Process for polymerizing trifunctional long-chain branched olefins

文档序号:1850595 发布日期:2021-11-16 浏览:26次 中文

阅读说明:本技术 聚合三官能长链支化烯烃的方法 (Process for polymerizing trifunctional long-chain branched olefins ) 是由 罗伯特·Dj·弗勒泽 R·E·M·布罗内尔 D·J·阿里欧拉 B·D·斯塔伯特 O·D·雷德 于 2020-03-27 设计创作,主要内容包括:合成长链支化聚合物的方法。所述方法包含任选地在存在氢气的情况下使一种或多种C-(2)-C-(14)烯烃单体、至少一种二烯、任选地溶剂和多链催化剂接触在一起,其中所述多链催化剂包括多个聚合位点;产生所述C-(2)-C-(14)烯烃单体的至少两条聚合物链,每条聚合物链在所述聚合位点之一处聚合;通过将所述两条聚合物链与所述二烯连接来合成所述长链支化聚合物,所述两条聚合物链的所述连接是在所述聚合期间以协同方式进行的;以及由所述二烯产生三官能长支链和四官能长支链,其中所述长链支化聚合物的三官能长支链与四官能长支链的比率为0.05:1到100:0;以及调整所述三官能长支链与四官能长支链的比率。所述二烯具有根据式(I)的结构:(A method of synthesizing a long chain branched polymer. The method comprises reacting one or more C, optionally in the presence of hydrogen 2 ‑C 14 Contacting together an olefin monomer, at least one diene, optionally a solvent, and a multi-chain catalyst, wherein the multi-chain catalyst comprises a plurality of polymerization sites; generating said C 2 ‑C 14 At least two polymer chains of olefin monomers, each polymer chain being polymerized at one of the polymerization sites; synthesizing the long-chain branched polymer by linking the two polymer chains to the diene, the linking of the two polymer chains being performed in a synergistic manner during the polymerization; and generating trifunctional from said dieneLong chain branches and tetrafunctional long chain branches, wherein the long chain branched polymer has a ratio of trifunctional long chain branches to tetrafunctional long chain branches of from 0.05:1 to 100: 0; and adjusting the ratio of trifunctional long-chain branches to tetrafunctional long-chain branches. The diene has a structure according to formula (I):)

1. A method of synthesizing a long chain branched polymer, the method comprising:

optionally reacting one or more C's in the presence of hydrogen2-C14Contacting together an olefin monomer, at least one diene, optionally a solvent, and a multi-chain catalyst, wherein the multi-chain catalyst comprises a plurality of polymerization sites, and wherein the diene has a structure according to formula (I):

wherein X is-C (R)2-、-Si(R)2-or-Ge (R)2-, wherein each R is independently C1-C12A hydrocarbyl group or-H;

generating said C2-C14At least two polymer chains of olefin monomers, each polymer chain being polymerized at one of the polymerization sites; and

synthesizing the long-chain branched polymer by linking the two polymer chains to the diene, the linking of the two polymer chains being performed in a synergistic manner during the polymerization, wherein the long-chain branched polymer has a ratio of trifunctional long-chain branches to tetrafunctional long-chain branches of from 0.05:1 to 100: 0; and

if the ratio of trifunctional long-chain branches to tetrafunctional long-chain branches deviates from the target ratio of trifunctional long-chain branches to tetrafunctional long-chain branches, then by varying the C2-C14The feed ratio of olefin monomer to hydrogen is adjusted to adjust the ratio.

2. The process according to claim 1, wherein X in formula (I) is-C (R)2-, and wherein each R is-H.

3. The process according to claim 1, wherein X in formula (I) is-C (R)2-, and wherein each R is C1-C12An alkyl group.

4. The process according to claim 1, wherein X in formula (I) is-Si (R)2-, and wherein each R is C1-C12An alkyl group.

5. The method of claim 4, wherein the diene is dimethyldivinylsilane.

6. The method of any preceding claim, wherein the long-chain branched polymer is an ethylene-based copolymer comprising at least 50 mol% ethylene.

7. The method of any one of the preceding claims, wherein the multi-chain catalyst is a surface concentration of metal atoms greater than or equal to 1.5 metal atoms per square nanometer (1.5 metals/nm)2) The heterogeneous catalyst of (1).

8. The method of any one of claims 1 to 6, wherein the multi-chain catalyst comprises two linked transition metals linked by a dianionic activator, wherein the distance between metal atoms is less than or equal to

9. The method of any one of the preceding claims, wherein the multi-chain catalyst comprises two or more transition metals covalently linked, wherein the distance between metal atoms is less than or equal to

10. The method of any one of the preceding claims, wherein the multi-chain catalyst comprises a monoanionic ligand and an IUPAC group IV metal selected from the group consisting of titanium, hafnium, or zirconium.

11. The method of any preceding claim, wherein the trifunctional long-chain branches occur at a frequency of at least 0.05 per 1000 carbon atoms.

12. The method of any preceding claim, wherein the trifunctional long-chain branches occur at a frequency of at least 0.1 per 1000 carbon atoms.

13. The method of any preceding claim, wherein the trifunctional long-chain branches occur at a frequency of at least 0.2 per 1000 carbon atoms.

14. The method of any preceding claim, wherein the ratio of trifunctional long-chain branches to tetrafunctional long-chain branches is greater than 0.1: 1.

15. The method of any preceding claim, wherein controlling the ratio of trifunctional long-chain branches to tetrafunctional long-chain branches comprises increasing H2To increase the amount of trifunctional long chain branching.

16. The method of any one of claims 1 to 15, wherein a control stationThe ratio of trifunctional long-chain branches to tetrafunctional long-chain branches comprises the reaction of said ethylene with H2Is adjusted to be greater than 999:1 such that the ratio of trifunctional long-chain branches to tetrafunctional long-chain branches is less than 0.001: 1.

17. The method of claims 1-15, wherein controlling the ratio of trifunctional long-chain branches to tetrafunctional long-chain branches comprises reacting the ethylene with H2Is adjusted to less than 25:1 such that the ratio of trifunctional long-chain branches to tetrafunctional long-chain branches is greater than 0.5: 1.

18. The method of claims 1-15, wherein controlling the ratio of trifunctional long-chain branches to tetrafunctional long-chain branches comprises reacting the ethylene with H2Is adjusted to be less than 50:50 such that the ratio of trifunctional long-chain branches to tetrafunctional long-chain branches is greater than 1: 1.

19. The polymer of any preceding claim, wherein the weight average molecular weight (M) of the long chain branched polymer as determined by gel permeation chromatography using a triple detectorw) Is less than 150,000 daltons.

20. The process of any of the preceding claims, wherein the polymerization occurs in a solution polymerization reactor, a slurry reactor, a gas phase reactor, a batch reactor, a continuous reactor, a mixed reactor, a non-back-mixed reactor, a series reactor, or a loop reactor.

21. The method of any preceding claim, wherein the long chain branched polymer has a weight average molecular weight divided by a number average molecular weight (M) as determined by gel permeation chromatography using a triple detectorw/Mn) The defined Molecular Weight Distribution (MWD) is less than 4.

Technical Field

Embodiments of the present disclosure generally relate to polymer compositions having trifunctional long chain branches and methods of synthesizing the polymer compositions.

Background

Olefin-based polymers such as polyethylene are produced by various catalyst systems. The selection of such catalyst systems for use in the polymerization process of olefin-based polymers is an important factor contributing to the characteristics and properties of such olefin-based polymers.

Polyethylene and polypropylene are manufactured for use in a variety of articles. Polyethylene and polypropylene polymerization processes may differ in many respects to produce a variety of resulting polyethylene resins having different physical properties that make the various resins suitable for different applications. The amount of long chain branching in a polyolefin affects the physical properties of the polyolefin. The effect of branching on the properties of polyethylene depends on the length and amount of branching. Short branches mainly affect mechanical and thermal properties. As the branch length increases, the branches can form lamellar crystals, and mechanical and thermal properties decrease. Small amounts of long chain branching can significantly alter polymer processing properties.

To form long chain branching, vinyl groups or terminal double bonds of the polymer chains are incorporated into new polymer chains. The reincorporation of vinyl terminated polymers and the introduction of diene comonomers are two mechanisms for incorporating the vinyl groups on the polymer chain into the second polymer chain. In addition, long chain branching is caused by free radicals. It is difficult to control the amount of branching in all three mechanisms. When long chain branching is initiated using free radicals or dienes, branching may become too much, causing gelling and reactor fouling. The reincorporation mechanism does not produce too much branching, and branching can only occur after the polymer chains are produced, further limiting the amount of branching that can occur.

Disclosure of Invention

Embodiments of the present disclosure include methods for synthesizing long chain branched polymers. In one or more embodiments, the method comprises reacting one or more C, optionally in the presence of hydrogen2-C14The olefin monomer, at least one diene, optionally a solvent, and a multi-chain catalyst are contacted together, wherein the multi-chain catalyst comprises a plurality of polymerization sites. Generating said C2-C14At least two polymer chains of olefin monomers, each polymer chain being polymerized at one of said polymerization sites. The long chain branched polymer is then synthesized by linking the two polymer chains to the diene. Said linking of said two polymer chains is performed in a synergistic manner during said polymerization. Generating trifunctional long-chain branches from the diene, wherein the long-chain branched polymer has a ratio of trifunctional long-chain branches to tetrafunctional long-chain branches of from 0.05:1 to 100: 0. If the ratio of trifunctional long-chain branches to tetrafunctional long-chain branches deviates from the target ratio of trifunctional long-chain branches to tetrafunctional long-chain branches, then C can be varied2-C14The feed ratio of olefin monomer to hydrogen is adjusted to adjust the ratio.

In one or more embodiments, the diene has a structure according to formula (I):

in the formula (I), X is CR2、SiR2Or GeR2Wherein each R is independently C1-C12A hydrocarbyl group or-H. In some embodiments, X in formula (I) is-C (R)2-, and wherein each R is-H, or each R is C1-C12An alkyl group. In other embodiments, X in formula (I) is-Si (R)2-, and wherein each R is C1-C12An alkyl group. In one or more embodiments, the diene is dimethyldivinyl silane.

Various embodiments of the process comprise polymerization that occurs in a solution polymerization reactor or a particle forming polymerization reactor, such as a slurry reactor or a gas phase reactor, wherein a molecular or solid supported catalyst is delivered to or formed in a reaction medium, wherein the reactor system is batch or continuous or hybrid, such as semi-batch, wherein the reactor residence time distribution is narrower as in a non-back-mixed reactor or broader as in a back-mixed reactor and in series and recycle reactors.

Drawings

FIG. 1 is a graphical depiction of the molecular weight of a polymer as the number of branched methines per 1000 carbons increases.

FIG. 2A is a plot of molecular weight increase versus diene bond.

Figure 2B is a plot of polydispersity versus diene bond.

FIG. 3 is a graphical depiction of the predicted dependence of Molecular Weight Distribution (MWD) curves on the level of trifunctional diene branching.

FIG. 4 is a graphical depiction of the predicted dependence of the relative peak of Molecular Weight (MW) on the level of trifunctional diene branching.

FIG. 5 is a graphical depiction of a MWD curve showing how the point of maximum slope is used to define a high MWD tail area metric.

FIG. 6: conventional (RI) GPC of Dimethyldiethylvinylsilane samples with increased diene content (examples 1.C and 1.1-1.7).

FIG. 7 is a global carbon NMR spectrum of dimethyldivinylsilane branched polyethylene (example 12.1).

FIG. 8 is Si (Me) of the carbon NMR spectrum of dimethyldivinylsilane branched polyethylene (example 12.1)2And (4) a region. Trifunctional LCB carbon 0.17Me/1000C and tetrafunctional LCB carbon 0.12 Me/1000C.

FIG. 9: the methine region of the carbon NMR spectrum of dimethyldivinylsilane branched polyethylene (example 12.1). Trifunctional LCB ═ 0.09CH/1000C, and tetrafunctional LCB ═ 0.16 CH/1000C.

FIG. 10: conventional (RI) and absolute (LS) GPC of linear PE (example 12.C) and dimethyldivinylsilane branched PE (example 12.1).

FIG. 11: extensional Viscosity Fixture (EVF) of dimethyldivinylsilane branched PE (example 12.1).

FIG. 12: melt strength plots of dimethyldivinylsilane branched polyethylene (example 12.1).

FIG. 13: DMS at 190 ℃ of dimethyldivinylsilane-branched polyethylene (example 12.1).

FIG. 14A is a graph comparing the absolute molecular weight distribution of conventional branched polymer samples for diene weight change.

Fig. 14B is a graph comparing conventional molecular weight distributions of conventional branched polymer samples for diene weight change.

Detailed Description

Specific examples of methods for synthesizing polymers and polymers synthesized by the methods of the present disclosure will now be described. It is to be understood that the methods for synthesizing polymers of the present disclosure may be embodied in different forms and should not be construed as limited to the specific embodiments set forth in the disclosure. Rather, embodiments are provided so that this disclosure will be thorough and complete and will fully convey the scope of the subject matter to those skilled in the art.

Definition of

The term "polymer" refers to a polymeric compound prepared by polymerizing monomers of the same or different types. Thus, the generic term polymer encompasses the term "homopolymer", which is commonly used to refer to polymers prepared from only one type of monomer, and "copolymer", which refers to polymers prepared from two or more different monomers. The term "interpolymer" as used herein refers to a polymer prepared by the polymerization of at least two different types of monomers. Thus, the generic term interpolymer encompasses copolymers, and polymers prepared from more than two different types of monomers, such as terpolymers.

