Organic nanoparticles with tunable size and rigidity by hyperbranching and cross-linking using microemulsion ATRP

Contributed by Krzysztof Matyjaszewski; received March 27, 2024; accepted June 13, 2024; reviewed by Brian C. Benicewicz, Russell J. Composto, and Tom Russell
July 10, 2024
121 (29) e2406337121

Significance

While the focus on grafting polymers from inorganic nanoparticle surfaces has been extensive, existing techniques feature only limited possibilities to alter the internal molecular structures of nanoparticle systems. This limits the variation of chemical functionality and assembly behavior of particle-based materials. The presented inimer-based microemulsion approach overcomes existing challenges by enabling highly uniform organic nanoparticle (oNP) systems with tunable mechanical properties (by variation of cross-link density) and functionality (by integration of functional small molecular entities). The macroinitiator characteristics of oNPs enable the versatile graft modification which allows the direct assembly of oNPs or their integration in matrix materials. These brush-tethered oNPs are thus anticipated to unlock innovative applications across a range of nanomaterial technologies.

Abstract

Unlike inorganic nanoparticles, organic nanoparticles (oNPs) offer the advantage of “interior tailorability,” thereby enabling the controlled variation of physicochemical characteristics and functionalities, for example, by incorporation of diverse functional small molecules. In this study, a unique inimer-based microemulsion approach is presented to realize oNPs with enhanced control of chemical and mechanical properties by deliberate variation of the degree of hyperbranching or cross-linking. The use of anionic cosurfactants led to oNPs with superior uniformity. Benefitting from the high initiator concentration from inimer and preserved chain-end functionality during atom transfer radical polymerization (ATRP), the capability of oNPs as a multifunctional macroinitiator for the subsequent surface-initiated ATRP was demonstrated. This facilitated the synthesis of densely tethered poly(methyl methacrylate) brush oNPs. Detailed analysis revealed that exceptionally high grafting densities (~1 nm−2) were attributable to multilayer surface grafting from oNPs due to the hyperbranched macromolecular architecture. The ability to control functional attributes along with elastic properties renders this “bottom-up” synthetic strategy of macroinitiator-type oNPs a unique platform for realizing functional materials with a broad spectrum of applications.
Nanoparticle brushes (i.e., polymer-grafted nanoparticles or “hairy particles”), characterized by functional polymer-tethered nanoparticle compounds, have emerged as nanomaterial platform to enable functional material systems with diverse applications (19). Depending on the density of grafted chains, steric crowding alters both the structure of polymers at the particle interface as well as the interactions and assembly behavior of brush particles in the solid state (1012). The resultant brush particle materials exhibit unique and synergistic features, and open avenues for versatile designs with applications spanning self-healing, colloidal crystal, dielectrics, solid-state electrolyte, and antibacterials (1321). The synthesis of brush nanoparticles requires surface-anchored functional groups or initiators as the prerequisite (2225). Polymer grafts can be tethered using “grafting-from,” where polymerization initiates from the surface-bound initiators (2628). This method leverages advanced surface-initiated atom transfer radical polymerization (SI-ATRP) strategy (2932), since ATRP enables the grafting of (co)polymers with precise control over predetermined molecular weight, low dispersity, preferred compositions, and high chain-end fidelity (3338).
While previous studies on nanoparticle brushes predominantly focused on polymer-tethered “inorganic” nanoparticles (SiO2, TiO2, ZnO, etc.) (3941), a new frontier has emerged in the form of organic nanoparticles (oNPs) (4245). Unlike their inorganic counterparts, materials composed of oNPs have lower density. Furthermore, oNPs offer interior tailorability, enabling the incorporation of various small molecules such as fluorescent compounds, semiconductors, and drugs. This internal customization empowers oNPs with multifaceted functionalities such as imaging, sensing, and drug delivery while ensuring environmental benignity and excellent biocompatibility (4649). The synthetic methods for oNPs diverge from the traditional “top-down” approach, which involves extracting oNPs from specific templates (50). Instead, the “bottom-up” approach employs self-assembly through noncovalent interactions, such as hydrogen bonding (51), van der Waals forces (52), and π–π interactions (53). However, unlike covalent bonding, these intermolecular interactions are typically labile, which may impact the overall stability and performance of oNPs under harsh conditions (44).
Polymerization in dispersed media offers enhanced control compared to solution-based methods, attributed to the confined propagation within emulsion compartments (5457). It provides advantages such as regulated high molecular weight polymer synthesis, prevention of agglomeration, and reduction in reaction medium viscosity (5760). By selecting an appropriate surfactant and employing suitable homogenization methods, a water and oil mixture can form a thermodynamically stable microemulsion (ME) with droplets of sub-100 nm diameter (6164). In previous studies, ME ATRP of inimer (i.e., initiator-containing monomer), 2-(2-bromoisobutyryloxy)ethyl methacrylate (BiBEM), resulted in the production of a uniform hyperbranched polymer with a spherical shape (65). This was in stark contrast to PBiBEM synthesized by solution polymerization, which exhibited high dispersity due to concurrent initiation and propagation. The uniform PBiBEM, characterized by its hyperbranched topology (66), high ATRP initiator concentration, and living chain-ends, was well-suited as oNP for further grafting brush layers. However, despite its potential as a nanoparticle macroinitiator, PBiBEM featured a rather low modulus (E ~ 10 to 100 MPa) in comparison to rigid inorganic compounds (E ~ tens of GPa). Softness limits the ability to organize oNPs into ordered assembly structures and thus may pose challenges for the application of oNPs, for example, as membrane materials.
In this study, we synthesized oNPs by copolymerizing BiBEM with ethylene glycol dimethacrylate (EGDMA) using ME ATRP with nonionic surfactant Brij 98 (Fig. 1A). Methacrylates were chosen for their higher glass transition temperatures compared to the acrylate counterparts (67), conferring improved tunability of mechanical properties to oNPs. The concurrent hyperbranching and cross-linking were aimed at enhancing the internal rigidity of oNP. The use of MEs as the dispersed media ensured that each oNP macromolecule was isolated within a droplet, resulting in spherical oNPs with nanoscale size, uniform size distribution, and minimal agglomeration. The increase of EGDMA content resulted in oNPs with smaller size, lower dispersity, and larger rigidity (Fig. 1B). Additionally, the introduction of sodium dodecyl sulfate (SDS) as a second anionic cosurfactant led to oNPs with less than 3% size dispersity (Fig. 1C). The high initiator concentration from inimer and retention of chain-end fidelity by ATRP, rendered the oNP an effective SI-ATRP macroinitiator. The representative grafting reactions of poly(methyl methacrylate) (PMMA) brushes from oNPs demonstrated high grafting density. The estimation of “active bromide” fractions suggested multilayer tethering of brushes, attributed to the peripheral hyperbranching of oNPs. Such oNP and related oNP brushes can expand the library of brush-modified nanoparticles and are envisioned as nanomaterials with specific functionalities customized for various application needs.
Fig. 1.
Schematic illustrations of oNPs synthesized via hyperbranching and cross-linking using ME ATRP. (A) Synthetic procedure; (B) tunable size achieved through varying cross-linker/inimer compositions; and (C) two different surfactant systems: nonionic Brij 98 as the sole surfactant and a combination of Brij 98 and anionic SDS as cosurfactants.

