@article{melillo_efimenko_genzer_2024, title={Water Droplet Equilibration on Silicone Elastomer Substrates Containing Mobile Silicones}, volume={10}, ISSN={["1520-5827"]}, url={https://doi.org/10.1021/acs.langmuir.4c02518}, DOI={10.1021/acs.langmuir.4c02518}, abstractNote={Understanding interfacial behavior at the water/silicone elastomer interface is vital in many applications, including microfluidics, antibiofouling, and self-cleaning surfaces. Silicone elastomers are not always static systems, however. Unreacted silicone molecules within a substrate may change the water-wetting behavior compared to fully reacted substrates. To investigate the impact of free silicone species at the interface, we systematically studied water wettability as a function of contact time with various silicone elastomer substrates. One set of Sylgard 184 substrates was prepared from curing at optimal stoichiometry and included doses of silicone molecules of varying molecular weights. Another set of substrates was made by mixing Sylgard 184 in imbalanced ratios to provide incomplete cross-linking. In the absence of added silicone oils, we observe a linear decrease in contact angle with time due to the evaporation of water. However, within certain molecular weight and loading levels of additional inert oils, nonlinear wetting behavior occurs as oils migrate to the interface and the system stabilizes. Similar nonlinear wetting behavior occurs without including silicone oils by mixing pure Sylgard 184 in very imbalanced mixing ratios, resulting in incomplete cross-linking. Such substrates contain unbound mobile silicone species that diffuse to the water-substrate interface. We show how the water contact angle technique can elucidate the presence of unbound silicone oils in the substrate.}, journal={LANGMUIR}, author={Melillo, Matthew J. and Efimenko, Kirill and Genzer, Jan}, year={2024}, month={Oct} } @article{liu_sui_harbinson_pudlo_perera_zhang_liu_ku_islam_liu_et al._2023, title={A scalable microstructure photonic coating fabricated by roll-to-roll “defects” for daytime sub-ambient passive radiative cooling}, volume={23}, ISSN={["1530-6992"]}, url={https://doi.org/10.1021/acs.nanolett.3c00111}, DOI={10.1021/acs.nanolett.3c00111}, abstractNote={The deep space's coldness (∼4 K) provides a ubiquitous and inexhaustible thermodynamic resource to suppress the cooling energy consumption. However, it is nontrivial to achieve subambient radiative cooling during daytime under strong direct sunlight, which requires rational and delicate photonic design for simultaneous high solar reflectivity (>94%) and thermal emissivity. A great challenge arises when trying to meet such strict photonic microstructure requirements while maintaining manufacturing scalability. Herein, we demonstrate a rapid, low-cost, template-free roll-to-roll method to fabricate spike microstructured photonic nanocomposite coatings with Al2O3 and TiO2 nanoparticles embedded that possess 96.0% of solar reflectivity and 97.0% of thermal emissivity. When facing direct sunlight in the spring of Chicago (average 699 W/m2 solar intensity), the coatings show a radiative cooling power of 39.1 W/m2. Combined with the coatings' superhydrophobic and contamination resistance merits, the potential 14.4% cooling energy-saving capability is numerically demonstrated across the United States.}, number={17}, journal={Nano Letters}, author={Liu, S. and Sui, C. and Harbinson, M. and Pudlo, M. and Perera, Himendra and Zhang, Zhenzhen and Liu, Ruguan and Ku, Zahyun and Islam, Md Didarul and Liu, Yuxuan and et al.}, editor={Ryu, JongEditor}, year={2023}, pages={7767–7774} } @article{davis_genzer_efimenko_abolhasani_2023, title={Continuous Ligand-Free Catalysis Using a Hybrid Polymer Network Support}, volume={7}, ISSN={["2691-3704"]}, url={https://doi.org/10.1021/jacsau.3c00261}, DOI={10.1021/jacsau.3c00261}, abstractNote={Although the pharmaceutical and fine chemical industries primarily utilize batch homogeneous reactions to carry out chemical transformations, emerging platforms seek to improve existing shortcomings by designing effective heterogeneous catalysis systems in continuous flow reactors. In this work, we present a versatile network-supported palladium (Pd) catalyst using a hybrid polymer of poly(methylvinylether-alt-maleic anhydride) and branched polyethyleneimine for intensified continuous flow synthesis of complex organic compounds via heterogeneous Suzuki–Miyaura cross-coupling and nitroarene hydrogenation reactions. The hydrophilicity of the hybrid polymer network facilitates the reagent mass transfer throughout the bulk of the catalyst particles. Through rapid automated exploration of the continuous and discrete parameters, as well as substrate scope screening, we identified optimal hybrid network-supported Pd catalyst composition and process parameters for Suzuki–Miyaura cross-coupling reactions of aryl bromides with steady-state yields up to 92% with a nominal residence time of 20 min. The developed heterogeneous catalytic system exhibits high activity and mechanical stability with no detectable Pd leaching at reaction temperatures up to 95 °C. Additionally, the versatility of the hybrid network-supported Pd catalyst is demonstrated by successfully performing continuous nitroarene hydrogenation with short residence times (<5 min) at room temperature. Room temperature hydrogenation yields of >99% were achieved in under 2 min nominal residence times with no leaching and catalyst deactivation for more than 20 h continuous time on stream. This catalytic system shows its industrial utility with significantly improved reaction yields of challenging substrates and its utility of environmentally-friendly solvent mixtures, high reusability, scalable and cost-effective synthesis, and multi-reaction successes.