"polyethylene" or "ethylene-based polymer" shall mean a polymer comprising greater than 50 mol% of units derived from ethylene monomers. This includes polyethylene homopolymers or copolymers (meaning units derived from two or more comonomers). Common forms of polyethylene known in the art comprise: low Density Polyethylene (LDPE); linear Low Density Polyethylene (LLDPE); ultra Low Density Polyethylene (ULDPE); very Low Density Polyethylene (VLDPE); a single-site catalyzed linear low density polyethylene comprising both a linear low density resin and a substantially linear low density resin (m-LLDPE); medium Density Polyethylene (MDPE); and High Density Polyethylene (HDPE).

"ethylene-diene-based polymer" shall mean a polymer comprising greater than 50 mol% of units derived from ethylene monomers and also comprising a minor component of a diene. The ethylene-diene-based polymer may optionally comprise a copolymer derived from one or more (C)3-C12) An alpha-olefin.

Embodiments of the present disclosure include methods for synthesizing long chain branched polymers. In one or more embodiments, the method comprises reacting one or more C, optionally in the presence of hydrogen2-C14The olefin monomer, at least one diene, optionally a solvent, and a multi-chain catalyst are contacted together, wherein the multi-chain catalyst comprises a plurality of polymerization sites. Generating said C2-C14At least two polymer chains of olefin monomers, each polymer chain being polymerized at one of said polymerization sites. The long chain branched polymer is then synthesized by linking the two polymer chains to the diene. The two poly(s)Said linking of the polymer chains is carried out in a synergistic manner during said polymerization.

In various embodiments, the long-chain branched polymer has a ratio of trifunctional long-chain branches to tetrafunctional long-chain branches of 0.05:1 to 100: 0.

In one or more embodiments, the trifunctional long-chain branches are generated from the diene, where the trifunctional long-chain branches occur at a frequency of at least 0.03 per 1000 carbon atoms.

The term "linked" when referring to "linking two polymer chains" means that the polymer chains are covalently linked.

In one or more embodiments, the ratio of trifunctional long-chain branches to tetrafunctional long-chain branches is adjusted if the ratio deviates from a target ratio of trifunctional long-chain branches to tetrafunctional long-chain branches. By adjusting C2-C14Amount of olefin monomer feed, amount of hydrogen feed, C2-C14The ratio of olefin monomer feed to hydrogen, reactor temperature, or a combination thereof.

In some embodiments, C2-C14The feed ratio of olefin monomer feed to hydrogen is from 100:0 to 1: 100. In one or more embodiments, the feed ratio is 3:1 to 1: 1. In various embodiments, the feed ratio is from 100:1 to 2:1, 10:2 to 1:2, or 3:1 to 1: 5.

In various embodiments, the diene has a structure according to formula (I):

in the formula (I), X is CR2、SiR2Or GeR2Wherein each R is independently C1-C12A hydrocarbyl group or-H. In some embodiments, X in formula (I) is-C (R)2-, and wherein each R is-H, or each R is C1-C12An alkyl group. In other embodiments, X in formula (I) is-Si (R)2-, and wherein each R is C1-C12An alkyl group. In one or more implementationsIn one embodiment, the diene is dimethyldivinyl silane.

In some embodiments, when R of formula (I) is C1-C12When alkyl, said C1-C12Alkyl is methyl, ethyl, 1-propyl, 2-propyl, 1-butyl, 2-methylpropyl, 1-dimethylethyl, 1-pentyl; 1-hexyl, 1-heptyl, n-octyl, t-octyl, nonyl, decyl, undecyl or dodecyl. The term "C1-C12Alkyl "means a saturated straight or branched chain hydrocarbon group of 1 to 12 carbon atoms.

In one or more embodiments, the process of the present disclosure produces a polymerization product comprising ethylene, at least one diene comonomer, and optionally at least one C3To C14A polymer of a comonomer. The polymer includes trifunctional long-chain branches generated from the diene, the trifunctional long-chain branches occurring at a frequency of at least 0.03 per 1000 carbon atoms of the polymer.

In various embodiments, the polymer produced according to the polymerization process of the present disclosure comprises trifunctional long-chain branches that occur at a frequency of at least 0.05 per 1000 carbon atoms. In one or more embodiments, the trifunctional long-chain branches of the polymers occur at a frequency of at least 0.1 per 1000 carbon atoms. In various embodiments, the trifunctional long-chain branches of the polymers occur at a frequency of at least 0.2 per 1000 carbon atoms.

The method of synthesizing polymers according to the present disclosure differs from conventional long chain branching or previous "ladder branching" described in the following applications: application No. PCTUS 2019/053524; number PCTUS 2019/053527; number PCTUS 2019/053529; and pct us2019/053537, each filed on 27/9/2019 and incorporated herein by reference in its entirety. The term "long chain branching" refers to branches having greater than 100 carbon atoms. "branched" refers to a portion of a polymer that extends from a tertiary carbon atom. When a branch extends from a tertiary carbon atom, there are two other branches that together may be a polymer chain from which the branch extends. In this disclosure, branching is defined as trifunctional long-chain branching in which the branch point has three polymer chains emanating from it. Conventionally, Long Chain Branching (LCB) may occur naturally during polymerization, as shown in scheme 1. Naturally occurring LCB can occur through vinyl termination of the polymer chain and reinsertion of the macromolecular vinyl group that produces trifunctional long chain branching. Depending on the degree of branching, various methods can determine the LCB (e.g., Nuclear Magnetic Resonance (NMR)), or differentiate the role of the LCB in the polymer. For example, the effect of LCB is observed in shear flow in the van gupu-palman analysis, shear viscosity also increases at low angular frequencies, and the intensity of shear thinning behavior can also be attributed to LCB. In extensional flow, the effect of LCB is generally identified in the degree of strain hardening or melt strength and the maximum deformation achieved. It is difficult to achieve high levels of natural LCB in polymers due to the limited concentration of vinyl terminated polymers (at most one per polymer chain) and the need to achieve high ethylene conversion to ensure LCB formation. To ensure high conversion, the ethylene concentration in the reactor is low, thus enabling the reinsertion of a large amount of vinyl terminated polymer into the second polymer chain.

Scheme 1: naturally occurring long chain branching: chain transfer events to produce vinyl terminated polymers

In scheme 1, "Cat" is the catalyst and "P" is the polymer chain.

The formation of long chain branches by naturally occurring branches is minimized. One way to enhance LCB is by adding an alpha, omega-diene to the polymerization system, whether in a free radical, heterogeneous or homogeneous process. Typically, the diene is added to the polymer chain in a similar manner to the alpha-olefin, but leaves pendant vinyl groups that can reinsert into the polymer chain to form the LCB, as shown in scheme 2. Generally, the diene length is not critical, so long as it can link two polymer chains together. In principle, the concentration of side vinyl groups can be controlled by the amount of diene added to the reactor. Thus, the extent of LCB can be controlled by the concentration of the pendant vinyl groups.

Scheme 2: long chain branching by diene incorporation

In scheme 2, "Cat" is the catalyst; "P" is a polymer chain; and the diene in this example is 1, 5-hexadiene.

Conventional methods of incorporating dienes into polymer synthesis systems suffer from substantial drawbacks of gel formation or reactor fouling. The kinetic model discussed in the following paragraphs may provide good predictive results that enable a better understanding of gel formation. For example, longer polymer chains have more intercalated olefins, and thus more intercalated dienes, and thus more pendant vinyl groups, which means that longer polymer chains will be more likely to reinsert into the catalyst to form LCBs. Thus, longer polymer chains preferentially reinsert to form tetrafunctional branches, which are even larger polymer molecules and lead to gel problems. As indicated in scheme 2, tetrafunctional LCBs have short segments (the number of carbon atoms between the two double bonds of the diene) that bridge two long chains on each side of the short segment. Weight average molecular weight (M) as a function of branching for polyethylene in a semi-batch reactor at constant pressurew) And number average molecular weight (M)n) The simulation of (a) is shown in fig. 1. In FIG. 1, MnFollowing only MwBecomes infinite with a slight increase. When M iswWhen the amount is increased to an amount greater than 200,000 grams per mole (g/mol), the polymer gels, gelation occurs, or reactor fouling is present.

The term "gel" or "gelling" refers to a solid composed of at least two components: the first is a three-dimensionally crosslinked polymer, and the second is a medium in which the polymer is not completely soluble. When the polymer gel is not completely dissolved, the reactor may be contaminated with the polymer gel.

The term "ladder-branched" polymer refers to a polymer formed by a "ladder-branching mechanism". As depicted in scheme 2, the polymer has a trifunctional long-chain branched structure. In addition, the terms "ladder-branched" polymer and "ladder-branching mechanism" also refer to trifunctional polymers and polymerization processes that produce trifunctional, long-chain branched polymers.

The process for synthesizing trifunctional long-chain branched polymers achieves long-chain branching and avoids gel formation or reactor fouling. Without intending to be bound by theory, it is believed that reactor fouling is avoided by reacting the two olefins of the diene in a consistent manner across the two proximal polymer chains. For example and as demonstrated by scheme 3, one olefin in a diene reacts before a second olefin, and the second olefin reacts before too many ethylene molecules are added to the polymer chain, thereby removing the immediate proximity of the second olefin to the reaction site. The reaction of a first olefin in a diene with one polymer and the reaction of a second olefin in a diene with an adjacent polymer chain before the insertion of a number of ethylene monomers is referred to as the concerted addition of the diene to the proximal polymer chain.

Scheme 3: the description of the incorporation of dienes in a synergistic manner (P being the polymer chain) is also referred to as a trifunctional "ladder branching" mechanism.

Depending on the catalyst or the diene, different intermediates may be produced from the diene reaction. Previous work has shown that it is also possible to form trifunctional LCBs from the addition of dienes to the double-stranded catalyst (scheme 3) in the formation of trifunctional long-chain branching. (scheme 4).

Scheme 4: the formation of trifunctional long-chain branches from the reaction of dienes is described.

The polymer chain is a linear segment of a polymer, or more specifically a copolymer, which is optionally linked at a terminal end by a branched attachment point. For example, a tetrafunctional branch attachment point connects the ends of four polymer chains, as opposed to a trifunctional branch attachment point, which connects the ends of three polymer chains as shown in scheme 1.

While not intending to be bound by theory, as explained in this section, the mechanism describes how the bis-chain catalyst can produce a unique trifunctional bridged molecular architecture upon polymerization of diene comonomers under desired conditions. The term "diene" refers to a monomer or molecule having two olefins. A schematic depiction of the mechanism is shown in scheme 5, where the catalyst center produces two polyolefin chains. Scheme 5 shows how the combination of diene bridging and chain transfer results in a diene "ladder branched" trifunctional polymer structure. The term diene "ladder-branched" polymer refers to long chain branching, wherein a short chain or step containing one to twelve carbon atoms links two polymer chains together. As shown, a metal-ligand catalyst having at least two polymer chain sites grows two separate polymer chains. One olefin of the diene is incorporated into one of the sites of the catalyst, and it is believed that due to the close proximity of the growth sites, the second olefin of the diene is then rapidly incorporated into the second polymer chain, forming a bridge or step. This continuous addition of diene is referred to as "synergistic" addition of diene, as distinguished from catalysts without two proximal chains, where diene addition results in a concentration of vinyl-containing polymer in the reactor that reacts at a later time. The term "step" refers to once the diene is incorporated into two separate polymer chains, linking the chains together. The first polymer chain and the second polymer chain continue to grow until the polymer is transferred to another catalyst, the polymer is released from the catalyst, the catalyst die, or another diene is added.

Scheme 5. Description of the trifunctional "ladder branching" mechanism comprising the resulting molecular architecture. Metal-ligand catalysts consisting of L-M+Collectively.

As depicted in scheme 5, trifunctional trapezoidal branching can occur upon introduction of hydrogen. The introduction of hydrogen terminates the polymer chain at one of the polymerization sites of the multi-chain catalyst. Upon termination, the polymer chains are disconnected, thereby producing a trifunctional polymer. The polymers of the present disclosure comprise trifunctional long chain branches produced from dienes of formula (I).

In one or more embodiments, if the ratio of trifunctional branches to tetrafunctional branches deviates from the target ratio of trifunctional long-chain branches to tetrafunctional long-chain branches, then by varying the C2-C14The feed ratio of olefin monomer to hydrogen is adjusted to adjust the ratio. In some embodiments, the feed ratio may be varied during the reaction. The ratio of trifunctional to tetrafunctional branches depends on the hydrogen. If the hydrogen concentration is increased, the ratio of trifunctional to tetrafunctional branches increases. In addition, when the polymerization process occurs in solution, the concentration of hydrogen in the solution can affect the ratio of trifunctional to tetrafunctional branching. In some embodiments, a specific amount of hydrogen may be introduced prior to initiating the polymerization reaction, and the temperature may be increased or decreased to adjust the ratio of trifunctional to tetrafunctional branches. The increase in temperature will cause the hydrogen and C to react2-C14The reactivity of the olefin monomer increases, which results in an increased amount of trifunctional long-chain branching.

In one or more embodiments, the ratio of trifunctional to tetrafunctional branches is controlled by adjusting the ethylene/hydrogen ratio or other reactor conditions, such as temperature, in the reactor. In some embodiments, the ratio of trifunctional long-chain branches to tetrafunctional long-chain branches is greater than 0.1:1 to about 100: 0.

