Chain Transfer with Dialkyl Zinc During Hafnium-Pyridyl Amido-Catalyzed Polymerization of 1-Octene: Relative Rates, Reversibility and Kinetic Models.

Similar documents
Supporting Information. for. Hexene. Clark R. Landis*,

Complex Promoted by Electron-Deficient Alkenes. Brian V. Popp and Shannon S. Stahl*

Supporting Information

SUPPLEMENTARY INFORMATION

SUPPORTING INFORMATION

Cationic Alkylaluminum-Complexed Zirconocene Hydrides as Participants in Olefin-Polymerization Catalysis. Supporting Information

Supporting Information

METHANOLYSIS OF ACETAL

Chemically recyclable alternating copolymers with low polydispersity from

Selective Reduction of Carboxylic acids to Aldehydes Catalyzed by B(C 6 F 5 ) 3

Bulk ring-opening transesterification polymerization of the renewable δ-decalactone using

Supplementary Material (ESI) for Chemical Communications This journal is (c) The Royal Society of Chemistry 2008

Optimizing Ion Transport in Polyether-based Electrolytes for Lithium Batteries

Induced Circular Dichroism of Stereoregular Vinyl Polymers

One-pot polymer brush synthesis via simultaneous isocyanate coupling chemistry and grafting from RAFT polymerization

Supporting Information.

Supporting information

Supporting Information

Mild and efficient bromination of Poly(hydroxyethyl)acrylate and its use towards Ionic-Liquid containing polymers

Combined metallocene catalysts: an efficient technique to manipulate long-chain branching frequency of polyethylene

Azaphosphatranes as Structurally Tunable Organocatalysts for Carbonate Synthesis from CO 2 and Epoxides

Determination of Polymer Modifier in Asphalt

Well-defined polyethylene-based random, block and bilayered molecular cobrushes

Supporting Information. Aqueous Dispersions of Polypropylene and Poly(1-butene) with Variable Microstructures formed with Neutral Ni(II)-Complexes

[(NHC)Au I ]-Catalyzed Acid Free Hydration of Alkynes at Part-Per-Million Catalyst Loadings

Supporting Information

Supporting Information

Molecular Weight Distribution of Living Chains in Polystyrene Pre-pared by Atom Transfer Radical Polymerization

Aziridine in Polymers: A Strategy to Functionalize Polymers by Ring- Opening Reaction of Aziridine

Supporting Information

Supporting Information for: End-capping ROMP Polymers with Vinyl. Lactones

ELECTRONIC SUPPORTING INFORMATION Pentablock star shaped polymers in less than 90 minutes via

Hyperbranched Poly(N-(2-Hydroxypropyl) Methacrylamide) via RAFT Self- Condensing Vinyl Polymerization

Supporting Information

Efficient Magnesium Catalysts for the Copolymerization of Epoxides and CO 2 ; Using Water to Synthesize Polycarbonate Polyols

Supporting Information

SUPPORTING INFORMATION

Supporting Information

ISOC Catalytic Mechanisms and High Pressure, Operando Kinetic Studies of Homogeneous Catalysts

RAFT /MADIX polymerization of N-vinylcaprolactam in water-ethanol solvent mixtures

SUPPORTING INFORMATION

Supporting Information

Chemistry Gas Chromatography: Separation of Volatile Organics

Supporting Information

Rediscovering the Isospecific Ring-Opening Polymerization of Racemic Propylene Oxide with Dibutylmagnesium

All-conjugated, all-crystalline donor-acceptor block. copolymers P3HT-b-PNDIT2 via direct arylation

Supporting Information. Bimetallic Oxidative Addition Involving Radical Intermediates in Nickel-Catalyzed Alkyl- Alkyl Kumada Coupling Reactions

Supporting information

Red Color CPL Emission of Chiral 1,2-DACH-based Polymers via. Chiral Transfer of the Conjugated Chain Backbone Structure

of Polystyrene 4-arm Stars Synthesized by RAFT- Mediated Miniemulsions.

