, 84, 89–114. [Google Scholar] [CrossRef]
  23. Credendino, R.; Busico, V.; Causà, M.; Barone, V.; Budzelaar, P.H.M.; Zicovich-Wilson, C. Periodic DFT Modeling of Bulk and Surface Properties of MgCl2. Phys. Chem. Chem. Phys. 2009, 11, 6525–6532. [Google Scholar] [CrossRef] [PubMed]
  24. Stukalov, D.V.; Zilberberg, I.L.; Zakharov, V.A. Surface Species of Titanium(IV) and Titanium(III) in MgCl2-Supported Ziegler-Natta Catalysts. A Periodic Density Functional Theory Study. Macromolecules 2009, 42, 8165–8171. [Google Scholar] [CrossRef]
  25. Stukalov, D.V.; Zakharov, V.A. Active Site Formation in MgCl2–Supported Ziegler-Natta Catalysts. A Density Functional Theory Study. J. Phys. Chem. C 2009, 113, 21376–21382. [Google Scholar] [CrossRef]
  26. Brambilla, L.; Zerbi, G.; Piemontesi, F.; Nascetti, S.; Morini, G. Structure of MgCl2–TiCl4 Complex in Co-milled Ziegler-Natta Catalyst Precursors with Different TiCl4 Content: Experimental and Theoretical Vibrational Spectra. J. Mol. Catal. A Chem. 2007, 263, 103–111. [Google Scholar] [CrossRef]
  27. D’Amore, M.; Credendino, R.; Budzelaar, P.H.M.; Causá, M.; Busico, V. A Periodic Hybrid DFT Approach (Including Dispersion) to MgCl2–Supported Ziegler-Natta Catalysts–1: TiCl4 Adsorption on MgCl2 Crystal Surfaces. J. Catal. 2012, 286, 103–110. [Google Scholar] [CrossRef]
  28. Breuza, E.; Antinucci, G.; Budzelaar, P.H.M.; Busico, V.; Correa, A.; Ehm, C. MgCl2-Supported Ziegler-Natta Catalysts: A DFT-D “Flexible-Cluster” Approach to Internal Donor Adducts. J. Phys. Chem. C 2018, 122, 9046–9053. [Google Scholar] [CrossRef]
  29. Vanka, K.; Singh, G.; Iyer, D.; Gupta, V.K. DFT Study of Lewis Base Interactions with the MgCl2 Surface in the Ziegler-Natta Catalytic System: Expanding the Role of the Donors. J. Phys. Chem. C 2010, 114, 15771–15781. [Google Scholar] [CrossRef]
  30. Cavallo, L.; Del Piero, S.; Ducéré, J.-M.; Fedele, R.; Melchior, A.; Morini, G.; Piemontesi, F.; Tolazzi, M. Key Interactions in Heterogeneous Ziegler-Natta Catalytic Systems:  Structure and Energetics of TiCl4–Lewis Base Complexes. J. Phys. Chem. C. 2007, 111, 4412–4419. [Google Scholar] [CrossRef]
  31. Cavallo, L.; Ducéré, J.M.; Fedele, R.; Melchior, A.; Mimmi, M.C.; Morini, G.; Piemontesi, F.; Tolazzi, M. Ziegler-Natta Catalytic Systems: Calorimetric and DFT Study on TiCl4-Lewis Base Interactions. J. Therm. Anal. Calorim. 2008, 91, 101–106. [Google Scholar] [CrossRef]
  32. Cavallo, L.; Fedele, R.; Morini, G.; Ducéré, J.-M.; Melchior, A.; Correa, A.; Piemontesi, F.; Tolazzi, M. An Empirical Correction Term to Density Functional Theory for the Description of the TiCl4-Lewis Base Complexes. Macromol. Symp. 2007, 260, 122–126. [Google Scholar] [CrossRef]
  33. Bahri-Laleh, N.; Nekoomanesh-Haghighi, M.; Mirmohammadi, S.A. A DFT Study on the Effect of Hydrogen in Ethylene and Propylene Polymerization Using a Ti-Based Heterogeneous Ziegler-Natta Catalyst. J. Organomet. Chem. 2012, 719, 74–79. [Google Scholar] [CrossRef]
  34. Bahri-Laleh, N. Interaction of Different Poisons with MgCl2/TiCl4 Based Ziegler-Natta Catalysts. Appl. Surf. Sci. 2016, 379, 395–401. [Google Scholar] [CrossRef]