Without intending to be bound by theory, it is believed that the molecular weight distribution associated with these proposed kinetics is inherently stable at high levels of branching when the diene bridging reaction is the sole source of branching. Molecular Weight Distribution (MWD) is the weight average molecular weight divided by the number average molecular weight (M)w/Mn) And (4) defining. Inherent stability of MWDMeaning weight average molecular weight (M), even at high branching levelsw) Only moderately increased, in contrast to conventional diene comonomer branching techniques, where M iswAnd Mw/MnBecomes infinite at moderate tetrafunctional branching levels.

The combination of the multi-chain catalyst and the diene affects the amount and type of branching. Embodiments of the present disclosure relate to controlling polymer properties such as: 1) the use of multiple diene species, 2) the use of multiple multi-chain catalyst species, 3) a combination of polymerization environments comprising multiple reactor zones or gradient zones, or 4) control and combination of multiple types of long chain branching, such as trifunctional long chain branching and tetrafunctional long chain branching.

Although a variety of catalysts are used, including single chain catalysts, conventional branching may be allowed. The use of various diene species also includes those dienes that do not branch or result in "conventional" LCB. The method of synthesizing polymers according to the present disclosure is different from conventional long chain branching. The term "long chain branching" refers to branches having greater than 100 carbon atoms. The term "branched" refers to a portion of a polymer that extends from a tertiary carbon atom. When a branch extends from a tertiary carbon atom, there are two other branches that together may be a polymer chain from which the branch extends. Long Chain Branching (LCB) may occur naturally during the polymerization process, as shown in scheme 1. This can occur through the termination of the polymer chain and the reinsertion of the macromolecular vinyl group that produces the trifunctional long-chain branch.

In one or more embodiments, the method for polymerizing a long chain branched polymer comprises a catalyst having at least two active sites in close proximity (a multi-chain catalyst). Close proximity comprises less thanIs less thanOr aboutThe distance of (c).

It is well known that modern computing techniques can reproduce with good precision the known experimental structure as a method of estimating the distance between the chains of catalyst. Diene structure according to formula (I), wherein X is-C (R)2-、-Si(R)2-or-Ge (R)2Where each R is independently hydrogen or a hydrocarbon group, allows the size of the diene to be estimated. According to X being-Ge (R)2The diene of formula (I) has an end-to-end distance from the diene of aboutThus, the polymerization site of the multiple chains may be atInternal, or in the case of bimetallic catalysts, two metals inAnd (4) the following steps.

For heterogeneous systems, the surface concentration of the metal can be estimated, typically at every square nanometer (M/nm)2) The metal atom of (2) measures the metal. This surface coverage provides an estimate of the available metal on the surface, which, if uniformly dispersed, can be converted to an M-M distance, which reflects the distance between the polymer chains. For extended surfaces, 1 metal/nm2Resulting in a distance between the metal atoms ofIn thatNext, 1.5 metal/nm can be determined2The coverage of (c).

Examples of catalysts having at least two active sites, wherein the active sites are in close proximity, include, but are not limited to: a bimetallic transition metal catalyst; a heterogeneous catalyst; a dianionic activator having two associated active catalysts; a linked transition metal catalyst having more than one growing polymer chain; a group IV olefin polymerization catalyst comprising a monoanionic group, a bidentate monoanionic group, a tridentate monoanionic group, or a monodentate, bidentate, or tridentate monoanionic group with an external donor.

The catalysts in table 1 are illustrative examples of the classes of catalysts previously described and the particular catalysts contemplated. The examples in table 1 are not intended to be limiting; rather, the examples are merely illustrative and specific examples of the previously mentioned classes of catalysts.

Table 1: catalyst with immediately adjacent more than one active site

While not intending to be bound by theory, as explained in this section, the mechanism describes how the bis-chain catalyst can produce a unique trifunctional bridged molecular architecture upon polymerization of diene comonomers under desired conditions. The term "diene" refers to a monomer or molecule having two olefins. A schematic depiction of the mechanism is shown in scheme 5, where the catalyst center produces two polyolefin chains. Scheme 5 shows how the combination of diene bridging and chain transfer results in a diene "ladder branched" trifunctional polymer structure. The term diene "ladder-branched" polymer refers to long chain branching, where a short chain or step links two polymer chains together. As shown, a metal-ligand catalyst having at least two polymer chain sites grows two separate polymer chains. One olefin of the diene is incorporated into one of the sites of the catalyst, and it is believed that due to the close proximity of the growth sites, the second olefin of the diene is then rapidly incorporated into the second polymer chain, forming a bridge or step. This continuous addition of diene is referred to as "synergistic" addition of diene, as distinguished from catalysts without two proximal chains, where diene addition results in a concentration of vinyl-containing polymer in the reactor that reacts at a later time. The term "step" refers to once the diene is incorporated into two separate polymer chains, linking the chains together. The first polymer chain and the second polymer chain continue to grow until the polymer is transferred to another catalyst, the polymer is released from the catalyst, the catalyst die, or another diene is added.

In one or more embodiments, the polymers of the present disclosure are ethylene-based copolymers comprising at least 50 mol% ethylene. In the present disclosure, "ethylene-based polymer" refers to homopolymers and/or interpolymers (including copolymers) of ethylene and optionally one or more comonomers, such as alpha-olefins, that may comprise at least 50 mole percent (mol%) of monomer units derived from ethylene. All individual values and subranges subsumed under "at least 50 mole percent" are disclosed herein as separate examples; for example, ethylene-based polymers, homopolymers and/or interpolymers (including copolymers) of ethylene, and optionally one or more comonomers such as alpha-olefins, can include: at least 60 mole percent of monomer units derived from ethylene; at least 70 mole percent of monomer units derived from ethylene; at least 80 mole percent of monomer units derived from ethylene; or 50 to 100 mole percent of monomer units derived from ethylene; or 80 to 100 mole percent of monomer units derived from ethylene.

Dynamics of

A mathematical model for tetrafunctional "ladder-branched" long chain branching was previously derived and described in the following applications: application No. PCTUS 2019/053524; number PCTUS 2019/053527; number PCTUS 2019/053529; and pct us2019/053537, each filed on 27/9/2019. Here, a model of trifunctional "ladder-branched" long chain branching was derived. The mathematical model will also be used to establish claim metrics and ranges. A mathematical model of the branching architecture as described in the present disclosure can be derived from the kinetic description of the proposed branching mechanism. This model is based on several assumptions that facilitate mathematical simplification, but these assumptions are not intended to limit the scope of the disclosure. The hypothesis follows the general industrial application of inactive addition of copolymers and additional hypothesis specific to the hypothesized diene branching mechanism. Common assumptions made include: (1) growth is much faster than chain transfer, so the average chain length is much longer than one monomer length; (2) only a single pure catalyst species is active; (3) the catalyst center produces many chains over its lifetime, and thus the chain lifetime is a fraction of the reaction or residence time; (4) when there is negligible compositional drift, the copolymerization can be approximated by a homopolymerization model.

Kinetics of diene trifunctional "ladder branching" theory

And (6) model derivation. The first step in the derivation of the model of the system is to write the kinetics in symbolic form, indicating the effect of each reaction on the molecular properties of interest. It is standard practice to use indices to indicate the number of repeating units associated with growing (living) or non-living polymer chains. Furthermore, it is also recognized that when homopolymer rate constants are considered to be effective composite copolymerization rate constants, the molecular architecture of addition copolymers can be accurately described by homopolymer kinetics and models (Tobita and Hamielec, polymers (polymers) 1991,32(14), 2641).

The kinetics of simple addition polymerization with a dual site catalyst are described below, where Pn,mIs an active catalyst center for growing two polymer molecules, the left molecule having n repeating units and the right molecule having m repeating units. Previous work has investigated the formation of tetrafunctional branches from dienes bridged across two resulting polymer molecules. The kinetics below consider the hypothesis of being trifunctional (b)3) Branched or difunctional bonds (b)2) Diene bridging results from a diene bridging across two growing molecules. Diene reactions such as cyclization or single insertion are considered non-productive and are ignored in the kinetic protocol.

Kinetics of the theory of trifunctional and bifunctional trapezoidal branching of dienes

a) Propagation of (left) Pn,m+M----->Pn+1,m kp(liter/mole/second)

(Right) Pn,m+M----->Pn,m+1 kp(liter/mole/second)

b) Chain transfer (left) Pn,m+A----->P0,m+Dn+kc ktra(liter/mole/second)

(Right) Pn,m+A----->Pn,0+Dm+kc ktra(liter/mole/second)

c) Diene bridge connection (left) Pn,m+D----->Pn+m,0+b3 2kd(liter/mole/second)

(Right) Pn,m+D----->P0,n+m+b3 2kd(liter/mole/second)

Pn,m+D----->P0,0+Dn+m+b2 4kg(liter/mole/second)

d) Reinitiation of Pn,0+M----->Pn,1Fast reaction

P0,m+M----->P1,mFast reaction

P0,0+2M----->P1,1Fast reaction

The kinetics of each of the two polymer molecules grown on the catalyst are depicted as labeled left and right. The result of propagation is that the molecule increases in size incrementally by one repeat unit, whether the propagation is to the left (P)n+1,m) Or on the right (P)n,m+1). The chain transfer reaction separates the chain from the catalyst and from the left (D)n) Side or right (D)m) Side-formed dead polymer molecules. Additional simple chain transfer type reactions such as hydrogenation or beta hydride elimination do not add to the complexity of the model.

Depicts the diene bridging reaction k on each catalyst sidedAnd due to the two reactive groups on the diene (D), the factor used per rate is 2. Since the diene bridging reaction k on both (2) sides is depicted oncegAnd diene (D) has two reactive sites, so it is used by a factor of 4. Thus, the rate constants for the diene consumption kinetics are defined on a group-by-group basis rather than on a molecular basis.

Standard practice is to assume that re-initiation of the polymer chains occurs very quickly and relatively infrequently with respect to propagation. By assuming transient re-initiation, the species P will be substantially reduced in the polymer populationn,0、P0,mAnd P0,0Excluded from consideration.

Population balance and rate. The kinetic scheme can be rendered as a series of equilibrium equations describing how each reaction affects the molecular architecture. In describing these balances, it is convenient to use shorthand nomenclature to express each reaction rate. These rate sets are defined as follows. The kinetic model can be extended to include other chain transfer reactions, such as by hydrogen (k) only by extending the definition of transfer termstrh) And beta hydride elimination (k)b) E.g. k ═ ktraA+ktrhH2+kb

Kinetic rate group definition: k ═ ktraA Ψ=kdD П=kgD Φ=kpM

The growing polymer (P) is described below using the kinetic set defined aboven,m) Molecule and non-living polymer molecule (D)n) The discrete population balance of (1) is that the molecular size is n is more than or equal to 1 and m is more than or equal to 1. These balances define the rate of change of the size of the molecular population pairs and can be modified to include additional convective terms if the balance is applied to a particular reactor environment or type. Using delta in these discrete balanceskThe term is meant to indicate that the term is only included when k is 0.

RPn,m=Φ(Pn-1,m+Pn,m-1–2Pn,m)–(2Ω+4Ψ+4П)Pn,mm-1(2ΨVn+ΩLn)+δn-1(2ΨVm+ΩRm)+δn-1δm-14Пξ0,0

Wherein:

other important population balances can be deduced from the above, such as left-hand (L)n) And right side (R)n) Growing polymer subspecies distribution and convolution distribution (V)n) Group balance of (1). Due to the definition of the kinetic protocolSymmetry is imposed so that the left and right growing polymer sub-species distributions are equal.

RLn=Φ(Ln-1-Ln)–(Ω+4Ψ+4П)Ln+2ΨVnn-1(Ω+2Ψ+4П)ξ0,0

RVn=2Φ(Vn-1-Vn)–(2Ω+4Ψ+4П)Vn+4ΨVn-1+2ΩLn-1n-24Пξ0,0

In which ξ0,0As a concentration of the total active catalyst,

wherein:

the first step in rendering a usable model is by assigning the relevant polymer sub-class Rates (RPs)n,m、RLn、RVn) The setting to zero is used to enforce a "steady state assumption" on the distribution of growing polymer species. This is a very common assumption in addition polymerization modeling when the growing chain lifetime is only a small fraction of the time period of interest. In most non-living industrial polymerizations of this type, the chain life is usually much less than one second, while the reactor residence time is at least a few minutes.

Method for predicting moments of MWD mean

Models describing the distribution moments of the chain lengths of the polymer species can generally be derived from population balances generated by kinetic protocols. Moment-based models can be used to predict molecular weight averages and polydispersity indices, but do not generally describe minor nuances in MWD, such as bimodality, peak MW, and tailing. The method of moments requires the definition of moments of chain length distribution for various subspecies of polymerization, as follows. Bulk polymer moment (λ)i) Reflecting bulk polymer properties, and solutions to models of bulk moments generally require solutions to various reactive polymer moments.

Living polymer moment:

bulk polymer MWD moment:

considering that the rate of change of the living species is assumed to be zero, the rate of change of the bulk moment is easily deduced from the population balance of the non-living polymers.

It is expected that any skilled polymer reaction engineer will be able to derive a moment model from a series of population balances. The dominant bulk polymer moment (. lamda.) is given below0,λ1,λ2) Wherein after applying the assumption that the kinetic chain is long, negligible terms are removed, and thus Φ>>Ω,Φ>>Ψ,Φ>>П。

0=(2Ω+4П)ξ0,01=(2Ω+8П)ξ1,02=(2Ω+8П)ξ2,0+8Пξ1,1

Evaluation of the rate of change of these bulk moments requires many living polymer sub-species moments. Due to the "steady state assumption", these living polymer moments are algebraic numbers and are given below. When predicting higher body moments such as λ3When this occurs, an additional moment of motion is required.