Electronic Supporting Information for

Supporting Information

Supporting Information

One-Shot Synthesis and Melt Self-Assembly of Bottlebrush Copolymers with a Gradient Compositional Profile

A TTFV pyrene-based copolymer: synthesis, redox properties, and aggregation behaviour

A contribution from the Department of Chemistry, Washington University, Campus Box 1134, One Brookings Drive, Saint Louis, Missouri 63130

Supporting Information

1,1,3,3-Tetramethylguanidine-Promoted Ring-Opening Polymerization of N-Butyl N-Carboxyanhydride Using Alcohol Initiators

Cationic Polymerization of Vinyl Ethers Controlled by Visible Light

SUPPLEMENTARY INFORMATION

Supporting Information for

Preparation of 1:1 alternating, nucleobase-containing copolymers for use in sequence-controlled polymerization

Gel Permeation Chromatography

Formation of Benzynes from 2,6-Dihaloaryllithiums: Mechanistic Basis of the Regioselectivity

The Aggregation Behaviour of Polyisoprene Pluronic Graft Copolymers in Selective Solvents

Supporting Information for

Efficient Pd-Catalyzed Amination of Heteroaryl Halides

Electronic Supplementary Information. for. Self-Assembly of Dendritic-Linear Block Copolymers With Fixed Molecular Weight and Block Ratio.

Experimental Section

Electronic Supplementary Material (ESI) for Chemical Communications This journal is The Royal Society of Chemistry 2012

Fluorescent nanoparticles from PEGylated polyfluorenes - Supporting Information

Experimental procedures and characterization of the supramolecular complexes

How to use GPC/SEC for compositional analysis

Supporting information S1. Procedures Materials and methods Preparation of the reagents Preparation of the calibration standard solutions

Supporting Information

Adam Wu, Jean Michel Lauzon, Indah Andriani, and Brian R. James*

Supporting Information

Supplementary Information

Catalytic hydrogenation of liquid alkenes with a silica grafted hydride. pincer iridium(iii) complex: Support for a heterogeneous mechanism

Field-Flow Fractionation of Macromolecules and Structures That Cannot be Characterized by Conventional GPC/SEC Techniques

Supporting Information

High Frequency sonoatrp of 2-Hydroxyethyl Acrylate in an Aqueous Medium

Supporting Information. Sequence-Regulated Copolymers via Tandem Catalysis of Living Radical Polymerization and In Situ Transesterification

Photolabile Protecting Groups: A Strategy for Making

Supporting Information. Synthesis of Sulfur-Rich Polymers: Copolymerization of Episulfide with Carbon Disulfide

Synthesis of Peptide-Grafted Comb Polypeptides via Polymerisation of NCA-Peptides

Chemical Communications SUPPORTING INFORMATION. Polymer Catenanes via Supramolecularly Templated ATRP Initiator

In situ reaction monitoring reveals a diastereoselective ligand exchange. Supporting Information

Supporting Information

SUPPORTING INFORMATION. Multireactive Poly(2-oxazoline) Nanofibers through Electrospinning with Crosslinking on the Fly. and Amitav Sanyal a,b *

Scheme 1: Reaction scheme for the synthesis of p(an-co-mma) copolymer

SUPPORTING INFORMATION

Analysis of Star Polymers Using the Agilent 1260 Infinity Multi-Detector GPC/SEC System

Electronic supporting information for

Supporting Information

Supporting Information. segments

Supporting Information

Supporting Information for

Block copolymers containing organic semiconductor segments by RAFT polymerization

Characterization of polyphenylene sulphide using the Agilent PL-GPC 220 High Temperature GPC System with triple detection

Transcription:

Supporting information Chain Transfer with Dialkyl Zinc During Hafnium-Pyridyl Amido-Catalyzed Polymerization of 1-Octene: Relative Rates, Reversibility and Kinetic Models. Heather C. Johnson, Eric S. Cueny, Clark R. Landis* Department of Chemistry, University of Wisconsin Madison, 1101 University Avenue, Madison, Wisconsin 53706, United States Corresponding Author * landis@chem.wisc.edu Contents General experimental considerations S-2 General polymerization procedure S-2 Reversibility of chain transfer to ZnEt 2 analyzed by 13 C NMR spectroscopy S-3 Direct comparison of MWDs of polyoctene obtained using ZnEt 2 and ZnMe 2 S-4 Approximate Mayo analysis S-4 Observing equilibrium between ZnMe 2 and Zn i Pr 2 S-4 Kinetic modeling general considerations S-5 Optimizations to determine k p and k d S-6 Model 1: preferred kinetic model incorporating ZnR 2 exchange equilibria S-8 Comparison of experimental and simulated UV GPC traces (Model 1) S-9 Model 2: reversible Hf-pol/Zn-pol` chain transfer S-11 Model 3: exchange of ethyl group in ZnEtPol S-13 Model 4 no additional chain transfers S-15 Effect of k 2,1 upon Model 4 S-17 Catalyst death vs dormancy from 2,1-misinsertion S-17 References S-20 S-1