Results

Synthetic Procedures.

The oNPs based on hyperbranching and cross-linking were prepared using a ME ATRP method (SI Appendix, Fig. S1). The aqueous phase contained hydrophilic Cu catalyst complex, [Br−CuII(TPMA)]+ [TPMA = tris(2-pyridylmethyl)amine], together with surfactants in water. Brij 98, i.e., polyoxyethylene (20) oleyl ether, served as the primary surfactant at the oil–water interface. Its nonionic nature ensured the formation of stable ME droplets with sub-100 nm hydrodynamic diameters (65). Additionally, SDS, an anionic surfactant, was selected as a cosurfactant (SDS: Brij 98 = 10: 90% mol/mol) to enhance electrostatic interactions for a better-stabilized ME. The presence of dodecyl sulfate anion (DS) also facilitated the transport of catalyst through interfaces via ion pair formation with Cu2+, thereby improving polymerization efficiency (68, 69). Sodium bromide (NaBr) was added to mitigate the dissociation of the Cu catalyst (70). The oil phase, composed of the premixed inimer BiBEM (SI Appendix, Fig. S2) and cross-linker EGDMA, was introduced dropwise into the aqueous phase without any organic solvent. This oil–water mixture was homogenized using a vortex mixer to form ME, setting the stage for copolymerization within ME compartments upon feeding ascorbic acid (AsAc) as the reducing agent.
The reaction reached high comonomer conversion after 2 h, as evidenced by the absence of vinyl proton peaks in proton NMR (1H NMR) spectra. The hydrodynamic sizes of ME droplets were analyzed using dynamic light scattering (DLS) at three distinct stages of the same reaction (SI Appendix, Fig. S3): immediately after vortex mixing (t = 0 min), at t = 3 min, and just before product precipitation (t = 2 h). Across these stages, the number distribution mode revealed consistent monomodal peaks, though with the progressively reduced droplet size. The volume distribution accentuates the presence of larger nanoparticles (SI Appendix, Eqs. S1S3) and the ME showed bimodal size distributions immediately after vortex mixing, despite having peak positions similar to those at the subsequent stages. This observation suggested that the initial mixing process resulted in a diverse range of droplet sizes, which may converge toward a more uniform size distribution as the reaction progressed. The purification of the final oNP products involved precipitation in methanol, dissolution in tetrahydrofuran (THF), centrifugation, and repeated dialysis. This process yielded oNPs that could be either maintained in THF as a stable solution or dried under a vacuum to form solids for further studies and subsequent SI-ATRP reactions. The stability of oNPs in acidic conditions was evaluated through a 24-h exposure to trifluoroacetic acid (TFA, 10 mL) with oNPs (B60E40, 41.6 mg) dissolved in dichloromethane (DCM, anhydrous, 50 mL). The impact of TFA on the size of the oNPs over this period was negligible (SI Appendix, Fig. S4), leading us to conclude that the oNPs are stable under acidic conditions.

Tunable Size and Rigidity of oNPs.

