You are on page 1of 9

Angewandte

Research Articles Chemie

International Edition: DOI: 10.1002/anie.201910828


Polymers German Edition: DOI: 10.1002/ange.201910828

Dimethyl Dihydroacridines as Photocatalysts in Organocatalyzed


Atom Transfer Radical Polymerization of Acrylate Monomers
Bonnie L. Buss, Chern-Hooi Lim, and Garret M. Miyake*

Abstract: Development of photocatalysts (PCs) with diverse Organocatalyzed atom transfer radical polymerization
properties has been essential in the advancement of organo- (O-ATRP) is a metal-free variant of photoredox-catalyzed
catalyzed atom transfer radical polymerization (O-ATRP). ATRP which eliminates the concern of trace metal contam-
Dimethyl dihydroacridines are presented here as a new family ination in the polymer product and is advantageous in
of organic PCs, for the first time enabling controlled polymer- electronic and biomedical applications, while also enabling
ization of challenging acrylate monomers by O-ATRP. Struc- opportunities for “greener” reaction design in polymer syn-
ture–property relationships for seven PCs are established, thesis.[5] Induced by light, O-ATRP relies on a strongly
demonstrating tunable photochemical and electrochemical reducing organic PC to mediate an oxidative quenching
properties, and accessing a strongly oxidizing 2PCC+ intermedi- catalytic cycle (Figure 1 a). O-ATRP processes following
ate for efficient deactivation. In O-ATRP, the combination of a reductive quenching pathway have also been reported but
PC, implementation of continuous-flow reactors, and promo- rely on the presence of stoichiometric quantities of sacrificial
tion of deactivation through addition of LiBr are critical to electron donors, which can also induce undesirable side
producing well-defined acrylate polymers with dispersities as reactions.[6] The proposed O-ATRP mechanism proceeds
low as 1.12. The utility of this approach is established through through four central steps.[7] Photoexcitation of a ground-state
demonstration of the oxygen-tolerance of the system and PC generates the singlet excited state 1PC*, which can
application to diverse acrylate monomers, including the syn- undergo intersystem crossing to produce a long-lived triplet
3
thesis of well-defined di- and triblock copolymers. PC*. Either 1PC* or 3PC* then directly reduces an alkyl

Introduction

The ability of photoredox catalysis to manipulate elec-


tron- or energy-transfer reactivity has revolutionized small-
molecule and macromolecular chemistry, presenting oppor-
tunities to develop new chemical transformations under mild
and energy efficient reaction conditions.[1] Recently, photo-
redox catalysis has been applied in controlled radical
polymerization (CRP) approaches for light-regulated syn-
thesis of well-defined polymers, most commonly in atom
transfer radical polymerization (ATRP) and reversible addi-
tion-fragmentation transfer (RAFT).[2] ATRP, the most
widely studied CRP methodology, is used to access polymers
with controlled properties, higher-order architectures, and
consequently diverse applications.[3] Traditionally, ATRP is
operated through activation of a CuI catalyst by heat to
promote an innersphere electron transfer to generate a prop-
agating radical species. However, in recent advances, new
light-driven ATRP processes have been reported using
photocatalysts (PCs) derived from copper, ruthenium, or
iridium.[4]

[*] B. L. Buss, Dr. C.-H. Lim, Dr. G. M. Miyake


Department of Chemistry, Colorado State University
Fort Collins, CO 80523 (USA)
E-mail: garret.miyake@colostate.edu
Dr. C.-H. Lim
New Iridium LLC
Boulder, CO 80303 (USA)
Supporting information and the ORCID identification number(s) for Figure 1. a) Proposed mechanism of O-ATRP. b) Previously reported
the author(s) of this article can be found under: PCs developed for O-ATRP (left) with the new dimethyl-dihydroacridine
https://doi.org/10.1002/anie.201910828. PC family investigated in this work (right).

Angew. Chem. Int. Ed. 2020, 59, 3209 – 3217 T 2019 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim 3209
Angewandte
Research Articles Chemie