}, journal={JACS AU}, author={Davis, Bradley A. and Genzer, Jan and Efimenko, Kirill and Abolhasani, Milad}, year={2023}, month={Jul} } @article{im_frey_lacks_genzer_dickey_2023, title={Enhanced Triboelectric Charge Stability by Air-Stable Radicals}, volume={9}, ISSN={["2198-3844"]}, url={https://doi.org/10.1002/advs.202304459}, DOI={10.1002/advs.202304459}, abstractNote={AbstractThis paper demonstrates that air‐stable radicals enhance the stability of triboelectric charge on surfaces. While charge on surfaces is often undesirable (e.g., static discharge), improved charge retention can benefit specific applications such as air filtration. Here, it is shown that self‐assembled monolayers (SAMs) containing air‐stable radicals, 2,2,6,6‐tetramethylpiperidin‐1‐yl)oxidanyl (TEMPO), hold the charge longer than those without TEMPO. Charging and retention are monitored by Kelvin Probe Force Microscopy (KPFM) as a function of time. Without the radicals on the surface, charge retention increases with the water contact angle (hydrophobicity), consistent with the understanding that surface water molecules can accelerate charge dissipation. Yet, the most prolonged charge retention is observed in surfaces treated with TEMPO, which are more hydrophilic than untreated control surfaces. The charge retention decreases with reducing radical density by etching the TEMPO‐silane with tetrabutylammonium fluoride (TBAF) or scavenging the radicals with ascorbic acid. These results suggest a pathway toward increasing the lifetime of triboelectric charges, which may enhance air filtration, improve tribocharging for patterning charges on surfaces, or boost triboelectric energy harvesting.}, journal={ADVANCED SCIENCE}, author={Im, Sooik and Frey, Ethan and Lacks, Daniel J. and Genzer, Jan and Dickey, Michael D.}, year={2023}, month={Sep} } @article{frey_im_bachmann_genzer_dickey_2023, title={Patterning of a High Surface Area Liquid Metal-Carbon Composite Film Using Laser Processing}, volume={9}, ISSN={["1616-3028"]}, url={https://doi.org/10.1002/adfm.202308574}, DOI={10.1002/adfm.202308574}, abstractNote={AbstractLiquid metal is a compelling material for making soft and stretchable devices due to its high electrical conductivity and extreme stretchability. One way to pattern liquid metal is to nebulize it into small droplets, spray it onto a surface as a film, and then use a laser to “sinter” it into circuit patterns. Here, it is shown that including poly(amic acid) in the spray‐deposited film has multiple benefits: it (1) allows the unsintered regions to be removed easily, (2) lowers the power required for sintering, (3) converts to carbon upon exposure to create a carbon‐metal composite, and (4) increases the surface area of the film by 2632% compared to bulk EGaIn. The conductive liquid metal‐carbon circuits can also be transferred to a soft substrate to produce stretchable conductors. The circuits slightly increase in conductivity up to ≈30% strain and then decrease such that by 100% strain, the resistance is only ≈1.02 times its initial resistance. Lastly, the film is highly reactive with water molecules in the air, increasing in resistance over time in humid conditions. The high reactivity and surface area of the film indicate potential applications in batteries, catalysts, and capacitors. Meanwhile, the facile patterning indicates potential applications in soft circuits.}, journal={ADVANCED FUNCTIONAL MATERIALS}, author={Frey, Ethan J. and Im, Sooik and Bachmann, Adam L. and Genzer, Jan and Dickey, Michael D.}, year={2023}, month={Sep} } @article{machikiti_pourdeyhimi_genzer_efimenko_2023, title={Tuning Interfacial Adhesion in Polyester/Polyamide Systems}, volume={7}, ISSN={["1520-5045"]}, url={https://doi.org/10.1021/acs.iecr.3c01517}, DOI={10.1021/acs.iecr.3c01517}, abstractNote={Polymer–polymer adhesion is critical in polymer processing and application areas where lamination, welding, composites, blending, and coextrusion are involved. Polyethylene terephthalate/polyamide (PET/PA) pairs are widely used to produce bicomponent fibers. The strong adhesion due to the formation of chemical bonds in the interfacial region limits post-production fiber processing. The adhesion strength in the PET/PA systems showed that PET/PA66 had the highest adhesion energy, followed by PET/PA6, PET/PA11, and PET/PA12 for all processing conditions. We developed a method that enables adhesion control by introducing a poly(octadecene-alt-maleic anhydride) (POMA) alternating copolymer by either direct interfacial modification or addition into the PET phase. Upon POMA introduction to the system, the reduction in interfacial adhesion strength between PET and polyamides is observed. We established the relationship between adhesion strengths in PET/polyamide systems, processing conditions, and concentration of the POMA modifier.