Impact of Reaction System Turbulence on the Dispersity and Activity of Heterogeneous Ziegler–Natta Catalytic Systems for Polydiene Production: Insights from Kinetic and CFD Analyses
Abstract
1. Introduction
1.1. Modern State of Research in the Field of Coordination (Co)Polymerization of Dienes in the Presence of Ziegler–Natta Catalytic Systems
1.1.1. Ziegler–Natta Catalytic Systems and the Stereoregulating Ability of Their Active Sites in Coordination (Co)Polymerization of Dienes
1.1.2. Cocatalysts in Ziegler–Natta Catalytic Systems and Their Role and Influence on Coordination (Co)Polymerization of Dienes
1.1.3. Multi-Site Nature of Ziegler–Natta Catalytic Systems and Methods for Determining the Structure and Number of Types of Their Active Sites
1.1.4. Influence of Diene Structure on Its Reactivity in Coordination (Co)Polymerization
1.1.5. Methods for Studying the Kinetics of Coordination (Co)Polymerization of Dienes and the Molecular Characteristics of the Resulting (Co)Polymers
1.1.6. Influence of Mechanical Effects on the Kinetics of Coordination (Co)Polymerization of Dienes and the Molecular Characteristics of the Resulting (Co)Polymers
1.1.7. Butadiene-Isoprene Copolymer (The Most Demanded Product of Coordination Copolymerization of Two Dienes) and Its Structure, Properties, and Applications
1.2. The Aims and Objectives of This Study
- 1.
- Development of a kinetic model of the butadiene-isoprene copolymerization, taking into account the multi-site nature of the TiCl4 + Al(i-Bu)3 catalytic system and the temporal dynamics of changes in concentrations of its active sites of different types. Verification of the kinetic model based on experimental data of the butadiene-isoprene copolymerization kinetics.
- 2.
- Development of a CFD model for the breakage of TiCl4 + Al(i-Bu)3 particles in a tubular turbulent apparatus of a diffuser–confuser design, which allows the description of changes in the distribution of catalytic system particles by equivalent radius as a result of their breakage. Verification of the CFD model based on experimental data on the distributions of TiCl4 + Al(i-Bu)3 catalytic system particles by equivalent radius before and after breakage.
- 3.
- Quantitative description of the relationship between the turbulence conditions of the reaction system, sizes of TiCl4 + Al(i-Bu)3 catalytic system particles after turbulent exposure, concentration of active sites, kinetics of butadiene-isoprene copolymerization, and molecular weight characteristics of the copolymer. Development of an integrated model of the butadiene-isoprene copolymerization that combines CFD model calculations and kinetic model calculations using auxiliary equations that link the output parameters of the CFD model with the input parameters of the kinetic model.
- 4.
- Conducting computational experiments. Accumulating data on the relationship between reaction system turbulence conditions, particle sizes TiCl4 + Al(i-Bu)3 of catalytic system after turbulent exposure, concentration of active sites, copolymerization kinetics, and copolymer molecular characteristics as predicted by the integrated model under various geometric parameters of the tubular turbulent apparatus and linear feed rates of the reaction system.
- 5.
- Analyzing the influence of turbulence conditions of the reaction system—average and maximum turbulent kinetic energy, and average residence time of TiCl4 + Al(i-Bu)3 catalyst particles in the tubular turbulent apparatus—on their dispersity, the copolymerization rate, and the molecular weight characteristics of the copolymer.
- 6.
- Formulating scientifically justified recommendations for controlling reaction system turbulence conditions to purposefully synthesize copolymers with desired molecular weight characteristics in the presence of heterogeneous Ziegler–Natta catalytic systems.
2. Materials and Methods
2.1. Experimental Data Used for the Development and Verification of the CFD Model of Breakage of TiCl4 + Al(i-Bu)3 Catalytic System Particles and the Kinetic Model of Butadiene-Isoprene Copolymerization in the Presence of TiCl4 + Al(i-Bu)3 Catalytic System
2.2. CFD Model of Breakage of Particles of the TiCl4 + Al(i-Bu)3 Catalytic System
2.3. Kinetic Model of Butadiene-Isoprene Copolymerization in the Presence of the TiCl4 + Al(i-Bu)3 Catalytic System
3. Results
3.1. Heuristic Solution of the Inverse Kinetic Problem of Butadiene-Isoprene Copolymerization in the Presence of TiCl4 + Al(i-Bu)3 Catalytic System
- 1.
- Chain propagation:
- 2.
- Chain transfer to monomer:
- 3.
- Chain transfer to cocatalyst:
- 4.
- Deactivation of active sites:
- 5.
- Interconversion of active sites of different types:
- 1.
- Large computational volume, since the size of the solution search space is proportional to dX, where d is a certain constant (the conditional range within which parameter values are sought) and X is the number of unknowns. Large computational volume leads to significant calculation time.
- 2.
- High solution uncertainty. Due to the incomplete nature of experimental data regarding the process, different sets of rate constant values can enable the kinetic model to satisfactorily describe the experimental data. This is particularly observed because the concentrations of intermediate unstable compounds often cannot be measured. A classic example is a chemical process at equilibrium. If time-dependent concentrations during the system’s transition to equilibrium are not experimentally measured, then only the equilibrium constant—the ratio of the forward and reverse reaction rate constants—can be determined precisely, but not their individual values. In other words, an infinite number of individual forward and reverse reaction rate constant pairs can equally well describe the chemical process at equilibrium. Unfortunately, the number of measurable process regularities usually grows more slowly than the number of unknown rate constants in the process as its mechanism complexity increases.
3.2. Assessment of the Adequacy of the Kinetic Model and the Accuracy of Its Parameter Estimation
3.3. Algorithm for Conducting Computational Experiments
- 1.