halide initiator through outersphere electron transfer to Additionally, a photoredox-catalyzed ATRP approach for
produce a propagating radical species, as well as the ion-pair acrylates was also performed using a precious metal based
2
PCC+X@ . Deactivation of the propagating chain-end occurs fac-Ir(ppy)3 PC following an oxidative quenching catalytic
through reinstallation of the halide, generating the PC and cycle, accessing control over MW growth to produce polymers
a dormant polymer. Central to success in O-ATRP, as with moderate Y.[23] To access this polymerization with an
determined by control over polymer molecular weight organic PC, we hypothesized that a PC candidate must
(MW) and dispersity (Y) close to 1.0, is the presence of possess both a sufficiently oxidizing 2PCC+ with a correspond-
a dynamic equilibrium between the activation and deactiva- ingly high E1/2 (2PCC+/1PC) potential to promote fast deacti-
tion steps, where the rate of deactivation (with rate constant vation (high kd) to compensate for a high kp value, but also
kd) must be higher than the rates of propagation (kp) and maintain a strongly reducing 3PC* with a sufficiently negative
activation (ka), limiting radical concentrations and undesir- E0(2PCC+/3PC*) value for efficient alkyl bromide activation.
able termination events by radical quenching. Using density-functional theory (DFT) calculations to
To date, advances in O-ATRP have been enabled through guide organic PC design, herein dimethyl dihydroacridines
development of strongly reducing organic PCs, which are are reported as a new class of organic PCs adept at controlled
capable of directly reducing an activated alkyl bromide ATRP polymerizations of acrylate monomers by O-ATRP. Given the
initiator or dormant polymer chain-end (& @0.8 V vs. SCE).[8] structural similarity to previously employed PCs in O-ATRP
In 2014, perylene and N-phenyl phenothiazine were reported and tunable donor-acceptor motifs, we sought to investigate
as strongly reducing PCs for the polymerization of meth- this class of molecules for use in photoredox-catalyzed
acrylate monomers by O-ATRP.[9] Since then, other organic processes, accessing tailored photo- and electrochemical
PCs derived from N,N-diaryl dihydrophenazine[10] and N-aryl properties. In this approach, well-defined acrylate polymers
phenoxazine[11] families, among others,[12] have been devel- with controlled MWs and low dispersities (Y , 1.20) were
oped (Figure 1 b). N-aryl phenothiazine PCs have been synthesized using a 365 nm light-emitting diode (LED) in
applied in diverse contexts[13] and have also been used for a continuous-flow reactor in conjunction with LiBr salt
mechanistic analysis.[14] Recently, PC structure–property additives, which are hypothesized to promote efficient
relationships have been studied using N,N-diaryl dihydrophe- deactivation.
nazine and N-aryl phenoxazine PCs.[15] Empirically, these
studies have determined key PC design principles for
effective catalytic performance in O-ATRP, among which Results and Discussion
are the ability of the PC to exhibit intramolecular charge-
transfer (CT) excited states, redox reversibility, and sufficient Photocatalyst Development
thermodynamic driving forces (redox potentials) to mediate
the oxidative quenching O-ATRP cycle.[16] N,N-diaryl dihy- Dimethyl dihydroacridines have previously been applied
drophenazine and N-aryl phenoxazine PCs have also been in the development of organic LEDs as thermally activated
studied in diverse polymerization-related contexts, including delayed fluorescence emitters. Tunable absorption profiles,
the effects of light intensity and solvent, adaptation of O- small S1 and T1 energy gaps (DEST < 0.4 eV), as well as CT
ATRP to continuous-flow reactors, synthesis of star polymers, characteristics were reported,[24] making this a promising
and demonstration of oxygen tolerance.[17a–e] In addition, structural motif for application in photoredox catalysis. Initial
these organic PCs were also applied in small-molecule DFT calculations for 9,9-dimethyl-10-(naphthalen-1-yl)-9,10-
reactions, including trifluoromethylation, C@N and C@S dihydroacridine (1 a) predicted an E0ox (2PCC+/1PC) value of
cross-couplings by a dual catalytic approach with NiII salts, 0.72 V vs. SCE. Corroborating the DFT prediction, 1 a was
and the reduction of carbon dioxide to methane using sunlight experimentally determined to have an Ep/2 value of 0.82 V vs.
for solar fuel generation.[17f–g] SCE (Figure 2 a). Although 1 a displayed a non-reversible
Despite advances in PC design, O-ATRP has largely been cyclic voltammogram, this relatively high Ep/2 value encour-
limited to polymerization of methacrylate monomers, but the aged further exploration of dimethyl dihydroacridines as
controlled polymerization of other monomers is highly potential PC candidates, given that current successful strongly
desired.[18] Poly(acrylates) possess disparate thermal and reducing PCs in O-ATRP only have E1/2 values of up to about
mechanical properties, enabling widespread industrial and 0.7 V vs. SCE.[16]
academic use, including drug delivery, superabsorbent mate- Notably, installation of biphenyl groups at the 2- and 7-
rials, coatings, adhesives and additive manufacturing.[19] As positions of 1 a imparted redox reversibility, a key require-
such, we sought to leverage current understanding of organic ment for catalyst turnover (Figure 2 a). Furthermore, relative
PC design to target the O-ATRP of acrylate monomers. The to 1 a, increasing the conjugation of the dimethyl dihydroa-
CRP of acrylates is inherently challenging because of high kp, cridine molecules also red-shifted the absorption profile by
with values ranging from 15 000 to 24 000 L mol@1 s@1,[20] an about 80 nm, presenting the possible use of milder irradiation
order of magnitude larger than methacrylates. Furthermore, conditions (Figure 2 b). These results motivated us to synthe-
acrylate chain-end groups containing bromides are more size a library of core-substituted derivatives and evaluate
difficult to reduce compared to the corresponding methacry- their catalytic potential within the previously established
lates, emphasizing the need for efficient PCs for activation.[21] design framework for O-ATRP PCs, including absorption
Photoinduced copper-catalyzed ATRP processes have properties, excited-state characteristics, and redox proper-
been reported for successful CRP of acrylate monomers.[22] ties.[7, 15a] Seven dimethyl dihydroacridine PCs with diverse

3210 www.angewandte.org T 2019 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2020, 59, 3209 – 3217
Angewandte
Research Articles Chemie

Figure 2. a) Cyclic voltammogram of the PCs 1 a and 1. b) UV-vis spectra of 1 a, 1, 2, and 3. c) Electrochemical series of experimentally-measured
excited state redox potentials E0*T1, exp. = E0(2PCC+/3PC*) and oxidation potentials E1/2 (2PCC+/1PC) of PCs investigated in this study. High- and low-
lying SOMO for PCs with electronically neutral N-aryl groups, overlays of the absorption profiles (purple), and emission profiles (teal) with the
experimentally determined Stokes shifts for each PC. photographs of the PCs dissolved in solvents with increasing polarity with lmax,emission for each
solvent for 1 (d), 4 (e), and 5 (f). All CV, UV-vis, and emission data collected in N,N-dimethylformamide. See the Supporting Information for full
experimental and computational details.

electron-poor (cyanophenyl), electron-rich (methoxyphenyl), Installation of either electron-donating or electron-with-


and highly conjugated (biphenyl or naphthalene) groups on drawing groups onto the core-substituent was found to
both the core and N-aryl positions were synthesized (Fig- strongly influence lmax,abs. For example, installation of 4-
ure 2 c). UV-vis spectroscopy was used to probe the ability to cyanophenyl groups (PC 3, 382 nm) red-shifted lmax,abs by
tune the frontier orbitals with various substitutions, and as 42 nm compared to 4-methoxyphenyl groups (2, 340 nm;
such tune the absorption profiles of the PC (Figure 2 b; see Figure 2 b). In a comparison of N-aryl modifications with
Figures S41–S47 in the Supporting Information). biphenyl core substituents, decreasing N-aryl conjugation
from a 1-naphthyl group (1) to a phenyl group (5) led to

Angew. Chem. Int. Ed. 2020, 59, 3209 – 3217 T 2019 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.angewandte.org 3211
Angewandte
Research Articles Chemie

Table 1: Summary of photophysical and electrochemical measurements for PCs 1–7.