The preparation of oNPs with varying compositional ratios of BiBEM to EGDMA (from 90:10 to 60:40 mol/mol) was carried out with either Brij 98 as the sole surfactant or Brij 98/SDS as cosurfactants (Table 1). These systems were denoted as BxE100-x and SBxE100-x, respectively, where x indicated the BiBEM molar ratio. Following purification, the oNP samples were characterized by size exclusion chromatography (SEC) with THF eluent. The apparent molecular weights (Mn, RI) and their distributions (i.e., Mw/Mn) were determined using the refractive index (RI) detector and PMMA standards. The absolute molecular weights (Mn, MALS) were measured by the multiangle light scattering (MALS) detector and calibrated using sample concentrations. Additionally, the hydrodynamic diameters of the ME droplets (after 2 h, prior to precipitation) and of the oNPs after redispersion in THF were analyzed by DLS using the volume-averaged distribution.
Table 1.
oNPs by Hyperbranching and cross-linking using ME ATRP
Entry*x, BiBEM (mol %)x, EGDMA (mol %)Mn, RI (×103)Mw/MnMn, MALS (×103)DME, water (nm)§DoNP, THF (nm)§Size change (% v/v)
B100E0#10001401.241,47027.036.0137.0
B90E1090102021.586,02832.9 ± 0.736.4 ± 0.335.4
B85E1585151811.565,75932.3 ± 0.535.4 ± 0.731.6
B80E2080201521.505,00629.3 ± 0.731.2 ± 0.320.7
B75E2575251401.414,26126.9 ± 0.527.2 ± 0.43.4
B70E3070301211.343,99025.0 ± 0.524.7 ± 0.3−3.6
B65E3565351041.253,50324.3 ± 0.223.8 ± 0.2−6.0
B60E406040911.232,85022.3 ± 0.720.9 ± 0.5−17.7
SB90E1090101371.233,45924.1 ± 0.626.5 ± 0.632.9
SB85E1585151271.143,99123.7 ± 0.425.5 ± 0.424.6
SB80E2080201071.113,60922.0 ± 0.322.7 ± 0.29.9
SB75E257525871.093,31121.0 ± 0.421.0 ± 0.20.0
SB70E307030761.082,60219.9 ± 0.418.8 ± 0.1−15.7
SB65E356535721.082,68819.4 ± 0.417.9 ± 0.3−21.4
SB60E406040671.072,21519.3 ± 0.317.7 ± 0.3−22.9
*
Reaction conditions: when Brij 98 functions as the sole surfactant (BxE100−x), [BiBEM + EGDMA]0: [CuBr2(TPMA)]0: [AsAc]0 = 100: 0.06: 0.3, BiBEM 0.5 × (x/100) g, Brij 98 0.5 g in 12 g water, AsAc (10 mg/mL in water, 0.095 mL) was injected, 50 °C for 2 h; with SDS and Brij 98 as cosurfactants (SBxE100−x), besides the aforementioned conditions, [SDS]0: [Brij 98]0 = 10: 90 (% mol/mol), [NaBr]0 = 0.1 M.
Apparent molecular weight (Mn, RI) and distribution (Mw/Mn) of oNPs were determined by THF SEC using the RI detector and calibrated by PMMA standards.
Absolute molecular weight of oNP was determined by THF SEC using MALS detector and calibrated using oNP sample concentration in THF (mg/mL).
§
Hydrodynamic diameters of the ME droplets (after 2 h, prior to precipitation) and of the oNPs redispersed in THF were determined by DLS.
Size changes (% v/v) between redispersion and ME were calculated through (VoNPs,THFVME,water) ÷ VME,water.
#
Sample information of hyperbranched PBiBEM was referred in a previous study (65).
Given the absence of solvents other than the comonomers within the reaction compartments, the original size of oNP was presumed to closely match the hydrodynamic diameter of the ME before product precipitation. However, the increasing cross-linking density contributed to significant size alterations upon redispersion in THF (Fig. 2A). For oNP with a lower cross-linker to inimer ratio, the redispersion led to the expansion of their hyperbranched structures. This resulted in a “swollen” effect, evidenced by the increased hydrodynamic diameters. Conversely, oNPs with a higher proportion of EGDMA demonstrated a “shrunk” effect upon redispersion. This diminished penetration by solvent molecules led to a decreased hydrodynamic size. Additionally, the size reduction could also be attributed to the loss of unreacted comonomers, which could have been incorporated to a cross-linked and vitrified matrix at high conversion. Therefore, despite the general trend of size reduction in both ME and redispersed states, oNPs with increased EGDMA content showed a gradual “size change (% v/v)” transition from “swollen” to “shrunk” effect (Fig. 2B). This dynamic interplay between the oNP compositions and its dissolution behavior highlighted the importance of macromolecular structure design of oNP. Hydrodynamic diameters of redispersed oNP products varied from 20.9 to 36.4 nm with Brij 98 as the sole surfactant and from 17.7 to 26.5 nm with inclusion of SDS as a cosurfactant (SI Appendix, Fig. S5). The smaller size of the latter were attributed to the improved electrostatic stability of ME provided by SDS. Using the same formulations, anionic surfactants produce smaller droplet sizes than nonionic analogs (71). The negatively charged ME droplets (due to anionic SDS) prevented coalescence; their electrostatic repulsions were stronger than the steric repulsion from the nonionic surfactant Brij 98. Notably, in both the “Brij 98 only” and “cosurfactant” systems, the “size changes” varied in a similar trend from approximately +35 to −20% v/v (Fig. 2C). This demonstrated that the compositional ratios of oNPs played a crucial role in regulating the internal architecture thereby influencing their dissolution behavior.
Fig. 2.
oNPs synthesized with tunable size by hyperbranching and cross-linking using ME ATRP (reaction conditions are listed in Table 1). (A) Proposed dissolution behaviors of “swollen” and “shrunk” oNPs when varying the cross-linker (EGDMA) to inimer (BiBEM) compositional ratios. Hydrodynamic diameters associated with different BiBEM compositions, measured by DLS of ME droplets before product precipitation (empty icons) and redispersed oNPs in THF (filled icons), where (B) Brij 98 functioned as the sole surfactant (green), and (C) SDS was incorporated with Brij 98 as cosurfactants (orange). Size changes (% v/v, blue stars) were calculated through (VoNPs,THFVME,water) ÷ VME,water. (D) Molecular weight distributions (Mw/Mn) of oNPs were determined by THF SEC RI detector and were calibrated using linear PMMA standards. (Normalized) THF SEC elution with time of oNPs using (E) Brij 98 as the sole surfactant (green) and (F) SDS/Brij 98 as cosurfactants (orange).
Since each oNP could be considered as a single macromolecule, the size distributions of oNPs were also characterized as the molecular weight distributions, as indicated by the THF SEC RI detector (Fig. 2D). The oNPs synthesized exclusively with Brij 98 exhibited a higher dispersity (Mw/Mn = 1.58) starting from B90E10 but demonstrated improved size uniformity with increasing amounts of cross-linker (Fig. 2E). A more uniform distribution with a lower dispersity (Mw/Mn = 1.23) was observed for B60E40, attributed to the overall denser architectures regulated by EGDMA. Moreover, with the introduction of a cosurfactant system, oNPs started with low dispersity at lower EGDMA compositions (Mw/Mn = 1.23 for SB90E10). The uniformity further improved as the EGDMA content increased, achieving a dispersity of 1.07 for SB60E40 (Fig. 2F). The enhancement in oNP uniformity was achievable either by elevating the cross-linker concentration or by adding SDS as a cosurfactant. The latter's effectiveness was twofold: First, the anionic nature of SDS promoted electrostatic stabilization of the ME, leading to more uniform size distributions. Second, SDS facilitated the transport of the copper catalyst across oil–water interfaces, enabling more controlled copolymerization and, consequently, more uniform chain growth.
The morphology of oNPs was analyzed using atomic force microscopy (AFM) for three specific BiBEM/EGDMA compositions: 90:10, 75:25, and 60:40. Compared with transmission electron microscopy, AFM provides height information and three-dimensional cross-sectional views, illustrating the morphology of oNPs more effectively. The oNP sample solutions (0.05 mg/mL in THF) were drop-cast onto a silicon wafer and placed in a vacuum oven at 45 °C for 24 h to facilitate drying and the removal of solvent residue. Samples synthesized with Brij 98 only exhibited larger size contrasts and were used for bulk film imaging (Fig. 3 AC), while the more uniform oNPs, synthesized using cosurfactants, were highly diluted (0.05 mg/mL) for dispersed sample imaging (Fig. 3 DF). Images of bulk films showed spherical oNPs with some particles partially merging due to the entanglement of interfacial hyperbranched structures. With increasing EGDMA compositions, higher numbers of oNPs were observed within the same area (800 × 800 nm), corroborating the size reduction observed in DLS. Furthermore, a change in uniformity was noted, with dispersed oNPs evenly distributed and showing minimal agglomeration. Despite a relatively low dispersity (Mw/Mn = 1.23) measured for SB90E10, a broad size distribution was evident, with both large and small oNPs visible. Enhanced size uniformity was observed with increased cross-linker content, as seen in the SB60E40 samples, which displayed sharp cone-like structures in 3D cross-sectional height images. The rigidity of oNPs was initially estimated by analyzing the height-to-width ratio, with the expected 1:1 ratio for “inorganic” NPs. For each uniformly dispersed SBxE100-x sample, 12 individual oNPs were selected for analysis (SI Appendix, Figs. S6–S12). After excluding the maximal and minimal values, the average height-to-width ratio was calculated, starting from 0.22 for SB90E10 with low EGDMA composition and increasing to a plateau around 0.45 for SB75E25 (SI Appendix, Fig. S13 and Table S2). This rough calculation suggested that a higher cross-linking density led to an increased oNP rigidity.
Fig. 3.
AFM planar and 3D cross-sectional height images of oNP bulk films: (A) B90E10, (B) B75E25, and (C) B60E40; and dispersed oNPs: (D) SB90E10, (E) SB75E25, and (F) SB60E40. Sample concentration: (0.05 mg/mL in THF); Image size: 800 × 800 nm; (Scale bar: 160 nm.)
The density of solid oNPs was measured using a helium pycnometer. This method involves displacing the skeletal volume of the oNPs with helium gas, with the density calculated based on the volume-pressure relationship dictated by Boyle’s Law. The solid densities recorded were (SI Appendix, Table S3) 1.4452 ± 0.0061 g/cm3 for B90E10, 1.4249 ± 0.0073 g/cm3 for B75E25, and 1.4368 ± 0.0027 g/cm3 for B60E40. These values, closely aligned with each other, are all significantly higher than densities of the respective comonomers (1.303 g/cm3 for BiBEM and 1.051 g/cm3 for EGDMA), indicating the dense architecture of the oNPs. The highest density, observed in B90E10, is attributed to a larger content of BiBEM. Meanwhile, B60E40 had the second-highest density due to a higher cross-linking content. This dense yet relatively low-density structure of the oNPs makes them advantageous for the fabrication of lightweight composite materials in comparison to their inorganic counterparts.
To investigate the intraparticle elastic properties, Brillouin light spectroscopy (BLS) was employed to further elucidate the adjustable internal rigidity of oNPs. This technique examines the inelastic scattering between light and acoustic phonons within a material. The latter cause a frequency shift in the scattered light (fB, GHz) from which the elastic modulus of a material can be deduced (7274). The study analyzed three oNP bulk films with variable cross-linker contents: B90E10, B75E25, and B60E40 (SI Appendix, Fig. S14). The frequency shift fB(q) of the scattered light, as a function of the scattering wavevector (q, nm−1), was plotted accordingly for these samples (Fig. 4). The observed linear acoustic relation yields the longitudinal sound velocity cL via fB=cLq/2π (where for the propagation parallel to the film, q=(4π/λ)sin(θ/2) with θ denoting the scattering angle and λ the wavelength of light) (75, 76). The calculated sound velocities increased with the EGDMA content in oNPs, as 2,412 m s−1 for B90E10, 2,520 m s−1 for B75E25, and 2,550 m s−1 for B60E40, respectively; the experimental error amounts to less than 1%. The increase in sound velocity corresponded to about 12% enhancement in the estimated longitudinal modulus (MoNP) of the oNP from B90E10 to B60E40 assuming the same density (MoNP=ρcL2, where ρ is the density). The bulk-like modulus revealed by BLS suggested that oNPs featured near bulk density. This indicated the effectiveness of the dual cross-link and hyperbranching mechanism underlying oNP formation.
Fig. 4.
Measured (hexagonal icons) and fitted (dash line) phonon dispersion relations of oNP films B90E10, B75E25, and B60E40 using BLS. For the scattering geometry used for q > 0.03 nm−1 (normal to the film), q = (4πn/λ)sin(θ/2). The linear acoustic behavior was revealed using a RI n = 1.53. The black arrow served to guide the eye, showing the increased frequency shift of the scattered light with higher EGDMA content in oNP. The Inset shows a digital photo of B75E25 oNP bulk film.