}, journal={INDUSTRIAL & ENGINEERING CHEMISTRY RESEARCH}, author={Machikiti, Zvikomborero and Pourdeyhimi, Behnam and Genzer, Jan and Efimenko, Kirill}, year={2023}, month={Jul} } @article{kannan_genzer_2023, title={UV- and thermally-active small bi-functional gelator for creating gradient polymer network coatings}, volume={18}, ISSN={["1559-4106"]}, DOI={10.1116/6.0002268}, abstractNote={We present a versatile one-pot synthesis method for creating surface-anchored orthogonal gradient networks using a small bi-functional gelator, 4-azidosulfonylphenethyltrimethoxysilane (4-ASPTMS). The sulfonyl azide (SAz) group of 4-ASPTMS is UV (≤254 nm) and thermally active (≥100 °C) and, thus, enables us to vary the cross-link density in orthogonal directions by controlling the activation of SAz groups via UV and temperature means. We deposit a thin layer (∼200 nm) of a mixture comprising ∼90% precursor polymer and ∼10% 4-ASPTMS in a silicon wafer. Upon UV irradiation or annealing the layers, SAz releases nitrogen by forming singlet and triplet nitrenes that concurrently react with any C–H bond in the vicinity leading to sulfonamide cross-links. Condensation among trimethoxy groups in the bulk connects 4-ASPTMS units and completes the cross-linking. Simultaneously, 4-ASPTMS near the substrate reacts with surface-bound –OH motifs that anchor the cross-linked polymer chains to the substrate. We demonstrate the generation of orthogonal gradient network coatings exhibiting cross-link density (or stiffness) gradients in orthogonal directions using such a simple process.}, number={1}, journal={BIOINTERPHASES}, author={Kannan, Pandiyarajan Chinnayan and Genzer, Jan}, year={2023}, month={Jan} } @article{machikiti_pourdeyhimi_genzer_efimenko_2022, title={Controlling PA6/PET adhesion to facilitate interfacial fracture}, volume={171}, ISSN={["1873-1945"]}, DOI={10.1016/j.eurpolymj.2022.111196}, abstractNote={Microfibers get often produced in the form of bicomponent polymer systems. The materials of choice are Nylon 6 (PA6) and poly(ethylene terephthalate) (PET). This combination of PA6 and PET is preferable because of its beneficial attributes (i.e., thermal stability, mechanical strength, etc.). PA6 and PET exhibit high adhesion when processed at elevated temperatures due to chemical bonds formation by aminolysis of the ester group in PET with a secondary amine in PA6. These fibers are split/fibrillated by mechanical energy (hydroentangling or needle punching). For energy input, it is desirable to have adhesion between the PA6 and PET materials that is not too strong to allow for easy polymer splitting. Therefore, we developed a method for tailoring the PA6/PET interface adhesion by adding modifiers that react preferentially with the PA6 component. The reactivity between PA6 and PET was investigated by spin coating thin films of PA6 and PET on silicon wafers and annealing them at high temperatures. The reaction between PET and small molecules containing secondary amines (i.e., caprolactam, diallyamine, diethylamine, and diisopropylamine) shows a chemical bond between the ester group in PET and the secondary amine group. The poly(styrene-alt-maleic anhydride) (PSMA) and poly(octadecene-alt-maleic anhydride) (POMA) were chosen as model polymer interfacial modifiers. The feasibility of modifying secondary amines is examined by reacting the two modifiers, PSMA and POMA, with small molecules containing secondary amine groups. PA6 and PET display high fracture toughness (i.e., adhesion strength) at elevated temperatures and longer annealing times because of strong interactions between the amine and ester groups in PA6 and PET, respectively. We then assess the adhesion strength between PA6 and PET modified with PSMA and POMA. Both modifiers reduce interfacial adhesion strength between PA6 and PET. Therefore, it is feasible to tailor adhesion at the PA6/PET interface, which could prove helpful in microfibers production.}, journal={EUROPEAN POLYMER JOURNAL}, author={Machikiti, Zvikomborero and Pourdeyhimi, Behnam and Genzer, Jan and Efimenko, Kirill}, year={2022}, month={May} } @article{mu_genzer_gorman_2022, title={Degradable Anti-Biofouling Polyester Coatings with Controllable Lifetimes}, volume={1}, ISSN={["0743-7463"]}, url={https://doi.org/10.1021/acs.langmuir.1c02822}, DOI={10.1021/acs.langmuir.1c02822}, abstractNote={To achieve degradable, anti-biofouling coatings with longer lifetimes and better mechanical properties, we synthesized a series of degradable co-polyesters composed of cyclic ketene acetals, di-(ethylene glycol) methyl ether methacrylate, and a photoactive curing agent, 4-benzoylphenyl methacrylate, using a radical ring-opening polymerization. The precursor co-polyesters were spin-coated on a benzophenone-functionalized silicon wafer to form ca. 60 nm films and drop-casted on glass to form ∼32 μm films. The copolymers were cross-linked via UV irradiation at 365 nm. The degradation of films was studied by immersing the specimens in aqueous buffers of different pH values. The results show that both the pH of buffer solutions and gel fractions of networks affect the degradation rate. The coatings show good bovine serum albumin resistance capability. By adjusting the fractions of monomers, the degradation rate and degree of hydration (e.g., swelling ratio) are controllable.}, number={4}, journal={LANGMUIR}, publisher={American Chemical Society (ACS)}, author={Mu, Gaoyan and Genzer, Jan and Gorman, Christopher B.}, year={2022}, month={Jan} } @article{davis_bennett_genzer_efimenko_abolhasani_2022, title={Intensified Hydrogenation in Flow Using a Poly(beta-cyclodextrin) Network-Supported Catalyst}, volume={11}, ISSN={["2168-0485"]}, url={https://doi.org/10.1021/acssuschemeng.2c05467}, DOI={10.1021/acssuschemeng.2c05467}, abstractNote={The intersection of heterogeneous catalysis and flow chemistry is of great importance for the emerging distributed manufacturing of specialty chemicals. Specifically, continuous production of aryl amines is an essential step for on-demand and on-site manufacturing of fine chemicals. This work presents a heterogeneous flow chemistry route for accelerated chemoselective hydrogenation of nitroarenes using a poly(β-cyclodextrin) network-supported palladium catalyst. The developed packed-bed flow reactor enables the selective hydrogenation of a rationally selected library of