- Based on the CFD model of breakage of TiCl4 + Al(i-Bu)3 catalytic system particles (the equations of which are presented in Table 1), calculations of the particle size distribution by equivalent radius dq/dr after exiting the tubular turbulent apparatus with the diffuser–confuser design were performed using the Fluent module of the ANSYS Workbench 17.1 platform. The composition of the reaction system in the calculations corresponded to the case of butadiene homopolymerization (q = 1, the extreme right point of the dependence ). For each new turbulence condition of the reaction system arising from changes in the geometric parameters of the tubular turbulent apparatus and the feed velocity of the reaction system, a separate CFD model calculation was carried out. Each such calculation yielded its own particle size distribution dq/dr.
- 2.
- Using Equation (10), the concentrations (at q = 1) were calculated for each obtained dq/dr distribution.
- 3.
- For each obtained concentration value (at q = 1), new parameters a and b of the multiplier were adjusted so that the discrepancy between the value obtained at step 2 of this algorithm and the value obtained from Equation (15) at q = 1 was minimized. This minimization was performed using the optimization algorithm embedded in the FindMinimum operator of the Mathematica 12.0 computer algebra system. Since for each concentration, two parameters (a and b) were sought based on a single value , an additional constraint was required to identify a and b. This condition was the minimal deviation between the new values a and b and the original values a0 and b0 from Equation (15). Therefore, at this stage of the algorithm, the following objective function was minimized to determine a and b:
- 4.
- For each obtained pair of a and b values, Equation (15) was used to calculate μ at q = 0.76 and q = 0.85. These monomer mixture compositions correspond to the production of BIR grades BIR-24 and BIR-15, respectively, and thus, these q values were chosen for the computational experiments.
- 5.
- For each obtained value , the initial concentrations of active sites of all types were calculated. From the equation
- 6.
- Using the kinetic model equations of butadiene-isoprene copolymerization in the presence of the TiCl4 + Al(i-Bu)3 catalytic system, presented in SI1, the time dependencies of copolymer yield U, number-average molecular weight Mn, and weight-average molecular weight Mw of the copolymer were calculated for each obtained set of parameter values
3.4. Results of Computational Experiments
- 1.
- Ratio of diffuser diameter dd to confuser diameter dc at constant diffuser diameter
- 2.
- Feed velocity v of the reaction system into the tubular turbulent apparatus
- 3.
- Diffuser opening angle α
- 4.
- Ratio of section length L to diffuser diameter dd at constant diffuser diameter
- 5.
- Number of sections of the tubular turbulent apparatus
- 6.
- Ratio of section length L to diffuser diameter dd at constant section length and constant ratio of confuser to diffuser diameter
4. Discussion
- 1.
- The kinetic model did not allow for tracking the influence of copolymerization conditions on the copolymer composition. Technically, the kinetic model permits calculating the copolymer composition (mole fraction of butadiene units in the copolymer, Q) using the equation:
- 2.
- The kinetic model equations do not allow for the calculation of the isomeric composition of copolymer units and their stereoregularity. This was done intentionally to avoid complicating the kinetic model equations. If such calculations become necessary, the kinetic model equations can be modified. Each chain propagation reaction can be split into separate reactions, each leading to the formation of a monomer unit with a specific structure. The rate constants of these reactions can be identified to describe the experimental isomeric composition of the copolymer units. The fractions of triads of units in the copolymer composition can be calculated based on the isomeric composition of its units within Markov statistics.
- 3.
- The kinetic model is formulated in a pseudo-homogeneous approximation, i.e., it does not take into account the diffusion stage of monomer molecules to the surface of catalytic system particles and inside their pores. Nevertheless, the kinetic model was able to describe the experimental data on the kinetics of butadiene-isoprene copolymerization, which is sufficient justification for the validity of using this pseudo-homogeneous approximation. In practice, this means that if monomer diffusion processes influence the copolymerization rate, then the obtained values of the reaction rate constants should be regarded as apparent values.
- 1.
- The verification of the CFD model was carried out on a very limited set of experimental data. Essentially, this consists of two distributions of particle size of the TiCl4 + Al(i-Bu)3 catalytic system by the equivalent radius dq/dr (before and after passing the reaction system through the tubular turbulent apparatus). The choice of equations for calculating the particle breakage frequency g(V) and the probability density function for particle breakage from V depended on the shape of these distributions. After testing various equations, it was established that the available experimental distributions dq/dr are most accurately described by the equations from study [52] (these equations are shown in Table 1). However, there are many CFD models for the breakage of solid particles in liquids [51,60,61,62]. With the accumulation of more experimental data on the influence of turbulence on dq/dr, the choice of the CFD model that most adequately describes this influence may change.
- 2.
- In the calculations, catalytic system particles are represented very simply as spheres, and the surface area considered does not take into account the surface area of their pores. This allowed us to demonstrate that the specific surface area of the particles, which determines the catalytic system activity, is inversely proportional to their equivalent radius r. This generally reproduces the experimentally observed trend (the smaller the particles, the more active they should be). However, quantitatively, taking into account active sites in the pores of catalytic system particles, the activity of these particles may be proportional to 1/rn, where n ≠ 1.
- 3.
- The cause of the destruction of solid catalytic system particles should be the mechanical stresses arising in them, caused by shear stresses in the liquid phase (reaction system). Mechanical stresses do not explicitly appear in the developed CFD model. Their influence on the particle breakage rate in the developed CFD model is effectively taken into account by the values of the constants and and the particle shape factor in the population balance equation. The values of C, C1, and C2 should themselves depend on the properties of the reaction system to avoid paradoxes. For example, with an increase in the viscosity of the reaction system, its turbulence should decrease (turbulence arises when the inertial forces of laminar flow tubes exceed the viscous friction forces between these flow tubes), but at the same time, the shear stresses in the liquid increase, which should increase the breakage rate of catalytic system particles. That is, with an increase in the viscosity of the reaction system (for example, with an increase in copolymer yield), the turbulent kinetic energy and its dissipation rate should decrease, while the particle breakage rate of the catalytic system should increase. This statement is based on experimental observation—in a previous study [39], it was shown that if, during butadiene polymerization in the presence of the TiCl4 + Al(i-Bu)3 catalytic system, the reaction system is passed through the tubular turbulent apparatus but not immediately, but after a certain polymerization duration, the catalytic system activity increases more strongly, and the longer the polymerization time, the stronger the increase. Unfortunately, this effect cannot be described by the developed CFD model with constant values of the constants FS, C1, and C2. The values of C, C1, and C2 are not universal.