Thermal Properties.

The thermal behavior of the oNPs was evaluated using differential scanning calorimetry (DSC) (SI Appendix, Fig. S15). This analysis involved comparing the thermal properties of PBiBEM and PEGDMA homopolymers, which were synthesized using a similar ME ATRP method, to those of the oNPs. The glass transition temperature (Tg) for each sample was identified by the peak position of the differential heat flow derivative (DDSC). The findings revealed that PBiBEM displayed a distinct and relatively narrow glass transition with a Tg around 34 °C. In contrast, the PEGDMA homopolymer did not exhibit a noticeable glass transition even when the temperature range was extended up to 195 °C (SI Appendix, Fig. S16), although previous studies estimated PEGDMA Tg to be around 144 °C using oxygen permeability data obtained from electron spin resonance (77). The oNP samples, in comparison to PBiBEM, all showed broader glass transitions which became less discernible with increasing EGDMA content. Additionally, their Tgs shifted from 37 °C for SB90E10 to 43 °C for SB60E40. This trend suggested that as the hyperbranched structure became more densely cross-linked, the internal macromolecular mobility was significantly restricted. Despite the expectation that highly cross-linked oNPs would exhibit their predicted Tgs at higher temperatures, the observed weak glass transitions and Tgs around 40 °C were attributed to the presence of remaining dangling chains. Further thermostability testing through thermogravimetric analysis (TGA) indicated that the oNPs were thermally stable up to 250 °C (SI Appendix, Fig. S17).

SI-ATRP Grafting from oNPs.