nitroarenes with >99% yield at room temperature and short residence times (1 min). Utilizing sodium borohydride as the hydrogen carrier in a pressurized packed-bed flow reactor allows safe and efficient delivery of hydrogen to nitroarene molecules. We demonstrate the robustness and versatility of the flow reactor packed with the network-supported catalyst through its consistently high reaction yield over a 3 day run and its reusability and stability in several solvent mixtures with a single-reactor aryl amine manufacturing throughput of up to 31.5 g/day. Furthermore, the catalytic packed-bed reactor is used in a case study for a two-step telescopic synthesis of a critical intermediate for the antibacterial drug linezolid, further supporting its utility as an industrially relevant catalyst for the broad application of catalytic hydrogenations in flow.}, journal={ACS SUSTAINABLE CHEMISTRY & ENGINEERING}, author={Davis, Bradley A. and Bennett, Jeffrey A. and Genzer, Jan and Efimenko, Kirill and Abolhasani, Milad}, year={2022}, month={Nov} } @article{ramesh_davis_roros_zhou_he_gao_menegatti_khan_genzer_2022, title={Nonwoven Membranes with Infrared Light-Controlled Permeability}, volume={9}, ISSN={["1944-8252"]}, DOI={10.1021/acsami.2c13280}, abstractNote={This study presents the development of the first composite nonwoven fiber mats (NWFs) with infrared light-controlled permeability. The membranes were prepared by coating polypropylene NWFs with a photothermal layer of poly(N-isopropylacrylamide) (PNIPAm)-based microgels impregnated with graphene oxide nanoparticles (GONPs). This design enables "photothermal smart-gating" using light dosage as remote control of the membrane's permeability to electrolytes. Upon exposure to infrared light, the GONPs trigger a rapid local increase in temperature, which contracts the PNIPAm-based microgels lodged in the pore space of the NWFs. The contraction of the microgels can be reverted by cooling from the surrounding aqueous environment. The efficient conversion of infrared light into localized heat by GONPs coupled with the phase transition of the microgels above the lower critical solution temperature (LCST) of PNIPAm provide effective control over the effective porosity, and thus the permeability, of the membrane. The material design parameters, namely the monomer composition of the microgels and the GONP-to-microgel ratio, enable tuning the permeability shift in response to IR light; control NWFs coated with GONP-free microgels displayed thermal responsiveness only, whereas native NWFs showed no smart-gating behavior at all. This technology shows potential toward processing temperature-sensitive bioactive ingredients or remote-controlled bioreactors.}, journal={ACS APPLIED MATERIALS & INTERFACES}, author={Ramesh, Srivatsan and Davis, Jack and Roros, Alexandra and Zhou, Chuanzhen and He, Nanfei and Gao, Wei and Menegatti, Stefano and Khan, Saad and Genzer, Jan}, year={2022}, month={Sep} } @article{rittschof_orihuela_genzer_efimenko_2022, title={PDMS networks meet barnacles: a complex and often toxic relationship}, volume={11}, ISSN={["1029-2454"]}, url={https://doi.org/10.1080/08927014.2022.2145471}, DOI={10.1080/08927014.2022.2145471}, abstractNote={Abstract The biological impact of chemical formulations used in various coating applications is essential in guiding the development of new materials that directly contact living organisms. To illustrate this point, an investigation addressing the impact of chemical compositions of polydimethylsiloxane networks on a common platform for foul-release biofouling management coatings was conducted. The acute toxicity of network components to barnacle larvae, the impacts of aqueous extracts of crosslinker, silicones and organometallic catalyst on trypsin enzymatic activity, and the impact of assembled networks on barnacle adhesion was evaluated. The outcomes of the study indicate that all components used in the formulation of the silicone network alter trypsin enzymatic activity and have a range of acute toxicity to barnacle larvae. Also, the adhesion strength of barnacles attached to PDMS networks correlates to the network formulation protocol. This information can be used to assess action mechanisms and risk–benefit analysis of PDMS networks.}, journal={BIOFOULING}, author={Rittschof, Daniel and Orihuela, Beatriz and Genzer, Jan and Efimenko, Kirill}, year={2022}, month={Nov} } @article{barbieri_cutright_ramesh_khan_efimenko_genzer_menegatti_2022, title={Potent Antibacterial Composite Nonwovens Functionalized with Bioactive Peptides and Polymers}, volume={8}, ISSN={["2196-7350"]}, url={https://doi.org/10.1002/admi.202201061}, DOI={10.1002/admi.202201061}, abstractNote={AbstractThis study presents a set of strategies for producing potent antibacterial fabrics by functionalizing nonwoven fabrics (NWFs) with antimicrobial peptides and polymers (AMPs). The incorporation of AMPs is initially optimized on 2D substrates by evaluating conjugation on a poly(maleic anhydride) copolymer coating versus adsorption on polycationic/anionic films and microgels. The evaluation of the resulting surfaces against S. aureus and E. coli highlights the superior antibacterial activity of poly‐ionic films loaded with daptomycin and polymyxin B as well as microgels featuring controlled release of bacitracin and polymyxin B. These formulations are translated onto spun‐bond polypropylene and poly(ethylene terephthalate) NWFs. The poly‐ionic coatings are either covalently anchored or physically adsorbed onto the surface of the fibers, while the microgels and antibacterial polymers are adsorbed and photo‐crosslinked thereon using a ultraviolet (UV)‐crosslinkable benzophenone‐based polymer. Selected formulations loaded with bacitracin and polymyxin B afford a 105‐fold reduction of Staphylococcus aureus (S. aureus) and Escherichia coli (E. coli) in artificial sweat, respectively, on par with commercial antibacterial NWFs. The proposed antibacterial fabric, however, outperforms its commercial counterparts in terms of biocompatibility, showing virtually no adverse effect on human epidermal keratinocytes. Collectively, these results demonstrate affordable and scalable routes for developing antimicrobial NWFs that efficiently eliminate resilient pathogenic bacteria.