PC lmax,abs. e lmax,S1 em. Ff ES1,exp. ET1,calc. ET1,77 K E1/2 E0*S1,exp. E0*T1,exp. 77 K E0*T1,calc.
[nm][a] [m@1 cm@1] [nm][b] [%] [eV][c] [eV][d] [eV][c,e] (2PCC+/1PC) (2PCC+/1PC*) (2PCC+/3PC*) (2PCC+/3PC*)
(V vs. SCE)[f ] (V vs. SCE)[g] (V vs. SCE)[e,g] (V vs. SCE)[d]
1 361 46 300 487 0.3 2.55 2.41 2.34 0.76 @1.79 @1.58 @1.86
2 340 38 600 509 0.1 2.44 2.36 2.33 0.71 @1.73 @1.62 @1.94
3 382 44 300 458 83 2.71 2.43 2.39 0.90 @1.81 @1.49 @1.69
4 360 49 800 453 3.8 2.74 2.30 2.34 0.77 @1.98 @1.57 @1.77
5 361 31 500 443 27 2.80 2.29 2.35 0.76 @2.04 @1.59 @1.75
6 363 43 100 444 70 2.79 2.28 2.34 0.75 @2.04 @1.59 @1.78
7 355 50 100 535 8.2 2.32 2.39 2.37 0.82 @1.50 @1.55 @1.78
[a] Absorption wavelength measured using ultraviolet-visible spectroscopy in DMF. [b] Emission wavelength measured using steady-state
fluorescence spectroscopy in DMF. [c] Energies were calculated using the maximum wavelength of emission. [d] DFT calculations performed at uM06/
6–311 + Gdp//uM06/6–31 + Gdp level of theory with CPCM-described solvation in aqueous solvent. [e] Spectral emission measured at 77 K after 1 ms
gate-delay. [f ] All measurements were performed in a three-compartment electrochemical cell with an Ag/AgNO3 reference electrode in MeCN (0.01 m)
and 0.1 m NBu4PF6 electrolyte solution with DMF analyte solution. Platinum was used at the working and counter electrodes. [g] Excited-state redox
potentials were calculated using energies estimated from the maximum wavelength of singlet or triplet emission and the experimentally measured E1/
0 0
2 ; E *S1,exp. = E1/2@ES1,exp. and E *T1, exp. 77K = E1/2@ET1,77 K. DMF = N,N-dimethylformamide; CPCM = conductor-like polarizable continuum model;
MeCN = acetonitrile.

a decrease in molar absorptivity (e; difference of withdrawing groups or an extended p system as the N-aryl
& 18 000 m @1 cm@1) but the same lmax,abs at 361 nm (Table 1, moiety, were predicted to have localization of the higher-lying
entry 1). Despite increasing p-system conjugation through SOMO onto the N-aryl substituent. The PCs 4, 5, and 6 all
core substitution, most PCs did not absorb beyond 400 nm. As showed a higher-lying SOMO localization onto one core
an exception, the presence of 4-cyanophenyl groups at either substituent, while 3 showed a higher-lying SOMO distributed
the core or N-aryl positions resulted in greater red-shifting of across both core substituents (see Figure S1).
lmax,abs, yielding some visible-light absorption. Notably, all the Experimentally, CT character can be observed through
core-modified dimethyl dihydroacridine PC candidates a large Stokes shift and visualized through solvatochromism,
showed strong light absorption compared to the non-core- where the polar 1PC* is progressively stabilized by increasing
modified 1 a, with e ranging from 31 500 to 50 140 m @1 cm@1. solvent polarity, resulting in lower-energy emission and
This high degree of efficiency in photoexcitation is also a corresponding red-shifted lmax,em. Evaluation of CT from
1
corroborated through computationally predicted oscillator PC* can estimate the CT character of 3PC*, as CT singlet and
strengths (f values ranging from 1.211 to 1.749), indicating triplet excited states are expected to be energetically degen-
high p–p* transition probabilities (see Table S2). erate, with low DEST.[24] A high fluorescence quantum yield
To study the nature of PC*, which governs alkyl bromide (Ff) can indicate a lack of CT, as CT states have been shown
activation, a combination of DFT calculations and experi- to minimize fluorescence and increase triplet yields.[27] Con-
mental studies were used to evaluate the ability of these PCs sistent with previous observations in N,N-diaryl dihydrophe-
to access CT excited states, which have been empirically nazine and N-aryl phenoxazine studies, for dimethyl dihy-
shown to be beneficial for good O-ATRP performance in droacridines subtle N-aryl substitution differences were
structurally similar N,N-diaryl dihydrophenazine and N-aryl significant in influencing the nature of the experimentally
phenoxazine PCs.[25, 15a] In these systems, electron density is observed CT character. Of these candidates, 1 (Figure 2 d), 2,
transferred from the electron-rich core to either the N-aryl or and 7 displayed the largest degree of CT through the largest
core substituent, generating a shift in charge density within measured Stokes shifts (ranging from 126 to 180 nm) paired
the molecule in its excited state, which is dictated by the with low Ff (0.1 % to 8.7 %), and the most dramatic
electron-accepting ability of the aryl “acceptor” as well as the solvatochromism spanning blue to yellow wavelengths of
environment surrounding the “donor” tricyclic core. The emission (Figure 2 d; see Figure S1). By the same analysis, 4
connection between PC CT and superior O-ATRP perfor- and 5 (Figure 2 e and f) displayed a moderate degree of CT,
mance was investigated in two separate studies, but different while 3 and 6 displayed the least amount of CT character (see
conclusions were drawn: For N-aryl phenoxazines, increasing Figure S1). Notably, Ff values ranging from 0.1 % (2) to 83 %
CT character has been shown to augment triplet yields, (3) can be obtained by modulating the core substitution.
positing that higher concentrations of 3PC* promotes fast Evaluation of excited-state redox potentials of these PCs
activation.[27] Conversely, studies with N,N-diaryl dihydro- was performed by DFT calculations to predict E0*T1,calc.
phenazine PCs suggest that CT lowers *PC reduction (2PCC+/3PC*) values (Table 1). Experimentally, E0*S1,exp.
potentials, slowing down activation, reducing radical concen- (2PCC+/1PC*) was determined by a modified Rehm–Weller
trations, and minimizing termination.[15c] equation: E0*S1,exp. = E1/2@ES1,exp. , where ES1,exp. was measured
Computationally, CT characteristics can be predicted from the maximum wavelength of steady-state emission at
through the presence of charge-separated singly occupied room temperature. Experimental triplet energies (ET1,exp.)
molecular orbitals (SOMOs) for 3PC*. Of the PCs evaluated were measured from PC phosphorescence at 77 K with a 1 ms
in this study, 1, 2, and 7, which possess either electron- gate-delay using time-resolved spectroscopy. ES1,exp. was also