ATRP provides macromolecules with high end-group fidelity for chain extensions (78, 79). The large fraction of bromide from BiBEM enabled oNPs to graft polymeric brushes directly from the surface (i.e., oNP-Br), bypassing additional surface modifications needed, for example, for silica particles. In a model SI-ATRP reaction, ethyl α-bromoisobutyrate (EBiB) acted as a “sacrificial initiator” to estimate the molecular weight of the grafted chains (Fig. 5A) (80, 81). Methyl methacrylate (MMA) was chosen as the monomer to determine the absolute molecular weight. Surface-initiated activators regenerated by electron transfer (SI-ARGET) ATRP method was employed, with tin(II) 2-ethylhexanoate [Sn(Oct)2] serving as the reducing agent (8284) The grafting process was quenched at relatively low viscosity to mitigate the radical coupling effects and cross-linking between brush particles. The mixtures of PMMA-grafted oNPs (oNP-g-PMMA) and linear PMMA (EBiB-PMMA) were purified through precipitation in methanol and redispersion in THF for further analysis. The THF SEC revealed that the elution time for particle brushes B75E25-g-PMMA shifted toward a higher molecular weight compared to the original oNP (Fig. 5B), confirming the retention of end-group functionality in the oNP-Br macroinitiator. Given the distinct separation between the peak positions, the molecular weight indicated by the peak of isolated EBiB-PMMA was attributed to the grafted PMMA brush layers. (SI Appendix, Figs. S18 and S19).
Fig. 5.
(A) Schematic illustration of SI-ARGET ATRP grafting from oNP-Br as macroinitiator with sacrificial initiator (EBiB). (B) (Normalized) THF SEC elution with time of oNP (B75E25, light green) and mixture of PMMA-grafted oNPs (B75E25-g-PMMA) with linear PMMA (EBiB-PMMA, dark green). (C) AFM planar and 3D cross-sectional height images of PMA-grafted oNPs (B60E40-g-PMA). Sample concentration: (0.05 mg/mL in THF); image size: 800 × 800 nm; (Scale bar: 160 nm.) Sample information: molecular weight and distribution of grafted PMA, Mn,app = 54,190 and Mw/Mn = 1.07; average number of active -Br, Nactive -Br = 902; fractions of “active -Br” per oNP-Br, factive -Br = 0.13; grafting density, σ = 0.66 chains nm−2.
Grafting density (σ, chains nm−2), a pivotal parameter for nanoparticle brushes, defines the number of polymer chains grafted per unit surface area (85) To accurately determine the grafting density of oNP-g-PMMA, extensive reaction metrics were exploited, including the hydrodynamic diameter (DoNP) and the calibrated molecular weight (Mn,MALS,oNP) of oNPs, monomer conversion of MMA (conv.), and the absolute molecular weight of grafted PMMA brushes (Mn, brush). The comprehensive calculation methodology was detailed in SI Appendix, Eqs. S4S12 and Table S4), with the consolidated findings presented in Table 2.
Table 2.
SI-ATRP from oNPs as macroinitiator (oNP-Br) with sacrificial initiator (EBiB)
Entry*DoNP (nm)Mn,MALS,oNP (×103)conv. (%)Mn, brush(×103)§Mw/Mn§[xoNP-Br]0Nactive -Brfactive -Br (%)σ (nm−2)
B100E0-g-PtBA#34.01,51027.013.21.32N/A1,50636.70.41
B90E10-g-PMMA36.46,02821.440.21.081.692,26511.30.54
B85E15-g-PMMA35.45,75919.740.11.081.471,90010.40.48
B80E20-g-PMMA31.25,00618.539.01.091.391,56110.20.51
B75E25-g-PMMA27.24,26120.541.81.091.471,39111.30.60
B70E30-g-PMMA24.73,99021.243.91.091.431,28311.70.67
B65E35-g-PMMA23.83,50319.844.91.091.2294610.40.53
B60E40-g-PMMA20.92,85015.937.41.091.1373310.60.53
*
Reaction conditions: the “active bromide” concentration of oNP-Br macroinitiator ([oNP-Br]0) was denoted as an unknown x; oNP-Br approximately 50 mg, MMA 56.3 mmol, 6 mL (50% v/v in anisole), EBiB 0.0113 mmol, [MMA]0: [oNP-Br]0: [EBiB]0 = 5,000: x: 1, [CuBr2]0 (0.005 g/mL in N,N-dimethylformamide, DMF) 200 ppm compared to monomer concentration, [CuBr2]0: [Me6TREN]0: [Sn(Oct)2]0 = 1: 3: 5, 50 °C reaction for 1 h.
The hydrodynamic diameters (in THF) and the absolute molecular weights of oNP-Br were the same as those in Table 1.
Monomer conversion was measured by a thermogravimetric method (SI Appendix, Table S1).
§
Absolute molecular weight and distribution of the grafted PMMA brush layer assisted by EBiB-PMMA were determined using THF SEC RI detector and calibrated by PMMA standards.
Calculated “active -Br” concentration (refer to the unknown x), average number and fractions of “active -Br” per oNP-Br, and the grafting density of oNP-g-PMMA were detailed in SI Appendix, Table S4.
#
Sample information of PtBA brushes grafting from polyinimer (containing a disulfide linker) was referred in a previous study (65).
The efficacy of oNP-Br macroinitiator with multiple initiating sites for SI-ATRP was evidenced by the average number of chains grafted or the number of “active bromides” (Nactive -Br) from the oNP surface. The term “active” was introduced to highlight the difference to “buried -Br” that were not able to initiate chains due to the steric hindrance and the lack of accessibility to activators and monomers in the interior or even at the peripheral areas of oNP-Br. Our findings revealed that the Nactive -Br for oNP-g-PMMA consistently decreased with an increase in EGDMA composition. Despite the reduction in hydrodynamic sizes, the grafting density (Nactive -Br per surface area) remained similar, ranging from 0.48 to 0.67 chains nm−2. These values were similar to the grafting densities for densely grafted inorganic nanoparticle brushes (86) For inorganic nanoparticles, polymer grafts can be initiated from a single layer from the solid surface. However, initiation from the oNP-Br surface could involve several Br-layers due to the hyperbranched structure at the fringe (SI Appendix, Fig. S20). Consequently, the proportion of active -Br (factive -Br) was approximately 11% based on the average total number of BiBEM molecules per oNP. The thickness of this multilayer (Δh) pertaining to active -Br was estimated, based on several simplified assumptions (SI Appendix, Eqs. S13 and S14 and Table S5): 1) the oNPs were considered to be perfectly spherical; 2) BiBEM and EGDMA, both being methacrylates, were assumed to have comparable reactivity ratios; and 3) there was no sterically hindered -Br at the oNP surface (87). For B90E10-g-PMMA, Δh was estimated at 0.71 nm. Considering the statistical segment length of a vinyl monomer unit as approximately 0.25 nm, the 0.71 nm thickness corresponded to about 2 to 3 layers of BiBEM containing active -Br. With an increase in cross-linking density, Δh gradually decreased to 0.38 nm, influenced by the reduced size and enhanced internal rigidity, yet still indicative of nearly a bilayer of active -Br.
The oNP-g-PMMA particle brushes were efficiently purified via ultrahigh-speed centrifugation at 16,000 × g for 1 h to separate linear polymers (EBiB-PMMA). Subsequent DLS measurement determined the intensity-averaged hydrodynamic sizes of the purified particle brushes (SI Appendix, Fig. S21). For instance, B75E25-g-PMMA dispersed in THF had an average diameter of 144 nm without any signs of aggregation. This measurement closely aligned with the approximated size expected for brush particles in a good solvent, assuming a degree of polymerization of 416 (SI Appendix, Eqs. S15S17 and Table S6). A representative morphology study using AFM was conducted on the additionally synthesized polymer-grafted oNPs (Fig. 5C), with poly(methyl acrylate) (PMA) brushes from oNP B60E40 (B60E40-g-PMA). This combination of oNP and polymer was selected to highlight the contrast between the most rigid oNP and the relatively soft PMA brushes (with Tg of approximately 10 °C). The linear PMA polymers, originating from sacrificial initiators (EBiB-PMA) within the product mixtures, were separated through ultrahigh-speed centrifugation. AFM height images depicted the nanoparticle brushes as evenly distributed, with high uniformity in size and no evidence of aggregation. Moreover, a distinct height contrast was observed suggesting a “core and shell” structure for each brush-tethered oNP, indicative of the interfaces between the oNP cores and the PMA brush layers.

Discussion

In summary, this work demonstrated the integrated synthesis and characterization of oNPs with tunable size and rigidity via concurrent hyperbranching and cross-linking using ME ATRP. The capability to finely adjust the properties of oNPs, such as their hydrodynamic diameters and internal rigidity, underscored the potential of this methodology in creating highly customized nanoparticle systems for specific applications. Notably, the high end-group fidelity, afforded by ATRP, enabled the efficient surface grafting of polymeric brushes, further enhancing the versatility of oNPs. The detailed investigation into the grafting density of brush-tethered oNPs highlighted their potential as uniform, self-assembled nanomaterials. This work contributed to the fundamental understanding of oNP synthesis and properties. Forthcoming research endeavors will focus on exploring functionalization opportunities (e.g., fluorescence) for this new class of oNPs and assessing their performance in practical applications, building upon the strong foundation established by this study.

Materials and Methods

2-Hydroxyethyl methacrylate (HEMA), the starting material to synthesize inimer BiBEM, was prewashed to remove the residual ethylene glycol dimethacrylate. Detailed HEMA prewashing procedures, inimer synthesis, oNP synthesis, surface-initiated ARGET ATRP, characterization methods, and instrument specifications, are provided in the experimental section of SI Appendix.

Data, Materials, and Software Availability

All study data are included in the article and/or supporting information.

Acknowledgments

The financial support from NSF (DMR 2202747, DMR 2209587), and the Department of Energy (DOE-BES−DESC0018784), is gratefully acknowledged.

Author contributions

A.K., M.R.B., and K.M. designed research; R.Y., J.T., A.G., J.J., X.H., Y.Z., H.W., Q.L., and G.F. performed research; G.F. and A.K. analyzed data; and R.Y., M.R.B., and K.M. wrote the paper.

Competing interests

The authors declare no competing interest.