After algebraic simplification of the evaluation of the moment ratio, the instantaneous number and the weight average chain length (DP) are provided belown,DPw). Average molecular weight (M), of coursen,Mw) Equal to the average chain length multiplied by the apparent monomer repeat unit weight in grams/mole.

Amount of branching

By some substitution, e.g. average linear kinetic chain length DP without dienenoEqual to Φ/Ω, further simplifying the expression of the model. Also, by expression from a dimensionless instantaneous branch metric, e.g. FbIs a fraction of the diene attachment points that is difunctional, and can further simplify the model. Using FbAs a measure, it is reasonable because at varying diene levels, F can be expectedbIs fairly constant but will certainly vary with catalyst choice and may vary with reaction conditions.

Fraction of bifunctional diene attachment points

Additional measures are needed to describe the relative levels of branching, and two options are presented here. One preferred option is to use RcIt is the ratio of diene attachment points to the original polymer molecules. RcOne advantage of (a) is that it is simply a scaling of diene attachment points and is expected to increase in proportion to diene. RkcThe disadvantage of (a) is that the original kinetic chain length or concentration can usually only be directly obtained when a series of data comprising a level of zero diene branching is measured.

Diene attachment point per original kinetic chain

Measure RnIs a measure of branching RkcWherein R isnIs the ratio of diene attachment points to polymer molecules. The use of R is facilitated by the measurability of the chain length or concentration as measured by GPC of the number average molecular weightnTo analyze the data. However, RnIs not simply proportional to the diene since the bifunctional point of attachment affects the polymer moleculeThe number of the cells. In the absence of bifunctional coupling (F)b0), two metrics RkcAnd RnAre the same.

Diene attachment point per polymer molecule

The average chain length and molecular weight are described below with respect to polydispersity, where the diene-free polydispersity index is 2 due to assumed simplicity and kinetic ideality.

Once the model is rendered as Fb、RkcAnd RnEtc., several simple conclusions can be drawn from the above model. For example, the model shows the weight average chain length (DP)w) Or molecular weight (M)w) It is possible to increase only a maximum of two-fold in the case of diene incorporation. Any bifunctional linkage is expected to lower DPnOr MnAnd alleviate DPwOr MwAny increase in (a). Polydispersity at high trifunctional branching level (Z) starting with a zero diene polydispersity of 2p) Up to 4 and the effect via any bifunctional bond is mitigated.

FIG. 2 demonstrates diene attachment point function (F)b) Influence on the molecular weight and polydispersity of the polymer. The model clearly shows that pure trifunctional diene bridging has a limited dual potential impact on molecular weight and polydispersity, and that the incremental effect is at high diene levels such as RkcDecrease > 3. In addition, such asIf desired, a moderate level of difunctional diene attachment points is Fb5% or 10%, it is possible that experimental data may not even prove diene levels with MwA positive correlation therebetween.

Model of the complete MWD curve

Population balances of molecular weight distribution curves can sometimes be solved. Explicit algebraic solutions are usually only available if the reaction rate has no spatial or temporal variation, as assumed in this case. Of particular interest is the polymer distribution function DnWhich was previously used to render model equations of the moments of the ontology MWD. Also, instantaneous bulk polymer chain length distribution XnCan be expressed as follows:

instantaneous body chain length distribution:

due to the assumption of long chains, all species can be distributed (X)n、Dn、Ln、VnEtc.) are considered to be continuous rather than discrete functions. When the differential terms are replaced by derivatives, the steady state polymer species population balance can be approximated closely by a differential equation in the continuous variable n. For example, LnThe steady state population balance of (A) contains a difference term Ln–Ln-1The differential term is replaced by a derivative as shown below.

Similar alternatives yield a series of Ordinary Differential Equations (ODEs) that can be integrated to yield various defined distributions of Living sub-species L (n), and V (n) of chain length distributions. The model is outlined below as an initial value problem, where the chain length distribution function is assumed to start with n-0. The lower limit n of the distribution function is chosen to be 0 for mathematical simplicity only and ultimately does not have a significant effect on model prediction when forming high polymers.

Analytical solutions exist for the continuous distribution functions L (n) and V (n), and these functions can be used to render a function of the continuous bulk polymer chain length distribution (X (n)).

The solution for x (n) is somewhat complex, but can be simply expressed as follows:

the various assignments are as follows:

the branch metric (F) previously applied to the instantaneous average chain length and molecular weight model can be usedb、Rkc、Rn) To render alternative assignments of the x (n) function terms. The following X (n) term is denoted as FbAnd RkcAnd can be substituted by applying substitution Rkc=Rn/(1-Fb Rn) To convert to use Rn

α=1+Rkc+1/2Fb Rkc

The integral of x (n) can be used to indicate number average and weight average chain length as well as polydispersity.

Wherein:

as expected, the integration of distribution equation x (n) is in perfect agreement with the previously presented instantaneous moment model of average degree of polymerization and average molecular weight. The x (n) distribution model provides no additional or conflicting predictions of MWD average and polydispersity. However, this complete chain length distribution model can be used to understand how subtle differences in MWD are affected by diene addition. In particular, the model is able to predict the morphology, steepness and tailing of the MWD as a function of diene incorporation level and incorporation mode.

Limit cases of MWD model

There are two limiting cases for the chain length distribution model. When F is presentbWhen 1 and the polymer is perfectly linear, an insignificant situation occurs, with the most likely MWD. The average chain length of the most likely MWD obtained decreases with diene bridging level, due to Fb1, the diene bridging level is therefore entirely difunctional.

The limit of more interest is when there is no bifunctional bond (F)b0) and the amount of branching RkcAnd RnAre the same. Since each of the diene bonds is a trifunctional branch point, specific ones can be usedIn the nomenclature of branched polymers.

For Fb=0,BnBranch point per polymer molecule ═ Rkc=Rn

For Fb=0,BcBranch point per linear segment (1+ B)n)/Bn

The pure trifunctional branched chain length distribution pair B is shown belownAnd Bc

For Fb=0,X(n)=e-n(1+Bn)/DPno

For Fb=0,

The above distribution function may be integrated to yield F, shown belowbMWD average of 0. The number average molecular weight of such trifunctional branched systems does not change with increasing diene incorporation, since the branching reaction does not change the number of polymer molecules in the system.

The above polydispersity (M)w/Mn) The relationship to the level of trifunctional branching shows no instability or divergence at any level of branching. Most surprisingly, at high levels of branching, the polydispersity is expected to stabilize at 4. Of course, this prediction is for ideal copolymerization and symmetric catalyst systems, where any undesirable condition is expected to increase polydispersity.

The chain length distribution function can again be used to construct a predicted MWD curve. FIG. 3 is a series of simulated SEC curves in which the level of trifunctional branching (B)cOr Bn) Are variable. The independent variables in fig. 3 are scaled by linear molecular weight or chain length, making the plot versatile and independent of starting molecular weight. The null-split case in fig. 3 is the well-known "most likely" MWD and is expected for linear addition copolymerizations performed under ideal homogeneous conditions. FIG. 4 is a graph of the relative peak MW of the trifunctional diene branching, showing a MWD peak at 0.2<Bn<0.9 or 0.17<Bc<The approximate range of 0.5 is most sensitive to branching levels at intermediate branching levels.

Conventional branching model

The purpose of this section is to compare various conventional diene branched and random polymer couplings to the "ladder branching" model. The comparison shows the inherent instability of conventional diene branching and random polymer coupling, as opposed to "ladder branching". The molecular architecture resulting from diene "trapezoidal branching" is different from (a) the conventional diene Continuous Stirred Tank Reactor (CSTR) branching model, (b) the conventional diene semi-batch branching model; (c) polymer CSTR coupling model; and (d) a polymer batch coupling model.

a) Conventional diene CSTR branching model Ver strat-1980 (g.ver strat, c.cozewith, w.w.graessey, journal of applied polymer science (j.app.polym.sci.) -1980, 25,59), Guzman-2010(j.d.guzman, d.j.ariola, t.karjala, j.gaubert, b.w.s.kolthammer, journal of the american society of chemical engineers (AIChE) 2010,56, 1325):

b) conventional diene semi-batch branching models, Cozewalk-1979 (C.Cozewalk, W.W.Graessley, G.Ver Strate, chem.Eng.Sci.). 1979,34,245) and d) Polymer batch coupling models, Cozewalk-1979, Flory-1953(P.J.Flory, "Principles of Polymer Chemistry (Principles of Polymer Chemistry), Kannel University Press (Cornell University Press),1953), Tobita-1995(H.Tobita, J.Polymer.Sci.). B1995, 33, 1191):

c) polymer CSTR coupling model:

characterization of trifunctional Long-chain branched polyolefins

Depending on the degree of branching, various methods can determine the LCB (e.g., Nuclear Magnetic Resonance (NMR)), or differentiate the role of the LCB in the polymer. For example, the effect of LCB is observed in shear flow in the van gupu-palman analysis, shear viscosity also increases at low angular frequencies, and the intensity of shear thinning behavior can also be attributed to LCB. In extensional flow, the effect of LCB is generally identified in the degree of hard strain change or melt strength and the maximum deformation achieved. Other figures, such as Mark Houwink (Mark-Houwink) plot and g 'of expanded Molecular Weight Distribution (MWD)'visThe map provides additional information about the LCB. It is difficult to achieve high levels of natural LCB in polymers because of the limited concentration of vinyl terminated polymers (at most one per polymer chain) and the need to go to high conversions to ensure LCB formation. To ensure high conversion, the ethylene concentration in the reactor is low, thus enabling capping of a large amount of vinyl groupsReinserting the polymer of (a) into a second polymer chain.

NMR analysis is optimal to distinguish between trifunctional long chain branching and tetrafunctional long chain branching. Some dienes allow diagnostic measures to be taken on trifunctional long chain branching and tetrafunctional long chain branching. The mechanism in scheme 6 depicts the difference between the formation of trifunctional long chain branches and the formation of tetrafunctional long chain branches. In this case, the ratio of branching can be controlled by the ratio of ethylene to hydrogen in the reactor. In this particular example, dimethyldivinyl silane has a diagnostic methyl group on the silicon atom that can be used to determine trifunctional long chain branching or tetrafunctional long chain branching (see scheme 6 and figures 7-9). The carbon on the silicon of the tetrafunctional branched polymer is shifted towards a high magnetic field relative to the carbon on the silicon of the trifunctional branched polymer (see fig. 8). Examples will show that it is possible to control the ratio of trifunctional long-chain branches to tetrafunctional long-chain branches.

Scheme 6: the formation of trifunctional long-chain branches from the reaction of dienes is described.

In addition to hydrogenolysis, capping events such as beta-hydride elimination may also cause trifunctional long chain branching. If β -hydride elimination is the key mechanism, the unsaturation will appear as shown, for example, as vinylidene in scheme 7. Temperature effects can generally be used to control intramolecular (β -hydride) versus bimolecular (ethylene propagation) processes.

Scheme 7: the formation of trifunctional long-chain branches from the reaction of dienes followed by elimination of β -hydrides is described.

Conventional methods of incorporating dienes into polymer synthesis systems suffer from the fundamental drawbacks of gel formation or reactor fouling at high levels of branching. The kinetic model discussed in the previous paragraph may provide for achieving a better understanding of the gelGood predictive results are formed. For example, longer polymer chains proportionally have more pendant vinyl groups and polymer chains containing more pendant vinyl groups will be more likely to reinsert into the catalyst to form LCBs. Thus, larger polymer chains preferentially reinsert to form tetrafunctional branches, which are even larger polymer molecules and cause gel problems or instability when LCB levels reach a threshold. Weight average molecular weight (M) as a function of conventional tetrafunctional branching for ethylene-based polymers in semi-batch reactors at constant pressurew) And number average molecular weight (M)n) The simulation of (a) is shown in fig. 1. In FIG. 1, MnFollowing only MwBecomes infinite with a slight increase. In this example, with MwIncreasing to amounts greater than 200,000 grams per mole (g/mol), the polymer Molecular Weight Distribution (MWD) becomes unstable and gel formation begins. MWD is determined by the weight average molecular weight MwDivided by number average molecular weight Mn(Mw/Mn) And (4) defining.

For the purposes of this disclosure, a polymer gel is narrowly defined as a polymer fraction that phase separates due to its high branching level and/or high molecular weight. Polymer gels can be observed in solution or in the melt, and tend to interfere with properties such as optical clarity and film and fiber performance. Polyethylene interpolymer gels can be measured by the insolubility of the polymer in hot xylene. Gel content is generally related to and therefore estimated by GPC polymer recovery. When the polymer gel is formed, the polymer gel may deposit inside the reactor and cause fouling.

Measures have been previously disclosed to describe the high molecular weight tail effect when adding dienes to the polymerization and are described in the following applications: number PCTUS 2019/053524; number PCTUS 2019/053527; number PCTUS 2019/053529; and pct us2019/053537, each filed on 27/9/2019. Tetrafunctional "ladder branched" polymers do not show this tailing effect. A series of metrics have been previously disclosed, namely G (79/29), G (96/08), AHeight ofAnd ATail partWhich isThe amount of high MW polymer was quantified (see figure 5). The term "high MW tail" or "high molecular weight tail" refers to the high molecular weight fraction as indicated by conventional GPC and absolute GPC. Depending on the catalyst-diene pairing and experimental conditions, it can be expected that the "trapezoidal branching" system will have some conventional branching, making the shape metric value higher than would be expected for pure "trapezoidal branching".