3212 www.angewandte.org T 2019 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2020, 59, 3209 – 3217
Angewandte
Research Articles Chemie

evaluated at 77 K with no gate delay, finding significant shifts ly).[17f, 7] In all cases these polymerizations were uncontrolled.
in emission profiles. Interestingly, the PCs 2 and 7, with the For 4Cz-IPN, Mn decreased over the course of polymerization
highest Stokes shifts, presented the lowest DEST values at 77 K from 170 to 63 kDa, with Y > 2.0 (see Figure S93). PhenN and
(0.39 and 0.34 eV; see Table S4). Furthermore, computational PhenO showed some characteristics of an O-ATRP process,
ET1 predictions corresponded well with experimental values, but suffered from nonlinear growth of the MW and high Y
with differences of less than 0.07 eV. (see Figures S91 and S92).
Like absorption, the electronics of core-substitution Initial application of 1–7 to O-ATRP of BA supported the
strongly influenced E0*T1,exp. , where the PC with the lowest- fact that CT was required for good O-ATRP performance,
energy absorption (3) also possessed the lowest E0*T1,exp. similar to what was observed in N,N-diaryl dihydrophena-
(@1.49 V vs. SCE). Correspondingly, 2, with a relatively zines.[25] For the PCs with the least amount of CT character
high-energy absorption, was predicted to have the most (PCs 3 and 6), the polymerization was uncontrolled, with
reducing E0*T1,exp. of @1.62 V vs. SCE. Despite diverse MWs decreasing with increasing monomer conversion and
absorption and CT properties, systematic study of various Y > 2.0. Excitingly, control over polymer MW growth, an
withdrawing, donating, and neutral N-aryl groups showed indication of efficient deactivation processes, was realized for
minimal influence on the reduction potential, with E0*T1,exp. 1, 2, 4, 5, and 7, with all I* values near 100 % (Table 2,
ranging from @1.59 to @1.55 V vs. SCE.
The stability of the deactivating species 2PCC+ was Table 2: Summary of results of O-ATRP of BA performed in batch and
measured by CV to determine E1/2 (2PCC+/1PC), finding flow using PCs 1–7.[a]
similar influence of the electronics of the core substituent as Entry PC Reactor Conv. Mn,calc, X I*
was observed for both measured absorption characteristics [%][b] [kDa][c] (Mw/Mn)[c] [%][d]
and E0*T1,exp. values. The presence of donating groups (2)
1 1 Batch 65 9.3 1.64 92
resulted in stabilization of 2PCC+ (E1/2 = 0.71 V vs. SCE).
2 2 Batch 77 10.6 1.53 96
Withdrawing groups (3) destabilized 2PCC+ and accessed 3 3 Batch 42 31.4 4.93 35
a strongly oxidizing 2PCC+ (E1/2 = 0.90 V vs. SCE). As in 4 4 Batch 68 9.6 1.62 93
evaluation of E0*T1,exp. , the nature of the N-aryl group has 5 5 Batch 59 8.7 1.70 90
minimal influence on E1/2, with a range of 0.75 to 0.82 V vs. 6 6 Batch 72 25.8 3.52 37
SCE. Computationally, the oxidizing ability of 2PCC+ (E0ox) 7 7 Batch 76 10.9 1.89 92
was predicted by DFT, with values systematically about 8 1 Flow 67 8.9 1.59 100
9 2 Flow 81 11.0 1.35 97
0.25 V lower than the experimentally measured E1/2, justifying
10 3 Flow 71 13.1 4.57 72
the difference in magnitude between E0*T1,exp. and E0*T1,calc. In 11 4 Flow 81 11.1 1.48 96
sum, the characterization of these dimethyl dihydroacridine 12 5 Flow 79 10.2 1.48 102
PC candidates shows that photophysical and electrochemical 13 6 Flow 82 11.4 3.58 94
properties can be tuned in an analogous fashion to other 14 7 Flow 73 9.6 1.54 100
established organic PCs with similar donor-acceptor motifs, [a] Reaction conditions: [1000]:[10]:[1] of [BA]:[DBMM]:[PC] with
while accessing more strongly oxidizing 2PCC+ characteristics 1.0 equiv. DMAc to BA by volume and irradiated by 365 nm LEDs. Batch
and offering opportunities for previously unaccessable reac- reactions are conducted under ambient temperatures and flow reactions
tivity. at 22 8C. [b] Determined by 1H NMR spectroscopy. [c] Measured using
GPC. [d] Calculated by (Conv. x [Mon]/[RX] x Mw,Mon)/1000. [e] Initiator
efficiency (I*) calculated by (Theo. Mn/calcd Mn) 100.