Supporting Information

Appendix 01 (PDF)

References

1
A. C. Balazs, T. Emrick, T. P. Russell, Nanoparticle polymer composites: Where two small worlds meet. Science 314, 1107–1110 (2006).
2
S. K. Kumar, N. Jouault, B. Benicewicz, T. Neely, Nanocomposites with polymer grafted nanoparticles. Macromolecules 46, 3199–3214 (2013).
3
C. M. Hui et al., Surface-initiated polymerization as an enabling tool for multifunctional (nano-)engineered hybrid materials. Chem. Mater. 26, 745–762 (2014).
4
G. Nie, G. Li, L. Wang, X. Zhang, Nanocomposites of polymer brush and inorganic nanoparticles: Preparation, characterization and application. Polym. Chem. 7, 753–769 (2016).
5
J. O. Zoppe et al., Surface-initiated controlled radical polymerization: State-of-the-art, opportunities, and challenges in surface and interface engineering with polymer brushes. Chem. Rev. 117, 1105–1318 (2017).
6
S. K. Kumar, B. C. Benicewicz, R. A. Vaia, K. I. Winey, 50th anniversary perspective: Are polymer nanocomposites practical for applications? Macromolecules 50, 714–731 (2017).
7
G. Morgese et al., Next-generation polymer shells for inorganic nanoparticles are highly compact, ultra-dense, and long-lasting cyclic brushes. Angew. Chem. Int. Ed. 56, 4507–4511 (2017).
8
J. Yan, M. R. Bockstaller, K. Matyjaszewski, Brush-modified materials: Control of molecular architecture, assembly behavior, properties and applications. Prog. Polym. Sci. 100, 101180 (2020).
9
T. Hueckel, X. Luo, O. F. Aly, R. J. Macfarlane, Nanoparticle brushes: Macromolecular ligands for materials synthesis. Acc. Chem. Res. 56, 1931–1941 (2023).
10
J. Choi et al., Effect of polymer-graft modification on the order formation in particle assembly structures. Langmuir 29, 6452–6459 (2013).
11
M. Schmitt et al., Processing fragile matter: Effect of polymer graft modification on the mechanical properties and processibility of (nano-) particulate solids. Soft Matter 12, 3527–3537 (2016).
12
J. Lee et al., Molecular parameters governing the elastic properties of brush particle films. Macromolecules 53, 1502–1513 (2020).
13
Y. Zhao et al., Copolymer brush particle hybrid materials with “recall-and-repair” capability. Chem. Mater. 35, 6990–6997 (2023).
14
G. A. Williams et al., Mechanically robust and self-healable superlattice nanocomposites by self-assembly of single-component “sticky” polymer-grafted nanoparticles. Adv. Mater. 27, 3934–3941 (2015).
15
K. Ohno, T. Morinaga, S. Takeno, Y. Tsujii, T. Fukuda, Suspensions of silica particles grafted with concentrated polymer brush: A new family of colloidal crystals. Macromolecules 39, 1245–1249 (2006).
16
M. Wåhlander et al., Tailoring dielectric properties using designed polymer-grafted ZnO nanoparticles in silicone rubber. J. Mater. Chem. A 5, 14241–14258 (2017).
17
C. A. Grabowski et al., Performance of dielectric nanocomposites: Matrix-free, hairy nanoparticle assemblies and amorphous polymer-nanoparticle blends. ACS Appl. Mater. Interfaces 6, 21500–21509 (2014).
18
L. L. Dienemann et al., Hybrid particle brush coatings with tailored design for enhanced dendrite prevention and cycle life in lithium metal batteries. ACS Appl. Energy Mater. 6, 11602–11612 (2023).
19
S. Li et al., Grafting polymer from oxygen-vacancy-rich nanoparticles to enable protective layers for stable lithium metal anode. Nano Energy 76, 105046 (2020).
20
P. Pageni et al., Charged metallopolymer-grafted silica nanoparticles for antimicrobial applications. Biomacromolecules 19, 417–425 (2018).
21
J. M. Kubiak, B. Li, M. Suazo, R. J. Macfarlane, Polymer grafted nanoparticle composites with enhanced thermal and mechanical properties. ACS Appl. Mater. Interfaces 14, 21535–21543 (2022).
22
S. Kango et al., Surface modification of inorganic nanoparticles for development of organic–inorganic nanocomposites—A review. Prog. Polym. Sci. 38, 1232–1261 (2013).
23
Y. Song et al., Bioinspired polydopamine (PDA) chemistry meets ordered mesoporous carbons (OMCs): A benign surface modification strategy for versatile functionalization. Chem. Mater. 28, 5013–5021 (2016).
24
J. Yan et al., A fatty acid-inspired tetherable initiator for surface-initiated atom transfer radical polymerization. Chem. Mater. 29, 4963–4969 (2017).
25
J. Han et al., Nanosized organo-silica particles with “built-in” surface-initiated atom transfer radical polymerization capability as a platform for brush particle synthesis. ACS Macro Lett. 9, 1218–1223 (2020).
26
R. K. O’Reilly, M. J. Joralemon, C. J. Hawker, K. L. Wooley, Facile syntheses of surface-functionalized micelles and shell cross-linked nanoparticles. J. Polym. Sci. Part A: Polym. Chem. 44, 5203–5217 (2006).
27
K. Matyjaszewski, H. Dong, W. Jakubowski, J. Pietrasik, A. Kusumo, Grafting from surfaces for “everyone”: ARGET ATRP in the presence of air. Langmuir 23, 4528–4531 (2007).
28
M. S. Messina, K. M. M. Messina, A. Bhattacharya, H. R. Montgomery, H. D. Maynard, Preparation of biomolecule-polymer conjugates by grafting-from using ATRP, RAFT, or ROMP. Prog. Polym. Sci. 100, 101186 (2020).
29
A. Khabibullin, E. Mastan, K. Matyjaszewski, S. Zhu, “Surface-initiated atom transfer radical polymerization” in Controlled Radical Polymerization at and from Solid Surfaces, P. Vana, Ed. (Springer International Publishing, Cham, 2016), pp. 29–76.
30
L. Wu, U. Glebe, A. Böker, Synthesis of hybrid silica nanoparticles densely grafted with thermo and pH dual-responsive brushes via surface-initiated ATRP. Macromolecules 49, 9586–9596 (2016).
31
J. Yan et al., Solution processable liquid metal nanodroplets by surface-initiated atom transfer radical polymerization. Nat. Nanotechnol. 14, 684–690 (2019).
32
Z. Wang et al., Synthesis of gradient copolymer grafted particle brushes by ATRP. Macromolecules 52, 9466–9475 (2019).
33
K. Matyjaszewski, N. V. Tsarevsky, Nanostructured functional materials prepared by atom transfer radical polymerization. Nat. Chem. 1, 276–288 (2009).
34
K. Matyjaszewski, Advanced materials by atom transfer radical polymerization. Adv. Mater. 30, 1706441 (2018).
35
R. Yin, Z. Wang, M. R. Bockstaller, K. Matyjaszewski, Tuning dispersity of linear polymers and polymeric brushes grown from nanoparticles by atom transfer radical polymerization. Polym. Chem. 12, 6071–6082 (2021).
36
R. Yin et al., Alternating methyl methacrylate/n-butyl acrylate copolymer prepared by atom transfer radical polymerization. ACS Macro Lett. 11, 1217–1223 (2022).
37
R. Yin et al., Composition-orientation induced mechanical synergy in nanoparticle brushes with grafted gradient copolymers. Macromolecules 56, 9626–9635 (2023).
38
X. Hu et al., Red-light-driven atom transfer radical polymerization for high-throughput polymer synthesis in open air. J. Am. Chem. Soc. 145, 24315–24327 (2023).
39
R. J. Tannenbaum et al., Activated gas transport in polymer-grafted nanoparticle membranes. Macromolecules 56, 3954–3961 (2023).
40
G. Zhang et al., Novel polysulfone hybrid ultrafiltration membrane prepared with TiO2-g-HEMA and its antifouling characteristics. J. Membr. Sci. 436, 163–173 (2013).
41
Z. Wang, M. R. Bockstaller, K. Matyjaszewski, Synthesis and applications of ZnO/polymer nanohybrids. ACS Mater. Lett. 3, 599–621 (2021).
42
K. Margulis-Goshen, S. Magdassi, Organic nanoparticles from microemulsions: Formation and applications. Curr. Opin. Colloid Interface Sci. 17, 290–296 (2012).
43
B. Guo, E. Middha, B. Liu, Solvent magic for organic particles. ACS Nano 13, 2675–2680 (2019).
44