- 1.
- An original method for solving the inverse kinetic problem for the butadiene-isoprene copolymerization was proposed, based on a heuristic decomposition of the solution. This method allowed the determination of the rate constants of the reactions in the considered process at a relatively low computational cost. Since each reaction rate constant was determined based on the experimental data on which it had the greatest influence, the obtained solution has a low degree of uncertainty. Equations were obtained linking the apparent values of the reaction rate constants (the values of the reaction rate constants within the framework of the kinetic model of single-site homopolymerization that allow this kinetic model to satisfactorily reproduce the experimental kinetic patterns of multi-site copolymerization) to their true values for multi-site copolymerization. The presented heuristic decomposition method for solving the inverse kinetic problem is a universal method by which reaction rate constants for other multi-site copolymerizations can be found. The preliminary description of experimental data using a single-site homopolymerization kinetic model is a convenient method for finding apparent reaction rate constants, which can then be used as initial approximations for the true reaction rate constants of multi-site copolymerization.
- 2.
- The influence of turbulence on the distribution of catalytic system particles by equivalent radius as a result of their breakage, the activity of the catalytic system, and the kinetic patterns of multi-site copolymerization in the presence of this catalytic system for one specific research object—the butadiene-isoprene copolymerization in the presence of the TiCl4 + Al(i-Bu)3 catalytic system—was theoretically described. The cause-and-effect relationships describing this influence are presented in Table 4. Despite the aforementioned limitations of the kinetic and CFD models, the forecasts obtained using this theoretical framework can be considered, as it is grounded in fundamental physical and chemical laws. In subsequent studies, this theoretical description may be extended to other research objects—(co)polymerization processes in the presence of heterogeneous catalytic systems—enabling a more detailed determination of the role of reaction system turbulence as a controlling factor of the particle size distribution and activity of heterogeneous Ziegler–Natta catalytic systems. Taking into account the specifics of each research object, individual parameters and equations in this theoretical framework may accordingly be modified. The limitations identified in the developed theoretical description will aid in understanding how it should be adapted when generalized.
- 3.
- It has been established that the influence of turbulence on the distribution of catalytic system particles by equivalent radius and on their activity is determined by two factors: the duration of turbulent exposure to the particles and the intensity of this exposure (which, in the present study, was characterized by the turbulence kinetic energy). For the research object considered, turbulence intensity proved to be a more significant factor than the exposure duration. Considering that the turbulent impact on catalytic system particles in the examined copolymerization process can be implemented without significant modifications or cost increases to the process technology, a tubular turbulent apparatus of a diffuser–confuser design is recommended to achieve this impact. Based on computational experiment results, to enhance the intensity of catalytic particle breakage and thereby increase catalytic activity, it is primarily recommended to increase the ratio of the diffuser diameter dd to the confuser diameter dc while maintaining a constant diffuser diameter, as well as to increase the feed velocity of the reaction system into the tubular turbulent apparatus. The stronger the catalytic system particles are dispersed, the more the molecular weights of the copolymer decrease (this effect not only follows from calculations but is also confirmed by experimental data shown in Figure 4). Consequently, to obtain a copolymer with high molecular weight, turbulent impact on catalytic system particles should be avoided.
5. Conclusions
- 1.
- A kinetic model of the butadiene-isoprene copolymerization has been developed, taking into account the multi-site nature of the TiCl4 + Al(i-Bu)3 catalytic system and the time changes in concentrations of different types of active sites. For the first time, a semi-analytical method for solving the inverse kinetic problem has been developed for the class of multi-site coordination copolymerization processes. The method is based on decomposing the solution of the inverse kinetic problem into four stages: determination of the chain propagation reaction rate constants kp (Stage I), interconversion of active sites (Stage II), deactivation of active sites kt (Stage III), and chain transfer to monomer molecules kM (Stage IV). As a result of solving the inverse kinetic problem, 32 nonzero values of the reaction rate constants for the considered copolymerization at 25 °C were identified.
- 2.
- A CFD model of breakage of TiCl4 + Al(i-Bu)3 catalytic system particles in a tubular turbulent apparatus of a diffuser–confuser design at the reaction system formation stage was developed using the Fluent module of ANSYS Workbench 17.1. This CFD model involves the numerical solution of the conservation equations for the mass, momentum, and energy of the reaction system, the K-ε turbulence model equations, and population balance equations for the catalytic system particles. The developed CFD model can be applied to describe the breakage of particles of other catalytic systems, provided that new parameters of the population balance equations are identified.
- 3.
- The integrated model of the butadiene-isoprene copolymerization has been developed, combining calculations from the CFD model and the kinetic model through auxiliary equations that link the output parameters of the CFD model with the input parameters of the kinetic model. Within these auxiliary equations, the concentrations of active sites are considered proportional to the specific surface area of the catalytic system particles, which is calculated based on the particle size distribution of the TiCl4 + Al(i-Bu)3 catalytic system by equivalent radius. The dependence of the concentrations of different types of active sites on the monomer mixture composition q is described using a kinetic model of active site formation via a two-stage mechanism (stage 1—adsorption of monomer molecules on adsorption sites; stage 2—formation of the Ti–C bond).
- 4.
- Using the integrated model of butadiene and isoprene copolymerization, it was established that the greater the turbulent kinetic energy K (attributable to the design features of the tubular turbulent apparatus or the magnitude of the feed velocity of the reaction system), the higher the dispersity of the TiCl4 + Al(i-Bu)3 catalytic system particles, the higher the concentration of active sites of this catalytic system, the higher the copolymerization rate, and the lower the number-average and weight-average molecular weights of the copolymer.
- 5.