General experimental considerations Unless otherwise specified, all manipulations were carried out in air- and moisture-free conditions using standard Schlenk line and glovebox techniques. 1 1 and 2 2 were prepared by literature methods. All other compounds were commercial products and used as received. 1, ZnEt 2 and ZnMe 2 were stored as stock solutions in toluene inside a N 2 -filled glovebox (0.0096 M and 0.034 M, respectively). General NMR spectra were collected using Bruker Avance 400 MHz instrument fitted with a Smartprobe, Avance 500 MHz instrument fitted with a DCH cryoprobe, or Avance 600 with TCI-F cryoprobe. Quantitative NMR spectra measuring monomer consumption or end groups were collected using a relaxation delay of 25 s. GPC analyses were performed using a Viscotek GPCmax/VE 2001 instrument fitted with PolyPore columns (2 300 7.5 mm) featuring 5 µm particle size from Polymer Laboratories. Samples were eluted with THF at a flow rate of 1 ml/min at 40 C. Polymers were characterized by differential refractive index (RI) and UV (λ 344 nm) detection using a Viscotek Model 302-050 Tetra Detector Array. Omnisec software (Viscotek, Inc.) was used for initial data processing such as positioning the baseline, setting limits, and applying the molecular weight calibration (relative to polystyrene standards). Further processing was carried out in Excel. General polymerization procedure In a typical experiment, 86 µl of a 0.0096 M toluene solution of 1 was combined with 120 µl of a 0.0076 M toluene solution of [Ph 3 C][B(C 6 F 5 ) 4 ] (i.e. a 1:1.1 mixture of 1:[Ph 3 C][B(C 6 F 5 ) 4 ]). The mixture was heated at 50 C for 3 minutes in an aluminum heating block to form 3. 20.6 µl of this solution was sampled, and added to a vial containing toluene, 1-octene and ZnR 2 (R = Me, Et) (if applicable), at 50 C, to initiate the polymerization. Polymerizations were quenched with a solution of 2 (0.0912 µmol) at selected time points. To the vials, CH 2 Ph 2 (36 µmol) was added as an internal standard. Aliquots were sampled for analysis by 1 H NMR spectroscopy in CDCl 3. The polymer samples (including the NMR samples) were prepared for GPC analysis by dissolving up to 10 ml in THF and diluted such that a 1 mg/ml concentration was achieved. Temperature effects on polymerization The experiments were carried out according to the general procedure above. In each case, 3 was formed at 50 C for 3 minutes. Then, the aluminum heating block was cooled or warmed to the desired temperature before initiating polymerization. Otherwise the protocol is identical to that of the general polymerization procedure. S-2

Reversibility of chain transfer to ZnEt 2 analyzed by 13 C NMR spectroscopy Polymerizations were conducted according to the general polymerization procedure described above ([1-octene] 0 = 0.503 M, [3] 0 = 0.083 mm, 120 s reaction) with either 2 mm or 0.2 mm ZnEt 2. The polymers were isolated by evaporation of nearly all solvent, then precipitated with MeOH, washed with MeOH, then dried in vacuo. The polymer samples were dissolved in CDCl 3 and 13 C{ 1 H} NMR spectra of the polymers were collected. The NMR spectra were compared with literature values. 3 If chain shuttling were occurring, m x (r)m y stereoerrors should be observed due to shuttling of the polyoctene chain between different enantiomers of the catalyst, as found by Busico and co-workers 4 who observed approximately two-fold increases in the probability for inversion when doubling [AlMe 3 ] in a related system. In our system, <5% stereoerrors were observed with no significant change on increasing [ZnEt 2 ] from 0.2 to 2 Mm ([2.3 ± 1.8]% and [2.9 ± 1.8]%, respectively the large errors are due to a low signal:noise ratio at these concentrations). Examples of the relevant region of the 13 C{ 1 H} NMR spectra are shown in Figure S-1. These data indicate that no significant reversible chain transfer occurs under our conditions. Figure S1 Methine region of 13 C{ 1 H} NMR spectra of polyoctene produced in the presence of 0.2 mm (top) or 2 mm (bottom) [ZnEt 2 ]. S-3