F. Fang, M. Li, J. Zhang, C.-S. Lee, Different strategies for organic nanoparticle preparation in biomedicine. ACS Mater. Lett. 2, 531–549 (2020).
45
M. Karg et al., Nanogels and microgels: From model colloids to applications, recent developments, and future trends. Langmuir 35, 6231–6255 (2019).
46
Y. Jiang, K. Pu, Advanced photoacoustic imaging applications of near-infrared absorbing organic nanoparticles. Small 13, 1700710 (2017).
47
J. Qi et al., Real-time and high-resolution bioimaging with bright aggregation-induced emission dots in short-wave infrared region. Adv. Mater. 30, 1706856 (2018).
48
S. Correa, N. Boehnke, E. Deiss-Yehiely, P. T. Hammond, Solution conditions tune and optimize loading of therapeutic polyelectrolytes into layer-by-layer functionalized liposomes. ACS Nano 13, 5623–5634 (2019).
49
B. J. Kim, H. Cheong, B. H. Hwang, H. J. Cha, Mussel-inspired protein nanoparticles containing iron(III)–DOPA complexes for pH-responsive drug delivery. Angew. Chem. Int. Ed. 54, 7318–7322 (2015).
50
J. Zhang et al., Green mass production of pure nanodrugs via an ice-template-assisted strategy. Nano Lett. 19, 658–665 (2019).
51
Z. Le et al., Hydrogen-bonded tannic acid-based anticancer nanoparticle for enhancement of oral chemotherapy. ACS Appl. Mater. Interfaces 10, 42186–42197 (2018).
52
Y. J. Kim, P. Guo, R. D. Schaller, Aqueous carbon quantum dot-embedded PC60-PC61BM nanospheres for ecological fluorescent printing: Contrasting fluorescence resonance energy-transfer signals between watermelon-like and random morphologies. J. Phys. Chem. Lett. 10, 6525–6535 (2019).
53
W.-R. Zhuang et al., Applications of π-π stacking interactions in the design of drug-delivery systems. J. Control. Release 294, 311–326 (2019).
54
P. B. Zetterlund, Y. Kagawa, M. Okubo, Controlled/living radical polymerization in dispersed systems. Chem. Rev. 108, 3747–3794 (2008).
55
K. Min, K. Matyjaszewski, Atom transfer radical polymerization in aqueous dispersed media. Cent. Eur. J. Chem. 7, 657–674 (2009).
56
R. Tomovska, J. C. de la Cal, J. M. Asua, “Reactions in heterogeneous media” in Monitoring Polymerization Reactions, W. F. Reed, A. M. Alb, Eds. (John Wiley & Sons, Inc., 2013), pp. 59–77.
57
Y. Wang, F. Lorandi, M. Fantin, K. Matyjaszewski, Atom transfer radical polymerization in dispersed media with low-ppm catalyst loading. Polymer 275, 125913 (2023).
58
J. K. Oh, C. Tang, H. Gao, N. V. Tsarevsky, K. Matyjaszewski, Inverse miniemulsion ATRP: A new method for synthesis and functionalization of well-defined water-soluble/cross-linked polymeric particles. J. Am. Chem. Soc. 128, 5578–5584 (2006).
59
M. Khan et al., Synthesis of multicompositional onion-like nanoparticles via RAFT emulsion polymerization. Angew. Chem. Int. Ed. 60, 23281–23288 (2021).
60
A. Cintora, F. Käfer, C. Yuan, C. K. Ober, Effect of monomer hydrophilicity on ARGET–ATRP kinetics in aqueous mini-emulsion polymerization. J. Polym. Sci. 60, 666–673 (2022).
61
K. Min, H. Gao, K. Matyjaszewski, Development of an ab initio emulsion atom transfer radical polymerization: From microemulsion to emulsion. J. Am. Chem. Soc. 128, 10521–10526 (2006).
62
S. Slomkowski et al., Terminology of polymers and polymerization processes in dispersed systems (IUPAC Recommendations 2011). Pure Appl. Chem. 83, 2229–2259 (2011).
63
K. Min et al., One-pot synthesis of hairy nanoparticles by emulsion ATRP. Macromolecules 42, 1597–1603 (2009).
64
M. Ciftci, M. A. Tasdelen, W. Li, K. Matyjaszewski, Y. Yagci, Photoinitiated ATRP in inverse microemulsion. Macromolecules 46, 9537–9543 (2013).
65
K. Min, H. Gao, New method to access hyperbranched polymers with uniform structure via one-pot polymerization of inimer in microemulsion. J. Am. Chem. Soc. 134, 15680–15683 (2012).
66
G. Shao, A. Li, Y. Liu, B. Yuan, W. Zhang, Branched polymers: Synthesis and application. Macromolecules 57, 830–846 (2024).
67
T. Li et al., Chain flexibility and glass transition temperatures of poly(n-alkyl (meth)acrylate)s: Implications of tacticity and chain dynamics. Polymer 213, 123207 (2021).
68
M. Fantin et al., Harnessing the interaction between surfactant and hydrophilic catalyst to control eATRP in miniemulsion. Macromolecules 50, 3726–3732 (2017).
69
R. Yin et al., Miniemulsion SI-ATRP by interfacial and ion-pair catalysis for the synthesis of nanoparticle brushes. Macromolecules 55, 6332–6340 (2022).
70
V. L. Teo, B. J. Davis, N. V. Tsarevsky, P. B. Zetterlund, Successful miniemulsion ATRP using an anionic surfactant: Minimization of deactivator loss by addition of a halide salt. Macromolecules 47, 6230–6237 (2014).
71
T. Al Najjar, N. K. Allam, E. N. El Sawy, Anionic/nonionic surfactants for controlled synthesis of highly concentrated sub-50 nm polystyrene spheres. Nanoscale Adv. 3, 5626–5635 (2021).
72
F. Kargar, A. A. Balandin, Advances in Brillouin-Mandelstam light-scattering spectroscopy. Nat. Photon. 15, 720–731 (2021).
73
H. Kim et al., Tunable hypersonic bandgap formation in anisotropic crystals of dumbbell nanoparticles. ACS Nano 17, 19224–19231 (2023).
74
P. Voudouris et al., Anisotropic elasticity of quasi-one-component polymer nanocomposites. ACS Nano 5, 5746–5754 (2011).
75
T. Horiguchi, K. Shimizu, T. Kurashima, M. Tateda, Y. Koyamada, Development of a distributed sensing technique using Brillouin scattering. J. Light. Technol. 13, 1296–1302 (1995).
76
D. Zhao, D. Schneider, G. Fytas, S. K. Kumar, Controlling the thermomechanical behavior of nanoparticle/polymer films. ACS Nano 8, 8163–8173 (2014).
77
D. Li, S. Zhu, A. E. Hamielec, E.s.r. study on permeation of oxygen in crosslinked polymers. Polymer 34, 1383–1387 (1993).
78
W. Jakubowski, B. Kirci-Denizli, R. R. Gil, K. Matyjaszewski, Polystyrene with improved chain-end functionality and higher molecular weight by ARGET ATRP. Macromol. Chem. Phys. 209, 32–39 (2008).
79
F. Nyström, A. H. Soeriyadi, C. Boyer, P. B. Zetterlund, M. R. Whittaker, End-group fidelity of copper(0)-meditated radical polymerization at high monomer conversion: An ESI-MS investigation. J. Polym. Sci. Part A: Polym. Chem. 49, 5313–5321 (2011).
80
G. Gazzola et al., Oxygen tolerance during surface-initiated photo-ATRP: Tips and tricks for making brushes under environmental conditions. ACS Macro Lett. 12, 1166–1172 (2023).
81
S. Wohlhauser, C. Rader, C. Weder, Facile method to determine the molecular weight of polymer grafts grown from cellulose nanocrystals. Biomacromolecules 23, 699–707 (2022).
82
L. Mueller, W. Jakubowski, W. Tang, K. Matyjaszewski, Successful chain extension of polyacrylate and polystyrene macroinitiators with methacrylates in an ARGET and ICAR ATRP. Macromolecules 40, 6464–6472 (2007).
83
Y. Kwak, A. J. D. Magenau, K. Matyjaszewski, ARGET ATRP of methyl acrylate with inexpensive ligands and ppm concentrations of catalyst. Macromolecules 44, 811–819 (2011).
84
Y. Zhao et al., Topologically induced heterogeneity in gradient copolymer brush particle materials. Macromolecules 55, 8846–8856 (2022).
85
L. Michalek, L. Barner, C. Barner-Kowollik, Polymer on top: Current limits and future perspectives of quantitatively evaluating surface grafting. Adv. Mater. 30, 1706321 (2018).
86
Z. Wang et al., Control of dispersity and grafting density of particle brushes by variation of ATRP catalyst concentration. ACS Macro Lett. 8, 859–864 (2019).
87
F. J. Arraez, P. H. M. Van Steenberge, J. Sobieski, K. Matyjaszewski, D. R. D'hooge, Conformational variations for surface-initiated reversible deactivation radical polymerization. Macromolecules 54, 8270–8288 (2021).