- It was established for the first time that the rates of butadiene and isoprene copolymerization, and consequently the activity values of the TiCl4-Al(i-Bu)3 catalytic system, correlate with the average or maximum turbulent kinetic energy of the reaction system in the tubular turbulent apparatus, but do not correlate with the mean residence time of the reaction system in the tubular turbulent apparatus. In other words, this study demonstrates for the first time that to achieve maximum efficiency in increasing the activity of the TiCl4-Al(i-Bu)3 catalytic system through turbulence exposure of the reaction system, it is necessary to increase the level of turbulent kinetic energy in the tubular turbulent apparatus rather than its volume.
- 6.
- Based on the results of computational experiments, it was established that to achieve more intensive breakage of catalytic system particles and increase its activity, it is primarily recommended to increase the ratio of the diffuser diameter dd to the confuser diameter dc while maintaining a constant diffuser diameter, along with increasing the feed velocity of the reaction system into the tubular turbulent apparatus.
Supplementary Materials
Author Contributions
Funding
Data Availability Statement
Acknowledgments
Conflicts of Interest
Abbreviations
BR | butadiene rubber |
CFD | computational fluid dynamics |
BIR | butadiene-isoprene rubber |
IR | isoprene rubber |
i-Bu | isobutyl |
NMR | nuclear magnetic resonance |
References
- Nath, C.D.; Shiono, T.; Ikeda, T. Copolymerization of 1,3-Butadiene and Isoprene with Cobalt Dichloride/Methylaluminoxane in the Presence of Triphenylphosphine. J. Polym. Sci. Part A Polym. Chem. 2002, 40, 3086–3092. [Google Scholar] [CrossRef]
- Li, P.; Liu, K.; Fu, Z.; Wang, Z.; Hua, H. Preparation of Butadiene-Isoprene Copolymer with High Vinyl Contents by Al(OPhCH3)(i-Bu)2/MoO2Cl2·TNPP. Polymers 2019, 11, 527. [Google Scholar] [CrossRef]
- Cornils, B.; Herrmann, W.A.; Xu, J.-H.; Zanthoff, H.-W. (Eds.) Catalysis from A to Z: A Concise Encyclopedia, 5th ed.; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2020. [Google Scholar] [CrossRef]
- Ricci, G.; Pampaloni, G.; Sommazzi, A.; Masi, F. Dienes Polymerization: Where We Are and What Lies Ahead. Macromolecules 2021, 54, 5879–5914. [Google Scholar] [CrossRef]
- Kissin, Y. Isospecific Polymerization of Olefins with Heterogeneous Ziegler-Natta Catalysts; Spinger: New York, NY, USA, 1985; 439p. [Google Scholar]
- Niu, O.; Zhang, J.; He, A. Trans-1,4-stereospecific copolymerization of isoprene and butadiene catalyzed by TiCl4/MgCl2 Ziegler-Natta catalyst: III Effect of alkylaluminium on monomer reactivity ratios. Polym. Int. 2021, 70, 1449–1455. [Google Scholar] [CrossRef]
- He, A.; Huang, B.; Jiao, S.; Hu, Y. Synthesis of a High-Trans 1,4-Butadiene/Isoprene Copolymers with Supported Titanium Catalysts. J. Appl. Polym. Sci. 2003, 89, 1800–1807. [Google Scholar] [CrossRef]
- Peng, W.; Xie, J.; Zhang, J.; Yang, X.; He, A. Isoprene polymerizations catalyzed by TiCl4/MgCl2 type Ziegler-Natta catalysts with different titanium contents. Mol. Catal. 2020, 494, 111110. [Google Scholar] [CrossRef]
- Nifant’ev, I.E.; Tavtorkin, A.N.; Korchagina, S.A.; Gavrilenko, I.F.; Glebova, N.N.; Kostitsyna, N.N.; Yakovlev, V.A.; Bondarenko, G.N.; Filatova, M.P. Neodymium tris-diarylphosphates: Systematic study of the structure–reactivity relationship in butadiene and isoprene polymerization. Appl. Catal. A Gen. 2014, 478, 219–227. [Google Scholar] [CrossRef]
- Huang, J.-M.; Yao, C.-G.; Li, S.-H.; Cui, D.-M. Yttrium-catalyzed cis-1,4-Selective Polymerization of 2-(4-Halophenyl)-1,3-Butadienes and Their Copolymerization with Isoprene. Chin. J. Polym. Sci. 2021, 39, 309–315. [Google Scholar] [CrossRef]
- Li, W.; Nie, H.; Shao, H.; Ren, H.; He, A. Synthesis, chain structures and phase morphologies of trans-1,4-poly(butadiene-coisoprene) copolymers. Polymer 2018, 156, 148–161. [Google Scholar] [CrossRef]
- Farias, F.R.L.; Vasconcelos, M.K.; Brandao, A.L.T. Influence of number of catalyst sites in 1,3-butadiene solution polymerizations catalyzed by titanium tetrachloride. Eng. Rep. 2020, 3, e12333. [Google Scholar] [CrossRef]
- Liu, B.; Wang, X.; Pan, Y.; Lin, F.; Wu, C.; Qu, J.; Luo, Y.; Cui, D. Unprecedented 3,4-Isoprene and cis-1,4-Butadiene Copolymers with Controlled Sequence Distribution by Single Yttrium Cationic Species. Macromolecules 2014, 47, 8524–8530. [Google Scholar] [CrossRef]
- Niu, Q.; Zhang, J.; Peng, W.; Fan, Z.; He, A. Effect of alkylaluminium on the regio- and stereoselectivity in copolymerization of isoprene and butadiene using TiCl4/MgCl2 type Ziegler-Natta catalyst. Mol. Catal. 2019, 471, 1–8. [Google Scholar] [CrossRef]
- Zhang, J.; Peng, W.; He, A. Influence of alkylaluminium on the copolymerization of isoprene and butadiene with supported Ziegler-Natta catalyst. Polymer 2020, 203, 122766. [Google Scholar] [CrossRef]
