You are on page 1of 13

pubs.acs.

org/Macromolecules Article

Backbone-Degradable Acrylate Latex: Toward Overcoming


Hydrolysis Limitations of Cyclic Ketene Acetal Monomers
Srinivasa Reddy Mothe,# Wenguang Zhao,# Alexander M. van Herk, and Praveen Thoniyot*
Cite This: https://doi.org/10.1021/acs.macromol.3c02474 Read Online

ACCESS Metrics & More Article Recommendations *


sı Supporting Information
Downloaded via NORTHWESTERN POLYTECHNICAL UNIV on March 20, 2024 at 00:55:04 (UTC).
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: Latex particles made with radical emulsion polymer-


ization contain all carbon-backbone polymer chains, which are
nonhydrolyzable and nonbiodegradable in the environment and
hence pose a huge challenge as persistent pollutants. Radical ring-
opening copolymerization (rROP) of cyclic ketene acetals (CKAs)
such as 2-methylene-1,3-dioxepane (MDO) with vinyl monomers
is an interesting approach being explored intensely for introducing
degradable ester groups in the polymer backbones, opening up the
potential of synthesizing biodegradable copolymers. However,
carrying out such reactions in an aqueous medium to obtain
biodegradable dispersions is very challenging due to the hydrolytic
instability of CKAs. There are conflicting reports on preparing degradable lattices via rROP of MDO with vinyl acetate (VAc)
monomers, but subsequent detailed reaction parameters were published, which show that under the right conditions, successful
copolymerizations can be performed. We report that the use of neutral surfactants, high pH, and stabilizing comonomers play a
crucial role in the successful incorporation of degradability in the polymer backbone during emulsion copolymerization. This is the
first report on the synthesis of acrylate copolymers with the incorporation of hydrolytic degradability (utilizing MDO) under
potentially more industrially feasible emulsion polymerization conditions. The findings underscore the potential for further
improvement in this area. The results in the current study provide further insight into the breadth of parameters that must be
carefully considered to produce biodegradable lattices via the rROP of CKAs.

■ INTRODUCTION
A wide variety of acrylate-based polymers have been developed
alone does not imply biodegradability; the latter is contingent
on the biodegradability of the formed oligomers.18
over the years with applications such as plastics, waterborne Later on, rROP of CKAs and vinyl monomers is routinely
coatings,1 adhesives,2 nanocomposites,3 drug carriers,4,5 and exploited18−22 and has been employed to prepare degradable
scale control agents.6 Acrylate copolymers are commonly used poly(styrene), 23 poly(methacrylate),24−26 various poly-
in personal or consumer care applications due to their tunable (acrylates),27−29 poly(vinyl ethers),30 and poly(vinyl ace-
functionalities that enable the incorporation of required tates).31 The degradable polymers synthesized via the
benefits in formulations. At the same time, it is almost incorporation of CKA units usually require high temperatures
unavoidable that these polymers wind up in the environment. and water-free conditions.32 These polymerizations are neither
With introducing degradable units in the carbon−carbon main suitable in combination with certain functional groups nor can
chain of these polymers, the accumulation of solid waste in the they be easily performed in water.33,34 Polymer dispersions of
environment,7 particularly the leakage of synthetic polymers degradable acrylates made by emulsion polymerization are of
from several consumer care and agricultural industries to our great interest due to the use of water as a nontoxic and green
oceans, might be alleviated. There has been a rapidly increasing reaction medium. Of course, the challenge is the hydrolysis of
interest in this area in recent years.8−10 The degradable version CKAs due to the diffusion of the monomer from the droplets
of acrylate copolymers is typically prepared by free-radical ring- to the aqueous phase into the particles. Miniemulsion
opening polymerization (rROP) of cyclic ketene acetal (CKA) polymerization does not require the transport of monomer
with vinyl acrylate monomers. The early work of CKA
polymerization by Bailey and co-workers in the 1980s unveiled
the synthesis of degradable copolymers by introducing Received: November 30, 2023
cleavable ester linkages into the backbone of traditional Revised: February 12, 2024
nondegradable polymers, which led to advanced applications Accepted: February 22, 2024
of these degradable plastics.11−17 It is noteworthy that the
degradability of these copolymers is often demonstrated under
alkaline conditions. However, it is crucial to note that this

© XXXX American Chemical Society https://doi.org/10.1021/acs.macromol.3c02474


A Macromolecules XXXX, XXX, XXX−XXX
Macromolecules pubs.acs.org/Macromolecules Article

Scheme 1. Emulsion Polymerization of MDO and MMA and HEA to Afford Main-Chain-Degradable Poly(methyl
methacrylate) Nanoparticles

through the aqueous phase, in contrast to regular emulsion control was not as strict in the latter case and the instantaneous
polymerization (see later for a further discussion).35 monomer conversions were not as close to starved conditions
Miniemulsion polymerization has led to a wide variety of (this is most likely due to a different way of administering the
surface functionalities and the encapsulation of different kinds redox initiation system).42 Despite the successful emulsion
of materials, e.g., drugs, dyes, inorganic materials, etc. In 2012, copolymerization of MDO and VAc, there are no reports on
the Landfester group published a miniemulsion process for the emulsion copolymerization of MDO and acrylates. The
making degradable polymeric nanoparticles via rROP of 5,6- main reason is the very unfavorable reactivity ratios of the two
benzo-2-methylene-1,3-dioxepane (BMDO) with methyl monomers (rMDO = 0.04, rMMA = 3.5)20 coupled with the
methacrylate (MMA) and styrene; this was the first report in strong hydrolytic instability. With BMDO, there are more
which water is the continuous phase.36 It must be remarked favorable reactivity ratios with most monomers, for example,
here that both the reactivity ratios and the hydrolytic stability BMDO and MMA (rBMDO = 0.53 and rMMA = 1.96) and a
of BMDO are more favorable than that of MDO. Although higher hydrolytic stability, for this combination in emulsion
miniemulsion polymerization has been explored for the copolymerization, about 6% insertion of ester bonds was
development of a wide range of polymer materials,37 its achieved and degradability was reported.43 Recently, we
scope is typically on very hydrophobic monomers. Currently further explored the MDO hydrolysis mechanism in water
commercial products derived from miniemulsion polymer- and evaluated conditions that enabled aqueous phase solution
ization are much less prevalent compared to emulsion copolymerization of MDO with hydroxyethyl acrylate and
polymerization products.38 After this miniemulsion work, acrylamide.44 In both cases, we could only observe very
almost a decade passed without significant progress in aqueous minimal incorporation of MDO in the copolymer (as
polymerization attempts in CKA chemistry, even though evidenced by the hydrolysis experiments of the resulting
success was reported in creating degradable latex with other copolymer), and the majority of MDO underwent hydrolysis.
chemistries that insert degradability, such as the example of However, this study provided the hope that under the right
dibenzo[c,e]oxepane-5-thione (DOT) copolymerization with conditions, MDO emulsion copolymerization could be carried
acrylates and styrene.39 There is still a need for the CKA-based out even with methacrylates, as predicted by Bailey who
breakable bond insertion since ester groups might be preferred pioneered the field decades ago.45 From this study, it became
over the thioester groups, where oligomer yellowing and smell clear that the hydrolysis of MDO is predominantly acid-
associated with thiol groups can be an issue in various catalyzed.44 In the case of VAc-MDO, the pH was limited by
applications, which will not be the case for carboxyl esters the occurrence of saponification of VAc and was optimized to
derived from CKA monomers. pH = 8. Even though VAc and MDO have favorable reactivity
A breakthrough in the aqueous phase polymerization of ratios, we chose to work with acrylates, as acrylates do not
CKAs was recently reported by Carter et al. where a high undergo saponification like VAc. Therefore, acrylates have less
incorporation of 2-methylene-1,3-dioxepane (MDO) was limitations to high pH conditions, which suitably slow down
achieved by emulsion copolymerization of MDO with vinyl the hydrolysis of CKA monomers.44,45 Following this
acetate (VAc) as the comonomer and the resulting latex was approach, we report the synthesis of degradable poly(MMA)
utilized for biodegradable single-use paper coatings.40 The nanoparticles via the batch emulsion copolymerization of
authors mentioned a favorable reactivity ratio, adequate pH MMA with MDO and 2-hydroxyethyl acrylate (HEA) with a
adjustment, and high instantaneous conversions (starved-feed 75:20:5 ratio, respectively. Polymers with up to 9 mol % of
addition of monomers) as the main reason for the high degradable units in the main-chain polymer backbone were
incorporation of MDO in the vinyl latex. The starved obtained. Subsequent hydrolytic degradation of the polymer
monomer feeding strategy resulted in a very short residence results in small oligomeric degradation products (Scheme 1).
time of MDO in its monomeric form in water, minimizing the We realize that the batch approach is not optimal for monomer
hydrolysis. However, subsequent work by Agarwal et al. failed pairs with significantly different reactivity ratios. Further
to reproduce the high incorporation of degradability in VAc optimization is deemed feasible, and a semibatch approach is
latex, indicating that reproducibility is still challenging.41 Their considered with a more optimal monomer addition profile for
kinetic hydrolysis studies on MDO in water under a variety of these types of monomers.29 This strategy will be investigated in
pH and polarity conditions with cosolvents have concluded our future studies to enhance the synthesis process, including
that preparing biodegradable emulsion polymers using the aspects such as a higher solid content, yield, and achieving
rROP of MDO and vinyl monomers (including VAc) is still a better control over reactivity, particularly regarding MDO
challenge. Following these conflicting reports, Carter et al. incorporation in the polymer backbone. Even though the
provided more in-depth data and quantitative analysis tools in competing hydrolysis could not be avoided fully under our
a recent report, giving further proof for the successful emulsion current reaction conditions, we expect this study to provide a
polymerization of VAc and MDO.42 The main differences framework of conditions under which degradable acrylate latex
between the work of Carter and that of Agarwal are that pH can be successfully prepared by rROP of CKA monomers.
B https://doi.org/10.1021/acs.macromol.3c02474
Macromolecules XXXX, XXX, XXX−XXX
Macromolecules pubs.acs.org/Macromolecules Article