Application of Dimethyl Dihydroacridines to O-ATRP of Acrylate


Monomers entries 1–7), where I* = (Mn,theo./Mn,GPC)x 100. The PC 2
provided the best results (I* = 96 %, 77 % conversion; Fig-
To test the ability of these dimethyl dihydroacridine PC ure 3 a), but produced a polymer with high Y at 1.53. For 1, 2,
candidates to catalyze O-ATRP of challenging acrylate 4, 5, and 7, bimodal gel permeation chromatography (GPC)
monomers, initial polymerizations were conducted using n- traces revealed an accompanying high MW p(BA) species
butyl acrylate (BA) as the monomer and diethyl 2-bromo-2- present in low quantities. Control experiments with BA in
methylmalonate (DBMM) as the initiator in N,N-dimethyla- DMAc under 365 nm irradiation revealed significant mono-
cetamide (DMAc) with 365 nm LED irradiation under batch mer autopolymerization (63 % conversion at 2.5 hours; see
reactor conditions. For comparison to dimethyl dihydroacri- Table S3) and thus likely contributed to the formation of this
dine PCs, other well-studied organic PCs with diverse redox undesired high MW species.
properties, including 3,7-di(4-biphenyl) 1-naphthalene-10- As the best performing PC under these reaction con-
phenoxazine (PhenO), 5,10-di-(2-naphthyl)-5–10-dihydro- ditions, 2 possessed the highest E0*T1,exp. but also the lowest E1/
phenazine (PhenN), and 1,2,3,5-tetrakis(carbazol-9-yl)-4,6- 2, we hypothesized that promoting efficient activation could
dicyanobenzene (4Cz-IPN) were applied in the O-ATRP of outcompete autopolymerization side-reactivity, leading to
BA under the same reaction conditions. PhenO and PhenN lower Y while maintaining control over MW growth. Notably,
both possess strongly reducing E0*T1,calc values (@1.70 and acrylate alkyl halide chain-end groups are more difficult to
@2.12 V vs. SCE) and somewhat stabilized E1/2 values (0.65 reduce compared to methacrylates, which can be observed by
and 0.21 V vs. SCE), while 4Cz-IPN is a moderate reductant CV measurements of representative alkyl bromide initiators
and strong oxidant (@1.06 and 1.50 V vs. SCE, respective- with methacrylate and acrylate functionalities. An acrylate

Angew. Chem. Int. Ed. 2020, 59, 3209 – 3217 T 2019 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.angewandte.org 3213
Angewandte
Research Articles Chemie

monium bromide salts showed some decrease in Y, with LiBr


providing the greatest effect (Y , 1.23; see Table S20). O-
ATRP of BA using lithium salts with diverse anions (LiCl, LiI,
LiPF6) was also performed. A complete loss of MW control
and Y > 2.0 in all cases was observed (see Table S20),
illustrating the significance of the bromide anion in deactiva-
tion. Analysis of relative LiBr concentrations showed the
conditions producing polymers with lowest Y (30 equiv. LiBr
relative to PC, or 83.3 mm) to also have the slowest overall
reaction rate (Figure 3 c and d), which paired with lower Y,
suggests LiBr concentration plays a key role in deactivation.
Next, a combination of DFT calculations and experimen-
tal studies were employed to explore the role of the LiBr salt
in this O-ATRP system, probing its potential influence on
both activation and deactivation mechanistic steps. One
possibility is a Lewis acid driven activation by Li+ of the
alkyl bromide ester polymer chain-end, facilitating rapid
activation. However, CV revealed no change in reduction Ep/2
(RX/RXC@) of methyl-2-bromopropionate upon addition of
Figure 3. Plots of Mn vs. conversion (blue) and X vs. conversion (red)
for O-ATRP of BA plotted against the theoretical Mn (dashed line) LiBr (Ep/2 = @0.92 V vs. SCE; see Figure S61). To evaluate
using 2 conducted in a batch reactor (a), in flow reactor (b), and in potential effects of LiBr on PC photophysical behavior,
a flow reactor with 30 equiv. LiBr relative to PC (c). d) First-order absorption and emission characterization of 2 was performed
kinetic plot of O-ATRP of BA with varying LiBr equivalents relative to in the presence of LiBr, thus revealing no changes in these
PC. Reaction conditions are [1000]:[10]:[1]:[x] of [BA]:[DBMM]:[PC 2]:- properties (see Figure S83 and S84).
[LiBr] with 1.0 equiv by volume of DMAc for (a) and 1.5 equiv for (b)
Formally, the deactivation step is a termolecular process
and (c), with irradiation using 365 nm LEDs.
requiring low concentrations of PnC, 2PCC+, and X@ to collide,
which is entropically unfavorable. Previous studies with
analogue, methyl 2-bromopropionate, was found to undergo dihydrophenazines have shown that decreasing solvent polar-
one-electron reduction at Ep/2 = @0.96 V vs. SCE, while ity promotes ion-pairing of 2PCC+X@ , with a subsequent
a methacrylate model, methyl a-bromoisobutyrate, was decrease in polymerization rate and improvement in perfor-
reduced at Ep/2 = @0.80 V vs. SCE, or a difference of 0.16 V mance.[25] As such, we propose that efficient deactivation
(see Figure S67), highlighting the need for high E0*T1,exp. for requires oxidation of PnC by a 2PCC+X@ ion-pair, which shifts
efficient activation of acrylate-derived chain-end groups. a formal three-body collisional event to a more favorable
One alternative to batch reactor systems are continuous- pseudo-two-body event.[11, 23] In addition, we hypothesize
flow reactors, which facilitate more uniform reaction irradi- a concerted mechanism for the oxidation of PnC by a 2PCC+X@
ation and have been shown to improve photoinduced CRP ion-pair, where the formation of Pn-X is tied to the nuclear
systems through enhancing PC photoexcitation, thus promot- coordinate of X@ , thus avoiding the thermodynamically
ing efficient alkyl bromide activation.[28] In this approach, unfavorable carbocation Pn+ intermediary species. Using
a commercially available temperature-controlled Hepato- DFT, the association of 2PCC+ + X@ ! 2PCC+X@ using 2 was
chem PhotoRedOx box, an 18 W EvoluChem 365 nm LED, predicted to be slightly endergonic (DG0complex = 0.6 kcal
and a 2 mL PFA flow reactor was employed. Importantly, mol@1). As such, we postulate that the presence of excess
control experiments with BA in DMAc under these con- bromide ions through LiBr addition increases the population
ditions did not produce any undesired autopolymerization of the proposed 2PCC+X@ deactivator species, promoting rapid
(see Table S14). The PCs 1–7 were then evaluated for the O- deactivation of the propagating chain-end and explaining the
ATRP of BA under continuous-flow conditions, showing observed decrease in polymerization rate and lower Y.
similar trends in performance as batch conditions, albeit with To further elucidate the relative importance of PC choice
consistently lower Y (Table 2, entries 8–14). Again, 2 proved and the optimized reaction conditions in successful O-ATRP
superior, however with initial high Y > 1.5 at conversions less of acrylates, polymerizations were also conducted using 4-
than 50 %, with Y = 1.35 (I* = 97 % at 81 % conversion). CzIPN, PhenO, and PhenN under continuous-flow conditions
By using 2, further reaction optimization analyzing the with LiBr as an additive. For 4Cz-IPN, no polymerization was
effects of initiator, PC loading, and reaction concentration observed in the presence of LiBr, while O-ATRP without
were conducted (see Figures S124–S131). Increasing the LiBr yielded an uncontrolled polymerization with Mn >
DMF:BA ratio (1:1 to 1.5:1 v/v) significantly improved 40 kDa and Y > 2.0. (see Table S21). The O-ATRP of BA
polymerization results, with Y < 1.5 at all monomer conver- using PhenO and PhenN was moderately improved under the
sions (Figure 3 b). To further improve the system, the addition optimized reaction conditions, showing some control over
of various bromide salts was investigated, as this has been MW growth and with Y < 1.5 at higher monomer conversions.
shown to promote deactivation in aqueous copper-catalyzed However, both polymerizations displayed bimodal MW
ATRP systems and in some photoredox-catalyzed ATRP distributions by GPC, with I* ranging from 31–109 %.
systems.[29] Analysis of LiBr, NaBr, KBr, and tetrabutylam-