}, journal={ADVANCED MATERIALS INTERFACES}, author={Barbieri, Eduardo and Cutright, Camden C. and Ramesh, Srivatsan and Khan, Saad A. and Efimenko, Kirill and Genzer, Jan and Menegatti, Stefano}, year={2022}, month={Aug} } @misc{ramesh_khan_park_ford_menegatti_genzer_2022, title={Self-healing and repair of fabrics: A comprehensive review of the application toolkit}, volume={54}, ISSN={["1873-4103"]}, DOI={10.1016/j.mattod.2021.11.016}, abstractNote={Self-healing fabrics respond to chemical and physical damage by restoring functional, structural, and morphological features. We present a comprehensive review of textile hybrids or composites capable of self-healing and repairing fabrics against damages across the micro- (µm), meso- (µm – mm), and macro-scale (>mm). The reviewed literature is organized in three sections presenting (i) the chemistry and fabrication principles of designing self-healing fabrics against increasing size scales of repair, (ii) stimuli-driven and autonomous healing, and (iii) the methods to characterize the recovery of wettability, barrier, morphological, mechanical, and other properties. The discussion of mainstream methods for developing self-healing fabrics focuses on coatings, composites, and specialized fabrication techniques required as the damage size grows from µm to mm to >mm. The section on stimuli-driven repair and autonomous recovery discusses the time scales associated with different damage repair, showing how external stimuli provide a higher driving force towards healing and accelerate material restoration than autonomous recovery. Finally, an array of optical, mechanical, and functional characterization techniques is discussed to evaluate the recovery yield and understand the repair mechanisms of the various fabrics. This review demonstrates the virtually limitless uses of next-generation self-healing systems, from separations to protective clothing, anti-fouling, and self-cleaning.}, journal={MATERIALS TODAY}, author={Ramesh, Srivatsan and Khan, Saad and Park, Yaewon and Ford, Ericka and Menegatti, Stefano and Genzer, Jan}, year={2022}, month={Apr}, pages={90–109} } @misc{shaw_kuriakose_cheeseman_dickey_genzer_christofferson_crawford_mcconville_chapman_truong_et al._2021, title={Antipathogenic properties and applications of low-dimensional materials}, volume={12}, ISSN={["2041-1723"]}, url={https://doi.org/10.1038/s41467-021-23278-7}, DOI={10.1038/s41467-021-23278-7}, abstractNote={AbstractA major health concern of the 21st century is the rise of multi-drug resistant pathogenic microbial species. Recent technological advancements have led to considerable opportunities for low-dimensional materials (LDMs) as potential next-generation antimicrobials. LDMs have demonstrated antimicrobial behaviour towards a variety of pathogenic bacterial and fungal cells, due to their unique physicochemical properties. This review provides a critical assessment of current LDMs that have exhibited antimicrobial behaviour and their mechanism of action. Future design considerations and constraints in deploying LDMs for antimicrobial applications are discussed. It is envisioned that this review will guide future design parameters for LDM-based antimicrobial applications.}, number={1}, journal={NATURE COMMUNICATIONS}, author={Shaw, Z. L. and Kuriakose, Sruthi and Cheeseman, Samuel and Dickey, Michael D. and Genzer, Jan and Christofferson, Andrew J. and Crawford, Russell J. and McConville, Chris F. and Chapman, James and Truong, Vi Khanh and et al.}, year={2021}, month={Jun} } @article{bennett_davis_ramezani_genzer_efimenko_abolhasani_2021, title={Continuous Ligand-Free Suzuki-Miyaura Cross-Coupling Reactions in a Cartridge Flow Reactor Using a Gel-Supported Catalyst}, volume={60}, ISSN={["0888-5885"]}, url={https://doi.org/10.1021/acs.iecr.1c01531}, DOI={10.1021/acs.iecr.1c01531}, abstractNote={The Suzuki–Miyaura cross-coupling reaction is one of the most important reactions for pharmaceutical and fine chemical synthesis, performed using both homogeneous and heterogeneous catalysis. In this work, we cross-link poly(methylhydrosiloxane) (PMHS) with tri(ethylene glycol divinyl ether) to create a versatile and readily accessible gel catalyst support for Suzuki–Miyaura cross-coupling reactions in a pseudoheterogeneous manner. The Si–H units present on the PMHS backbone act dually as the cross-linking site and the reducing agent to anchor and reduce palladium(II) acetate to active palladium(0). The PMHS-supported Pd catalyst is then packed into a stainless-steel flow reactor to create a cartridgelike reactor for the continuous operation of a model Suzuki–Miyaura cross-coupling reaction. We systematically investigate the role of reaction temperature, catalyst loading, cross-linking density, and gel particle size on the transient and steady-state behavior of the cartridge flow reactor through an automated flow chemistry platform. The PMHS-supported catalytic particles demonstrate minimal deactivation and leaching over a continuous (80 h) Suzuki–Miyaura cross-coupling reaction at a 30 min nominal residence time at a relatively high reaction temperature of 95 °C. The developed modular flow chemistry strategy equipped with the cartridge flow reactor enables accelerated studies of the fundamental and applied characteristics of gel-supported catalysts while providing increased safety, higher throughput, and removal of the separation step needed for catalyst recovery compared to homogeneous cross-coupling reactions in batch reactors.