  35. Guo, X.; Shao, Y.; Luo, J.; Liu, Z.; Liu, B. The Atomic Defects on the (104) and (110) Surfaces of the MgCl2-Supported Ziegler-Natta Catalyst: A Periodic DFT Study. Catal. Sci. Technol. 2022, 12, 6761–6770. [Google Scholar] [CrossRef]
  36. Credendino, R.; Liguori, D.; Fan, Z.; Morini, G.; Cavallo, L. Toward a Unified Model Explaining Heterogeneous Ziegler-Natta Catalysis. ACS Catal. 2015, 5, 5431–5435. [Google Scholar] [CrossRef]
  37. Guo, X.; Cui, L.; Wang, Y.; Yi, J.; Sun, J.; Liu, Z.; Liu, B. Mechanistic Study on Effect of Electron Donors in Propylene Polymerization Using the Ziegler-Natta Catalyst. J. Phys. Chem. C 2021, 125, 8533–8542. [Google Scholar] [CrossRef]
  38. Guo, X.; Cui, L.; Yi, J.; Liu, Z.; Liu, B. Understanding the Role of Sulfonyl Amine Donors in Propylene Polymerization Using MgCl2-Supported Ziegler-Natta Catalyst. J. Phys. Chem. C 2022, 126, 8655–8666. [Google Scholar] [CrossRef]
  39. Guo, X.; Liu, Z.; Fan, Z.; Liu, B. New Insights into the Nature of Ti(II) and Ti(III) Active Sites in the Heterogeneous Ziegler-Natta Catalyst. J. Phys. Chem. C 2023, 127, 5720–5730. [Google Scholar] [CrossRef]
  40. Kumawat, J.; Kumar Gupta, V.; Vanka, K. The Nature of the Active Site in Ziegler-Natta Olefin Polymerization Catalysis Systems—A Computational Investigation. Eur. J. Inorg. Chem. 2014, 2014, 5063–5076. [Google Scholar] [CrossRef]
  41. Taniike, T.; Terano, M. Coadsorption Model for First-Principle Description of Roles of Donors in Heterogeneous Ziegler-Natta Propylene Polymerization. J. Catal. 2012, 293, 39–50. [Google Scholar] [CrossRef]
  42. Ratanasak, M.; Parasuk, V. Understanding the Roles of Novel Electron Donors in Ziegler-Natta Catalyzed Propylene Polymerization. RSC Adv. 2016, 6, 112776–112783. [Google Scholar] [CrossRef]
  43. Milanesi, M.; Piovano, A.; Wada, T.; Zarupski, J.; Chmamingkwan, P.; Taniike, T.; Groppo, E. Influence of the Synthetic Procedure on the Properties of Three Ziegler-Natta Catalysts with the same 1,3-Diether Internal Donor. Catal. Today 2023, 418, 114077. [Google Scholar] [CrossRef]
  44. Nikolaeva, M.; Mikenas, T.; Matsko, M.; Zakharov, V. Effect of AlEt3 and an External Donor on the Distribution of Active Sites According to Their Stereospecificity in Propylene Polymerization over TiCl4/MgCl2 Catalysts with Different Titanium Content. Macromol. Chem. Phys. 2016, 217, 1384–1395. [Google Scholar] [CrossRef]
  45. Vittoria, A.; Antinucci, G.; Zaccaria, F.; Cipullo, R.; Busico, V. Monitoring the Kinetics of Internal Donor Clean-up from Ziegler-Natta Catalytic Surfaces: An Integrated Experimental and Computational Study. J. Phys. Chem. C 2020, 124, 14245–14252. [Google Scholar] [CrossRef]
  46. Khatri, V.; Sahoo, U.; Kaur, S.; Rani, R.; Singh, G.; Kapur, G.S.; Kashyap, H. Control of Ziegler-Natta Catalyst Activity by the Structural Design of Alkoxysilane-Based External Donors. New J. Chem. 2020, 44, 6845–6852. [Google Scholar] [CrossRef]
  47. Raj, K.; Gupta, V.; Vanka, K. The Potential Role of Lewis Acid–Base Adducts in Enhancing Stereoselectivity in Ziegler-Natta Catalysts: A DFT Study. J. Phys. Chem. C 2023, 127, 7220–7229. [Google Scholar] [CrossRef]
  48. US9156927B2; Catalyst Component for Olefin Polymerization Reaction and Catalyst Comprising Same. BASF Corporation: Ludwigshafen, Germany, 2015.