Direct comparison of MWDs of polyoctene obtained using ZnEt 2 and ZnMe 2 Figure S2 RI vs retention volume plots for polymers produced in the presence of 1.6 mm ZnEt 2 or ZnMe 2 at 10 s, 30 s and 120 s (similar conversions). Approximate Mayo analysis The chain transfer constant k ex /k p (k ex = rate constant for chain transfer, k p = rate constant for propagation) can be estimated using a Mayo analysis. The chain transfer constant is the slope of a plot of 1/M n vs [ZnMe 2 ]/(112.24*[1-octene]), as shown in the Mayo equation (eq. S-1). In these experiments, polymerizations were conducted in the presence of various amounts of ZnMe 2 and quenched using 2 after several seconds of reaction (ca. 10% conversion). At these short reaction times, we assume that [1-octene] ~ [1-octene] 0. The obtained chain transfer constant is 20 ± 4. 1 1 112.24 1-octene (S1) M n = M n in the absence of ZnMe 2. 112.24 = molecular weight of 1-octene in g mol -1. Observing equilibrium between ZnMe 2 and Zn i Pr 2 Zn i Pr 2 (39 µl of a commercial 1.0 M solution in toluene), CH 2 Ph 2 in d 8 -toluene (internal standard, 4 µmol) and d 8 -toluene (367 µl) were added to a NMR tube capped with a septum. ZnMe 2 (74 µl of a commercial 1.2 M solution in toluene) was stored in a gas-tight syringe with the needle piercing the septum of a N 2 -filled vial. A second NMR tube containing ZnMe 2 (74 µl) and CH 2 Ph 2 (4 µmol) and d 8 -toluene (406 µl) was prepared. The NMR spectrometer was warmed to 50 C, and spectra were taken to establish the T 1 values of each species and, then, the concentration of the Zn species. T 1 of ZnMe 2 and Zn i Pr 2 were determined as 10 s and 16 s, respectively. A d 1 value of 30 s was therefore selected for quantitative concentration measurement. The initial concentrations of Zn i Pr 2 and ZnMe 2 were 2.49 and 1.82 mm, respectively. ZnMe 2 in the gastight syringe was injected into the NMR tube containing Zn i Pr 2, the tube was inverted, and inserted into the NMR probe. Within the first timepoint (ca. 3 minutes due to time taken inserting the sample, shimming and the 30 s d 1 ) equilibrium had been reached. The equilibrium constant [Zn i PrMe] 2 /([ZnMe 2 ][Zn i Pr 2 ]) = 4.1. S-4

For 1-octene polymerization, the value of k ex2 obtained via fitting of the MWDs (vide infra) is 0.34 M -1 s -1. The ZnMe 2 /Zn i Pr 2 equilibrium can be modelled using the experimentally observed equilibrium constant of 4.1 and the optimized value of k ex2. This simulation (Figure S-3) shows that equilibrium is reached within several minutes. Thus, being unable to observe the approach to equilibrium within minutes by NMR spectroscopy is not unreasonable (if the simulated rate constant is valid on this model system). These experiments show that a value of 0.34 M -1 s -1 for k ex2 is consistent with the data. Figure S3 Simulated approach to equilibrium for ZnMe 2 /Zn i Pr 2 using k ex2 =0.34 M -1 s -1 K eq =[Zn i PrMe] 2 /([ZnMe 2 ][Zn i Pr 2 ) = 4.1. and Kinetic modeling general considerations Landis and co-workers 5 previously showed two ways to simulate molecular weight distributions of (EBI)ZrMe 2 /B(C 6 F 5 ) 3 -catalyzed 1-hexene polymerization that accounts for experimental band broadening: the exponentially-modified Gaussian (EMG) approach, and the band broadening correction (BBC) approach. The EMG approach convolutes discrete calculated concentrations of each polymer i with an EMG function, producing a continuous simulated GPC trace. In contrast, the BBC approach produces discrete values for the RI signal of polymer i through multiplication of the calculated RI signal (km i c i ) by an additional mass term, m i. This mass correction adequately reproduces the effect of band broadening in experimental GPC traces. In the kinetic simulations of (EBI)ZrMe 2 /B(C 6 F 5 ) 3 -catalyzed 1-hexene polymerization, each approach determined the same values, within error, for the optimized rate constants. Therefore, this work uses the simpler BBC approach. Incorporating the BBC approach yields the following equations for the calculation of RI or UV signal corresponding to i monomers: RI i = α 1 * MW i 2 * [Pol i ] UV i = α 2 * MW i * [Pol i ] (S-2) (S-3) Where α 1 and α 2 are scaling factors, MW i = molecular weight of polymer i, and [Pol i ] = concentration of polymer i. The kinetic models were designed in Mathematica, converted to.xml files and imported into the Copasi modelling suite. All optimizations and simulations were performed using Copasi. Molecular weight data from the GPC instrument were treated as previously described. 5 Since S-5