}, number={26}, journal={INDUSTRIAL & ENGINEERING CHEMISTRY RESEARCH}, publisher={American Chemical Society (ACS)}, author={Bennett, Jeffrey A. and Davis, Bradley A. and Ramezani, Mahdi and Genzer, Jan and Efimenko, Kirill and Abolhasani, Milad}, year={2021}, month={Jul}, pages={9418–9428} } @article{ko_truong_woo_dickey_hsiao_genzer_2021, title={Counterpropagating Gradients of Antibacterial and Antifouling Polymer Brushes}, volume={12}, ISSN={["1526-4602"]}, url={https://doi.org/10.1021/acs.biomac.1c01386}, DOI={10.1021/acs.biomac.1c01386}, abstractNote={We report on the formation of counterpropagating density gradients in poly([2-dimethylaminoethyl] methacrylate) (PDMAEMA) brushes featuring spatially varying quaternized and betainized units. Starting with PDMAEMA brushes with constant grafting density and degree of polymerization, we first generate a density gradient of quaternized units by directional vapor reaction involving methyl iodide. The unreacted DMAEMA units are then betainized through gaseous-phase betainization with 1,3-propanesultone. The gas reaction of PDMAEMA with 1,3-propanesultone eliminates the formation of byproducts present during the liquid-phase modification. We use the counterpropagating density gradients of quaternized and betainized PDMAEMA brushes in antibacterial and antifouling studies. Completely quaternized and betainized brushes exhibit antibacterial and antifouling behaviors. Samples containing 12% of quaternized and 85% of betainized units act simultaneously as antibacterial and antifouling surfaces.}, journal={BIOMACROMOLECULES}, publisher={American Chemical Society (ACS)}, author={Ko, Yeongun and Truong, Vi Khanh and Woo, Sun Young and Dickey, Michael D. and Hsiao, Lilian and Genzer, Jan}, year={2021}, month={Dec} } @article{clark_thacker_mcgill_miles_westmoreland_efimenko_genzer_santiso_2021, title={DFT Analysis of Organotin Catalytic Mechanisms in Dehydration Esterification Reactions for Terephthalic Acid and 2,2,4,4-Tetramethyl-1,3-cyclobutanediol}, volume={125}, ISSN={["1520-5215"]}, url={https://doi.org/10.1021/acs.jpca.1c00850}, DOI={10.1021/acs.jpca.1c00850}, abstractNote={Polyesters synthesized from 2,2,4,4-tetramethyl-1,3-cyclobutanediol (TMCD) and terephthalic acid (TPA) are improved alternatives to toxic polycarbonates based on bisphenol A. In this work, we use ωB97X-D/LANL2DZdp calculations, in the presence of a benzaldehyde polarizable continuum model solvent, to show that esterification of TMCD and TPA will reduce and subsequently dehydrate a dimethyl tin oxide catalyst, becoming ligands on the now four-coordinate complex. This reaction then proceeds most plausibly by an intramolecular acyl-transfer mechanism from the tin complex, aided by a coordinated proton donor such as hydronium. These findings are a key first step in understanding polyester synthesis and avoiding undesirable side reactions during production.}, number={23}, journal={JOURNAL OF PHYSICAL CHEMISTRY A}, publisher={American Chemical Society (ACS)}, author={Clark, Jennifer A. and Thacker, Pranav J. and McGill, Charles J. and Miles, Jason R. and Westmoreland, Phillip R. and Efimenko, Kirill and Genzer, Jan and Santiso, Erik E.}, year={2021}, month={Jun}, pages={4943–4956} } @article{castillo_dickey_gorman_genzer_efimenko_2021, title={Deposition of silicate coatings on poly(ethylene terephthalate) for improved scratch and solvent resistance}, volume={10}, ISSN={["1097-4628"]}, url={https://doi.org/10.1002/app.51800}, DOI={10.1002/app.51800}, abstractNote={AbstractThe surface of poly(ethylene terephthalate) (PET) films crystallizes when exposed to a broad class of organic solvents (protic polar, aprotic polar, and non‐polar). This phenomenon leads to irreversible changes in transparency and mechanical properties. It also limits PET films and coatings in applications including electronics, medicine, packaging, and construction. This work describes a method to impart solvent resistance to the PET surface by depositing a silicate film. This coating enables many post‐functionalization options, including those that tune wettability or impart antifouling characteristics to the parent PET surface. It also endows PET surface resistance to organic solvent exposure without affecting the overall mechanical and optical properties of the PET. Additionally, the silicate layer improves the scratch resistance of PET surfaces.}, journal={JOURNAL OF APPLIED POLYMER SCIENCE}, publisher={Wiley}, author={Castillo, Gilbert A. and Dickey, Michael D. and Gorman, Christopher B. and Genzer, Jan and Efimenko, Kirill}, year={2021}, month={Oct} } @article{poblete_mondal_ma_dickey_genzer_zhu_2022, title={Direct measurement of rate-dependent mode I and mode II traction-separation laws for cohesive zone modeling of laminated glass}, volume={279}, ISSN={["1879-1085"]}, DOI={10.1016/j.compstruct.2021.114759}, abstractNote={This paper reports a combined experimental-modeling study on the adhesion at glass/poly(vinyl butyral) (PVB) interfaces. PVB is a critical component in laminated glasses that provides impact resistance and prevents spallation in the event of glass breakage. We characterize the mode I (normal) and mode II (shear) interfacial fracture behaviors in glass/PVB/glass laminates and obtained independent cohesive (traction–separation) laws for the two modes. We observe a pronounced rate dependence of both traction-separation laws. Specifically, with increasing loading rate, the interfacial stiffness, the peak stress, and the fracture toughness increase while the critical opening or shear displacement decreases. These measured traction-separation laws are used as inputs in finite element analysis to predict the mechanical behavior of the peel test, which agrees reasonably well with the experimental results. Finite element analysis of a glass laminate subjected to impact loading demonstrates how the measured interfacial properties can be used to predict the mechanical behavior and failure of laminated glass.