■ EXPERIMENTAL SECTION
Materials. All chemicals were purchased from Sigma-Aldrich and
Table 1. Optimization of Reaction Conditions for the
Synthesis and Characterization of Poly(MDO-co-MMA)
used as received unless stated otherwise. MDO12 and PEG-b-PCL46 Nanoparticles through Emulsion Polymerization to Obtain
were synthesized as previously reported. Monomers: ethylene glycol Stable Dispersions (EM3, EM5−EM7, and EM9) and with
dimethacrylate (EGDMA), MMA, HEA, and 2-hydroxyethyl meth- Improved MDO% Incorporation (EM4)a
acrylate (HEMA), n-butyl acrylate (BA), ethyl acrylate (EA), 2-
(dimethylamino)ethyl methacrylate (DMAEMA), methyl acrylate initial comonomer MDO
(MA), n-butyl methacrylate (BMA), 2-ethylhexyl acrylate (EHA), and mixture of incorporation in
MMA:MDO:HEA the final polymer
allyl methyl acrylate (AMA) were purchased from Sigma-Aldrich and experiment (mol %) conditions (mol %)b
passed through basic Al2O3 to remove inhibitors before polymer- c
EM1 80:20:0 SDS (0.125 g),
ization. Initiators: ammonium persulfate (APS) and 2,2′-azobis(2- TBAB (0.2 g) c
methylpropionamidine) dihydrochloride (V-50) were purchased from EM2 75:20:5
Sigma-Aldrich and recrystallized from methanol before use. EM3 75:20:5 triton X-100 3 ± 0.07
(0.125 g),
Surfactants: Triton X-100, sodium dodecyl sulfate (SDS), Poloxamer, TBAB (0.2 g)
Ryoto sugar ester, Span 80, Span 85, Triton X-405, Tween 20, Tween
EM4 75:20:5 triton X-100 9 ± 0.07
80, dihexyl sulfosuccinate sodium, sodium dodecylbenzene sulfonate (0.125 g)
EM5d 75:20:5 7 ± 0.07
(SDBS), cetyltrimethylammonium bromide (CTAB), and benzetho-
nium chloride (hyamine) were purchased from Sigma-Aldrich and EM6e 75:20:5 7 ± 0.02
used as received. Sodium hydroxide (NaOH), potassium hydroxide EM7 75:20:5 poloxamer-188 6 ± 0.02
(KOH), and tetra-n-butylammonium bromide (TBAB) were (0.125 g)
purchased from Sigma-Aldrich and used as received. 5 M NaOH EM8 80:20:0 triton X-100 <1
(0.125 g)
was prepared and stored as a stock solution. Chloroform-d (CDCl3)
was purchased from Cambridge Isotope Laboratories (CIL) and used EM9f 75:20:5 PEG-b-PCL 6
(0.125 g)
as received. The solvents tetrahydrofuran (THF; contains 0.005% of a
water), methanol (MeOH; contains 0.01% of water), and chloroform All polymerizations were conducted using APS initiator (0.3 g),
(CHCl3) were purchased from VWR and used as received. water (12 g), and aq. NaOH (8 g), maintaining a pH > 10 at 70 °C
Chloroform-d (CDCl3) was passed through anhydrous Na2CO3 to for 4 h. bDetermined by 1H NMR analysis, comparing the protons
remove any traces of internally formed hydrochloric acid, which may from the ring-opening of MDO with those from the methyl ester of
cause hydrolysis of polyester and unreacted MDO, ensuring a more MMA. cObserved only poly(MMA) and MDO hydrolyzed product.
d
straightforward analysis process. Repeat of EM4 to check the reproducibility. eUsed HEMA in place
Methods. Preparation of Poly(MDO-co-MMA) (EM1 and EM8) of HEA. fObserved low solid content (∼2 wt %) due to a significant
and Poly(MDO-co-MMA-co-HEA) (EM2) Nanoparticles. In a 50 mL amount of particle aggregation during the reaction.
single-neck round-bottom flask, a stirred solution of SDS (0.125 g; for
EM1 and EM2) and TBAB (0.2 g; for EM1 and EM2) or Triton X- prepared, added to the reaction mixture in one portion, and purged
100 (0.125 g; for EM8) in deionized (DI) water (12.0 g) was with argon gas for another 1 min. The resulting emulsion was
prepared (refer to Table 1). 5 M NaOH (5 M, 4.0 g) was then added monitored to maintain a pH > 10 throughout the reaction by adding
dropwise under argon gas for 2 min. The solution was purged with an additional 5 M NaOH dropwise. It was observed that the reaction
argon gas for 30 min at 70 °C with a stirring speed of 705 rpm. APS pH dropped rapidly, necessitating the periodic addition of 5 M NaOH
(0.3 g) was dissolved in 0.4 mL of DI water, and then argon gas was to maintain the desired pH. Approximately, a total of 8.0 g of 5 M
dropwise added to the reaction mixture and purging was continued NaOH, including the initial 4.0 g feed, was consumed throughout the
for an additional 5 min. A monomer mixture (4 g of MMA, 1 g of reaction. The reaction was stirred at 70 °C for 4 h. After this time,
MDO for EM1 and EM8 or 3.75 g of MMA, 1 g of MDO and 0.25 g sampling was done to estimate the reaction conversion using 1H
of HEA for EM2) was prepared and introduced to the reaction NMR analysis. Later, the reaction was quenched by rapid cooling and
mixture in one portion, followed by an argon gas purging for another purified with DI water through three consecutive centrifugations at
min. The resulting emulsion was monitored to maintain a pH > 10 7830 rpm for 45 min each, removing any residual monomers,
throughout the reaction by the periodic dropwise addition of 5 M hydrolyzed MDO, NaOH, and surfactant. In each purification step,
NaOH. Approximately 8.0 g of 5 M NaOH, including the initial 4.0 g the sedimented solid particles were separated from supernatants by
feed, was consumed during the reaction. The reaction proceeded at 70 decantation. The supernatant contained residual monomers, hydro-
°C for 4 h, after which sampling was performed to estimate the lyzed MDO, NaOH, and surfactant. Following purification, the
reaction conversion by 1H NMR analysis. The reaction was quenched resulting white solid nanoparticles were dried in a vacuum oven at 40
by rapid cooling and purified with DI water through three consecutive °C for 16 h and were analyzed for MDO incorporation with 1H NMR
centrifugations at 7830 rpm for 45 min each, removing any residual analysis. The solid content was 5 wt % for EM3−EM7 and 2% for
monomers, hydrolyzed MDO, NaOH, and surfactant. In each EM9.
purification step, the sedimented solid particles were separated from Preparation of Poly(MDO-co-MMA-co-HEA) Nanoparticles
supernatants by decantation. The supernatant contained residual (EM10−EM15) with Various Nonionic Surfactants. The following
monomers, hydrolyzed MDO, NaOH, and surfactant. Following surfactants were utilized in the synthesis of EM10−EM15 nano-
purification, the resulting white solid nanoparticles, obtained after particles (refer to Table 4): Ryoto sugar ester for EM10, Span 80 for
drying in a vacuum oven at 40 °C for 16 h, were analyzed for MDO EM11, Triton X-405 for EM12, Span 85 for EM13, Tween 80 for
incorporation with 1H NMR analysis. EM14, and Tween 20 for EM15. The experimental procedure
Preparation of Poly(MDO-co-MMA-co-HEA) Nanoparticles employed for EM3−EM7 and EM9 was consistent, and the same was
(EM3−EM7 and EM9). To a stirred solution of Triton X-100 applied to these formulations as well. For EM12 and EM13, no
(0.125 g; for EM3−EM6) or Poloxamer-188 (0.125 g; for EM7 only) detectable MDO incorporation was observed by the 1H NMR
or PEG-b-PCL (0.125 g; for EM9 only) and TBAB (0.2 g; for EM3 analysis. The solid content was 5 wt % for EM10 and EM11 and 4%
only) in DI water (12.0 g) in a 50 mL single-neck round-bottom flask for EM14 and EM15.
was added 5 M NaOH (4.0 g) dropwise under argon gas for 2 min Representative Accelerated Hydrolytic Degradation of Poly-
(refer to Table 1). The resulting solution was purged with an argon (MDO-co-MMA-co-HEA) (EM5). To a stirred solution of poly(MDO-
gas for 30 min at 70 °C with 705 rpm. APS (0.3 g) dissolved in 0.4 co-MMA-co-HEA) EM5 (200 mg) in THF (12 mL) was added a
mL of DI water was then added dropwise to the reaction mixture and solution of KOH (400 mg) in MeOH (4 mL) at room temperature.
purged with argon gas for another 5 min. Then, the monomer mixture The reaction mixture was stirred at room temperature for 72 h. The
(3.75 g of MMA, 1 g of MDO and 0.25 g of HEA or HEMA) was reaction mixture was acidified to pH 4−5 with 6 N HCl at 0 °C.