the BBC approach provides discrete calculated RI signals calculated as above, rather than a continuous spectrum, 24 discrete points along the MWD (from logmw = 4.353 to logmw = 6.050) were chosen to simulate the GPC spectra. Where possible, these were spaced ca. 0.05 logmw units apart to simulate the entire distribution equally. However, since the model considers 200 monomer insertions per step, each step in the model constitutes a MW difference of 22448 gmol -1. Therefore, owing to the logarithmic scale, the spacing between the first few points in the distribution is larger (e.g. the 1 st 200 insertions leads to logmw = 4.353; the next 200 lead to logmw = 4.653), so the lower MW portions of the MWD are less well fit than the later ones. For this reason, only molecular weight data from 20 s to 120 s (>30% conversion) were considered in the models. The data weighting scheme used relates to the mean square weightings, and the number of data points for each type of data set. Within each heterogeneous data set (e.g. RI values) the data are weighted consistently. Unless otherwise stated, the weightings are: [monomer] = 1, RI = 5x10-6, [vinylene] = 6 x 10 8. In all models, k i is fixed at 10 5 M -1 s -1 as no induction period was observed experimentally by our methods. This means that essentially all 3 becomes active catalyst. Therefore, the concentration of [3] is equal to the active site concentration. This approximation removes the need to have separate steps in which active catalyst is formed from 3. Since the experimental active site concentration (by quench-labeling) appears constant within error throughout the polymerization time course, this is a reasonable approximation. In all optimizations, [3] is fixed at 45% of experimentally injected catalyst (i.e. 0.037 mm if 0.083 mm 3 was injected). For the models involving either k ex2 or k ex3 (vide infra), two steps in optimization were often required to achieve satisfactory fits since, when fitting all rate constants, optimizations would often terminate in local minima providing poor fits to experimental data. Using monomer consumption data and experimentally determined concentrations of active sites (fixed across all runs as 45%, 0.037 mm), the rate constants k p (propagation) and k d (catalyst death) were found. Then, the values of k p and k d were fixed and rate constants for exchange (k ex1 and k ex2, k ex3 or k ex4, where applicable) and chain termination (k 2,1 ) were optimized against experimental data including monomer consumption, molecular weight distributions (RI) at various time points, and [vinylene] after 120 s. Addition of zinc reagents do not affect the rate of polymerization, so this is an appropriate way to model the data. Optimizations to determine k p and k d The values for k p and k d were optimized across a range of concentration regimes. As discussed above, the value of the active catalyst was set to 45% of the experimentally added catalyst, and k i was set to 10 5 M -1 s -1. Excellent agreement between the fits and simulations is observed (Figure S-4). Only the steps shown in the model in Scheme S-1 were considered. S-6

initiation k i 3 + mon Hf-Ins k Hf-Ins + mon i Hf-Pol[1] propagation Hf-Pol[n] + 200*mon catalyst death k p 200 Hf-Pol[n+1] Hf-Pol[n] k d Hf-Inact[n] Scheme S1 Initiation, propagation and catalyst death steps Figure S4 Plots of experimental (blue circles) and simulated (red diamonds) [monomer] vs time. Rate constants k p = 892 ± 18 M -1 s -1, (k p 200 = 4.46 ± 0.09 M -1 s -1 ), k d = 0.0156 ± 0.0009 s -1. Experimental conditions: (a) [1-octene] 0 = 0.251 M, [3] 0 = 0.083 M; (b) [1-octene] 0 = 0.503 M, [3] 0 = 0.083 M; (c) [1-octene] 0 = 1.006 M, [3] 0 = 0.083 mm; (d) [1-octene] 0 = 0.503 M, [3] 0 = 0.166 mm. S-7

Model 1: preferred kinetic model incorporating ZnR 2 exchange equilibria The kinetic model that best fits the data is shown below (Scheme S-2). The molecular weight distribution fits across 4 different concentration regimes are shown in Figure S-5. The rate constants for propagation and catalyst death were optimized as above. The rate constants obtained are: k p = 892 ± 18 M -1 s -1, k 2,1 = 0.26 ± 0.01 M -1 s -1, k ex1 = 7493 ± 95 M -1 s -1, k d = 0.0156 ± 0.0009 s -1, k ex2 = 0.34 ± 0.03 M -1 s -1. Several variables were fixed (k i = 10 5 M -1 s -1 ; [3] = 45% active sites, the equilibrium constant K 2 (k ex2 /k ex-2 ) was fixed at 0.24.). The optimized value for the concentration of vinylene end groups is 0.12 mm matches the experimental value of ca. 0.13 mm ([1-octene] 0 = 0.503 M, [3] 0 = 0.083 mm). The objective value (sum of squared deviations) is 4.63. Scheme S2 Model 1. k p 200 = k p /200 S-8