- Liu, J.; Zhao, A.; Zhou, R.; Yang, X.; He, A. Effect of the Degree of Cluster TiCl4 on the Polymerization of Isoprene over Ziegler-Natta Catalysts: A DFT Study. J. Phys. Chem. C 2024, 128, 6658–6671. [Google Scholar] [CrossRef]
- Taibulatov, P.A.; Mingaleev, V.Z.; Zakharov, V.P.; Ionova, I.A.; Monakov, Y.B. Kinetic inhomogeneity of copolymerization sites of butadiene and isoprene on titanium catalyst. Pol. Sci. Ser. B. 2010, 52, 450–458. [Google Scholar] [CrossRef]
- Yanborisov, E.V.; Spivak, S.I.; Yanborisov, V.M.; Monakov, Y.B. Calculation of molecular weight distributions of polymers synthesized on multisite catalytic systems. Dokl. Chem. 2010, 432, 148–150. [Google Scholar] [CrossRef]
- Leicht, H.; Göttker-Schnetmann, I.; Mecking, S. Stereoselective Copolymerization of Butadiene and Functionalized 1,3-Dienes. ACS Macro Lett. 2016, 5, 777–780. [Google Scholar] [CrossRef]
- Niu, Q.; Li, W.; Liu, X.; Wang, R.; He, A. Trans-1,4- stereospecific copolymerization of isoprene and butadiene catalyzed by TiCl4/MgCl2 type Ziegler-Natta catalyst II. Copolymerization Kinetics and Mechanism. Polymer 2018, 143, 173–183. [Google Scholar] [CrossRef]
- Zhang, Q.F.; Jiang, X.B.; He, A.H. Synthesis and Characterization of trans-1,4-Butadiene/Isoprene Copolymers: Determination of Monomer Reactivity Ratios and Temperature Dependence. Chin. J. Polym. Sci. 2014, 32, 1068–1076. [Google Scholar] [CrossRef]
- Wang, F.; Zhang, M.; Liu, H.; Hu, Y.; Zhang, X. Randomly coordinative chain transfer copolymerization of 1,3-butadiene and isoprene: A highly atom-economic way for accessing butadiene/isoprene rubber. Ind. Eng. Chem. Res. 2020, 59, 10754–10762. [Google Scholar] [CrossRef]
- Ren, W.; You, F.; Zhai, J.; Kang, X.; So, Y.-M.; Shi, X. Living (Co)Polymerization of Isoprene and Butadiene with Unparallel Stereoselectivity Catalyzed by Single Rare-Earth Metal Cationic Species. Macromolecules 2022, 55, 10640–10650. [Google Scholar] [CrossRef]
- Leicht, H.; Göttker-Schnetmann, I.; Mecking, S. Stereoselective Copolymerization of Butadiene and Functionalized 1,3-Dienes with Neodymium-Based Catalysts. Macromolecules 2017, 50, 8464–8468. [Google Scholar] [CrossRef]
- Li, W.; Peng, W.; Ren, S.; He, A. Synthesis and Characterization of trans-1,4-Poly(butadiene-co-isoprene) Rubbers (TBIR) with Different Fraction and Chain Sequence Distribution and Its Influence on the Properties of Natural Rubber/TBIR/Carbon Black Composites. Ind. Eng. Chem. Res. 2019, 58, 10609–10617. [Google Scholar] [CrossRef]
- Shao, H.; Ren, S.; Wang, R.; He, A. Temperature rising elution fractionation and fraction characterization of Trans-1, 4-poly (isoprene-co-butadiene). Polymer 2020, 186, 122015. [Google Scholar] [CrossRef]
- Doerr, A.M.; Burroughs, J.M.; Gitter, S.R.; Yang, X.; Boydston, A.J.; Long, B.K. Advances in Polymerizations Modulated by External Stimuli. ACS Catal. 2020, 10, 14457–14515. [Google Scholar] [CrossRef]
- Zhou, Y.-N.; Li, J.-J.; Wu, Y.-Y.; Luo, Z.-H. Role of External Field in Polymerization: Mechanism and Kinetics. Chem. Rev. 2020, 120, 2950–3048. [Google Scholar] [CrossRef] [PubMed]
- Ponomarenko, A.T.; Tameev, A.R.; Shevchenko, V.G. Action of Mechanical Forces on Polymerization and Polymers. Polymers 2022, 14, 604. [Google Scholar] [CrossRef]
- Zakharov, V.P.; Zakirova, I.D.; Mingaleev, V.Z.; Zakharova, E.M. The influence of the hydrodynamic conditions of reaction mixture flow on the thermomechanical properties of copolymers of butadiene and isoprene. Int. Polym. Sci. Technol. 2014, 41, 17189. [Google Scholar] [CrossRef]
- Xia, P.; Shao, H.; He, A. Excellent Oxygen Barrier Property of Unfilled Natural Rubber/trans-Butadiene-co-Isoprene Rubber Vulcanizates under the Synergistic Effect of Crosslinking Density and Crystallization. Polymers 2024, 16, 345. [Google Scholar] [CrossRef]
- Jiang, X.-b.; Zhang, Q.-f.; He, A.-h. Synthesis and Characterization of Trans-1,4-butadiene/Isoprene Copolymers: Determination of Sequence Distribution and Thermal Properties. Chin. J. Polym. Sci. 2015, 33, 815–822. [Google Scholar] [CrossRef]
- Wang, H.; Wang, R.-G.; Ma, Y.-S.; Luan, B.; He, A.-H. The Influence of Trans-1,4-poly(butadiene-co-isoprene) Copolymer Rubbers (TBIR) with Different Molecular Weights on the NR/TBIR Blends. Chin. J. Polym. Sci. 2019, 37, 966–973. [Google Scholar] [CrossRef]
- Fedorova, A.F.; Davydova, M.L.; Shadrinov, N.V.; Borisova, A.A.; Khaldeeva, A.R. Development of elastomeric materials based on butadiene-isoprene rubber. Pet. Eng. 2021, 19, 131–141. [Google Scholar] [CrossRef]
- Zhang, X.; Cui, H.; Song, L.; Ren, H.; Wang, R.; He, A. Elastomer nanocomposites with superior dynamic mechanical properties via trans-1, 4-poly (butadiene-co-isoprene) incorporation. Compos. Sci. Technol. 2018, 158, 156–163. [Google Scholar] [CrossRef]
- Li, N.; Zong, X.; Li, H.; He, A.; Zhang, X. Aging behaviors of trans-1, 4-poly (isoprene-co-butadiene) copolymer rubber. Polym. Degrad. Stab. 2021, 183, 109456. [Google Scholar] [CrossRef]