C https://doi.org/10.1021/acs.macromol.3c02474
Macromolecules XXXX, XXX, XXX−XXX
Macromolecules pubs.acs.org/Macromolecules Article

Figure 1. 1H NMR (CDCl3) analysis for (a) EM3 (after purification) and (b) EM2 (after purification).

Solvents were evaporated under reduced pressure, and the obtained fraction of 75 mol % of MMA and 20 mol % of MDO. The
solids were dried under a high vacuum for 2 h to remove solvent remaining MDO from the initial comonomer mixture under-
traces. CHCl3 (10 mL) was added to the mixture of solids and stirred went hydrolysis. 1H NMR analysis is shown in Figure 1 for
at room temperature for 2 h. Undissolved salts were removed by
syringe filtration, and CHCl3 was removed under reduced pressure to
comparison showing EM2 together with EM3 to distinguish
yield a pale yellow viscous degraded polymer mixture (200 mg, the incorporated poly(MDO) protons in the polymer
crude). backbone. The important factors for success are to obtain a
stable latex, minimize MDO hydrolysis, and incorporate MDO
■ RESULTS AND DISCUSSION
The copolymerization of (meth)acrylates and cyclic ketene
in the copolymer. Regarding the stability, we observed that
neutral surfactants were improving stability; regarding MDO
hydrolysis, the use of high pH was important as it is well-
acetals offers a fascinating route for free-radical ring-opening
known that the hydrolysis of MDO is minimized at high pH
copolymerization. This process creates interesting degradable
compared to neutral and acidic pH conditions.41,42,44 Of
polymers by incorporating degradable units into the acrylate
backbone in a single step. course, currently in this study, the batch process is used, which
We attempted emulsion copolymerizations of MMA and is not optimal for the reactivity ratios at hand, so further
MDO with varying stabilizers at a pH > 10, and the obtained optimization is certainly possible via a semibatch process.29 We
results are summarized in Table 1. In emulsion polymerization attempted the worst-case scenario with using a batch reaction;
in water, with MMA/MDO (molar fraction of 80:20 mol %) or furthermore, semibatch would necessitate the presence of
MMA/MDO/HEA (molar fraction of 75:20:5 mol %), APS as larger amounts of MDO at the beginning of the reaction,
a radical initiator, TBAB as a stabilizer, and SDS as an anionic potentially increasing hydrolysis levels, warranting further
surfactant resulted in the exclusive production of poly(MMA) investigation. Based on these current observations, it is
with over 99% conversion of MMA monomer (EM1 and hypothesized that the neutral surfactant Triton X-100 in
EM2) (refer to Figure 1b for representative 1H NMR analysis combination with the hydrophilic nature of the surface
of EM2). The MDO (4 g for EM1 and 3.75 g for EM2) was hydroxyl groups from HEA (or HEMA) at high pH plays a
found to undergo complete hydrolysis during the reaction. crucial role in stabilizing the latex particles. At this high pH and
When replacing SDS with Triton X-100, a neutral surfactant, while neutralizing with aqueous NaOH, the ionic strength is
while retaining other conditions, poly(MDO-co-MMA-co- high and ionic stabilization is challenging.41 Another factor
HEA) was obtained with 97 mol % of MMA and 3 mol % could be the transport of monomer from droplets to latex
of MDO incorporation in the polymer backbone (EM3). This particles. Although molecular monomer diffusion is generally
was achieved from an initial comonomer mixture with a molar accepted, there is evidence that monomer transport could also
D https://doi.org/10.1021/acs.macromol.3c02474
Macromolecules XXXX, XXX, XXX−XXX
Macromolecules pubs.acs.org/Macromolecules Article

Table 2. Characterization of Poly(MDO-co-MMA-co-HEA) Nanoparticles (EM3−EM7) and Their Corresponding Oligomeric


Degradation Products by Gel Permeation Chromatography (GPC)
polymer
compositiona polymer before degradation polymer after degradation
conversion of MDO (mol %) in ĐM Mntheo Mn Mw ĐM
experiment MMA/MDO (%)b polymera MMA MDO Mn (Da) Mw (Da) (Mw/Mn) (Da) (Da) (Da) (Mw/Mn)
EM3 99/99 3 0.97 0.03 49,500 283,500 5.72 3300 4600 16,900 3.65
EM4 99/99 9 0.91 0.09 48,500 233,400 4.80 1100 3000 5100 1.67
EM5 99/99 7 0.93 0.07 69,600 312,700 4.48 1400 1000 2700 2.57
EM6 99/99 7 0.93 0.07 46,700 325,900 6.97 1400 2400 3700 1.55
EM7 99/99 6 0.94 0.06 109,900 288,900 2.63 1700 2500 6000 2.34
a
Determined by 1H NMR analysis of the purified sample. bDetermined by 1H NMR analysis of the crude polymeric latex.

occur through direct collisions between monomer droplets and monomers, hydrolyzed MDO, NaOH, and surfactant, ensuring
latex particles. Most evidence is found in miniemulsion the purity of the samples before NMR analysis (refer to the
polymerizations,47,48 but there are several papers that also Experimental Section for the detailed procedure). From 1H
claim that collisions could be an important monomer transport NMR of poly(MDO-co-MMA-co-HEA) EM3 or EM5, shown
mechanism in regular emulsion polymerization (refer to Figure in Figures 1a and 3b, respectively, peak e at 3.97 ppm and
5).49,50 It was observed that the collision mechanism is favored peaks d between 2.7 and 2.2 ppm are corresponding to
when using neutral surfactants (over charged surfactants).50 methylene protons (−CH2−C(O)O−CH2−) and −C(O)O−
Although the collision mechanism is useful for completely CH2−, respectively, of the degradable units of polyester. The
water-insoluble species,48 in our case, where avoiding direct methylene protons adjacent to carbonyl (−CH2−C(O)O−)
contact between the monomer and water is desirable, the are often observed as a single peak in the homopolymerization
collision mechanism facilitated by neutral surfactants is likely of MDO22 but it was seen here as two different peaks at 2.7
to be beneficial. and 2.2 ppm. The splitting is due to differences in
During the course of optimizing reaction conditions, we configuration, and the shift supports a copolymer with a
observed a significant improvement in MDO incorporation (9 random distribution.52,53 The higher shifted peak at 2.7 ppm
mol %) when TBAB was not included (EM4). To confirm the corresponds to a proton d (Figure 1a) where MDO directly
reproducibility of EM4 results, we have repeated the reaction connected to acrylate and the lower shifted peak at 2.2 ppm
and found not much difference in polymer composition corresponds to the same proton d where MDO monomers
(EM5) (for a representative picture of the emulsion polymer connected within it. Peaks e and/or d are used to estimate the
EM5, refer to Figure S26 in the Supporting Information, SI). amount of heterodyads, i.e., the mol % of MDO monomer
Subsequent attempts to modify the reaction conditions by coupled to acrylate monomer with the comparison of proton c
switching from HEA to HEMA or changing the surfactant from (methyl ester of MMA).29 ,52 Peaks 178.2, 177.9, and 177.1
Triton X-100 to Poloxamer-188 or PEG-b-PCL (EM6, EM7, ppm from 13C NMR (Figure 3d) correspond to the carboxylate
and EM9, respectively) did not result in significant alterations of MMA, MDO, and HEA, respectively, and which are also
in polymer compositions, except for a lower solid content used to estimate the amount of heterodyads. For 1H NMR
observed in the case of PEG-b-PCL. This lower solid content analysis of EM6 and EM7, refer to Figures S1 and S3 in the SI.
was likely caused by extensive particle aggregation during the It is important to highlight the purity comparison between
reaction. To further investigate the impact of HEA, we EM3 and EM5 through 1H NMR characterization (Figures 1a
conducted a control experiment (EM8), which revealed that in and 3b). In EM3, two additional peaks appear around peaks e
the absence of HEA, less than 1% of MDO incorporation was + k, whereas they are absent in EM5. It should be noted that
detected along with the formation of some particle aggregates. EM3 was synthesized during the optimization of reaction
For EM3−EM9 lattices, in addition to MDO incorporation conditions, involving the use of the stabilizer TBAB.
(3−9 mol % out of the initial 20 mol % MDO monomer in the Consequently, EM3 may not be as pure as EM5, despite
recipe), the remaining MDO underwent hydrolysis. applying the same purification procedure to eliminate residual
At this juncture, it is important to highlight the significant monomers, surfactants, stabilizers, and NaOH in both. We
role of the hydroxyl groups from HEA in maintaining colloidal hypothesize that these two peaks around e + k could be
stability without imparting charge to the particles. We posit attributed to the presence of small (polymer) molecules
that maintaining a low particle charge, especially at this resulting from the polymerization. The peak area under e + k
elevated ionic strength and pH, enhances the transport of and d in EM3 is consistent with that of EM5 (refer to the peak
monomers through collisions and reduces the chances of integration in Figures 1a and 3b).
MDO hydrolysis. There are alternative stabilizing hydrophilic An approximate error value for MDO incorporation in the
acid monomers, such as acrylic acid (AA) or methacrylic acid polymer backbone (expressed as a percentage of MDO
(MAA), which could potentially sustain colloidal stability.51 incorporation values) was calculated from the signal-to-noise
However, owing to the high reactivity of acid groups with ratio (SNR) of the methylene protons (−CH2−C(O)O−
MDO and the potential hindrance to MDO incorporation in CH2−) of MDO (refer to Tables 1, 4, 6, and 7 in the
the polymer backbone, we have not employed them in the manuscript and Table S1 in the SI).29
current study.33,34 To demonstrate the degradability of poly(MDO-co-MMA-
After polymerization, all polymeric dispersions were purified co-HEA), EM3−EM7 copolymers were subjected to accel-
with DI water three times by centrifugation at 7830 rpm for 45 erated hydrolytic degradation by treating with an excess of
min each at room temperature. This process removed residual KOH in a 1:3 mixture of MeOH and THF solvent
E https://doi.org/10.1021/acs.macromol.3c02474
Macromolecules XXXX, XXX, XXX−XXX
Macromolecules pubs.acs.org/Macromolecules Article