  49. Li, C.; Feng, H.; Liu, H.; Zhuang, Z.; Zhou, J.; Liu, D. Effect of Dioldibenzoate Isomers as Electron Donors on the Performances of Ziegler-Natta Polypropylene Catalysts: Experiments and Calculations. J. Phys. Chem. C 2023, 127, 2294–2302. [Google Scholar] [CrossRef]
  50. GB 3682.1–2018; Plastics, Thermoplastic—Determination of Melt Mass Flow Rate (MFR) and Melt Volume Flow Rate (MVR), Part 1: Standard Method. China Standard Press: Beijing, China, 2018.
  51. Adamo, C.; Barone, V. Toward Reliable Density Functional Methods without Adjustable Parameters: The PBE0 Model. J. Chem. Phys. 1999, 110, 6158–6170. [Google Scholar] [CrossRef]
  52. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef]
  53. Grimme, S. Semiempirical GGA-type Density Functional Constructed with a Long-Range Dispersion Correction. J. Comput. Chem. 2006, 27, 1787–1799. [Google Scholar] [CrossRef]
  54. Grimme, S. Supramolecular Binding Thermodynamics by Dispersion-Corrected Density Functional Theory. Chem.-Eur. J. 2012, 18, 9955. [Google Scholar] [CrossRef]
  55. Lu, T.; Chen, X. Shermo: A General Code for Calculating Molecular Thermochemistry Properties. Comput. Theor. Chem. 2021, 1200, 113249. [Google Scholar] [CrossRef]
  56. Chadwick, J.C.; Heere, J.J.R.; Sudmeijer, O. Factors Influencing Chain Transfer with Monomer and with Hydrogen in Propene Polymerization Using MgCl2-Supported Ziegler-Natta Catalysts. Macromol. Chem. Phys. 2000, 201, 1846–1852. [Google Scholar] [CrossRef]
  57. Kojoh, S.; Tsutsui, T.; Kashiwa, N.; Itoh, M.; Mizuno, A. Effect of An External Donor upon Chain-transfer Reactions in Propylene Polymerization with a MgCl2-Supported Titanium Catalyst System. Polymer 1998, 39, 6309–6313. [Google Scholar] [CrossRef]
  58. Cavallo, L.; Guerra, G. A Density Functional and Molecular Mechanics Study Of β-Hydrogen Transfer in Homogeneous Ziegler-Natta Catalysis. Macromolecules 1996, 29, 2729–2737. [Google Scholar] [CrossRef]
  59. Talarico, G.; Budzelaar, P.H.M. Variability of Chain Transfer to Monomer Step in Olefin Polymerization. Organometallics 2008, 27, 4098–4107. [Google Scholar] [CrossRef]
  60. Busico, V.; Cipullo, R.; Pellecchia, R.; Ronca, S.; Roviello, G.; Talarico, G. Design of Stereoselective Ziegler-Natta Propene Polymerization Catalysts. Proc. Natl. Acad. Sci. USA 2006, 103, 15321–15326. [Google Scholar] [CrossRef] [PubMed]
  61. Gibson, V.; Spitzmesser, S.K. Advances in Non-Metallocene Olefin Polymerization Catalysis. Chem. Rev. 2003, 103, 283–315. [Google Scholar] [CrossRef] [PubMed]
  62. Yu, Y.; Fu, Z.; Fan, Z. Chain Transfer Reactions of Propylene Polymerization Catalyzed by AlEt3 Activated TiCl4/MgCl2 Catalyst Under very Low Monomer Addition Rate. J. Mol. Catal. A Chem. 2012, 363–364, 134–139. [Google Scholar] [CrossRef]
  63. Kissin, Y.V.; Rishina, L.A.; Vizen, E.I. Hydrogen Effects in Propylene Polymerization Reactions with Titanium-Based Ziegler-Natta Catalysts. II. Mechanism of the Chain-Transfer Reaction. J. Polym. Sci. Pol. Chem. 2002, 40, 1899–1911. [Google Scholar] [CrossRef]
Figure 1. The optimized DFT geometries for the HDDB stereoisomers in the free state (red: oxygen; grey: carbon; hydrogen is omitted).