Figure S5 Plots of experimental (blue circles) and simulated (red diamonds) RI vs logmw molecular weight distributions for Model 1. Experimental conditions: (a) [1-octene] 0 = 0.503 M, [3] 0 = 0.083 mm; [ZnEt 2 ] 0 = 1.6 mm (b) [1-octene] 0 = 0.503 M, [3] 0 = 0.166 mm; [ZnEt 2 ] 0 = 1.6 mm; (c) [1-octene] 0 = 0.503 M, [3] 0 = 0.083 M; [ZnEt 2 ] 0 = 0.8 mm; (d) [1-octene] 0 = 1.006 M, [3] 0 = 0.083 M; [ZnEt 2 ] 0 = 1.6 mm. Comparison of experimental and simulated UV GPC traces (Model 1) As an independent test of our favored kinetic model, simulation of the expected UV-GPC distribution was conducted according to equation S-3 (kinetic modeling general considerations section): the UV signal for polymer i was calculated by multiplying the sum of the concentrations of Hf-bound species [Hf-Pol[i]] and [Hf-Inact[i]] by the molecular weight of i (to correct for the effect of band broadening upon the GPC traces) and a scaling factor (that converts this expression to the 1/transmittance as output from the GPC instrument). To ensure this is an S-9

independent test, no additional optimizations were performed. Comparisons between UV expt and UV sim at 20 s and 60 s over a range of experimental conditions is shown in Figure S-6. Overall, there is a good match with experimental data, supporting the robustness of the kinetic model. Note: the dilution required to reach 1 mg/ml polyoctene concentration for GPC analysis means that samples produced at higher conversions (e.g. after 60 s) contain ca. 0.004 mm [chromophore], so noise is increased at higher conversions. Figure S6 Plots of experimental (yellow circles) and simulated (green diamonds) UV vs logmw molecular weight distributions. Experimental conditions: (a) [1-octene] 0 = 0.503 M, [3] 0 = 0.083 mm; [ZnEt 2 ] 0 = 1.6 mm (b) [1-octene] 0 = 0.503 M, [3] 0 = 0.166 mm; [ZnEt 2 ] 0 = 1.6 mm; (c) [1- octene] 0 = 0.503 M, [3] 0 = 0.083 M; [ZnEt 2 ] 0 = 0.8 mm; (d) [1-octene] 0 = 1.006 M, [3] 0 = 0.083 M; [ZnEt 2 ] 0 = 1.6 mm. S-10

Model 2: reversible Hf-pol/Zn-pol` chain transfer An alternative kinetic model considered is shown in Scheme S-3. The molecular weight distribution fits across 4 different concentration regimes are shown in Figure S-7. The rate constants for propagation and catalyst death were optimized as above. The rate constants obtained are: k p = 892 ± 18 M -1 s -1, k 2,1 = 0.26 ± 0.01 M -1 s -1, k ex1 = 7473 ± 92 M -1 s -1, k d = 0.0156 ± 0.0009 s -1, k ex3 = 13.3 ± 1.0 M -1 s -1. Several variables were fixed (k i = 10 5 M -1 s -1 ; [Hf] = 45% active sites). The optimized value for the concentration of vinylene end groups is 0.12 mm matches the experimental value of ca. 0.13 mm ([1-octene] 0 = 0.503 M, [3] 0 = 0.083 mm). The objective value (sum of squared deviations) is 4.51, similar to that of Model 1. We favor Model 1, despite a slightly higher objective function, since it allows for > 1 chain per Zn center, while Model 2 only allows for a maximum of 1 chain per Zn. Scheme S3 Model 2. k p 200 = k p /200 S-11

Figure S7 Plots of experimental (blue circles) and simulated (red diamonds) RI vs logmw molecular weight distributions for Model 2. Experimental conditions: (a) [1-octene] 0 = 0.503 M, [3] 0 = 0.083 mm; [ZnEt 2 ] 0 = 1.6 mm (b) [1-octene] 0 = 0.503 M, [3] 0 = 0.166 mm; [ZnEt 2 ] 0 = 1.6 mm; (c) [1-octene] 0 = 0.503 M, [3] 0 = 0.083 M; [ZnEt 2 ] 0 = 0.8 mm; (d) [1-octene] 0 = 1.006 M, [3] 0 = 0.083 M; [ZnEt 2 ] 0 = 1.6 mm. S-12