combination at room temperature for 72 h, producing short- molecular weight distribution was observed before (Mn =
chain methacrylate-based oligomers (Table 2 and Figures 2 28 100 Da) and after (Mn = 27 500 Da) being subjected to
hydrolytic degradation conditions (refer to Figure S17 in the
SI). This finding was further corroborated by 1H NMR analysis
of poly(MMA) before and after hydrolytic degradation,
revealing that the side chains of methyl esters of poly(MMA)
underwent trans-esterification instead of hydrolysis (refer to
Figure S16 in the SI). This observation, indicating no change
in the molecular weight distribution, was consistent with the
absence of hydrolytic degradation in the backbone and side
chains of methyl esters in poly(MMA). Similarly, trans-
esterification was also evident in the oligomeric degradation
products from EM5 during the hydrolytic degradation process
(refer to peak c in Figure 3a).
Aligned with the anticipated outcomes, our observations
from the Supporting Data, including 1H and 13C NMR (Figure
3) and GPC analyses (Table 2 and Figure 2), reveal
noteworthy changes. Specifically, there is a shift in peak e
from 3.9 (Figure 3b) to 4.8 (Figure 3a) in 1H NMR (zoomed
in 6−4 ppm). Additionally, a subtle shift is noted in the
carbonyl peaks of the carboxylate groups, shifting from 178.2,
177.9, and 177.1 ppm (Figure 3d) to 178.1, 177.8, and 176.9
Figure 2. GPC analysis of poly(MDO-co-MMA-co-HEA) nano-
particles (EM3−EM7) before and after accelerated hydrolytic ppm (Figure 3c) for MMA, MDO, and HEA, respectively.
degradation. Moreover, the shift in the peak corresponding to the
methylene protons (−C(O)O−CH2−) of MDO from 64.6
(Figure 3d) to 70.3 ppm (Figure 3c) in 13C NMR provides a
and 3). The copolymers were degraded by the cleavage of ester clear indication of the degradation of its polymeric backbone.
bonds in the polymeric backbone, as revealed by a decrease in However, peaks d in the oligomers (Figure 3a) exhibited slight
molar mass during hydrolysis. Gel permeation chromatography shifts, attributable to the presence of the newly formed
(GPC) analysis (Table 2 and Figure 2) was performed before neighboring group carboxylic acid. The GPC results provide
and after hydrolytic degradation of the copolymers (EM3− further corroboration of hydrolytic degradation (Figure 2).
EM7). Copolymers with higher molar mass (Mn = 46 700 Da− In the rROP of CKA chemistry, the potential side reaction
109 900 Da) were degraded into their corresponding involving ring-retaining may impact the degradability of the
oligomers with lower molar mass (Mn = 1000 Da−4600 Da) resulting copolymer as it does not insert an ester bond in the
based on the percentage of degradable ester units in the polymer backbone.22 The formation of a ring-retained
polymer backbone. The theoretical molar masses of the structure in poly(MDO-co-MMA-co-HEA) occurs through
oligomers (Mn = 1100 Da−3300 Da) were calculated based on vinyl addition of the propagating radical at the MDO. This
the mol % of ester units present in the polymer backbone from side product is identifiable by the characteristic peak from the
1
H NMR (3−9 mol %). The observed molar mass of oligomers quaternary carbon of the acetal ring, typically observed around
from copolymer EM5 closely matched the theoretical molar 100 ppm in the 13C NMR spectra of the copolymers.
mass (Mn = 1000 Da and Mn = 1400 Da, respectively), Remarkably, under our current reaction conditions, this
indicating a more or less uniform distribution of cleavable distinctive peak was conspicuously absent (see Figure 3d).
bonds within the copolymer chains and across different chains. This absence suggests a potential advantage for the
However, for copolymers EM3, EM4, EM6, and EM7, the applicability of our approach for industrial conditions, where
observed molar mass was significantly higher than the the mitigation of this side reaction contributes to the
theoretical molar mass (Table 2). This inconsistency in the robustness of the polymerization process.
observed molar mass of oligomers, as observed in the GPC Next, we turned our attention to characterizing nano-
analysis of copolymers, could be due to a composition drift particles of poly(MDO-co-MMA-co-HEA) (EM3, EM4, and
during the batch process. Additionally, the strong interactions EM7) using various instruments, including dynamic light
of the hydroxyl groups of HEA with the GPC column could scattering (DLS), scanning electron microscopy (SEM), and
play a role.42,54 Furthermore, the disparity between EM4 and transmission electron microscope (TEM). The particle sizes,
its repeating experiment EM5 regarding their theoretical and represented by the Z-average and number mean, along with the
observed molar mass of oligomers may suggest a batch-to- polydispersity index (PDI), are presented in Table 3,
batch variation, a common occurrence in emulsion polymer- complemented by SEM and TEM images (Figure 4). The
ization, especially under manual control at the lab scale. This dispersions of EM3, EM4, and EM7 exhibited particle sizes of
variability could arise from minor fluctuations in the 237.9, 194.9, and 266.6 nm by Z-average and 218, 146.7, and
temperature, stirring, and pH control. Polydispersity (Đ) 183.5 nm by number mean, respectively, showcasing a narrow
values were recorded as 2.63−6.97 for purified polymers and size distribution. SEM and TEM analyses confirmed the
1.55−3.65 for oligomers. A control experiment of non- spherical shape and smooth surface of all three latex particles in
degradable poly(MMA) with a known molecular weight the dispersions. Moreover, the stability of these nanoparticles
distribution (Mn = 28 100 Da) was also subjected to the was noteworthy, with EM3, EM4, and EM7 displaying good
same hydrolytic degradation conditions to confirm its long-term colloidal stability for over 6 months, showing no
nondegradability nature. As anticipated, no major change in significant change in particle size distribution and no
F https://doi.org/10.1021/acs.macromol.3c02474
Macromolecules XXXX, XXX, XXX−XXX
Macromolecules pubs.acs.org/Macromolecules Article

Figure 3. (b, d) 1H and 13C NMR (CDCl3) analyses for EM5. (a, c) 1H and 13C NMR (CDCl3) analyses for oligomeric degradation products
resulted from the hydrolytic degradation of EM5.