}, journal={COMPOSITE STRUCTURES}, author={Poblete, Felipe R. and Mondal, Kunal and Ma, Yinong and Dickey, Michael D. and Genzer, Jan and Zhu, Yong}, year={2022}, month={Jan} } @article{ramesh_davis_roros_eiben_fabiani_smith_reynolds_pourdeyhimi_khan_genzer_et al._2021, title={Dual-Responsive Microgels for Structural Repair and Recovery of Nonwoven Membranes for Liquid Filtration}, volume={3}, ISSN={["2637-6105"]}, url={https://doi.org/10.1021/acsapm.0c01360}, DOI={10.1021/acsapm.0c01360}, abstractNote={This study presents dual-responsive colloidal microgels to repair nonwoven fiber mats (NWFs) and recover their native morphological and functional properties. The formulation comprises poly(N-isopr...}, number={3}, journal={ACS APPLIED POLYMER MATERIALS}, publisher={American Chemical Society (ACS)}, author={Ramesh, Srivatsan and Davis, Jack and Roros, Alexandra and Eiben, Justin and Fabiani, Thomas and Smith, Ryan and Reynolds, Lewis and Pourdeyhimi, Behnam and Khan, Saad and Genzer, Jan and et al.}, year={2021}, month={Mar}, pages={1508–1517} } @article{mu_pandiyarajan_lu_weaver_genzer_gorman_2021, title={Dynamic Surfaces-Degradable Polyester Networks that Resist Protein Adsorption}, volume={37}, ISSN={["0743-7463"]}, url={https://doi.org/10.1021/acs.langmuir.1c00890}, DOI={10.1021/acs.langmuir.1c00890}, abstractNote={We synthesized a series of novel degradable alternating copolyesters composed of diglycolic anhydride (DGA) and two epoxides, epoxymethoxytriethylene glycol (ETEG) and a photoactive crosslinking agent epoxy benzophenone (EBP). After UV crosslinking, soaking the films in a good solvent (tetrahydrofuran) removed uncrosslinked material, and the resulting film gel fractions were calculated. These network films were then degraded in buffer solutions of varying pH values. The degradation of networks with lower gel fraction (fewer crosslinks) was faster and followed first-order kinetics. In contrast, the denser network degraded slower and followed zeroth-order kinetics. The lower gel fraction networks possess a higher swelling ratio and resist bovine serum albumin (BSA) adsorption better by entropic shielding and faster degradation. In comparison, higher gel fraction networks with higher EBP mole fractions adsorb more BSA due to hydrophobic interactions and slower degradation.}, number={30}, journal={LANGMUIR}, publisher={American Chemical Society (ACS)}, author={Mu, Gaoyan and Pandiyarajan, C. K. and Lu, Xiuyuan and Weaver, Matt and Genzer, Jan and Gorman, Christopher B.}, year={2021}, month={Aug}, pages={8978–8988} } @article{zboray_efimenko_jones_genzer_2021, title={Functional Gels Containing Hydroxamic Acid Degrade Organophosphates in Aqueous Solutions}, volume={60}, ISSN={["0888-5885"]}, url={https://doi.org/10.1021/acs.iecr.1c01374}, DOI={10.1021/acs.iecr.1c01374}, abstractNote={We synthesized poly(maleic anhydride-co-methyl vinyl ether) (PMAMVE) gels and functionalized them to form hydroxamic acid functional groups. We evaluated the performance of the gels in decomposing dimethyl nitrophenyl phosphate (DMNP). We monitored organophosphate degradation kinetics as a function of gel chemical composition, cross-linking density, and solution pH and applied Thiele modulus analysis to determine the importance of transport phenomena related to the particulate gel size. The decomposition of DMNP in the maleic anhydride gels followed pseudo-first-order kinetics for all studied conditions. The performance was influenced by the spatial confinement of the hydroxamic acid groups inside the gel. The gels made of PMAMVE copolymers modified with hydroxamic acid offer a robust new system with high degradation efficiency, scalability, and preparation simplicity.}, number={24}, journal={INDUSTRIAL & ENGINEERING CHEMISTRY RESEARCH}, publisher={American Chemical Society (ACS)}, author={Zboray, Steven and Efimenko, Kirill and Jones, Jacob L. and Genzer, Jan}, year={2021}, month={Jun}, pages={8799–8811} } @article{liu_islam_ku_boyd_zhong_urbas_smith_derov_nguyen_kim_et al._2021, title={Novel computational design of high refractive index nanocomposites and effective refractive index tuning based on nanoparticle morphology effect}, volume={223}, ISSN={["1879-1069"]}, url={https://doi.org/10.1016/j.compositesb.2021.109128}, DOI={10.1016/j.compositesb.2021.109128}, abstractNote={This study introduces a method to predict the refractive index (RI) of nanocomposites with the Finite Elements Analysis (FEA) based on the Fabry-Pérot interference. The efficacy was verified by comparing the estimated composites' RI with the available data in the literature. In the experimental verification, the FEA-based prediction showed closer results with the measurement as compared to the effective medium approximation (EMA) approaches, which are prevalently used to predict the physical properties of nanocomposites. Due to the modeling capability, the FEA-method could investigate the effect of the nanoparticle morphology (particle size, shape, and orientation) and distribution. Large particle size, particle agglomeration in high electric-field amplitude region, and particle elongation along the light oscillating direction are found to be the major factors to enhance the RI of composites. The underlying mechanism of RI changing is attributed to the light scattering by embedded nanoparticles, which provides one potential real-time RI tuning schematic.