Figure 1. The optimized DFT geometries for the HDDB stereoisomers in the free state (red: oxygen; grey: carbon; hydrogen is omitted).
Polymers 16 00559 g001
Figure 2. The trans-a adsorption modes of HDDB stereoisomers and TiCl4 on the (110)-facet of MgCl2 (red: oxygen; grey: carbon; green: chlorine; yellow: magnesium; hydrogen is omitted).
Figure 2. The trans-a adsorption modes of HDDB stereoisomers and TiCl4 on the (110)-facet of MgCl2 (red: oxygen; grey: carbon; green: chlorine; yellow: magnesium; hydrogen is omitted).
Polymers 16 00559 g002
Figure 3. Structures of HDDB–catalyst–propylene complexes (red: oxygen; grey: carbon; green: chlorine; yellow: magnesium; hydrogen is omitted).
Figure 3. Structures of HDDB–catalyst–propylene complexes (red: oxygen; grey: carbon; green: chlorine; yellow: magnesium; hydrogen is omitted).
Polymers 16 00559 g003
Figure 4. The transition state structures of HDDB–catalyst–propylene (si-insertion) complexes (red: oxygen; grey: carbon; green: chlorine; yellow: magnesium; hydrogen is omitted).
Figure 4. The transition state structures of HDDB–catalyst–propylene (si-insertion) complexes (red: oxygen; grey: carbon; green: chlorine; yellow: magnesium; hydrogen is omitted).
Polymers 16 00559 g004
Figure 5. The changes of macroscopic activity for the active centers with (a) PDDB and (b) HDDB ID when increasing the stereoisomer content. The residuals are distributed according to the relative content percent of the RS or SS stereoisomers. The slope represents the catalytic activity for the active centers per unit.
Figure 5. The changes of macroscopic activity for the active centers with (a) PDDB and (b) HDDB ID when increasing the stereoisomer content. The residuals are distributed according to the relative content percent of the RS or SS stereoisomers. The slope represents the catalytic activity for the active centers per unit.
Polymers 16 00559 g005
Figure 6. The transition state structures of HDDB stereoisomer catalysts in the hydrogen transfer reaction (red: oxygen; grey: carbon; green: chlorine; yellow: magnesium; hydrogen is omitted).
Figure 6. The transition state structures of HDDB stereoisomer catalysts in the hydrogen transfer reaction (red: oxygen; grey: carbon; green: chlorine; yellow: magnesium; hydrogen is omitted).
Polymers 16 00559 g006
Table 1. The polymerization results for the ZN catalysts with different HDDB stereoisomers as IDs.
Table 1. The polymerization results for the ZN catalysts with different HDDB stereoisomers as IDs.
EntryMesomer (wt/%)ID c
(wt/%)
AC
(kgPP/gcat)
I.I.