G https://doi.org/10.1021/acs.macromol.3c02474
Macromolecules XXXX, XXX, XXX−XXX
Macromolecules pubs.acs.org/Macromolecules Article

Table 3. Dynamic Light Scattering (DLS) Analysis of Table 5). Even when we attempted one of the reactions with a
Poly(MDO-co-MDO-co-HEA) Nanoparticles (EM3, EM4, different initiator, V-50, the outcome was still unsuccessful
and EM7) (EM19, Table 5). In all of these reactions (EM16−EM20),
despite the slower rate of hydrolysis in the alkaline medium
sample ID Z-average (nm) number mean (nm) PDI
with a pH > 10, MDO was observed to undergo complete
EM3 237.9 ± 1.6 218.0 ± 2.1 0.05 ± 0.1 hydrolysis during the reaction, resulting in no incorporation of
EM4 194.9 ± 2.0 146.7 ± 5.0 0.17 ± 0.1 MDO in the polymer backbone.41,42,44
EM7 266.6 ± 2.2 183.5 ± 15.5 0.27 ± 0.1 Subsequently, our focus shifted toward understanding the
mechanism of MDO incorporation in emulsion polymerization
sedimentation of particles (for representative results, refer to in water (Figure 5). Successful incorporation of MDO while
Figure S26 and Table S2 in the SI). In contrast, the polymeric utilizing neutral micelles could be the result of a series of
dispersion EM9 showed instability within days (refer to Figure intricate mechanistic aspects governing the processes of
S27 in the SI). We also investigated the chemical stability of monomer and oligomer transport to the micelles as well as
poly(MDO-co-MMA-co-HEA) EM6 nanoparticles after 5 copolymer growth within the micelles. Based on aqueous
months of storage at ambient conditions, observing no change solubility, stability, and concentration considerations, there is
in chemical composition (refer to Figure S2 in the SI) also a possibility of homogeneous nucleation of MMA or HEA
compared to the freshly prepared sample (refer to Figure S1 in in parallel with micellar nucleation in the system. It is likely
the SI), as confirmed by 1H NMR analysis. that during homogeneous nucleation, there will be substantial
To investigate the influence of other neutral surfactants, we hydrolysis of MDO. In conventional emulsion polymerization
conducted further experiments determining the mol % of (Figure 5a), where monomers are present in the aqueous phase
MDO incorporation in the main-chain methacrylate polymer in surfactant-stabilized micelles and droplets, the majority of
(Table 4). All polymerization conditions remained consistent MMA and MDO is present in the monomer droplets. It then
with the conditions of EM4 (MMA/MDO/HEA = 75:20:5), gets transported to the micelles, enabling propagation in the
with the only variation being the surfactant used. Reactions micelle. We anticipate diffusion of MDO through the aqueous
employing Ryoto sugar ester, Span 80, Tween 80, and Tween phase to lead to the main hydrolytic loss of MDO due to its
20 provided copolymer nanoparticles with 3−4 mol % of hydrolytic instability. This is because MDO, being less stable,
MDO incorporation (EM10, EM11, EM14, and EM15). In has to diffuse through the aqueous phase and overcome a high
contrast, the use of Triton X-405 and Span 85 led to the charge density barrier at the micellar interphase to participate
formation of only a poly(MMA) homopolymer with complete in a polymerization reaction with nonfavorable reactivity ratios.
hydrolysis of MDO (EM12 and EM13). Notably, copolymers The high charge density at the micellar water interphase might
EM10, EM11, EM14, and EM15 exhibited nanoparticle also enhance the stabilization of the polar transition state of
instability within a day. For 1H NMR analysis of EM10, hydrolysis, leading to complete hydrolysis. In our present
EM11, EM14, and EM15, refer to Figures S4−S7 in the SI. reaction conditions (Figure 5b), the neutral micelle interface
Furthermore, we explored the feasibility of employing other does not have a highly polar double layer. Furthermore, as
ionic surfactants (both anionic and cationic) as alternatives to discussed before, the neutral micelles might allow MDO to be
Triton X-100 under the EM4 reaction conditions. However, to transported into the micelle through direct collisions with the
our surprise, none of these experiments resulted in the monomer droplets without undergoing hydrolysis. The chains
formation of poly(MDO-co-MMA-co-HEA); instead, only containing HEA stabilize the micelle, further maintaining the
poly(MMA) homopolymer formation was observed, as neutral interface. In summary, to get MDO incorporation, it
confirmed by 1H NMR analysis (EM16−EM18 and EM20, must overcome the hydrolysis challenge by entering micelles

Figure 4. Scanning electron microscopy (SEM) and transmission electron microscope (TEM) analyses of poly(MDO-co-MDO-co-HEA)
nanoparticles. SEM images for (a) EM3, (b) EM4, and (c) EM7. TEM images for (d) EM4 and (e) EM7. Scale bars: (a−c) 1 μm and (d−e) 2 μm.

H https://doi.org/10.1021/acs.macromol.3c02474
Macromolecules XXXX, XXX, XXX−XXX
Macromolecules pubs.acs.org/Macromolecules Article

Table 4. Screening of Various Neutral Surfactants to Investigate the MDO Incorporation in Poly(MDO-co-MDO-co-HEA)
Nanoparticles (EM10−EM15)
final polymer compositionb
experiment surfactant conversion of MMA/MDO (%)a MDO incorporation (mol %)b MMA MDO
EM10 Ryoto sugar ester 99/99 4± 0.15 0.96 0.04
EM11 Span 80 99/99 3± 0.03 0.97 0.03
EM12 Triton X-405 99/99 n.d. 1.0 0
EM13 Span 85 99/99 n.d. 1.0 0
EM14 Tween 80 99/99 3± 0.03 0.97 0.03
EM15 Tween 20 99/99 4± 0.01 0.96 0.04
a
Determined by 1H NMR analysis of the crude polymeric emulsion. bDetermined by 1H NMR analysis of the purified sample. n.d. = not detected.

Table 5. Screening of Anionic and Cationic Surfactants to Investigate MDO Incorporation in Poly(MMA-co-MDO-co-HEA)
Nanoparticles (EM16−EM20)
final polymer composition
a
experiment initiator surfactant conversion of MMA/MDO (%) MDO incorporation (mol %) MMA MDO
EM16 APS DSS 99/99 n.d. 100 0
EM17 APS SDBS 99/99 n.d. 100 0
EM18 APS CTAB 99/99 n.d. 100 0
EM19 V-50 CTAB 99/99 n.d. 100 0
EM20 APS hyamine 99/99 n.d. 100 0
a
Determined by 1H NMR analysis of the crude polymeric emulsion. DSS, dihexyl sulfosuccinate sodium; SDBS, sodium dodecylbenzene sulfonate;
CTAB, cetyltrimethylammonium bromide; hyamine, benzethonium chloride.

Figure 5. Proposed mechanism of MDO transport via (a) diffusion or (b) collision while utilizing charged or neutral surfactants in emulsion
polymerization.

Table 6. Screening of Emulsion Polymerization Conditions Using Various Acrylate Monomersa


Z-average (d MDO incorporation in the final polymer latex
sample initial monomer feeding composition (mol %) nm) PDI (mol %)b
3BA1MDO 71%BA24%MDO 5%HEA 0.2%EGDMA 300.9 0.25 9 ± -.--c
3EA1MDO 71%EA 24%MDO 5%HEA 0.2%EGDMA 347.5 0.08 9 ± -.--c
3MA1MDO 71%MA 24%MDO 5%HEA 0.2%EGDMA 713.1 0.13 5 ± 0.04
1MMA3BA20MDO 19%MMA 56%BA 20%MDO 5%HEA 0.2%AMA 275.6 0.02 10 ± -.--c
1MMA3BA10MDO 21%MMA 64%BA 10%MDO 5%HEA 0.2%AMA 227.5 0.06 3 ± -.--c
3BA1MDO5DMAEMA 71%BA24%MDO 5%DMAEMA 0.2%EGDMA 273.9 0.09 “42 ± -.--c”
3BMA1MDO 71%BMA 24%MDO 5%HEA 0.2%EGDMA 244.3 0.04 10 ± -.--c
1BA2BMA1MDO 24%BA 47%BMA 24%MDO 5%HEA 0.2% 228.2 0.07 “5 ± -.--c”
EGDMA
3EHA1MDO 71%EHA 24%MDO 5%HEA 0.2%EGDMA aggregate n.d.

a
All polymerizations were conducted using APS initiator (0.3 g), Triton X-100 (0.125 g), water (12 g), and aq. NaOH (pH > 10, 8 g) at 70 °C for
4 h. bDetermined by 1H NMR analysis of the purified sample. BA, n-butyl acrylate; EA, ethyl acrylate; MA, methyl acrylate; DMAEMA, 2-
(dimethylamino)ethyl methacrylate; BMA, n-butyl methacrylate; EHA, 2-ethylhexyl acrylate; AMA, allyl methyl acrylate; and EGDMA, ethyl
methyl acrylate. cError is smaller than 0.01.