- Zong, X.; Wang, S.; Li, N.; Li, H.; Zhang, X.; He, A. Regulation effects of trans-1, 4-poly (isoprene-co-butadiene) copolymer on the processability, aggregation structure and properties of chloroprene rubber. Polymer 2021, 213, 123325. [Google Scholar] [CrossRef]
- Wang, H.; Zhang, X.; Nie, H.; Wang, R.; He, A. Multi-block copolymer as reactive multifunctional compatibilizer for NR/BR blends with desired network structures and dynamical properties: Compatibility, co-vulcanization and filler dispersion. Compos. Part A 2018, 116, 197–205. [Google Scholar] [CrossRef]
- Zakharov, V.; Ulitin, N.; Tereshchenko, K.; Zakharova, E.M. Turbulent Technologies in the Synthesis of Polydienes Using Ziegler-Natta Catalytic Components; Bashkir Encyclopedia: Ufa, Russia, 2018; p. 280. (In Russian) [Google Scholar]
- Tereshchenko, K.A.; Shiyan, D.A.; Ziganshina, A.S.; Ganiev, G.M.; Zakharov, V.P.; Ulitin, N.V. Control of Molar Mass Characteristics of Polybutadiene—A Component of Sticky Glue—By Physical Modification of the Catalytic System in Turbulent Flows. Polymer Sci. Ser. D 2020, 13, 250–257. [Google Scholar] [CrossRef]
- Ziganshina, A.S.; Shiyan, D.A.; Ganiev, G.M.; Tereshchenko, K.A.; Zakharov, V.P.; Ulitin, N.V. Controlling the Activity of Particles of TiCl4–Al(i-C4H9)3 Catalytic System by Changing Their Dispersion Composition in the Process of Producing Low-Molar-Mass Polybutadiene—A Component of Sticky Glue. Polym. Sci. Ser. D 2020, 13, 365–371. [Google Scholar] [CrossRef]
- Ganiev, G.M.; Tereshchenko, K.A.; Shiyan, D.A.; Ziganshina, A.S.; Zakharov, V.P.; Ulitin, N.V. Relationship of Molecular-Mass Characteristics of Polyisoprene, Component of Vulcanized Sealant, with Particle Sizes of Catalytic System TiCl4–Al(i-C4H9)3 in Isoprene Polymerization. Polym. Sci. Ser. D 2021, 14, 392–395. [Google Scholar] [CrossRef]
- Tereshchenko, K.A.; Ziganshina, A.S.; Zakharov, V.P.; Ulitin, N.V. Modeling of the Physicochemical Hydrodynamics of the Synthesis of Butadiene Rubber on the TiCl4−Al(i-C4H9)3 Catalytic System Modified in Turbulizing Flows. Russ. J. Phys. Chem. B 2017, 11, 504–512. [Google Scholar] [CrossRef]
- ANSYS. Fluent Theory Guide; ANSYS, Inc.: Canonsburg, PA, USA, 2013; p. 780. [Google Scholar]
- Aryafar, S.; Sheibat-Othman, N.; Mckenna, T.F.L. Coupling of CFD simulations and population balance modeling to predict brownian coagulation in an emulsion polymerization reactor. Macromol. React. Eng. 2017, 11, 1600054. [Google Scholar] [CrossRef]
- Karimi, Y.; Nazar, A.R.S.; Motavasel, M. CFD simulation of nanofluid heat transfer considering the aggregation of nanoparticles in population balance model. J. Therm. Anal. Calorim. 2020, 143, 671–684. [Google Scholar] [CrossRef]
- Atmaca, M.; Ezgi, C. Three-dimensional CFD modeling of a steam ejector. Energy Sources Part A Recovery Util. Environ. Eff. 2019, 44, 2236–2247. [Google Scholar] [CrossRef]
- Ahmad, M.; Li, B. A comparative analysis of turbulence models in FLUENT for high-lift airfoils at low Reynolds number. In Proceedings of the 2022 International Conference on Unmanned Aircraft Systems (ICUAS), Dubrovnik, Croatia, 21–24 June 2022. [Google Scholar] [CrossRef]
- Lehnigk, R.; Bainbridge, W.; Liao, Y.; Lucas, D.; Niemi, T.; Peltola, J.; Schlegel, F. An open-source population balance modeling framework for the simulation of polydisperse multiphase flows. AIChE J. 2021, 68, e17539. [Google Scholar] [CrossRef]
- Sarkar, J.; Shekhawat, L.K.; Loomba, V.; Rathore, A.S. CFD of mixing of multi-phase flow in a bioreactor using population balance model. Biotechnol. Prog. 2016, 32, 613–628. [Google Scholar] [CrossRef]
- ANSYS. Fluent Population Balance Module Manual; ANSYS, Inc.: Canonsburg, PA, USA, 2013; p. 66. [Google Scholar]
- Coulaloglou, C.A.; Tavlarides, L.L. Description of interaction processes in agitated liquid-liquid dispersions. Chem. Eng. Sci. 1977, 32, 1289–1297. [Google Scholar] [CrossRef]
- Wu, Y.; Ding, M.; Wang, J.; Zhao, B.; Wu, Z.; Zhao, P.; Tian, D.; Ding, Y.; Hu, A. Controlled Step-Growth Polymerization. CCS Chem. 2020, 2, 64–70. [Google Scholar] [CrossRef]
- Fazakas-Anca, I.S.; Modrea, A.; Vlase, S. Determination of Reactivity Ratios from Binary Copolymerization Using the k-Nearest Neighbor Non-Parametric Regression. Polymers 2021, 13, 3811. [Google Scholar] [CrossRef]
- Lansdowne, S.W.; Gilbert, R.G.; Napper, D.H.; Sangster, D.F. Relaxation studies of the seeded emulsion polymerization of styrene initiated by γ-radiolysis. J. Chem. Soc. Faraday Trans. 1 Phys. Chem. Condens. Phases 1980, 76, 1344–1355. [Google Scholar] [CrossRef]
- Podzimek, S. Light Scattering, Size Exclusion Chromatography and Asymmetric Flow Field Flow Fractionation: Powerful Tools for the Characterization of Polymers, Proteins and Nanoparticles; Wiley & Sons, Inc.: Hoboken, NJ, USA, 2011; 359. [Google Scholar]