From AHeight ofOr ATail partThe defined value increases sharply with increasing conventional branching levels. However, the "ladder branching" model (tetrafunctional or trifunctional) predicts a high MW area metric (A)Height ofOr ATail part) Hardly affected by the level of "trapezoidal branching". Most likely MWD AHeight ofAnd ATail partThe values were about 0.07 and 0.015, respectively. Exemplary MWD data will show that diene-free linear polymers tend to have a slightly higher A due to the non-ideal aspects of polymerizationHeight ofAnd ATail partThe value is obtained. The example data also shows various highly branched "ladder-branched" polymers in which substantially no high MW tail exceeds the expectation of the most likely MWD. The high MW area metric can also diagnose a slight level of high MW tail formation that a "ladder-branched" polymer can exhibit when accompanied by a degree of conventional branching. Measure ATail partLess than A affected by linear MWD non-idealitiesHeight of. However, in theory, AHeight ofAnd ATail partThe metric likewise indicates high MW tail formation.

Trifunctional long-chain branched polyolefins

As described in scheme 4, polymers resulting from "ladder branching" are included in the present disclosure.

In some embodiments, the polymers of the present disclosure have a trifunctional long chain branching level of greater than 0.1 per 1000 carbon atoms. In some embodiments, the polymers of the present disclosure have a level of trifunctional long chain branching greater than 0.2 per 1000 carbon atoms, greater than 0.3 per 1000 carbon atoms, greater than 0.4 per 1000 carbon atoms, or greater than 0.5 per 1000 carbon atoms.

In embodiments, the ethylene-based polymer of the present disclosure comprises a melt viscosity ratio at 190 ℃ of at least 10 orRheology ratio (V)0.1/V100) In which V is0.1Is the viscosity of the ethylene-based polymer at 190 ℃ at an angular frequency of 0.1 rad/sec, and V100Is the viscosity of the ethylene-based polymer at 190 ℃ at an angular frequency of 100 rad/sec. In one or more embodiments, the melt viscosity ratio is at least 14, at least 20, at least 25, or at least 30. In some embodiments, the melt viscosity ratio is greater than 50, at least 60, or greater than 100. In some embodiments, the melt viscosity ratio is 14 to 200.

"rheology ratio" and "melt viscosity ratio" from V at 190 ℃0.1/V100Definition of wherein V0.1Is the viscosity of the ethylene-based polymer at 190 ℃ at an angular frequency of 0.1 rad/sec, and V100Is the viscosity of the ethylene-based polymer at 190 ℃ at an angular frequency of 100 rad/sec.

In one or more embodiments, the ethylene-based polymers of the present disclosure have an average g 'of less than 0.86, wherein the average g' is the intrinsic viscosity ratio determined by gel permeation chromatography using a triple detector. In some embodiments, the ethylene-based polymer of the present disclosure has an average g' of 0.55 to 0.86. All individual values and subranges encompassed "0.55 to 0.86" are disclosed herein as separate embodiments; for example, the average g' of the ethylene-based polymer may be in the range of 0.64 to 0.75, 0.58 to 0.79, or 0.65 to 0.83. In one or more embodiments, the average g' is 0.55 to 0.84, 0.59 to 0.82, or 0.66 to 0.80.

In various embodiments, the melt strength of the ethylene-based polymers of the present disclosure can be greater than 6cN (Rheotens apparatus, 190 ℃, 2.4 mm/s)2120mm from the die exit to the wheel center, and an extrusion rate of 38.2s-1Capillary die length 30mm, diameter 2mm and entrance angle 180 °). In some embodiments, the melt strength of the ethylene-based polymer may be greater than 10 cN.

In embodiments, the ethylene-based polymer may have an area measurement a by MWDTail partA quantified molecular weight tail, and ATail partLess than or equal to 0.04. All individual values and subranges subsumed "less than or equal to 0.04" are disclosed herein as separate embodiments. For example, in some embodiments, a of the ethylene-based polymers of the present disclosureTail partGreater than 0 and less than or equal to 0.03 as determined according to gel permeation chromatography using a triple detector.

In one or more embodiments, the weight average molecular weight (M) of the polymers of the present disclosure as determined by gel permeation chromatography using a triple detectorw) And may be less than or equal to 800,000 daltons. In various embodiments, the weight average molecular weight (M) of the polymer as determined by gel permeation chromatography using a triple detectorw) And may be less than or equal to 400,000 daltons, less than or equal to 200,000 daltons, or less than or equal to 150,000 daltons.

In one or more embodiments, M of a polymer of the present disclosure as determined by gel permeation chromatography using a triple detectorw/Mn(weight average molecular weight/number average molecular weight) may be 6 or less. In various embodiments, the M of the polymer as determined by gel permeation chromatography using a triple detectorw/MnAnd may be less than 5 or less than 4. In some embodiments, the MWD of the long chain branching polymer is 1 to 3; and other embodiments comprise a MWD of 1.5 to 2.5.

As previously discussed, each Mw0And Mp0Is a measure of the polymer resin during polymerization without the addition of diene to the reactor. Each subsequent addition of diene produces a polymer resin from which a metric M can be determinedwOr Mp. The amount of diene incorporated into the reactor is small compared to the other reactants in the reactor. Thus, the addition of diene does not affect the total amount of comonomer, ethylene and solvent in the reactor.

In various embodiments, the ethylene-based polymer has a gpcBR branching index from 0.1 to 3.0. All individual values and subranges encompassed "0.10 to 3.00" are disclosed herein as separate embodiments; for example, the ethylene-based polymer may comprise a gpcBR branching index of 0.10 to 2.00, 0.10 to 1.00, 0.15 to 0.65, 0.20 to 0.75, or 0.10 to 0.95.

Olefins, mainly ethylene and propylene, are polymerized using the long chain branching polymerization process described in the preceding paragraph. In some embodiments, only a single type of olefin or a-olefin is present in the polymerization scheme, resulting in a homopolymer that is essentially a diene comonomer with a small amount of incorporation. However, additional alpha-olefins may be incorporated into the polymerization procedure. The additional alpha-olefin comonomer typically has no more than 20 carbon atoms. For example, the alpha-olefin comonomer may have 3 to 10 carbon atoms or 3 to 8 carbon atoms. Exemplary alpha-olefin comonomers include, but are not limited to, propylene, 1-butene, 1-pentene, 1-hexene, 1-heptene, 1-octene, 1-nonene, 1-decene, 4-methyl-l-pentene, and ethylidene norbornene. For example, the one or more alpha-olefin comonomers may be selected from the group consisting of propylene, 1-butene, 1-hexene, and 1-octene; or in the alternative, selected from the group consisting of 1-hexene and 1-octene.

Long chain branched polymers, such as homopolymers and/or interpolymers (including copolymers) of ethylene and optionally one or more comonomers such as alpha-olefins, may comprise at least 50 weight percent of units derived from ethylene. All individual values and subranges subsumed under "at least 50 weight percent" are disclosed herein as separate embodiments; for example, ethylene-based polymers, homopolymers and/or interpolymers (including copolymers) of ethylene, and optionally one or more comonomers such as alpha-olefins, can include: at least 60 weight percent of units derived from ethylene; at least 70 weight percent of units derived from ethylene; at least 80 weight percent of units derived from ethylene; or from 50 to 100 weight percent of units derived from ethylene; or 80 to 100 weight percent of units derived from ethylene.

In some embodiments of the ethylene-based polymer, the ethylene-based polymer comprises an additional alpha-olefin. An amount of additional alpha-olefin in the ethylene-based polymer is less than or equal to 50 mole percent (mol%); in other embodiments, the amount of additional alpha-olefin comprises at least 0.01 mol% to 25 mol%; and in further embodiments, the amount of additional alpha olefin comprises at least 0.1 mol% to 10 mol%. In some embodiments, the additional alpha olefin is 1-octene.

In some embodiments, the long chain branched polymer may comprise at least 50 mole percent of units derived from ethylene. All individual values and subranges from at least 90 mole percent are included herein and disclosed herein as separate examples. For example, the ethylene-based polymer may comprise at least 93 mole percent of units derived from ethylene; at least 96 mole percent units; at least 97 mole percent of units derived from ethylene; or in the alternative, from 90 to 100 mole percent of units derived from ethylene; 90 to 99.5 mole percent of units derived from ethylene; or 97 to 99.5 mole percent of units derived from ethylene.

In some embodiments of the long chain branched polymer, the amount of additional alpha olefin is less than 50%; other embodiments comprise at least 1 mole percent (mol%) to 20 mol%; and in further embodiments, the amount of additional alpha-olefin comprises at least 5 mol% to 10 mol%. In some embodiments, the additional alpha olefin is 1-octene.

Any conventional polymerization method may be employed to produce the long chain branched polymer. Such conventional polymerization processes include, but are not limited to, for example, solution polymerization processes, gas phase polymerization processes, slurry phase polymerization processes, and combinations thereof using one or more conventional reactors, such as loop reactors, isothermal reactors, fluidized bed gas phase reactors, stirred tank reactors, batch reactors in parallel or series, or any combination thereof.

In one embodiment, the ethylene-based polymer may be produced via solution polymerization in a dual reactor system, such as a single loop reactor system, wherein ethylene and optionally one or more alpha-olefins are polymerized in the presence of a catalyst system as described herein and optionally one or more co-catalysts. In another embodiment, the ethylene-based polymer can be produced by solution polymerization in a dual reactor system (e.g., a dual loop reactor system), wherein ethylene and optionally one or more alpha olefins are polymerized in the presence of the catalyst system and as described herein and optionally one or more other catalysts. The catalyst system as described herein may be used in the first reactor or the second reactor, optionally in combination with one or more other catalysts. In one embodiment, the ethylene-based polymer can be produced by solution polymerization in a dual reactor system (e.g., a dual loop reactor system), wherein ethylene and optionally one or more alpha olefins are polymerized in both reactors in the presence of a catalyst system as described herein.

In another embodiment, the long chain branched polymer may be produced via solution polymerization in a single reactor system, for example a single loop reactor system, wherein ethylene and optionally one or more a-olefins are polymerized in the presence of a catalyst system as described within this disclosure and optionally one or more co-catalysts as described in the preceding paragraph. In some embodiments, a long chain branching polymerization process for producing a long chain branched polymer comprises polymerizing ethylene and at least one additional alpha olefin in the presence of a catalyst system.

The long chain branched polymer may further comprise one or more additives. Such additives include, but are not limited to, antistatic agents, color enhancers, dyes, lubricants, pigments, primary antioxidants, secondary antioxidants, processing aids, UV stabilizers, and combinations thereof. The ethylene-based polymer may contain any amount of additives. The ethylene-based polymer may comprise from about 0 percent to about 10 percent of the total weight of such additives, by weight of the ethylene-based polymer and the one or more additives. The ethylene-based polymer may further include a filler, which may include, but is not limited to, organic or inorganic fillers. The long chain branched polymer may contain from about 0 to about 20 weight percent of a filler, such as calcium carbonate, talc, or Mg (OH), based on the combined weight of the ethylene-based polymer and all additives or fillers2. The ethylene-based polymer may be further blended with one or more polymers to form a blend.

In some embodiments, a long chain polymerization process for producing a long chain branched polymer may comprise polymerizing ethylene and at least one additional alpha olefin in the presence of a catalyst having two polymer production sites. The density of the long chain branched polymer produced from this catalyst having two polymer production sites can be, for example, 0.850g/cm according to ASTM D792 (incorporated herein by reference in its entirety)3To 0.960g/cm3、0.880g/cm3To 0.920g/cm3、0.880g/cm3To 0.910g/cm3Or 0.880g/cm3To 0.900g/cm3

In another embodiment, the long chain branched polymer produced by the long chain polymerization process may have a melt flow ratio (I) of 5 to 10010/I2) Wherein melt index I is measured according to ASTM D1238 (incorporated herein by reference in its entirety) at 190 ℃ and 2.16kg load2And the melt index I is measured according to ASTM D1238 at 190 ℃ and under a load of 10kg10. In other embodiments, the melt flow ratio (I)10/I2) From 5 to 50, and in other embodiments from 5 to 25, and in other embodiments from 5 to 9.

Gel permeation chromatography (conventional GPC)

The chromatographic system consisted of a PolymerChar GPC-IR (valencia, spain) high temperature GPC chromatograph equipped with an internal IR5 infrared detector (IR5) coupled to a Precision detector company (Precision Detectors) (now Agilent Technologies) 2-angle laser Light Scattering (LS) detector model 2040 and a 4-capillary viscometer (DV). For all absolute light scattering measurements, a 15 degree angle was used for the measurement. The autosampler oven chamber was set at 160 ℃ and the column chamber at 150 ℃. The column used was a 4 Agilent "Mixed a" 30cm 20 micron linear Mixed bed column. The chromatographic solvent used was 1,2, 4-trichlorobenzene and contained 200ppm of Butylated Hydroxytoluene (BHT). The solvent source was nitrogen sparged. The injection volume used was 200 microliters and the flow rate was 1.0 milliliters/minute.