I https://doi.org/10.1021/acs.macromol.3c02474
Macromolecules XXXX, XXX, XXX−XXX
Macromolecules pubs.acs.org/Macromolecules Article

Table 7. GPC Analysis of Copolymer Molar Mass Distribution before and after Hydrolytic Degradation
polymer before degradation polymer after degradation
sample MDO incorporation (mol %) Mn (Da) Mw (Da) Đ (Mw/Mn) Mntheo (Da) Mn (Da) Mw (Da) Đ (Mw/Mn)
3BA1MDO 9 ± 0.00 32,900 105,200 3.12 1400 700 1000 1.36
3EA1MDO 9 ± 0.00 21,100 83,100 3.92 1100 800 950 1.18
3MA1MDO 5 ± 0.04 45,800 108,500 2.36 1700 700 970 1.39
1MMA3BA20MDO 10 ± 0.00 8,300 27,700 3.31 1200 400 580 1.41
1MMA3BA10MDO 3 ± 0.00 14,500 42,000 2.90 4000 600 700 1.11
3BA1MDO5DMAEMA 42 ± 0.00a 26,800 90,200 3.36 300 310 1.10
3BMA1MDO 10 ± 0.00 17,700 73,200 4.13 1400 3400 9000 2.64
1BA2BMA1MDO 5 ± 0.00a 11,380 32,000 2.81 2600 6500 2.47
a
Determination of MDO incorporation was conducted through Mn of oligomeric degradation products by GPC analysis, as it proved challenging
by 1H NMR due to overlapping peaks.

rapidly, where, in the hydrophobic environment, it takes part in 3BMA1MDO, with the observed MDO incorporation of 9,
polymerization rather than hydrolysis.55 The neutral micelle 10, and 10 mol %, respectively. However, when the monomer
also promotes an alternate mechanism of transport of mixtures become excessively hydrophobic, early aggregation
monomers through collisions47−49 (Figure 5b). occurs, accompanied by hardly any incorporation of MDO in
It is to be noted that the hydrophilic−lipophilic balance the polymer backbone. This phenomenon is exemplified by
(HLB) or the orientation of hydrophilic and lipophilic groups 3EHA1MDO. In conclusion, compositions containing linear
in a surfactant plays a crucial role in stabilizing and growing and more hydrophobic monomers, such as BA, EA, and BMA,
radical polymer particles in the continuous phase of an tend to provide higher percentages of MDO incorporation into
emulsion polymerization system. The HLB value, measured on their corresponding polymer backbone. Upon examining
an arbitrary scale from 0 to 20, indicates the size and strength samples 1MMA3BA20MDO and 1MMA3BA10MDO, it was
of the polar portion relative to the nonpolar portion of the found that the recipes with a lower level of MDO also showed
nonionic surfactant molecule, where 0 represents complete a lower amount of MDO incorporation in the final polymer
lipophilicity and 20 denotes complete hydrophilicity. For oil- latex, as revealed by 1H NMR analysis. This suggests that less
in-water (o/w) emulsions, the surfactant HLB is ideally in the MDO is copolymerized, and consequently, a significant
range of 8−16.56 Within this range, the hydrophilic part of the portion of MDO underwent hydrolysis. For samples
surfactant aligns toward the water, while the lipophilic part 3BA1MDO5DMAEMA and 1BA2BMA1MDO, the calculation
aligns toward the oil, effectively stabilizing the o/w emulsion. of MDO was challenging by 1H NMR as the corresponding
Surfactants with HLB values outside this range (8−16) may peaks were found to be overlapping. Hence, the level of MDO
not be able to form stable o/w emulsions. Upon scrutinizing incorporation was roughly estimated by the molecular weight
Table 4 and comparing it with Triton X-100, it becomes of the oligomeric degradation products (Table 7). For 1H
apparent that the HLB value of Triton X-100 (∼13.5) is better NMR analysis of samples from Table 6, refer to Figures S8−
suited for o/w emulsions, contributing to improved MDO S15 in the SI.
incorporation and long-term stable emulsion. In contrast, Span Next, we treated the copolymers from Table 6 with an excess
80 (∼4.3), Triton X-405 (∼16.7), Span 85 (∼1.8), Tween 80 of KOH in a 1:3 mixture of MeOH and THF solvent
(∼15), and Tween 20 (∼16.7) either fall outside or closely combination at room temperature for 72 h, resulting in short-
approach the necessary HLB range. This observation might be chain acrylate-based oligomers. GPC analysis (Figure 6 and
an explanation for the limited or negligible incorporation of Table 7) was performed before and after the hydrolytic
MDO into the polymer backbone for the latter surfactants. degradation of copolymers. A decrease in molar mass (Mn)
We have conducted further reactions using other acrylate from 8300−45,800 to 300−3400 Da was observed during
monomers (BA, EA, MA, MMA/BA, and BMA/BA mixtures), hydrolysis as shown in the GPC analysis (Figure 6 and Table
and the results are summarized in Table 6. All polymerization 7). It is to be noted that the observed molar masses of the
conditions remained unchanged from the conditions of EM4 oligomeric degradation products did not closely match their
except for switching monomer types and compositions and theoretical molar masses, possibly due to some interaction of
performing with cross-linkers (AMA and EGDMA). As the resulting carboxyl groups with the GPC column.
discussed in the Introduction section, poly(VAc) containing Nevertheless, these findings suggest that even in batch
MDO-derived esters in emulsions has already been reactions with unfavorable reactivity ratios, some degree of
reported,40,42 and we did not explore MDO incorporation MDO incorporation remains achievable in emulsion copoly-
under our conditions since the high basicity of the medium merizations, provided specific conditions are met. These
might be complicating the system due to hydrolysis of the conditions include the presence of neutral surfactants, a
acetate side groups. hydroxy-functional comonomer, and a high pH environment.57
When different acrylate monomers are compared, it For GPC analysis of individual samples from Figure 6, refer to
becomes evident that more hydrophobic acrylate monomers Figures S18−S25 in the SI.
tend to achieve higher MDO% incorporation into their
polymer backbone. This trend is noticeable in the comparison
between 3BA1MDO and 3MA1MDO, where the former
■ CONCLUSIONS
We report the first batch emulsion copolymerization of MMA
exhibits a 9 mol % MDO incorporation, while the latter shows and MDO to obtain degradable latex at high pH and utilize a
a 5 mol % MDO incorporation. Similar trends were obtained neutral surfactant and 5 mol % of hydroxy-functional
in the cases of 3EA1MDO, 1MMA3BA20MDO, and costabilizing monomers. NMR spectral analysis and chemical
J https://doi.org/10.1021/acs.macromol.3c02474
Macromolecules XXXX, XXX, XXX−XXX
Macromolecules pubs.acs.org/Macromolecules Article

monomer transport by collisions could be favored under these


conditions, whereby we reduce the chance of hydrolysis. This
is all adding to the present know-how in the field of emulsion
copolymerization of CKA and specifically MDO. To the best
of our knowledge, the monomer transport by collisions has not
been invoked before for hydrophilic monomers. The breadth
of parameters considered in this paper can be expanded to the
rROP of CKAs to include monomers with highly unfavorable
reactivity ratios. We are currently exploring industrial-scale
studies on this sustainable latex polymerization process and
developing biodegradable substitutes for free-radical emulsion
polymers.


*
ASSOCIATED CONTENT
sı Supporting Information
The Supporting Information is available free of charge at
https://pubs.acs.org/doi/10.1021/acs.macromol.3c02474.
Experimental details; characterization data; 1H NMR
(CDCl3) of poly(MDO-co-MMA-co-HEA); and GPC of
poly(MMA) (a) before and (b) after subjected to
hydrolytic degradation, and representative picture for
the purified poly(MDO-co-MMA-co-HEA) (PDF)

■ AUTHOR INFORMATION
Corresponding Author
Praveen Thoniyot − Institute of Sustainability for Chemicals,
Energy and Environment (ISCE2), Agency for Science,
Technology and Research (A*STAR), Singapore 627833,
Republic of Singapore; Department of Chemical Engineering
and Chemistry, Eindhoven University of Technology, 5600
MB Eindhoven, The Netherlands; orcid.org/0000-0001-
5062-1831; Email: p.thoniyot@tue.nl
Authors
Srinivasa Reddy Mothe − Institute of Sustainability for
Chemicals, Energy and Environment (ISCE2), Agency for
Science, Technology and Research (A*STAR), Singapore
Figure 6. GPC analysis of copolymers derived from various vinyl 627833, Republic of Singapore; orcid.org/0000-0001-
monomers and their corresponding oligomeric degradation products. 6415-3311
Wenguang Zhao − Institute of Sustainability for Chemicals,
Energy and Environment (ISCE2), Agency for Science,
degradation experiments indicate up to 9 mol % of MDO Technology and Research (A*STAR), Singapore 627833,
incorporation in the emulsion polymer backbone. Chemical Republic of Singapore
degradation studies, followed by GPC analysis, confirm the Alexander M. van Herk − Department of the Built
incorporation of degradable units via MDO, corroborating the Environment, Building Materials, Eindhoven University of
results of NMR analysis. Use of charged surfactants results in Technology, 5600 MB Eindhoven, The Netherlands
no incorporation of MDO, leading to its wastage via complete
hydrolysis. Our study indicates that with proper design of the Complete contact information is available at:
surfactant system, pH control, and stabilizing monomer https://pubs.acs.org/10.1021/acs.macromol.3c02474
system, emulsion copolymerization can be successful for
acrylate and methacrylate systems. Comparing these new Author Contributions
#
results with the MDO-VAc system by the Carter group, it is S.R.M. and W.Z. contributed equally to this work. S.R.M.:
clear that for MDO-VAc, the favorable reactivity ratios in experiments, methodology, formal analysis, and data curation.
combination with starved conditions led to success. In our Z.W.: experiments, methodology, formal analysis, and data
work, we combine MDO with other monomers with curation. A.M.v.H.: conceptualization, data interpretation,
unfavorable reactivity ratios, and we do not apply starved review, and editing. P.T.: conceptualization, resources,
conditions. This makes this work much more challenging and validation, writing�review and editing, supervision, and
also more relevant for industrial applications. The essence of funding acquisition.
our findings is that in order to minimize hydrolysis under batch Notes
conditions, we need to keep the charge density on the latex The authors declare the following competing financial
particles as low as possible by utilizing neutral surfactants and a interest(s): A*STAR has filed a patent application
neutral costabilizing monomer in combination with a high pH. (WO2023075696A3) with Praveen Thoniyot, Srinivasa
In the system of Carter, this is not needed. We propose that Reddy MOTHE, and Wenguang Zhao as inventors for
K https://doi.org/10.1021/acs.macromol.3c02474
Macromolecules XXXX, XXX, XXX−XXX
Macromolecules pubs.acs.org/Macromolecules Article