3214 www.angewandte.org T 2019 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2020, 59, 3209 – 3217
Angewandte
Research Articles Chemie

Using these optimized reaction conditions, MW control in radical termination products (see Figure S161). Notably, no
the O-ATRP of BA catalyzed by 2 was demonstrated through significant difference in chain-end groups between polymers
adjustment of reaction stoichiometry, accessing MWs ranging synthesized with and without LiBr were observed (see
from 2 to 26 kDa with Y < 1.36 (Table 2, entries 1–5). Figure S162).
However, targeting high MWs proved challenging, with I* > To validate the chain-end group fidelity of the system and
100 %, perhaps from a confluence of increased termination demonstrate the ability of this system to produce complex
events and viscosity changes within the reactor. The O-ATRP polymeric materials, chain-extensions were performed under
of other alkyl and glycol acrylate monomers was performed continous-flow conditions for an isolated p(n-BA) macro-
using 2, with excellent control over MW and Y (entries 7–11). initiator (Mn = 4.6 kDa, Y = 1.26) with ethyl acrylate (EA) to
O-ATRP of methyl methacrylate (MMA) was also accom- produce a block-copolymer, p(n-BA)-b-p(EA), with Y = 1.16
plished, realizing low Y but with slightly bimodal GPC traces (12 kDa). This polymer was then again reintroduced as
(entry 12). As expected, given the lower kp value, the O- a macroinitiator and further extended with tert-butyl acrylate
ATRP of MMA was significantly slower than BA, requiring to produce a well-defined triblock copolymer (Mn = 20 kDa,
600 minutes of residence time for 72 % monomer conversion, Y = 1.44; Figure 4). An 1H NMR spectrum for the p(n-BA)
as compared to about 100 minutes for acrylates. macroinitiator with detailed assignments (see Figure S158)
Under batch conditions, no monomer conversion was also shows the presence of a- and w-chain-end groups.
observed using 2 at ambient atmosphere (see Table S5).
Recently, N-aryl phenoxazines were found to perform a well-
controlled O-ATRP under ambient conditions when no vial
headspace was present.[17e] To further test the oxygen-
tolerance of dimethyl dihydroacridines, the O-ATRP of BA
was performed under optimized flow conditions using re-
agents and solvents previously exposed to ambient conditions
proceeded efficiently with no observed induction period, with
Y , 1.24 and I* close to 100 % during polymerization
(Table 3, entry 6).
To demonstrate the temporal control characteristic of an
O-ATRP process, pulsed irradiation polymerization experi-
ments were performed under batch conditions, showing no Figure 4. Results of sequential chain-extension experiments with corre-
monomer conversion during periods of no irradiation (see sponding GPC traces to produce a p(n-BA)-b-p(EA)-b-p(t-BA) triblock
Figure S115). Analysis of bromide chain-end group retention copolymer.
of p(BA) was performed through MALDI mass spectrometry,
revealing the presence of H-terminated and bimolecular
Conclusion

Table 3: Summary of results of O-ATRP of acrylate monomers conducted using PC 2.[a] In summary, we have re-
ported the first successful O-
ATRP of acrylate monomers
with controlled MW and low
Y, enabled by rational de-
sign and development of
a new family of dimethyl
Entry Monomer [Monomer]:[DBMM] Res. Time Conv. Mn,calc, Mw,calc, X I*
dihydroacridine photocata-
[min] [%] (kDa) (kDa) (Mw/Mn) [%] lysts. Relationships between
PC properties and catalytic
1 n-BA [1000]:[2.5] 90 89 26.4 45.7 1.35 173
performance are discussed,
2 n-BA [1000]:[5] 120 88 15.2 22.8 1.32 150
3 n-BA [1000]:[10] 120 73 9.4 9.7 1.20 103 finding an interplay of pho-
4 n-BA [1000]:[20] 180 80 7.3 5.4 1.19 74 tophysical and electrochem-
5 n-BA [1000]:[40] 180 67 5.4 2.4 1.18 44 ical properties is necessary to
6[b] n-BA [1000]:[10] 120 85 10.3 11.1 1.24 108 achieve the desired reactivi-
7 t-BA [1000]:[10] 120 92 13.8 12.1 1.23 88 ty, as is supported by the best
8 MA [1000]:[10] 90 92 10.0 8.1 1.30 81
performing PC having the
9 EA [1000]:[10] 90 76 7.5 7.8 1.19 105
10 2-EHA [1000]:[10] 90 90 16.3 16.8 1.53 104
highest E0*T1,exp. but the low-
11 EGMEA [1000]:[10] 60 92 10.5 12.3 1.37 117 est E1/2. A combination of
12 MMA [1000]:[10] 600 72 8.8 7.4 1.17 85 reactor choice and LiBr ad-
ditives is found to be signifi-
[a] Reaction conditions: [1000]:[x]:[1]:[30] of [monomer]:[DBMM]:[PC 2]:[LiBr] with 1.5 mL DMAc relative
to 1 mL monomer performed in continuous-flow and irradiated by 18 W 365 nm LED at 22 8C. cant in achieving a well-con-
[b] Reaction components sparged with air for 30 minutes before polymerization. See the Supporting trolled system, and are pro-
Information for full experimental details. posed to promote the respec-