- Andrushkevich, T.V.; Bukhtiyarov, V.I. Scientific heritage of Georgii Konstantinovich Boreskov. Kinet. Catal. 2019, 60, 123–136. [Google Scholar] [CrossRef]
- Tereshchenko, K.A.; Ismagilov, R.T.; Ulitin, N.V.; Lyulinskaya, Y.L.; Novikov, A.S. Effect of Monomer Mixture Composition on TiCl4-Al(i-C4H9)3 Catalytic System Activity in Butadiene–Isoprene Copolymerization: A Theoretical Study. Computation 2025, 13, 184. [Google Scholar] [CrossRef]
- Mingaleev, V.Z. Effect of the particle size of the TiCl4-Ali-Bu3 catalyst on the contribution from mono- and bimetallic active sites to the polymerization of isoprene. Kinet. Catal. 2016, 57, 52–60. [Google Scholar] [CrossRef]
- Luo, H.; Svendsen, H.F. Theoretical Model for Drop and Bubble Breakup in Turbulent Dispersions. AIChE J. 1996, 42, 1225–1233. [Google Scholar] [CrossRef]
- Lehr, F.; Millies, M.; Mewes, D. Bubble-Size Distributions and Flow Fields in Bubble Columns. AIChE J. 2002, 48, 2426–2443. [Google Scholar] [CrossRef]
- Moreno-Atanasio, R.; Ghadiri, M. Mechanistic Analysis and Computer Simulation of Impact Breakage of Agglomerates Effect of Surface Energy. Chem. Eng. Sci. 2006, 61, 2476–2481. [Google Scholar] [CrossRef]
Equations | Explanation of Symbols Used in the Equations |
---|---|
1. Law of conservation of mass for the reaction system [44,45] | —average mass velocity of the reaction system, m/s (k = 1—reaction system; k = 2—particles of the TiCl4 + Al(i-Bu)3 catalytic system) [46]; αk—volume fraction of the k-th phase; ρk—density of the k-th phase, kg/m3; —velocity of the k-th phase, m/s; —density of the reaction system, kg/m3 [46]; |
2. Momentum conservation equation for the reaction system [44,45] | t—time, s; p—pressure of the reaction system, Pa; —viscosity of the reaction system, Pa·s [46]; µk—viscosity of the k-th phase, Pa·s; —turbulent viscosity of the reaction system, Pa·s [46,47]; The subscript T refers to quantities caused by turbulent fluctuations; The superscript T indicates matrix transpose; —constant; K—turbulent kinetic energy, J/kg; ε—turbulent kinetic energy dissipation rate, m2/s3; —gravitational acceleration, m/s2; |
3. Energy conservation equation for the reaction system [44,48] | Ek—enthalpy of the k-th phase, J/kg; —effective thermal conductivity, W/(m2·K); λk—thermal conductivity of the k-th phase, W/(m2·K); —molar heat capacity of the reaction system, J/(kg·K); cpk—molar heat capacity of the k-th phase, J/(kg·K); —turbulent Prandtl number; T—temperature of the reaction system, K; |
4. Equations for calculating the relative velocity of phases [44] , , , , , , . | d2—diameter of dispersed phase particles (catalytic system particles), m; fdrag—drag function; Re = ρmvmd2/μm—Reynolds number; —acceleration of dispersed phase particles, m/s2; —turbulent diffusion coefficient, m2/s; γγ—parameter reflecting the inertia of the catalytic system particles; |
5. Equation for calculating the change in volume fraction of the dispersed phase [44,49] | |
6. K-ε turbulence model equations (these equations are necessary to close the system of equations, specifically for calculating µT according to the formula ) [44,46,47,50] | (W/m3) and (kg/(m·s4))—source terms caused by turbulent interaction between the dispersed phase and the dispersion medium; —generation of turbulent kinetic energy, W/m3 [46]; and —turbulent Prandtl numbers for K and ε, respectively [46,47]; and —standard constants of the K-ε turbulence model [46,47]; |
7. Population balance model equations (these equations are necessary for calculating the concentrations N of the catalytic system particles with volumes V) [49,50,51,52] | N—concentration of catalytic system particles, m−3; ch—number of particles formed by breakage from one original particle (assumed equal to 2); g(V)—frequency of breakage of catalytic system particles; —density of the dispersion medium; C1 and C2 (kg/s2)—constants; —probability density function of breakage from a particle of volume V’ to a particle of volume V, m−3; FS—particle shape factor. |
Stage Number | Rate Constants Determined at This Stage | Experimental Data on the Basis of Which the Rate Constants Are Determined | Equations Relating the Values of the Rate Constants to the Experimental Data |
---|---|---|---|
1 | the initial slope of the time dependence of the copolymer’s Mn at the beginning of copolymerization | ||
2 | The time dependences of the relative activities of active sites of various types, Sx | ||
3 | |||
4 | the limiting values of the number-average molecular weight Mn and weight-average molecular weight Mw of the copolymer |
Notation of Reaction Rate Constants | Active Sites Type Numbers x | |||
---|---|---|---|---|
1 | 2 | 3 | 4 | |
Chain propagation reactions 1 | ||||
0 | 0 | |||
Active sites interconversion reactions 2 | ||||
– | 0 | 0 | 0 | |
0 | – | |||
– | ||||
– | ||||
– | 0 | 0 | 0 | |
0 | – | 0 | ||
0 | – | 0 | ||
0 | 0 | 0 | – | |
Active sites deactivation reactions 2 | ||||
0 | 0 | |||