potential future commercialization of the materials produced (17) Bailey, W. J.; Kuruganti, V. K.; Angle, J. S.Biodegradable
by the method. Polymers Produced by Free-Radical Ring-Opening Polymerization In
ACS Symposium Series; ACS Publications, 1990; Vol. 433, pp 149−

■ ACKNOWLEDGMENTS
The authors gratefully acknowledge the Agency for Science,
160.
(18) Pesenti, T.; Nicolas, J. 100th anniversary of macromolecular
science viewpoint: Degradable polymers from radical ring-opening
Technology and Research (A*STAR), Science and Engineer- polymerization: Latest advances, new directions, and ongoing
ing Research Council (SERC), under Specialty Chemicals challenges. ACS Macro Lett. 2020, 9, 1812−1835.
(19) Tardy, A.; Nicolas, J.; Gigmes, D.; Lefay, C.; Guillaneuf, Y.
AME IAF-PP Programme Grant (Project refs: A1786a0025 Radical ring-opening polymerization: Scope, limitations, and
and A20G1a0046). The authors extend their gratitude to application to (bio) degradable materials. Chem. Rev. 2017, 117,
Lohitha Rao Chennamaneni for the monomer supply and to 1319−1406.
Jean-Baptiste Lena for their assistance in calculating the (20) Lena, J.-B.; Herk, A. M. V. Toward biodegradable chain-growth
approximate error value for MDO incorporation. polymers and polymer particles: re-evaluation of reactivity ratios in
copolymerization of vinyl monomers with cyclic ketene acetal using

■ REFERENCES
(1) López, A. B.; de la Cal, J. C.; Asua, J. M. Highly hydrophobic
nonlinear regression with proper error analysis. Ind. Eng. Chem. Res.
2019, 58, 20923−20931.
(21) Jackson, A. W. Reversible-deactivation radical polymerization of
coatings from waterborne latexes. Langmuir 2016, 32, 7459−7466. cyclic ketene acetals. Polym. Chem. 2020, 11, 3525−3545.
(2) Gower, M. D.; Shanks, R. A. The effect of varied monomer (22) Reddy Mothe, S.; Tan, J. S. J.; Chennamaneni, L. R.; Aidil, F.;
composition on adhesive performance and peeling master curves for Su, Y.; Kang, H. C.; Lim, F. C. H.; Thoniyot, P. A systematic
acrylic pressure-sensitive adhesives. J. Appl. Polym. Sci. 2004, 93, investigation of the ring size effects on the free radical ring-opening
2909−2917. polymerization (rROP) of cyclic ketene acetal (CKA) using both
(3) Haloi, D. J.; Singha, N. K. Synthesis of poly (2-ethylhexyl experimental and theoretical approach. J. Polym. Sci. 2020, 58, 1728−
acrylate)/clay nanocomposite by in situ living radical polymerization. 1738.
J. Polym. Sci., Part A: Polym. Chem. 2011, 49, 1564−1571. (23) Jackson, A. W.; Mothe, S. R.; Chennamaneni, L. R.; Herk, A.
(4) Duan, S. F.; Cai, S.; Xie, Y. M.; Bagby, T.; Ren, S. Q.; Forrest, M. V.; Thoniyot, P. Unraveling the History and Revisiting the Synthesis
L. Synthesis and characterization of a multiarm poly (acrylic acid) star of Degradable Polystyrene Analogues via Radical Ring-Opening
polymer for application in sustained delivery of cisplatin and a nitric Copolymerization with Cyclic Ketene Acetals. Materials 2020, 13,
oxide prodrug. J. Polym. Sci., Part A: Polym. Chem. 2012, 50, 2715− No. 2325.
2724. (24) Wickel, H.; Agarwal, S.; Greiner, A. Homopolymers and
(5) Connal, L. A.; Li, Q.; Quinn, J. F.; Tjipto, E.; Caruso, F.; Qiao, random copolymers of 5, 6-benzo-2-methylene-1, 3-dioxepane and
G. G. pH-responsive poly (acrylic acid) core cross-linked star methyl methacrylate: Structural characterization using 1D and 2D
polymers: morphology transitions in solution and multilayer thin NMR. Macromolecules 2003, 36, 2397−2403.
films. Macromolecules 2008, 41, 2620−2626. (25) Wais, U.; Chennamaneni, L. R.; Thoniyot, P.; Zhang, H.;
(6) Wallace, A. D.; Al-Hamzah, A.; East, C. P.; Doherty, W. O. S.; Jackson, A. W. Main-chain degradable star polymers comprised of
Fellows, C. M. Effect of poly (acrylic acid) end-group functionality on pH-responsive hyperbranched cores and thermoresponsive poly-
inhibition of calcium oxalate crystal growth. J. Appl. Polym. Sci. 2010, ethylene glycol-based coronas. Polym. Chem. 2018, 9, 4824−4839.
116, 1165−1171. (26) Jackson, A. W.; Chennamaneni, L. R.; Thoniyot, P. Main-chain
(7) Hoornweg, D.; Bhada-Tata, P.; Kennedy, C. Waste production degradable, pH-responsive and covalently cross-linked nanoparticles
must peak this century. Nature 2013, 502, 615−617. via a one-step RAFT-based radical ring-opening terpolymerization.
(8) Derraik, J. G. B. The pollution of the marine environment by Eur. Polym. J. 2020, 122, No. 109391.
plastic debris: a review. Mar. Pollut. Bull. 2002, 44, 842−852. (27) Huang, J. Y.; Gil, R.; Matyjaszewski, K. Synthesis and
(9) Barnes, D. K. A.; Galgani, F.; Thompson, R. C.; Barlaz, M. characterization of copolymers of 5, 6-benzo-2-methylene-1, 3-
Accumulation and fragmentation of plastic debris in global environ- dioxepane and n-butyl acrylate. Polymer 2005, 46, 11698−11706.
ments. Philos. Trans. R. Soc., B 2009, 364, 1985−1998. (28) Sun, L. F.; Zhou, R. X.; Liu, Z. L. Synthesis and enzymatic
(10) Jambeck, J. R.; Geyer, R.; Wilcox, C.; Siegler, T. R.; Perryman, degradation of 2-methylene-1, 3-dioxepane and methyl acrylate
M.; Andrady, A.; Narayan, R.; Law, K. L. Plastic waste inputs from copolymers. J. Polym. Sci., Part A: Polym. Chem. 2003, 41, 2898−2904.
land into the ocean. Science 2015, 347, 768−771. (29) Lena, J.-B.; Jackson, A. W.; Chennamaneni, L. R.; Wong, C. T.;
(11) Bailey, W. J.; Ni, Z.; Wu, S.-R. Free radical ring-opening Lim, F.; Andriani, Y.; Thoniyot, P.; Herk, A. M. V. Degradable Poly
polymerization of 4, 7-dimethyl-2-methylene-1, 3-dioxepane and 5, 6- (alkyl acrylates) with Uniform Insertion of Ester Bonds, Comparing
benzo-2-methylene-1, 3-dioxepane. Macromolecules 1982, 15, 711− Batch and Semibatch Copolymerizations. Macromolecules 2020, 53,
714. 3994−4011.
(12) Bailey, W. J.; Ni, Z.; Wu, S.-R. Synthesis of poly-ϵ-caprolactone (30) Tardy, A.; Honore, J. C.; Tran, J.; Siri, D.; Delplace, V.; Bataille,
via a free radical mechanism. Free radical ring-opening polymerization I.; Letourneur, D.; Perrier, J.; Nicoletti, C.; Maresca, M.; Lefay, C.;
of 2-methylene-1, 3-dioxepane. J. Polym. Sci., Polym. Chem. Ed. 1982, Gigmes, D.; Nicolas, J.; Guillaneuf, Y. Radical copolymerization of
20, 3021−3030. vinyl ethers and cyclic ketene acetals as a versatile platform to design
(13) Bailey, W. J.; Wu, S.-R.; Ni, Z. Free radical ring-opening functional polyesters. Angew. Chem., Int. Ed. 2017, 56, 16515−16520.
polymerization of 4-n-hexyl-and 4-n-decyl-2-methylene-1, 3-dioxo- (31) Hedir, G. G.; Arno, M. C.; Langlais, M.; Husband, J. T.;
lanes. J. Macromol. Sci., Part A 1982, 18, 973−986. O’Reilly, R. K.; Dove, A. P. Poly (oligo (ethylene glycol) vinyl
(14) Bailey, W. J.; Wu, S.-R.; Ni, Z. Synthesis and free radical ring- acetate) s: A Versatile Class of Thermoresponsive and Biocompatible
opening polymerization of 2-methylene-4-phenyl-1, 3-dioxolane. Polymers. Angew. Chem., Int. Ed. 2017, 56, 9178−9182.
Makromol. Chem. 1982, 183, 1913−1920. (32) Allen, G.; Bevington, J. C. Comprehensive Polymer Science;
(15) Bailey, W. J.; Endo, T.; Gapud, B.; Lin, Y.-N.; Ni, Z.; Pan, C.-Y.; Pergamon Press: Oxford, 1989; Vol. 5.
Shaffer, S. E.; Wu, S.-R.; Yamazaki, N.; Yonezawa, K. Synthesis of (33) Wu, Z.; Stanley, R. R.; Pittman, C. U. Selective diesterification
functionally-terminated oligomers by free radical ring-opening of diols through cyclic ketene acetal intermediates. J. Org. Chem. 1999,
polymerization. J. Macromol. Sci., Part A 1984, 21, 979−995. 64, 8386−8395.
(16) Bailey, W. J. Free radical ring-opening polymerization. Polym. J. (34) Ren, L.; Speyerer, C.; Agarwal, S. Free-Radical Copolymeriza-
1985, 17, 85−95. tion Behavior of 5, 6-Benzo-2-methylene-1, 3-dioxepane and