Angew. Chem. Int. Ed. 2020, 59, 3209 – 3217 T 2019 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.angewandte.org 3215
Angewandte
Research Articles Chemie

tive activation and deactivation steps of the O-ATRP cycle. [9] a) G. M. Miyake, J. C. Theriot, Macromolecules 2014, 47, 8255 –
We envision that dimethyl dihydroacridines, with distinct 8261; b) N. J. Treat, H. Sprafke, J. W. Kramer, P. G. Clark, B. E.
properties and new reactivities, will further enable the Barton, J. R. Alaniz, B. P. Fors, C. J. Hawker, J. Am. Chem. Soc.
2014, 136, 16096 – 16101.
advancement of organic photoredox catalysis in both small-
[10] a) J. C. Theriot, C.-H. Lim, H. Yang, M. D. Ryan, C. B.
molecule and macromolecular syntheses, aiding in the
Musgrave, G. M. Miyake, Science 2016, 352, 1082 – 1086.
replacement of precious-metal catalysts for sustainable pho- [11] a) R. M. Pearson, C.-H. Lim, B. G. McCarthy, C. B. Musgrave,
toredox processes. G. M. Miyake, J. Am. Chem. Soc. 2016, 138, 11399 – 11407.
[12] a) V. K. Singh, C. Yu, S. Badgujar, Y. Kim, Y. Kwon, D. Kim, J.
Lee, T. Akhter, G. Thangavel, L. S. Park, J. Lee, P. C. Nandajan,
Acknowledgements R. Wannemacher, B. Milian-Medina, L. Luer, K. S. Kim, J.
Gierschner, M. S. Kwon, Nat. Catal. 2018, 1, 794 – 804; b) E. H.
This work was supported by Colorado State University and Discekici, A. Anastasaki, J. Read de Alaniz, C. J. Hawker,
Macromolecules 2018, 51, 7421 – 7434.
the NIH (R35GM119702). This work used the Extreme
[13] a) B. Narupai, Z. A. Page, N. J. Treat, A. J. McGrath, C. W.
Science and Engineering Discovery Environment (XSEDE), Pester, E. H. Discekici, N. D. Dolinski, G. F. Meyers, J. Read
which is supported by National Science Foundation Grant de Alaniz, C. J. Hawker, Angew. Chem. Int. Ed. 2018, 57, 13433 –
ACI-1548562. C.-H.L. acknowledges a National Institutes of 13438; Angew. Chem. 2018, 130, 13621 – 13626; b) E. H. Dis-
Health F32 Postdoctoral Fellowship (F32GM122392). The cekici, C. W. Pester, N. J. Treat, J. Lawrence, K. M. Mattson, B.
authors thank Dr. Matthew Ryan, Dr. Dianfeng Chen, and Narupai, E. P. Toumayan, Y. Luo, A. J. McGrath, P. G. Clark, J.
Max Kudisch for technical assistance and helpful discussions. Read de Alaniz, C. J. Hawker ACS Macro Lett. 2016, 5, 258 –
262.
[14] X. Pan, C. Fang, M. Fantin, N. Malhotra, W. Y. So, L. A. Peteanu,
A. A. Isse, A. Gennaro, P. Liu, K. Matyjaszewski, J. Am. Chem.
Conflict of interest Soc. 2016, 138, 2411 – 2425.
[15] a) B. G. McCarthy, R. M. Pearson, C.-H. Lim, S. M. Sartor, N. H.
The authors declare no conflict of interest. Damrauer, G. M. Miyake, J. Am. Chem. Soc. 2018, 140, 5088 –
5101; b) J. P. Cole, C. R. Federico, C.-H. Lim, G. M. Miyake,
Keywords: microreactors · monomers · organocatalysis · Macromolecules 2019, 52, 747 – 754; c) D. Koyama, H. J. A.
photochemistry · polymerization Dale, A. J. Orr-Ewing, J. Am. Chem. Soc. 2018, 140, 1285 – 1293.
[16] a) D. A. Corbin, C.-H. Lim, G. M. Miyake, Aldrichimica Acta
2019, 52, 7 – 21.
[17] a) M. D. Ryan, R. M. Pearson, T. A. French, G. M. Miyake,
Macromolecules 2017, 50, 4616 – 4622; b) M. D. Ryan, J. C.
How to cite: Angew. Chem. Int. Ed. 2020, 59, 3209 – 3217 Theriot, C.-H. Lim, H. Yang, A. G. Lockwood, N. G. Garrison,
Angew. Chem. 2020, 132, 3235 – 3243 S. R. Lincoln, C. B. Musgrave, G. M. Miyake, J. Polym. Sci. Part
A 2017, 55, 3017 – 3027; c) B. L. Ramsey, R. M. Pearson, L. R.
Beck, G. M. Miyake, Macromolecules 2017, 50, 2668 – 2674;
[1] a) N. A. Romero, D. A. Nicewicz, Chem. Rev. 2016, 116, 10075 – d) B. L. Buss, L. R. Beck, G. M. Miyake, Polym. Chem. 2018, 9,
10166; b) C. K. Prier, D. A. Rankic, D. W. C. MacMillan, Chem. 1658 – 1665; e) B. McCarthy, G. M. Miyake, ACS Macro Lett.
Rev. 2013, 113, 5322 – 5363; c) J. M. R. Narayanam, C. R. J. 2018, 7, 1016 – 1021; f) Y. Du, R. M. Pearson, C.-H. Lim, S. M.
Stephenson, Chem. Soc. Rev. 2011, 40, 102 – 113; d) L. Buzzetti, Sartor, M. D. Ryan, H. Yang, N. H. Damrauer, G. M. Miyake,
G. E. M. Crisenza, P. Melchiorre, Angew. Chem. Int. Ed. 2019, Chem. Eur. J. 2017, 23, 10962 – 10968; g) H. Rao, C.-H. Lim, J.
58, 3730 – 3747; Angew. Chem. 2019, 131, 3768 – 3786; e) N. Bonin, G. M. Miyake, M. Robert, J. Am. Chem. Soc. 2018, 140,
Corrigan, S. Shanmugam, J. Xu, C. Boyer, Chem. Soc. Rev. 2016,
17830 – 17834.
45, 6165 – 6212.
[18] a) X. Pan, M. Lamson, J. Yan, K. Matyjaszewski, ACS Macro
[2] a) F. A. Leibfarth, K. M. Mattson, B. P. Fors, H. A. Collins, C. J.
Lett. 2015, 4, 192 – 196; b) D.-F. Chen, B. M. Boyle, B. G.
Hawker, Angew. Chem. Int. Ed. 2013, 52, 199 – 210; Angew.
McCarthy, C. H.- Lim, G. M. Miyake, J. Am. Chem. Soc. 2019,
Chem. 2013, 125, 210 – 222; b) M. Chen, M. Zhong, J. A.
141, 13268 – 13277; c) G. Ramakers, A. Krivkov, V. Trouillet, A.
Johnson, Chem. Rev. 2016, 116, 10167 – 10211; c) N. Corrigan,
Welle, H. Mobius, T. Junkers, Macromol. Rapid Commun. 2017,
J. Yeow, P. Judzewitsch, J. Xu, C. Boyer, Angew. Chem. Int. Ed.
38, 1700423.
2019, 58, 5170 – 5189; Angew. Chem. 2019, 131, 5224 – 5243.
[3] a) K. Matyjaszewski, Adv. Mater. 2018, 30, 1706441; b) T. G. [19] M. Destarac, Polym. Chem. 2018, 9, 4947 – 4967.
Ribelli, F. Lorandi, M. Fantin, K. Matyjaszewski, Macromol. [20] W. Tang, K. Matyjaszewski, Macromolecules 2007, 40, 1858 –
Rapid Commun. 2019, 40, 1800616; c) M. Ouchi, T. Terashima, 1863.
M. Sawamoto, Chem. Rev. 2009, 109, 4963 – 5050. [21] S. Beuermann, M. Buback, Prog. Polym. Sci. 2002, 27, 191 – 254.
[4] a) Q. Yang, F. Dumur, F. Morlet-Savary, J. Poly, J. Lalevee, [22] A. Anastasaki, V. Nikolaou, Q. Zhang, J. Burns, S. R. Samanta,
Macromolecules 2015, 48, 1972 – 1980; b) G. Zhang, I. Y. Song, C. Waldron, A. J. Haddleton, R. McHale, D. Fox, V. Percec, P.
K. H. Ahn, T. Park, W. Choi, Macromolecules 2011, 44, 7594 – Wilson, D. M. Haddleton, J. Am. Chem. Soc. 2014, 136, 1141 –
7599; c) B. P. Fors, C. J. Hawker, Angew. Chem. Int. Ed. 2012, 51, 1149.
8850 – 8853; Angew. Chem. 2012, 124, 8980 – 8983. [23] N. J. Treat, B. P. Fors, J. W. Kramer, M. Christianson, C.-Y. Chiu,
[5] S. Shanmugam, C. Boyer, Science 2016, 352, 1053 – 1054. J. Read de Alaniz, C. J. Hawker, ACS Macro Lett. 2014, 3, 580 –
[6] G. Yilmaz, Y. Yagci, Polym. Chem. 2018, 9, 1757 – 1762. 584.
[7] J. C. Theriot, B. G. McCarthy, C.-H. Lim, G. M. Miyake, Macro- [24] a) M. Y. Wong, E. Zysman-Colman, Adv. Mater. 2017, 29,
mol. Rapid Commun. 2017, 38, 1700040. 1605444; b) W. Li, B. Li, X. Cai, L. Gan, Z. Xu, W. Li, K. Liu,
[8] H. G. Roth, N. A. Romero, D. A. Nicewicz, Synlett 2016, 27, D. Chen, S.-J. Su, Angew. Chem. Int. Ed. 2019, 58, 11301 – 11305;
714 – 723. Angew. Chem. 2019, 131, 11423 – 11427.