Information & Authors

Information

Published in

The cover image for PNAS Vol.121; No.29
Proceedings of the National Academy of Sciences
Vol. 121 | No. 29
July 16, 2024
PubMed: 38985759

Classifications

Data, Materials, and Software Availability

All study data are included in the article and/or supporting information.

Submission history

Received: March 27, 2024
Accepted: June 13, 2024
Published online: July 10, 2024
Published in issue: July 16, 2024

Keywords

  1. ATRP
  2. nanoparticle
  3. hyperbranched polymer
  4. microemulsion polymerization
  5. particle brushes

Acknowledgments

The financial support from NSF (DMR 2202747, DMR 2209587), and the Department of Energy (DOE-BES−DESC0018784), is gratefully acknowledged.
Author contributions
A.K., M.R.B., and K.M. designed research; R.Y., J.T., A.G., J.J., X.H., Y.Z., H.W., Q.L., and G.F. performed research; G.F. and A.K. analyzed data; and R.Y., M.R.B., and K.M. wrote the paper.
Competing interests
The authors declare no competing interest.

Notes

Reviewers: B.C.B., University of South Carolina; R.J.C., University of Pennsylvania; and T.R., University of Massachusetts Amherst.

Authors

Affiliations

Department of Chemistry, Carnegie Mellon University, Pittsburgh, PA 15213
Jirameth Tarnsangpradit
Department of Materials Science and Engineering, Carnegie Mellon University, Pittsburgh, PA 15213
Akhtar Gul
Department of Chemical and Biomolecular Engineering, University of Houston, Houston, TX 77204
Jaepil Jeong
Department of Chemistry, Carnegie Mellon University, Pittsburgh, PA 15213
Xiaolei Hu
Department of Chemistry, Carnegie Mellon University, Pittsburgh, PA 15213
Yuqi Zhao
Department of Materials Science and Engineering, Carnegie Mellon University, Pittsburgh, PA 15213
Hanshu Wu
Department of Chemistry, Carnegie Mellon University, Pittsburgh, PA 15213
Qiqi Li
Max Planck Institute for Polymer Research, Mainz 55128, Germany
Institute of Electronic Structure and Laser (IESL), Foundation for Research and Technology-Hellas (FORTH), Heraklion 70013, Greece
Max Planck Institute for Polymer Research, Mainz 55128, Germany
Institute of Electronic Structure and Laser (IESL), Foundation for Research and Technology-Hellas (FORTH), Heraklion 70013, Greece
Department of Chemical and Biomolecular Engineering, University of Houston, Houston, TX 77204
Department of Materials Science and Engineering, Carnegie Mellon University, Pittsburgh, PA 15213
Department of Chemistry, Carnegie Mellon University, Pittsburgh, PA 15213

Notes

1
To whom correspondence may be addressed. Email: [email protected] or [email protected].

Metrics & Citations

Metrics

Note: The article usage is presented with a three- to four-day delay and will update daily once available. Due to ths delay, usage data will not appear immediately following publication. Citation information is sourced from Crossref Cited-by service.


Altmetrics




Citations

Export the article citation data by selecting a format from the list below and clicking Export.

Cited by

    Loading...

    View Options

    View options

    PDF format

    Download this article as a PDF file

    DOWNLOAD PDF

    Login options

    Check if you have access through your login credentials or your institution to get full access on this article.

    Personal login Institutional Login

    Recommend to a librarian

    Recommend PNAS to a Librarian

    Purchase options

    Purchase this article to access the full text.

    Single Article Purchase

    Organic nanoparticles with tunable size and rigidity by hyperbranching and cross-linking using microemulsion ATRP
    Proceedings of the National Academy of Sciences
    • Vol. 121
    • No. 29

    Figures

    Tables

    Media

    Share

    Share

    Share article link

    Share on social media