Calibration of a GPC column set was performed with at least 20 narrow molecular weight distribution polystyrene standards having molecular weights ranging from 580 to 8,400,000 and arranged in 6 "cocktail" mixtures with at least ten times the separation between individual molecular weights. The standards were purchased from Agilent Technologies. Polystyrene standards are prepared in an amount of 0.025 grams in 50 milliliters of solvent having a molecular weight equal to or greater than 1,000,000, and in an amount of 0.05 grams in 50 milliliters of solvent having a molecular weight less than 1,000,000. Gently stir for 30 minutes at 80 ℃ to dissolve the polystyrene standards. Polystyrene standard peak molecular weights were converted to polyethylene molecular weights using equation 48 (as described by Williams and Ward in polymer science press (sci., polymer.let.), volume 6, page 621 (1968)):

Mpolyethylene=A×(MPolystyrene)B (48)

Where M is molecular weight, A has a value of 0.4315 and B equals 1.0.

A polynomial between 3 and 5 orders is used to fit the corresponding polyethylene equivalent calibration points. A small adjustment (from about 0.415 to 0.44) was made to a to correct for column resolution and band broadening effects, resulting in NIST standard NBS 1475 at 52,000 Mw.

Plate counts of the GPC column set were performed with eicosane (0.04 g prepared in 50 ml TCB and dissolved for 20 minutes with slow stirring). Plate count (equation 49) and symmetry (equation 50) were measured at 200 μ l injection according to the following equations:

where RV is the retention volume in milliliters, the peak width in milliliters, the peak maximum is the maximum height of the peak, and 1/2 height is the 1/2 height of the peak maximum.

Wherein RV is the retention volume in milliliters and the peak width is in milliliters, the peak maximum is the maximum position of the peak, one tenth of the height is 1/10 of the height of the peak maximum, and wherein the posterior peak refers to the tail of the retention volume later than the peak maximum, and wherein the anterior peak refers to the retention volume earlier than the peak of the peak maximum. The plate count of the chromatography system should be greater than 24,000 and the degree of symmetry should be between 0.98 and 1.22.

The samples were prepared in a semi-automated fashion using PolymerChar "Instrument Control" software, with the target weight of the sample set at 2mg/ml, and the solvent (containing 200ppm BHT) was added by a PolymerChar high temperature autosampler to a pre-nitrogen sparged vial capped with a septum. The sample was dissolved at 160 ℃ for 2 hours under "slow" shaking.

PolymerChar GPCOne was used based on GPC results using an internal IR5 detector (measurement channel) of a PolymerChar GPC-IR chromatograph according to equations 51-53TMSoftware, IR chromatograms at each equally spaced data collection point (i) minus baseline, and polyethylene equivalent molecular weights obtained from narrow standard calibration curves for point (i) according to equation 1n(GPC)、Mw(GPC)And Mz(GPC)And (4) calculating.

To monitor the time-varying deviation, a flow rate marker (decane) was introduced into each sample via a micropump controlled with a PolymerChar GPC-IR system. This flow rate marker (FM) was used to linearly correct the pump flow rate for each sample by(flow rate (nominal)): the RV of the corresponding decane peak within the sample (RV (FM sample)) was compared to the retention volume of the decane peak within the narrow standard calibration (RV (FM calibration)). Then, it was assumed that any change in decane marker peak time was related to a linear change in flow rate (effective)) throughout the run. To facilitate the highest accuracy of RV measurements of the flow marker peaks, the peaks of the flow marker concentration chromatogram were fitted to a quadratic equation using a least squares fitting procedure. The true peak position is then solved using the first derivative of the quadratic equation. After calibrating the system based on flow marker peaks, the effective flow rate (calibrated against narrow standards) is calculated as in equation 7. By PolymerChar GPCOneTMThe software completes the processing of the flow marker peak. Acceptable flow rate corrections are such that the effective flow rate should be within +/-2% of the nominal flow rate.

Flow rate (effective) to flow rate (nominal) (RV)(calibrated FM)/RV(FM sample)) (54)

Triple detector GPC (absolute GPC)

The chromatography system, operating conditions, column set-up, column calibration and calculation of the conventional molecular weight moments and distributions were performed according to the methods described in Gel Permeation Chromatography (GPC).

To determine the offset of the viscometer and light scatter detector relative to the IR5 detector, the systematic method for determining the multi-detector offset is performed in a manner consistent with that disclosed by Balke, Mourey et al. (Chalk 12 (1992) chromatographic Polymer (chromatographic Polymer), Mourey and Balke) (Balke, Thiiritsaku, Lew, Cheung, Mourey, chromatographic Polymer 13 (1992)), using PolymerChar GPCOneTMSoftware optimized polyethylene standards from broad homopolymer (M)w/Mn>3) And the results of the triple detector logarithm (MW and IV) with the results of the narrow standard column calibration from the narrow standard calibration curve.

Absolute molecular weight data in comparison to Zimm (Zimm, b.h., journal of chemico-physical (j.chem.phys.), (16, 1099) (1948)) and Kratochvil (Kratochvil, p., "classic Light Scattering of polymer solutions" (classic Light Scattering fr)From Polymer Solutions, Elsevier, Oxford, New York (1987)) in a manner consistent with that published in Polymer char GPCOneTMAnd (4) obtaining software. The total injected concentration for determining molecular weight is obtained from the mass detector area and the mass detector constant from one of a suitable linear polyethylene homopolymer or a polyethylene standard of known weight average molecular weight. Calculated molecular weight (using GPCOne)TM) Obtained using the light scattering constant and index concentration coefficient dn/dc 0.104 from one or more polyethylene standards mentioned below. In general, the mass detector response (IR5) and the light scattering constant (using GPCOne)TMDetermined) can be determined by linear standards having molecular weights in excess of about 50,000 grams/mole. Viscometer calibration (using GPCOne)TMDetermination) may be accomplished using the methods described by the manufacturer, or alternatively, by using published values for suitable linear Standards, such as Standard Reference Material (SRM)1475a, available from the National Institute of Standards and Technology, NIST. Calculation of viscometer constants (using GPCOne)TMObtained) that correlates the specific viscosity area (DV) and injection mass for the calibration standard to its intrinsic viscosity. The chromatographic concentrations were assumed to be low enough to eliminate the effect of the 2 nd viral coefficient (effect of concentration on molecular weight).

Absolute weight average molecular weight (M)w(Abs)) Light Scattering (LS) integrated chromatogram (determined by light scattering constants) divided by mass recovered from mass constants and mass detector (IR5) region (using GPCOne)TM) And (4) obtaining. Molecular weight and intrinsic viscosity response at the chromatographic end where signal to noise ratio becomes low (using GPCOne)TM) And (6) linear extrapolation. From equations 55-56, the other respective moments M are calculated as followsn(Abs)And Mz(Abs)

Dynamic mechanical spectrum (or small angle oscillation shear)

The complex viscosity (. eta.), modulus (G', G "), tan. delta. and phase angle (. delta.) were obtained by dynamic oscillatory frequency sweep testing at 190 ℃ in a frequency range of 0.1 to 100 rad/sec. The strain level was set in the linear viscoelastic range as identified by strain sweep testing at 100 rad/sec at 190 ℃. The tests were performed on a strain controlled rheometer ARES-G2 from thermal Analyzer, Inc. (TA Instruments) using 25mm diameter stainless steel parallel plates. Samples 3.3mm thick were extruded and then trimmed in two steps before actual testing. In the first step, the sample was melted for 2.5 minutes, squeezed into a 3mm gap and trimmed. After an additional 2.5 minutes of soaking at 190 ℃, the sample was squeezed into a 2mm gap and excess material was trimmed. The method has an additional built-in five minute delay to allow the system to reach thermal equilibrium. The test was performed under a nitrogen atmosphere.

gpcBR branching index by Triple Detector GPC (TDGPC)

The gpcBR branching index is determined by: as previously described, the light scattering, viscosity and concentration detectors are first calibrated. The baseline was then subtracted from the light scattering, viscometer and concentration chromatogram. An integration window is then set to ensure that all low molecular weight retention volume ranges are integrated into the light scattering and viscometer chromatograms indicating the presence of polymer detectable from the refractive index chromatogram. Polyethylene and polystyrene Markov Hooke constants were then established using linear polyethylene standards. After obtaining the constants, these two values are used to construct two linear reference conventional calibration values for polyethylene molecular weight and polyethylene intrinsic viscosity as a function of elution volume, as shown in equations (57) and (58):

the gpcBR branching index is a robust method for characterizing long chain branching, as described in Yau, Wallace w., "an example of polyolefin Characterization Using 3D-GPC-TREF" (samples of Using 3D-GPC-TREF for Poly-olefin Characterization) "," the discussion of macromolecules (macromol. symp.), "2007, 257, 29-45. The index avoids the "layer-by-layer" TDGPC calculations traditionally used to determine g' values and branching frequency calculations, favoring the entire polymer detector area. From the TDGPC data, the absolute weight average molecular weight (M) of the sample bulk can be obtained by a Light Scattering (LS) detector using a peak area methodwAbs). The method avoids a "layer-by-layer" ratio of light scatter detector signal to concentration detector signal, as required in conventional g' determination. In the case of TDGPC, the sample intrinsic viscosity can also be obtained independently using equation (59). In this case, the area calculation provides greater accuracy because as a whole sample area it is less sensitive to detector noise and variations caused by TDGPC settings over baseline and integration limits. More importantly, the peak area calculation is not affected by detector volume shifts. Similarly, a high accuracy of the Intrinsic Viscosity (IV) of the sample is obtained by the area method in equation (59):

in equation (59), DPi represents the differential pressure signal monitored directly from the in-line viscometer. To determine the gpcBR branching index, the light scattering elution area of the sample polymer was used to determine the molecular weight of the sample. The viscosity detector elution area of the sample polymer is used to determine the intrinsic viscosity (IV or eta) of the sample. First, the molecular weight and intrinsic viscosity of linear polyethylene standard samples (such as SRM1475a) or equivalents were determined using conventional calibration ("cc") for both molecular weight and intrinsic viscosity as a function of elution volume:

equation (61) is used to determine the gpcBR branching index:

wherein [ eta ]]Is the measured intrinsic viscosity, [ eta ]]ccIs the intrinsic viscosity from a conventional calibration (or conv GPC), Mw is the measured weight average molecular weight, and Mw,ccIs the weight average molecular weight of the conventional calibration. The weight average molecular weight by Light Scattering (LS) is commonly referred to as the "absolute weight average molecular weight" or "Mw(abs) ". M from molecular weight calibration curves using conventional GPC ("conventional calibration")w,ccCommonly referred to as "polymer chain backbone molecular weight", "conventional weight average molecular weight", and "Mw(conv)”。

All statistical values with the "cc or conv" subscripts were determined using their respective elution volumes, the corresponding conventional calibrations and concentrations (Ci) described previously. The non-subscript values are based on measurements of mass detector, LALLS, and viscometer area. Iterative adjustment KPEUntil the gpcBR measurement of the linear reference sample is zero. For example, in this particular case, the final values of α and Log K for determining gpcBR are 0.725 and-3.355 for polyethylene, and 0.722 and-3.993 for polystyrene, respectively. Once the K and alpha values are determined using the procedure in question.

Previously, the procedure was repeated using branched samples. The branched samples were analyzed using the final Markov Hooke constant as the best "cc" calibration.

The explanation for gpcBR is straightforward. For linear polymers, gpcBR will be close to zero since the values measured by LS and viscometry will be close to conventional calibration standards. For branched polymers, gpcBR will be higher than zero, especially at higher levels of long chain branching, since the measured polymer molecular weight will be higher than the calculated Mw,ccAnd calculated IVccWill be higher than the measured polymer IV. In fact, the gpcBR value represents the fraction due to polymer branchingFractional IV changes due to sub-size contraction effects. A gpcBR value of 0.5 or 2.0 would mean the molecular size shrinking effect of IV at the level of 50% and 200%, respectively, relative to an equivalent weight of linear polymer molecules. For these particular examples, the advantage of using gpcBR is the higher precision of gpcBR compared to traditional "g' index" and branching frequency calculations. All parameters used in the gpcBR index determination achieve high accuracy and are not adversely affected by the low TDGPC detector response from the concentration detector under high molecular weight conditions. Errors in detector volume alignment will not affect the accuracy of the gpcBR index determination.

NMR analysis

And (4) preparing a sample. The original polymer sample containing solvent and catalyst residues must be removed prior to NMR measurements for unsaturation and branching. The polymer was first dissolved in Tetrachloroethane (TCE) at 120-125 deg.C and then precipitated using 3-propanol (IPA) and cooled to room temperature. The polymer was separated by centrifugation. This process of washing the polymer was repeated at least 3 times. The resulting polymer was then dried in a vacuum oven at 50 ℃.

Approximately 70mg of the washed and dried polymer was placed in a 10mm NMR tube with 2.8ml of TCE. The sample was purged with nitrogen through the bubbling chamber of the sample for 15 minutes. The purged sample was then placed in an aluminum heating block at 125 ℃.

For branching analysis, a single pulse of sample was taken at 120 ℃ using a 600MHz Bruker Avance III HD mass spectrometer equipped with a 10mm 13C/1H DUL CryoProbe13C NMR spectra to collect between 1400 and 5000 scans with 90 pulses and a total relaxation delay (AQ + D1) of 10 seconds.