L https://doi.org/10.1021/acs.macromol.3c02474
Macromolecules XXXX, XXX, XXX−XXX
Macromolecules pubs.acs.org/Macromolecules Article

Methacrylic Acid via the in Situ Generation of 3-Methyl-1, 5- polymerization modification. Biomacromolecules 2015, 16, 2049−
dihydrobenzo [e][1, 3] dioxepin-3-yl Methacrylate and 2-(Acetox- 2058.
ymethyl) benzyl Methacrylate. Macromolecules 2007, 40, 7834−7841. (54) Hedir, G.; Stubbs, C.; Aston, P.; Dove, A. P.; Gibson, M. I.
(35) Landfester, K. Miniemulsionspolymerisation und Struktur von Synthesis of degradable poly (vinyl alcohol) by radical ring-opening
Polymer-und Hybridnanopartikeln. Angew. Chem. 2009, 121 (25), copolymerization and ice recrystallization inhibition activity. ACS
4556−4576, DOI: 10.1002/ange.200900723. Macro Lett. 2017, 6, 1404−1408.
(36) Siebert, J. M.; Baumann, D.; Zeller, A.; Mailänder, V.; (55) Van Herk, A. M. Chemistry and Technology of Emulsion
Landfester, K. Synthesis of Polyester Nanoparticles in Miniemulsion Polymerisation; John Wiley & Sons, 2013.
Obtained by Radical Ring-Opening of BMDO and Their Potential as (56) Agredo, P.; Rave, M. C.; Echeverri, J. D.; Romero, D.;
Biodegradable Drug Carriers. Macromol. Biosci. 2012, 12, 165−175. Salamanca, C. H. An evaluation of the physicochemical properties of
(37) Asua, J. M. Miniemulsion polymerization. Prog. Polym. Sci. stabilized oil-in-water emulsions using different cationic surfactant
2002, 27, 1283−1346. blends for potential use in the cosmetic industry. Cosmetics 2019, 6,
(38) Asua, J. M. Challenges for industrialization of miniemulsion No. 12.
polymerization. Prog. Polym. Sci. 2014, 39, 1797−1826. (57) Thoniyot, P.; Mothe, S. R.; Zhao, W. Method for Preparing a
(39) Galanopoulo, P.; Gil, N.; Gigmes, D.; Lefay, C.; Guillaneuf, Y.; Polymer Having Ester Functionality and Said Polymer Prepared
Lages, M.; Nicolas, J.; Lansalot, M.; d’Agosto, F. One-Step Synthesis Therefrom. WO Patent. WO075696A32023.
of Degradable Vinylic Polymer-Based Latexes via Aqueous Radical
Emulsion Polymerization. Angew. Chem., Int. Ed. 2022, 61,
No. e202117498.
(40) Carter, M. C. D.; Hejl, A.; Woodfin, S.; Einsla, B.; Janco, M.;
DeFelippis, J.; Cooper, R. J.; Even, R. C. Backbone-degradable vinyl
acetate latex: Coatings for single-use paper products. ACS Macro Lett.
2021, 10, 591−597.
(41) Kordes, B. R.; Ascherl, L.; Rüdinger, C.; Melchin, T.; Agarwal,
S. Competition between Hydrolysis and Radical Ring-Opening
Polymerization of MDO in Water. Who Makes the Race? Macro-
molecules 2023, 56, 1033−1044.
(42) Carter, M. C. D.; Hejl, A.; Janco, M.; DeFelippis, J.; Yang, P.;
Gallagher, M.; Liang, Y. Emulsion Polymerization of 2-Methylene-1,
3-Dioxepane and Vinyl Acetate: Process Analysis and Character-
ization. Macromolecules 2023, 56, 5718−5729.
(43) Galanopoulo, P.; Sinniger, L.; Gil, N.; Gigmes, D.; Lefay, C.;
Guillaneuf, Y.; Lages, M.; Nicolas, J.; d’Agosto, F.; Lansalot, M.
Degradable vinyl polymer particles by radical aqueous emulsion
copolymerization of methyl methacrylate and 5, 6-benzo-2-methyl-
ene-1, 3-dioxepane. Polym. Chem. 2023, 14, 1224−1231.
(44) Mothe, S. R.; Chennamaneni, L. R.; Tan, J.; Lim, F. C.; Zhao,
W.; Thoniyot, P. A Mechanistic Study on the Hydrolysis of Cyclic
Ketene Acetal (CKA) and Proof of Concept of Polymerization in
Water. Macromol. Chem. Phys. 2023, 224, No. 2300221.
(45) Bailey, W. J. Carboxy-Functional Polymers and Their Use as
Detergent Additives. U.S. Patent, US4,923,9411990.
(46) Thoniyot, P. Synthesis of Polyester Based Polymers without
Use of Organic Solvents. U.S. Patent, US17/438,7732022.
(47) Asua, J. M.; Rodriguez, V. S.; Silebi, C. A.; El-Aasser, M. S.
Miniemulsion copolymerization of styrene-methyl methacrylate:
Effect of transport phenomena. Makromol. Chem., Macromol. Symp.
1990, 35-36 (1), 59−85, DOI: 10.1002/masy.19900350107.
(48) Smeets, N. M. B.; Jansen, T. G. T.; Van Herk, A. M.; Meuldijk,
J.; Heuts, J. P. A. Mass transport by collisions in emulsion
polymerization: why it is possible to use very hydrophobic catalysts
for efficient molecular weight control. Polym. Chem. 2011, 2, 1830−
1836.
(49) Boscán, F.; Paulis, M.; Barandiaran, M. J. Towards the
production of high-performance lauryl methacrylate based polymers
through emulsion polymerization. Eur. Polym. J. 2017, 93, 44−52.
(50) Tripathi, A.; Wie, C.; Tauer, K. Swelling of latex particles-
towards a solution of the riddle. Colloid Polym. Sci. 2017, 295, 189−
196.
(51) Ceska, G. W. Carboxyl-stabilized emulsion polymers. J. Appl.
Polym. Sci. 1974, 18, 2493−2499.
(52) Undin, J.; Illanes, T.; Finne-Wistrand, A.; Albertsson, A. C.
Random introduction of degradable linkages into functional vinyl
polymers by radical ring-opening polymerization, tailored for soft
tissue engineering. Polym. Chem. 2012, 3, 1260−1266.
(53) Hedir, G. G.; Bell, C. A.; O’Reilly, R. K.; Dove, A. P. Functional
degradable polymers by radical ring-opening copolymerization of
MDO and vinyl bromobutanoate: synthesis, degradability and post-

M https://doi.org/10.1021/acs.macromol.3c02474
Macromolecules XXXX, XXX, XXX−XXX

You might also like