Model 3: direct exchange of ethyl group in ZnEtPol with Hf-Pol Another kinetic model considered is shown in Scheme S-4. The molecular weight distribution fits across 4 different concentration regimes are shown in Figure S-8. The rate constants for propagation and catalyst death were optimized as above. The rate constants obtained are: k p = 892 ± 18 M -1 s -1, k 2,1 = 0.17 ± 0.01 M -1 s -1, k ex1 = 8799 ± 128 M -1 s -1, k d = 0.0156 ± 0.0009 s -1, k ex4 = 37.5 ± 1.0 M -1 s -1. Several variables were fixed (k i = 10 5 M -1 s -1 ; [Hf] = 45% active sites). The fit between the simulated and experimental data are visually worse than for the models involving exchange steps k ex2 or k ex3. The optimized value for the concentration of vinylene end groups is 0.08 mm, lower than the experimental value of ca. 0.13 mm ([1-octene] 0 = 0.503 M, [3] 0 = 0.083 mm). This model was not further explored. Scheme S4 Model 3. k p 200 = k p /200 S-13

Figure S8 Plots of experimental (blue circles) and simulated (red diamonds) RI vs logmw molecular weight distributions for Model 3. Experimental conditions: (a) [1-octene] 0 = 0.503 M, [3] 0 = 0.083 mm; [ZnEt 2 ] 0 = 1.6 mm (b) [1-octene] 0 = 0.503 M, [3] 0 = 0.166 mm; [ZnEt 2 ] 0 = 1.6 mm; (c) [1-octene] 0 = 0.503 M, [3] 0 = 0.083 M; [ZnEt 2 ] 0 = 0.8 mm; (d) [1-octene] 0 = 1.006 M, [3] 0 = 0.083 M; [ZnEt 2 ] 0 = 1.6 mm. S-14

Model 4 no additional chain transfers Model 4 features only k ex1 as the only step involving ZnEt 2 (Scheme S-5). The molecular weight distribution fits across 4 different concentration regimes are shown in Figure S-9. The rate constants for propagation and catalyst death were optimized as above. The rate constants obtained are: k p = 892 ± 18 M -1 s -1, k 2,1 = 0.31 ± 0.01 M -1 s -1, k ex1 = 7200 ± 92 M -1 s -1, k d = 0.0156 ± 0.0009 s -1. Several variables were fixed (k i = 10 5 M -1 s -1 ; [Hf] = 45% active sites). The fit between the simulated and experimental data are worse than for the Model 1 or Model 2 with an objective value of 5.69, and the simulated data at 120 s are more bimodal than the data. The optimized value for the concentration of vinylene end groups is 0.15 mm (experimentally, 0.13 ± 0.3 mm) ([1-octene] 0 = 0.503 M, [3] 0 = 0.083 mm). Lowering the weighting of [vinylene] results in a lower objective function but raises the simulated [vinylene] to 0.18 mm, outside experimental error. This suggests that the additional chain transfer events considered in Models 1 3 provide subtle yet necessary means of smoothing the molecular weight distributions. Scheme S5 Model 4. k p 200 = k p /200 S-15

Figure S9 Plots of experimental (blue circles) and simulated (red diamonds) RI vs logmw molecular weight distributions for Model 4. Experimental conditions: (a) [1-octene] 0 = 0.503 M, [3] 0 = 0.083 mm; [ZnEt 2 ] 0 = 1.6 mm (b) [1-octene] 0 = 0.503 M, [3] 0 = 0.166 mm; [ZnEt 2 ] 0 = 1.6 mm; (c) [1-octene] 0 = 0.503 M, [3] 0 = 0.083 M; [ZnEt 2 ] 0 = 0.8 mm; (d) [1-octene] 0 = 1.006 M, [3] 0 = 0.083 M; [ZnEt 2 ] 0 = 1.6 mm. S-16