For the quantification of tetrafunctional and trifunctional long chain branches in polymers using dimethyldivinylsilane as diene, a single pulse of sample acquisition at 120 ℃ was performed using a 600MHz Bruker Avance III HD Mass spectrometer equipped with a 10mm 13C/1H DUL CryoProbe13C NMR spectra to collect 960 to 5000 scans with 90 pulses and a total relaxation delay (AQ + D1) of 12 seconds. Alternatively, QA-RIN is acquired using a relaxation delay of 7 secondsEPT spectra (J.Hou, Y.He, X.Qiu, Macromolecules (Macromolecules), 2017,50, 2407-2414). The parameters QA-RINEPT are selected to match the methyl to total carbon ratio of QA-RINEPT to the single pulse data.

Data processing and assignment method. All NMR data were processed using Mnova with a proton spectrum line broadening of 0.5HZ and a carbon spectrum line broadening of 3 HZ. The quality data was used as a reference for 5.99ppm TCE solvent resonance. Carbon Spectrum as the predominant CH of 29.99ppm of the Polymer2Reference to (3).

The hypothetical assignment of trifunctional LCB silyl-methyl (-3.36ppm) and tetrafunctional LCB silyl-methyl (-4.06ppm) was obtained using the ACD CNMR predictor and was found to be closely consistent with the observed resonance. Using a branched methine resonance (24.9ppm tetrafunctional, 25.7ppm trifunctional) with CH2These assignments were confirmed by a quantitative relationship between the resonances of carbon α to Y-branched silicon (about 15.3 ppm).

Batch reactor polymerization procedure

Batch reactor polymerization at 2L ParrTMIn a batch reactor. The reactor was heated by an electrical heating mantle and cooled by an internal serpentine cooling coil containing cooling water. Passing through CamileTMThe TG process computer controls and monitors the reactor and heating/cooling system. The bottom of the reactor was fitted with a dump valve that emptied the reactor contents into a stainless steel dump kettle. The decanters are pre-filled with a catalyst kill solution (typically 5mL of an Irgafos/Irganox/toluene mixture). The dump kettle was discharged into a 30 gallon blowdown tank, with both the kettle and tank purged with nitrogen. All solvents used for polymerization or catalyst make-up were run through a solvent purification column to remove any impurities that could affect polymerization. 1-octene and IsoparE were passed through two columns, the first containing a2 alumina and the second containing Q5. Ethylene was passed through two columns, the first containing A204 alumina andmolecular sieves, second column containing Q5 reactant. N used for transfer2Through oxidation containing A204Aluminum, and aluminum, and aluminum, and aluminum,molecular sieves and a single column of Q5.

Depending on the reactor loading, the reactor is first loaded from a jet tank that may contain IsoparE solvent and/or 1-octene. The spray canister is filled to the load set point by using a laboratory scale with the spray canister installed. After addition of the liquid feed, the reactor was heated to the polymerization temperature set point. If ethylene is used, it is added to the reactor while it is at the reaction temperature to maintain the reaction pressure set point. The ethylene addition was monitored by a Micro-Motion flow meter (Micro Motion). For some experiments, the standard conditions at 150 ℃ were 585g IsoparE with 13g ethylene, 15g 1-octene, 240psi hydrogen, and the standard conditions at 150 ℃ were 555g IsoparE with 15g ethylene, 45g 1-octene, 200psi hydrogen.

The procatalyst and activator are mixed with the appropriate amount of purified toluene to obtain the desired molarity solution. The procatalyst and activator were processed in an inert glove box, drawn into a syringe and transferred under pressure to a catalyst injection tank. The syringe was rinsed three times with 5mL of toluene. Running the timer was started immediately after the catalyst was added. If ethylene is used, ethylene is added through Camile to maintain the reaction pressure set point in the reactor. The polymerization was run for 10 minutes, then the agitator was stopped and the bottom dump valve was opened to empty the reactor contents into the dump pot. The contents of the pour pan were poured into a tray and placed in a laboratory fume hood where the solvent was evaporated overnight. The tray containing the remaining polymer was transferred to a vacuum oven where it was heated to 140 ℃ under vacuum to remove any residual solvent. After the tray was cooled to ambient temperature, the yield of polymer was weighed to measure efficiency and submitted for polymer testing.

Examples of batch reactors

Batch reactor example 1

In table 2, the polymer properties of the comparative linear polymer sample (1.C) are compared to the branched polymer of the batch reactor. The polymerization took place at a temperature of 150 ℃ and 585g of ISOPAR-ETM15g of octene neutralized 240psi of hydrogen pressure (. DELTA.H)2) The following steps. 14g of ethylene were loaded and the pressure maintained in the presence of 0.3. mu. mol of catalyst 1, 0.36. mu. mol of cocatalyst A (methyldietetradecylammonium tetrakis (pentafluorophenyl) borate) and 10. mu. mol of MMAO-3A. The addition of the dienyldivinylsilane was as indicated in the table.

Table 2: example 1 and comparative batch reactor polymer run and properties.

Table 2: continuously for

Data for comparative example 1.C and other diene examples 1.1-1.7 are collected in Table 2. NMR data confirmed that both trifunctional and tetrafunctional LCBs and LCB levels increased with increasing diene.

Fig. 6 depicts the conventional molecular weight distribution of the examples with varying amounts of diene.

Batch reactor example 2

In table 3, the polymer properties of the comparative linear polymer sample (2.C) are compared to the branched polymer of the batch reactor. The polymerization took place at a temperature of 150 ℃ and 585g of ISOPAR-ETM15g of octene neutralized 240psi of hydrogen pressure (. DELTA.H)2) The following steps. In the presence of 0.3. mu. mol of catalyst 1, 0.36. mu. mol of cocatalyst A (methyldietetradecylammonium tetrakis (pentafluorophenyl) borate) and 10. mu. mol of MMAO-3A, 15g of ethylene were charged and the pressure was maintained. As indicated in the tableAdding diene dimethyl divinyl silane.

Table 3: example 2 and comparative batch reactor polymer run and characteristics.

Table 3: continuously for

Data for comparative example 2.C and diene example 2.1 are collected in table 3. NMR data confirmed a 1.4:1 ratio of trifunctional (0.15LCB/1000C) to tetrafunctional (0.10LCB/1000C) to tetrafunctional (tri: tetraLCB).

The dynamic mechanical spectrum of branched example 2.1 was measured and the results are recorded in table 3. The viscosity at 0.1 rad/sec was determined to be 609,361Pa s and the viscosity at 100 rad/sec was measured to be 2,453Pa s, providing a rheology ratio (V) of 248.40.1/V100)。

Batch reactor example 3

In table 4, the polymer properties of the comparative linear polymer sample (3.C) are compared to the branched polymer of the batch reactor. The polymerization took place at a temperature of 140 ℃ and 585g of ISOPAR-ETM15g of octene neutralized 240psi of hydrogen pressure (. DELTA.H)2) The following steps. In the presence of 0.3. mu. mol of catalyst 1, 0.36. mu. mol of cocatalyst A (methyldietetradecylammonium tetrakis (pentafluorophenyl) borate) and 10. mu. mol of MMAO-3A, 10g of ethylene were loaded and the pressure was maintained. The addition of the dienyldivinylsilane was as indicated in the table.

Table 4: example 3 and comparative batch reactor polymer run and properties.

Table 4: continuously for

Data for comparative example 3.C and other dienes example 1.1 are collected in table 4. NMR data confirmed a 1.8:1 ratio of trifunctional (0.23LCB/1000C) to tetrafunctional (0.13LCB/1000C) to both.

The dynamic mechanical spectrum of branched example 3.1 was measured and the results are recorded in table 4. The viscosity at 0.1 rad/sec was determined to be 515,022 pas and the viscosity at 100 rad/sec was measured to be 2,140 pas, providing a rheology ratio (V) of 240.60.1/V100)。

Batch reactor example 4

In table 5, the polymer properties of the comparative linear polymer sample (4.C) are compared to the branched polymer of the batch reactor. The polymerization took place at a temperature of 150 ℃ and 585g of ISOPAR-ETM15g of octene neutralized 160psi of hydrogen pressure (. DELTA.H)2) The following steps. In the presence of 0.4. mu. mol of catalyst 1, 0.48. mu. mol of cocatalyst A (methyldietetradecylammonium tetrakis (pentafluorophenyl) borate) and 10. mu. mol of MMAO-3A, 15g of ethylene were charged and the pressure was maintained. The addition of the dienyldivinylsilane was as indicated in the table.

Table 5: example 4 and comparative batch reactor polymer run and characteristics.

Table 5: continuously for

Data for comparative example 4.C and diene example 4.1 are collected in table 5. NMR data confirmed a 1.03:1 ratio of trifunctional (0.31LCB/1000C) to tetrafunctional (0.30LCB/1000C) to tetrafunctional (tri: tetraLCB).

MeasuringThe dynamic mechanical spectrum of branching example 4.1 is shown and the results are reported in table 5. The viscosity at 0.1 rad/sec was determined to be 867,379 pas and the viscosity at 100 rad/sec was measured to be 2,818 pas, providing a rheology ratio (V) of 307.80.1/V100)。

Batch reactor example 5

In table 6, the polymer properties of the comparative linear polymer sample (5.C) are compared to the branched polymer of the batch reactor. The polymerization took place at a temperature of 160 ℃ and 585g of ISOPAR-ETM15g of octene neutralized 80psi of hydrogen pressure (. DELTA.H)2) The following steps. In the presence of 0.4. mu. mol of catalyst 1, 0.48. mu. mol of cocatalyst A (methyldietetradecylammonium tetrakis (pentafluorophenyl) borate) and 10. mu. mol of MMAO-3A, 15g of ethylene were charged and the pressure was maintained. The addition of the dienyldivinylsilane was as indicated in the table.

Table 6: example 5 and comparative batch reactor polymer run and characteristics.

Table 6: continuously for

Data for comparative example 5.C and diene example 5.1 are collected in table 6. NMR data confirmed a 0.8:1 ratio of trifunctional (0.25LCB/1000C) to tetrafunctional (0.30LCB/1000C) to both.

The dynamic mechanical spectrum of branching example 5.1 was measured and the results are recorded in table 6. The viscosity at 0.1 rad/sec was determined to be 813,746Pa s and the viscosity at 100 rad/sec was measured to be 2,742Pa s, providing a rheology ratio (V) of 296.70.1/V100)。

Batch reactor example 6

In Table 7, the polymerization of example 6.1 took place at a temperature of 150 ℃ with 585g of ISOPAR-ETM15g of octene neutralized 240psi of hydrogen pressure (. DELTA.H)2) The following steps. 13g of ethylene were loaded and the pressure maintained in the presence of 0.3. mu. mol of catalyst 1, 0.36. mu. mol of cocatalyst A (methyldietetradecylammonium tetrakis (pentafluorophenyl) borate) and 10. mu. mol of MMAO-3A. The addition of the dienyldivinylsilane was as indicated in the table. In Table 7, the polymerization of example 6.2 took place at a temperature of 160 ℃ and 600g of ISOPAR-ETMMedium, no octene and a hydrogen pressure (Δ H) of 240psi2) The following steps. 13g of ethylene were loaded and the pressure maintained in the presence of 0.4. mu. mol of catalyst 1, 0.48. mu. mol of cocatalyst A (methyldietetradecylammonium tetrakis (pentafluorophenyl) borate) and 10. mu. mol of MMAO-3A. The addition of the dienyldivinylsilane was as indicated in the table. In Table 7, the polymerization of example 6.3 took place at a temperature of 160 ℃ and 600g of ISOPAR-ETMMedium, no octene and a hydrogen pressure (Δ H) of 240psi2) The following steps. 13g of ethylene were loaded and the pressure maintained in the presence of 0.4. mu. mol of catalyst 2, 0.48. mu. mol of cocatalyst A (methyldietetradecylammonium tetrakis (pentafluorophenyl) borate) and 10. mu. mol of MMAO-3A. The addition of the dienyldivinylsilane was as indicated in the table.

Table 7: polymer run and characteristics of the batch reactor polymer of example 6.

Table 7: continuously for

Data for diene examples 6.1, 6.2 and 6.3 are collected in table 7. Examples 6.2 and 6.3 indicate that different catalysts can form different amounts of trifunctional LCB and different ratios of trifunctional LCB to tetrafunctional LCB. For example, under the same conditions, the trifunctional LCB levels of catalyst 1 (example 6.2) and catalyst 2 (example 6.3) were 0.26LCB/1000C and 0.07LCB/1000C, respectively, and the trifunctional to tetrafunctional LCB ratios were 2.2:1 and 0.4:1, respectively. The amount of trifunctional LCB and the ratio of trifunctional to tetrafunctional LCB depends largely on the catalyst.

Examples 6.1 and 6.2 indicate that the polymerizations were run under comparable conditions, with the key difference being that example 6.1 contained octene, whereas example 6.2 did not contain octene. The amount of trifunctional LCB and the ratio of trifunctional to tetrafunctional LCB were very similar in both runs.

The dynamic mechanical spectrum of branching example 6.1 was measured and the results are recorded in table 7. The viscosity at 0.1 rad/sec was determined to be 306,441Pa s and the viscosity at 100 rad/sec was measured to be 1,754Pa s, providing a rheology ratio (V) of 174.70.1/V100)。

Batch reactor example 7

In table 8, the polymer properties of the comparative linear polymer sample (7.C) are compared to the branched polymer of the batch reactor. The polymerization took place at a temperature of 160 ℃ and 580g of ISOPAR-ETM20g of octene contained no hydrogen. 11g of ethylene were loaded and the pressure maintained in the presence of 0.7. mu. mol of catalyst 1, 0.84. mu. mol of cocatalyst A (methyldietetradecylammonium tetrakis (pentafluorophenyl) borate) and 10. mu. mol of MMAO-3A. The addition of the dienyldivinylsilane was as indicated in the table.