(%)
MFR
(g/10 min)
MnMwPD
Compound aCatalyst b
197.199.314.954.198.51.869,509482,9466.9
263.094.613.252.698.42.262,469436,2877.0
348.187.610.347.197.53.557,078424,1127.4
434.680.28.644.797.14.654,724414,1437.6
525.068.27.440.296.56.053,043406,6317.7
616.352.55.132.695.47.348,547382,9677.9
71.415.92.420.492.39.447,726380,2918.0
a relative content of meso-HDDB in the same weight HDDB compound added in preparation of catalyst. b relative content of meso-HDDB in catalyst. c content of total HDDB (including mesomer and racemate) in catalyst.
Table 2. Comparison between the calculated stereoselectivity for the transition state in the si-/re-insertion reaction of propylene, and the isotactic index obtained from the experiments.
Table 2. Comparison between the calculated stereoselectivity for the transition state in the si-/re-insertion reaction of propylene, and the isotactic index obtained from the experiments.
IDTS (re) − TS (si) a (kcal/mol)Calculated Stereoselectivity (%)I.I. (%, Experiment)
RS-PDDB3.899.598.9
SS-PDDB0.875.591.1
RS-HDDB3.299.098.7
SS-HDDB1.182.791.2
a The data of PDDB have little difference to Ref. [49] because of the quasi-RRHO model used in Gibbs free energy calculation.
Table 3. Comparison between the calculated activation energy (kcal/mol) of si-insertion for the PDDB and HDDB isomers, and the calculated relative activity of SS/RS.
Table 3. Comparison between the calculated activation energy (kcal/mol) of si-insertion for the PDDB and HDDB isomers, and the calculated relative activity of SS/RS.
IDActivation Energy of si-InsertionCalculated Relative Activity of SS/RS
RS-PDDB11.31.2
SS-PDDB11.2
RS-HDDB11.83.2
SS-HDDB11.0
Table 4. The relative ratio of the activity regression values of individual active centers.
Table 4. The relative ratio of the activity regression values of individual active centers.
IDASSARSRegression CoefficientASS/ARS
PDDB6.34.90.99921.3
HDDB9.73.70.99942.6
Table 5. The adsorption energy (kcal/mol) of internal electron donors on the (110)-facet of MgCl2 and TiCl4/MgCl2.
Table 5. The adsorption energy (kcal/mol) of internal electron donors on the (110)-facet of MgCl2 and TiCl4/MgCl2.
Adsorption StructurePDDB aHDDB
MgCl2/2RS96.397.9
MgCl2/2SS88.587.7
TiCl4/MgCl2/2RS100.5103.2
TiCl4/MgCl2/2SS94.494.5
a The data of PDDB have little difference to reference [49] because of the different models of processing frequency used in the Gibbs free energy calculation (calculated at 298.15 K and 1 atm).
Table 6. Activation energy (kcal/mol) of hydrogen transfer reaction.
Table 6. Activation energy (kcal/mol) of hydrogen transfer reaction.
IDGHGH − ∆Gp
RS-PDDB16.565.22
SS-PDDB16.365.16
RS-HDDB16.925.09
SS-HDDB15.964.94
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

Feng, H.; Li, C.; Zhou, J.; Zhang, X.; Tang, S.; Xu, X.; Song, Z. Experimental and Theoretical Insights into the Effect of Dioldibenzoate Isomers on the Performance of Polypropylene Catalysts. Polymers 2024, 16, 559. https://doi.org/10.3390/polym16040559

AMA Style

Feng H, Li C, Zhou J, Zhang X, Tang S, Xu X, Song Z. Experimental and Theoretical Insights into the Effect of Dioldibenzoate Isomers on the Performance of Polypropylene Catalysts. Polymers. 2024; 16(4):559. https://doi.org/10.3390/polym16040559

Chicago/Turabian Style

Feng, Huasheng, Changxiu Li, Junling Zhou, Xiaofan Zhang, Shuxuan Tang, Xiangya Xu, and Zhihui Song. 2024. "Experimental and Theoretical Insights into the Effect of Dioldibenzoate Isomers on the Performance of Polypropylene Catalysts" Polymers 16, no. 4: 559. https://doi.org/10.3390/polym16040559

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