Chain transfer reactions to monomers 1 | ||||
0 | 0 | |||
0 | 0 | |||
0 | 0 |
A complex concept that can be quantitatively expressed by characteristics of varying detail (scalar characteristics, distribution functions, fields, etc.) | Simple scalar quantitative characteristic used in the calculation |
Conditions for the formation of the reaction system | Geometric parameters of the tubular turbulent apparatus and the linear feed velocity of the reaction system into it |
Conservation laws of mass, momentum, and energy of the reaction system considering turbulent fluctuations of its velocity, pressure, and temperature | CFD model including the two-parameter K-ε turbulence model with specified boundary conditions |
Turbulence of the reaction system | Turbulent kinetic energy K, J/kg |
Equations of solid mechanics and strength of materials | Population balance model equations for particles, based on the dependence of particle breakage frequency on the turbulent kinetic energy dissipation rate ε, m2/s3 |
Dispersity of particles of the heterogeneous catalytic system TiCl4 + Al(i-Bu)3 | Specific surface area of these particles—particle surface area normalized to their mass Surfsp, m2/g |
Various theories of heterogeneous catalysis | Semi-empirical dependence reflecting the kinetics of active sites formation according to the Langmuir monomolecular adsorption theory |
Activity and kinetic heterogeneity of the TiCl4-Al(i-Bu)3 catalytic system | Concentrations of active sites of various types μxy00, mol/L. |
The law of mass action, the method of statistical moments, the method of generating functions | Kinetic model of butadiene-isoprene copolymerization developed in this work |
Kinetic patterns of butadiene-isoprene copolymerization | Copolymerization rate, number-average Mn and weight-average molecular weight Mw of the copolymer |
Parameter Name | Value of the Parameter |
---|---|
Number of sections of the tubular turbulent apparatus | 1, 2, 3, 4, 5, 6, 7, 8 |
Feed velocity of the reaction system into the tubular turbulent apparatus v, m/s | 0.3, 0.6, 0.9, 1.2, 1.5 |
Ratio of the section length L to the diffuser diameter dd at constant diffuser diameter | 4/3, 5/3, 6/3, 7/3, 8/3 |
Ratio of the section length L to the diffuser diameter dd at constant section length and constant ratio of confuser diameter to diffuser diameter | 12/4, 12/5, 12/6, 12/7, 12/8 |
Ratio of the diffuser diameter dd to the confuser diameter dc at constant diffuser diameter | 8/3, 8/4, 8/5, 8/6, 8/7 |
Diffuser opening angle α, ° | 15, 30, 45, 60, 75, 90 |
v, m/s | Kmax × 101, J/kg | Km × 101, J/kg | τ, s | (1 + a + b)[A]0 × 104, mol/L | μ00 × 105, mol/L at q = 0.76 | μ00 × 105, mol/L at q = 0.85 |
---|---|---|---|---|---|---|
0.3 | 0.45 | 0.14 | 0.463 | 3.90 | 5.35 | 7.63 |
0.6 | 0.75 | 0.35 | 0.232 | 4.84 | 6.87 | 9.70 |
0.9 | 1.05 | 0.68 | 0.154 | 6.91 | 10.19 | 14.23 |
1.2 | 1.80 | 1.12 | 0.116 | 9.60 | 14.50 | 20.10 |
1.5 | 3.00 | 1.64 | 0.093 | 12.35 | 18.93 | 26.13 |
dd/dc | Kmax × 101, J/kg | Km × 101, J/kg | τ, s | (1 + a + b)[A]0 × 104, mol/L | μ00 × 105, mol/L at q = 0.76 | μ00 × 105, mol/L at q = 0.85 |
---|---|---|---|---|---|---|
8/3 | 3.80 | 1.82 | 0.144 | 12.99 | 19.95 | 27.52 |
8/4 | 1.80 | 1.02 | 0.150 | 8.81 | 13.24 | 18.39 |
8/5 | 1.20 | 0.68 | 0.154 | 6.97 | 10.29 | 14.36 |
8/6 | 0.80 | 0.48 | 0.158 | 6.03 | 8.78 | 12.30 |
8/7 | 0.40 | 0.30 | 0.159 | 5.21 | 7.47 | 10.51 |
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. |
© 2025 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (https://creativecommons.org/licenses/by/4.0/).
Share and Cite
Tereshchenko, K.A.; Ulitin, N.V.; Ismagilov, R.T.; Novikov, A.S. Impact of Reaction System Turbulence on the Dispersity and Activity of Heterogeneous Ziegler–Natta Catalytic Systems for Polydiene Production: Insights from Kinetic and CFD Analyses. Compounds 2025, 5, 39. https://doi.org/10.3390/compounds5040039
Tereshchenko KA, Ulitin NV, Ismagilov RT, Novikov AS. Impact of Reaction System Turbulence on the Dispersity and Activity of Heterogeneous Ziegler–Natta Catalytic Systems for Polydiene Production: Insights from Kinetic and CFD Analyses. Compounds. 2025; 5(4):39. https://doi.org/10.3390/compounds5040039
Chicago/Turabian StyleTereshchenko, Konstantin A., Nikolai V. Ulitin, Rustem T. Ismagilov, and Alexander S. Novikov. 2025. "Impact of Reaction System Turbulence on the Dispersity and Activity of Heterogeneous Ziegler–Natta Catalytic Systems for Polydiene Production: Insights from Kinetic and CFD Analyses" Compounds 5, no. 4: 39. https://doi.org/10.3390/compounds5040039
APA StyleTereshchenko, K. A., Ulitin, N. V., Ismagilov, R. T., & Novikov, A. S. (2025). Impact of Reaction System Turbulence on the Dispersity and Activity of Heterogeneous Ziegler–Natta Catalytic Systems for Polydiene Production: Insights from Kinetic and CFD Analyses. Compounds, 5(4), 39. https://doi.org/10.3390/compounds5040039