3216 www.angewandte.org T 2019 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2020, 59, 3209 – 3217
Angewandte
Research Articles Chemie

[25] C.-H. Lim, M. D. Ryan, B. G. McCarthy, J. C. Theriot, S. M. Angew. Chem. Int. Ed. 2012, 51, 4144 – 4147; Angew. Chem. 2012,
Sartor, N. H. Damrauer, C. B. Musgrave, G. M. Miyake, J. Am. 124, 4220 – 4223.
Chem. Soc. 2017, 139, 348 – 355. [29] a) A. Simakova, S. E. Averick, D. Konkolewicz, K. Matyjaszew-
[26] D. M. Arias-Rotondo, J. K. McCusker, Chem. Soc. Rev. 2016, 45, ski, Macromolecules 2012, 45, 6371 – 6379; b) M. Fantin, A. A.
5803 – 5820. Isse, A. Gennaro, K. Matyjaszewski, Macromolecules 2015, 48,
[27] S. M. Sartor, B. G. McCarthy, R. M. Pearson, J. M. Miyake, N. H. 6862 – 6875.
Damrauer, J. Am. Chem. Soc. 2018, 140, 4778 – 4781.
[28] a) R. L. Hartman, J. P. McMullen, K. F. Jenson, Angew. Chem.
Int. Ed. 2011, 50, 7502 – 7519; Angew. Chem. 2011, 123, 7642 – Manuscript received: August 23, 2019
7661; b) B. L. Buss, G. M. Miyake, Chem. Mater. 2018, 30, 3931 – Revised manuscript received: November 12, 2019
3942; c) T. Junkers, B. Wenn, React. Chem. Eng. 2016, 1, 60; Accepted manuscript online: November 26, 2019
d) J. W. Tucker, Y. Zhang, T. F. Jamison, C. R. J. Stephenson, Version of record online: January 21, 2020

Angew. Chem. Int. Ed. 2020, 59, 3209 – 3217 T 2019 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.angewandte.org 3217

You might also like