Table 8: example 7 and comparative batch reactor polymer run and characteristics.

Table 8: continuously for

Data for comparative example 7.C and diene example 7.1 are collected in table 8. NMR data confirmed that, in the absence of hydrogen, there is no trifunctional in this example. Tetrafunctional LCB is present (0.14LCB/1000C) and the ratio of trifunctional to tetrafunctional is zero.

The dynamic mechanical spectrum of example 7.1 was measured and the results are recorded in table 8. The viscosity at 0.1 rad/sec was determined to be 475,848Pa s and the viscosity at 100 rad/sec was measured to be 1,982Pa s, providing a rheology ratio (V) of 240.10.1/V100)。

Batch reactor example 8

In table 9, the polymer properties of the comparative linear polymer sample (8.C) are compared to the branched polymer of the batch reactor. The polymerization took place at a temperature of 160 ℃ and 580g of ISOPAR-ETM20g of octene neutralized 28psi of hydrogen pressure (. DELTA.H)2) The following steps. 13g of ethylene were loaded and the pressure maintained in the presence of 0.7. mu. mol of catalyst 1, 0.84. mu. mol of cocatalyst A (methyldietetradecylammonium tetrakis (pentafluorophenyl) borate) and 10. mu. mol of MMAO-3A. The addition of the dienyldivinylsilane was as indicated in the table.

Table 9: example 8 and comparative batch reactor polymer run and characteristics.

Table 9: continuously for

Data for comparative example 8.C and diene example 8.1 are collected in table 9. NMR data confirmed a ratio of 0.8:1 (tri: tetra LCB) for both trifunctional (0.09LCB/1000C) and tetrafunctional (0.11 LCB/1000C).

The dynamic mechanical spectrum of example 8.1 was measured and the results are recorded in table 9. The viscosity at 0.1 rad/sec was determined to be 721,022Pa s and the viscosity at 100 rad/sec was measured to be 2,297Pa s, providing a rheology ratio (V) of 313.80.1/V100)。

Batch reactor example 9

In Table 10, the polymerization of the linear polymer samples (9.C) will be comparedThe properties were compared to the branched polymer of a batch reactor. The polymerization took place at a temperature of 160 ℃ and 580g of ISOPAR-ETM20g of octene neutralized 46psi of hydrogen pressure (Δ H)2) The following steps. 13g of ethylene were loaded and the pressure maintained in the presence of 0.7. mu. mol of catalyst 1, 0.84. mu. mol of cocatalyst A (methyldietetradecylammonium tetrakis (pentafluorophenyl) borate) and 10. mu. mol of MMAO-3A. The addition of the dienyldivinylsilane was as indicated in the table.

Table 10: example 9 and comparative batch reactor polymer run and characteristics.

Table 10: continuously for

Data for comparative example 9.C and diene example 9.1 are collected in table 10. NMR data confirmed a 0.9:1 ratio of trifunctional (0.09LCB/1000C) to tetrafunctional (0.10LCB/1000C) to both.

The dynamic mechanical spectrum of example 9.1 was measured and the results are recorded in table 10. The viscosity at 0.1 rad/sec was determined to be 697,565 pas and the viscosity at 100 rad/sec was measured to be 2,782 pas, providing a rheology ratio (V) of 250.70.1/V100)。

Batch reactor example 10

In table 11, the polymer properties of the comparative linear polymer sample (10.C) are compared to the branched polymer of the batch reactor. The polymerization took place at a temperature of 160 ℃ and 575g of ISOPAR-ETM25g of octene and 83psi of hydrogen pressure (. DELTA.H)2) The following steps. 14g of ethylene were loaded and the pressure maintained in the presence of 0.6. mu. mol of catalyst 1, 0.72. mu. mol of cocatalyst A (methyldietetradecylammonium tetrakis (pentafluorophenyl) borate) and 10. mu. mol of MMAO-3A. Addition of diene dimethyl as indicated in the Table(ii) a divinylsilane.

Table 11: example 10 and comparative batch reactor polymer run and properties.

Table 11: continuously for

Data for comparative example 10.C and diene example 10.1 are collected in table 11. NMR data confirmed a 0.7:1 ratio of trifunctional (0.07LCB/1000C) to tetrafunctional (0.10LCB/1000C) to both.

The dynamic mechanical spectrum of example 10.1 was measured and the results are recorded in table 11. The viscosity at 0.1 rad/sec was determined to be 240,894 pas and the viscosity at 100 rad/sec was measured to be 1,642 pas, providing a rheology ratio (V) of 146.70.1/V100)。

Batch reactor example 11

In table 12, the polymer properties of the comparative linear polymer sample (11.C) are compared to the branched polymer of the batch reactor. The polymerization took place at a temperature of 150 ℃ and 575g of ISOPAR-ETM25g of octene neutralized 160psi of hydrogen pressure (. DELTA.H)2) The following steps. 14g of ethylene were loaded and the pressure maintained in the presence of 0.4. mu. mol of catalyst 1, 0.48. mu. mol of cocatalyst A (methyldietetradecylammonium tetrakis (pentafluorophenyl) borate) and 10. mu. mol of MMAO-3A. The addition of the dienyldivinylsilane was as indicated in the table.

Table 12: example 11 and comparative batch reactor polymer run and properties.

Table 12: continuously for

Data for comparative example 11.C and diene example 11.1 are collected in table 12. NMR data confirmed a 1.5:1 ratio of trifunctional (0.09LCB/1000C) to tetrafunctional (0.06LCB/1000C) to both.

The dynamic mechanical spectrum of branched example 11.1 was measured and the results are recorded in table 12. The viscosity at 0.1 rad/sec was determined to be 203,979 pas and the viscosity at 100 rad/sec was measured to be 1,523 pas, providing a rheology ratio (V) of 133.90.1/V100)。

The polymer of example 11.1 had a measured melt strength of 18cN and an extensibility of 32 mm/sec.

Batch reactor example 12

In table 13, the polymer properties of the comparative linear polymer sample (12.C) are compared to the branched polymer of the batch reactor. The polymerization took place at a temperature of 150 ℃ and 570g of ISOPAR-ETM30g of octene and 240psi of hydrogen pressure (. DELTA.H)2) The following steps. 20g of ethylene were loaded and the pressure maintained in the presence of 0.3. mu. mol of catalyst 1, 0.36. mu. mol of cocatalyst A (methyldietetradecylammonium tetrakis (pentafluorophenyl) borate) and 10. mu. mol of MMAO-3A. The addition of the dienyldivinylsilane was as indicated in the table.

Table 13: example 12 and comparative batch reactor polymer run and characteristics.

Table 13: continuously for

Data for comparative example 12.C and diene example 12.1 are collected in table 13. NMR data confirmed a 1.3:1 (tri: tetra LCB) ratio of both trifunctional (0.08LCB/1000C) and tetrafunctional (0.06LCB/1000C) (see FIGS. 7-9).

The conventional and absolute molecular weights of examples 12.C and 12.1 are plotted in fig. 10.

The Extensional Viscosity Fixture (EVF) of example 12.1 is shown in fig. 11.

The measured melt strength of the polymer of example 12.1 was 19cN and the extensibility was 32 mm/sec (see fig. 12).

The dynamic mechanical spectrum of branched example 12.1 was measured and the results are recorded in table 13. The viscosity at 0.1 rad/sec was calculated to be 395,948 pas and the viscosity at 100 rad/sec was measured to be 2,075 pas, providing a rheology ratio (V) of 190.80.1/V100) (see FIG. 13).

Batch reactor example 13

In table 14, the polymer properties of the comparative linear polymer sample (13.C) are compared to the branched polymer of the batch reactor. The polymerization took place at a temperature of 150 ℃ and 585g of ISOPAR-ETM15g of octene neutralized 240psi of hydrogen pressure (. DELTA.H)2) The following steps. 11g of ethylene were loaded and the pressure maintained in the presence of 0.4. mu. mol of catalyst 1, 0.48. mu. mol of cocatalyst A (methyldietetradecylammonium tetrakis (pentafluorophenyl) borate) and 10. mu. mol of MMAO-3A. The addition of the dienyldivinylsilane was as indicated in the table.

Table 14: example 13 and comparative batch reactor polymer run and characteristics.

Table 14: continuously for

Data for comparative example 13.C and diene example 13.1 are collected in table 14. NMR data confirmed a 1.7:1 ratio of trifunctional (0.12LCB/1000C) to tetrafunctional (0.07LCB/1000C) to both.

The dynamic mechanical spectrum of branching example 13.1 was measured and the results are recorded in table 14. The viscosity at 0.1 rad/sec was determined to be 53,390 pas and the viscosity at 100 rad/sec was measured to be 887 pas, providing a rheology ratio (V) of 60.20.1/V100)。

The polymer of example 13.1 had a measured melt strength of 10cN and an extensibility of 65 mm/sec.

Examples 7 (table 8) to 13 (table 14) indicate that for a given catalyst and comparable conditions, control of the ratio of trifunctional to tetrafunctional LCB can be controlled by the ratio of hydrogen to ethylene in the reactor, supporting the mechanism in scheme 6.

Examples 7 (table 8) to 13 (table 14) indicate that as the ratio of trifunctional to tetrafunctional LCB increases, Mw/Mw0Is systematically lowered, where MwIs the weight average molecular weight of the diene branched sample, and Mw0Is the weight average molecular weight of the unbranched comparative sample. As the ratio of trifunctional branching involved increases, MwIn support of the trifunctional kinetic model herein and the previously derived tetrafunctional model described in the following applications: number PCTUS 2019/053524; number PCTUS 2019/053527; number PCTUS 2019/053529; and pct us2019/053537, each filed on 27/9/2019.

Guzman (Guzman) -2010 demonstrates and analyzes the MWD and physical properties resulting from conventional diene branching in a steady state CSTR. Ethylene, 1-octene and 1, 9-decadiene were copolymerized in a highly mixed one gallon reactor system using a Constrained Geometry Catalyst (CGC). The specific CGC catalyst used by gutzmann-2010 is described in detail in U.S. patent No. 5,965,756 (structure IX) and U.S. patent No. 7,553,917 (example 3). The gutzmann-2010 catalyst was designed to grow a single chain from the center of the catalyst. Data for the Gutzmann-2010 was collected at steady state while operating the CSTR at a pressure of 525psig and a temperature of 155 ℃ over a range of diene feed concentrations. Various steady state polymer samples collected from the gutzmann-2010 samples did not contain measurable levels of gel or insoluble materials. However, at the highest diene feed levels, some minor internal reactor fouling was observed, and it was expected that higher diene feed levels would lead to gel formation or reactor MWD instability.

In the case of the gutzmann-2010, a series of selected data are summarized over a range of continuous diene feed levels for otherwise fixed reactor conditions. Throughout the series, the ethylene and 1-octene feed concentrations were set at 13.8 wt% and 3.6 wt%, respectively. The catalyst feed rate was continuously adjusted to maintain a constant ethylene conversion of 79% throughout the series, resulting in a fixed polymer production of 2.2 kg/hr. The polymer density (a measure of copolymer composition) was constant at about 0.922 g/cc.

The data in Gutzmann-2010 shows how changes in the level of conventional diene branching affect the average molecular weight and polydispersity, as well as properties such as viscosity, I2And I10As reflected. The effect of conventional diene branching on molecular weight is shown for both absolute and conventional MWD measurement techniques. While absolute MWD measurement is the preferred method for branched polymers, it is not always available. Thus, the gutzmann-2010 also contains the molecular weight as measured by conventional techniques using a refractive index detector. The results in Table 33 show that the weight average molecular weight (M) increases from zero to 923ppm as the diene feed is increased by either measurement techniquew) Substantially increased.

Although not reported in the gutzmann-2010, MWD curves for absolute and conventional GPC measurement techniques were also found and plotted in fig. 14A and 14B, respectively. The MWD curve data in FIG. 14 shows that the expected high M resulting from conventional diene branching occurswThe tail is formed. It is also apparent from the MWD curve that the peak MW does not shift significantly as diene branching increases.

The molecular weight distribution data in fig. 14A and 14B are simplified to simple metrics describing the evolution of MWD curve position and shape as more diene monomer is fed into the CSTR. The data shows these MWD metrics for absolute and conventional MWD measurements of polymer samples from gutzmann-2010. Absolute MWD measurements show that when the 1, 9-decadiene feed is in the range of 0 to 923ppm, minThe increase in the quantum is 87% at most. Such as MpAs indicated, the peak molecular weight change was not significantly changed for the molecular weight method, which is not consistent with the "ladder branched" polymer results. The form factor is summarized in Table 34 (Gutzmann-2010) and is not consistent with a "ladder branched" polymer, since G79/29And ATail partWith diene feed level and MwIs increased.

It should be apparent to those skilled in the art that various modifications to the described embodiments can be made without departing from the spirit and scope of the claimed subject matter. Thus, it is intended that the present specification cover the modifications and variations of the described embodiments provided they come within the scope of the appended claims and their equivalents.

51页详细技术资料下载
上一篇:一种医用注射器针头装配设备
下一篇:树脂颗粒及其制备方法和含有该树脂颗粒的亲水性赋予剂

网友询问留言

已有0条留言

还没有人留言评论。精彩留言会获得点赞!

精彩留言,会给你点赞!