Effect of k 2,1 upon Model 4 To demonstrate that small changes in k 2,1 can strongly affect the MWDs, simulations were conducted using Model 4 and the conditions [1-octene] 0 = 0.503 M, [ZnEt 2 ] 0 = 1.6 mm, [3] 0 = 0.037 mm (i.e. 45% active sites) rate constants k p = 892 ± 18 M -1 s -1, k 2,1 = 0.31 ± 0.01 M -1 s -1, k ex1 = 7200 ± 92 M -1 s -1, k d = 0.0156 ± 0.0009 s -1, k i = 10 5 M -1 s -1. The values of k 2,1 were fixed at 0, 0.2 M -1 s -1 and 0.4 M -1 s -1. An overlay of the resulting MWDs (and the experimental data) is provided in Figure S-10, showing the significant effect of modest k 2,1 (compared with k p and k ex1 ) values upon the MWD. Figure S10 Simulation of MWD data (at 60 s of polymerization) using Model 4 and altering the value of k 2,1. Catalyst death vs dormancy from 2,1-misinsertion As an alternative to irreversible catalyst death, catalyst dormancy (following a 2,1-misinsertion) was explored and compared with Model 4 (additional chain transfer events are not considered here). The model used is shown in Scheme S-6. There is no catalyst death step. Instead, after undergoing a 2,1-misinsertion of 1-octene (rate constat k mis ), the catalyst lies dormant with the polymer chain intact (Hf-dormant[n]). Then, Hf-dormant[n] can undergo β-hydride elimination to form Hf-Ins and the polymer chain Pol[n] with the rate constant k elim. The model was optimized against experimental data to find the rate constants k p 200, k mis, k elim and k ex1. [3] was fixed at 45% active sites. The resulting rate constants from the optimization are: k p = 1020 ± 20 M -1 s -1 (k p 200 = 5.1 ± 0.1 M -1 s -1 ), k ex1 = 8148 ± 164 M -1 s -1, k mis = 0.41 ± 0.01 M - 1 s -1, k elim = 0.38 ± 0.02 s -1. The MWD fits are adequate (see Figure S-12) but the fits for the monomer consumption show that 1-octene is too rapidly consumed in this model (three different [1-octene] 0 vs time plots are shown in Figure S-11). Additionally, the amount of [vinylene] is overestimated (e.g. 0.19 mm vs 0.13 mm experimentally when [1-octene] 0 = 0.503 M). S-17

initiation 3 + mon Hf-Ins k i k Hf-Ins + mon i Hf-Pol[1] propagation Hf-Pol[n] + 200*mon k p 200 Hf-Pol[n+1] chain transfer to Zn Hf-Pol[n] + ZnEt 2 2,1-misinsertion Hf-Pol[n] + mon k mis k ex1 Hf-Ins + Hf-Dormant[n] EtZn Pol[n] elimination Hf-Dormant[n] k elim Hf-Ins + Pol[n] Scheme S6 Catalyst dormancy model Figure S11 Plots of experimental (blue circles) and simulated (red diamonds) [monomer] vs time for (a) [1-octene] 0 = 0.251 M, (b) [1-octene] 0 = 0.503 M, and (c) [1-octene] 0 = 1.006 M S-18

Figure S12 Plots of experimental (blue circles) and simulated (red diamonds) RI vs logmw molecular weight distributions for the dormant catalyst model. Experimental conditions: (a) [1- octene] 0 = 0.503 M, [3] 0 = 0.083 mm; [ZnEt 2 ] 0 = 1.6 mm (b) [1-octene] 0 = 0.503 M, [3] 0 = 0.166 mm; [ZnEt 2 ] 0 = 1.6 mm; (c) [1-octene] 0 = 0.503 M, [3] 0 = 0.083 M; [ZnEt 2 ] 0 = 0.8 mm; (d) [1- octene] 0 = 1.006 M, [3] 0 = 0.083 M; [ZnEt 2 ] 0 = 1.6 mm. S-19

References (1) Frazier, K.; Boone, H.; Vosejpka, P.; Stevens, J.US20040220050 A1,2004 (2) Cueny, E. S.; Johnson, H. C.; Anding, B. J.; Landis, C. R., Mechanistic Studies of Hafnium-Pyridyl Amido-Catalyzed 1-Octene Polymerization and Chain Transfer Using Quench- Labeling Methods. J. Am. Chem. Soc. 2017, 139, 11903-11912. (3) Galland, G. B.; Da Silva, L. F.; Nicolini, A., Tacticity of Poly-α-olefins from Poly-1-hexene to Poly-1-octadecene. J. Polym. Sci., Part A: Polym. Chem. 2005, 43, 4744-4753. (4) Alfano, F.; Boone, H. W.; Busico, V.; Cipullo, R.; Stevens, J. C., Polypropylene Chain Shuttling at Enantiomorphous and Enantiopure Catalytic Species: Direct and Quantitative Evidence from Polymer Microstructure. Macromolecules 2007, 40, 7736-7738. (5) Nelsen, D. L.; Anding, B. J.; Sawicki, J. L.; Christianson, M. D.; Arriola, D. J.; Landis, C. R., Chromophore Quench-Labeling: An Approach to Quantifying Catalyst Speciation As Demonstrated for (EBI)ZrMe 2 /B(C 6 F 5 ) 3 -Catalyzed Polymerization of 1-Hexene. ACS Catal. 2016, 6, 7398-7408. S-20