}, journal={COMPOSITES PART B-ENGINEERING}, publisher={Elsevier BV}, author={Liu, Sipan and Islam, Md Didarul and Ku, Zahyun and Boyd, Darryl A. and Zhong, Yaxu and Urbas, Augustine M. and Smith, Evan and Derov, John and Nguyen, Vinh Q. and Kim, Woohong and et al.}, year={2021}, month={Oct} } @article{cutright_harris_ramesh_khan_genzer_menegatti_2021, title={Surface-Bound Microgels for Separation, Sensing, and Biomedical Applications}, volume={8}, ISSN={["1616-3028"]}, url={https://doi.org/10.1002/adfm.202104164}, DOI={10.1002/adfm.202104164}, abstractNote={AbstractThis study presents a comprehensive survey of microgel‐coated materials and their functional behavior, describing the complex interplay between the physicochemical and mechanical properties of the microgels and the chemical and morphological features of substrates. The cited literature is articulated in four main sections: i) properties of 2D and 3D substrates, ii) synthesis, modification, and characterization of the microgels, iii) deposition techniques and surface patterning, and iv) application of microgel‐coated surfaces focusing on separations, sensing, and biomedical applications. Each section discusses – by way of principles and examples – how the various design parameters work in concert to deliver functionality to the composite systems. The case studies presented herein are viewed through a multi‐scale lens. At the molecular level, the surface chemistry and the monomer make‐up of the microgels endow responsiveness to environmental and artificial physical and chemical cues. At the micro‐scale, the response effects shifts in size, mechanical, and optical properties, and affinity towards species in the surrounding liquid medium, ranging from small molecules to cells. These phenomena culminate at the macro‐scale in measurable, reversible, and reproducible effects, aiming in a myriad of directions, from lab‐scale to industrial applications.}, number={47}, journal={ADVANCED FUNCTIONAL MATERIALS}, publisher={Wiley}, author={Cutright, Camden C. and Harris, Jacob L. and Ramesh, Srivatsan and Khan, Saad A. and Genzer, Jan and Menegatti, Stefano}, year={2021}, month={Aug} } @article{woo_pandiyarajan_genzer_2021, title={Tuning the Properties of Surface-Anchored Polymer Networks by Varying the Concentration of a Thermally Activated Cross-Linker, Annealing Time, and Temperature in a One-Pot Reaction}, volume={3}, ISSN={["2637-6105"]}, DOI={10.1021/acsapm.1c00890}, abstractNote={We investigate the properties of surface-anchored polymer networks created via one-pot synthesis using thermally active 6-azidosulfonylhexyltriethoxysilane (6-ASHTES). 6-ASHTES is a bifunctional gelator that undergoes cross-linking and surface-anchoring reactions when annealed above 100 °C. We employ a poly(vinylpyrrolidone) (PVP) with different molecular weights (10–1300 kDa) as a model system to examine the effect of 6-ASHTES concentration, annealing time, and annealing temperature on gel formation. A thin film of PVP/6-ASHTES mixture is deposited on a clean silicon wafer and annealed to form network layers. Spectroscopic ellipsometry measures the film thickness of the cross-linked layers from which the gel fraction and swelling ratio are determined. The gel fraction of PVP in the network can be "dialed in" by varying the annealing time, temperature, and concentration of 6-ASHTES in the PVP/6-ASHTES mixture. We use a simple Monte Carlo simulation model to describe cross-linking as a function of cross-linker concentration, reaction rate, reaction time, and polymer length. The trends obtained from the model simulations are in qualitative agreement with the experimental data.}, number={11}, journal={ACS APPLIED POLYMER MATERIALS}, author={Woo, Sun Young and Pandiyarajan, C. K. and Genzer, Jan}, year={2021}, month={Nov}, pages={5568–5577} } @article{pandiyarajan_genzer_2021, title={UV- and Thermally-Active Bifunctional Gelators Create Surface-Anchored Polymer Networks}, volume={6}, ISSN={["1521-3927"]}, DOI={10.1002/marc.202100266}, abstractNote={AbstractA versatile one‐step synthesis of surface‐attached polymer networks using small bifunctional gelators (SBG), namely 4‐azidosulfonylphenethyltrimethoxysilane (4‐ASPTMS) and 6‐azidosulfonylhexyltriethoxysilane (6‐ASHTES) is reported. A thin layer (≈200 nm) of a mixture comprising ≈90% precursor polymer and 10% of 4‐ASPTMS or 10% 6‐ASHTES on a silicon wafer is deposited. Upon UV irradiation (≈l–254 nm) or annealing (>100 °C) layers, sulfonyl azides (SAz) release nitrogen by forming singlet and triplet nitrenes that concurrently react with any C─H bond in the vicinity resulting in sulfonamide crosslinks. Condensation among tri‐alkoxy groups (i.e., methoxy or ethoxy) in bulk connects the SBG units, which completes the crosslinking. Concurrently, when such functionalities react with hydroxyl groups at the surface, which enable the covalent attachment of the crosslinked polymer chains. A systematic investigation on reaction mechanism and gel formation using spectroscopic ellipsometry (SE) and Fourier‐transform infrared spectroscopy in the attenuated total reflection mode (FTIR‐ATR) is performed. Analogous thermally initiated gelation for both 4‐ASPTMS and 6‐ASHTES is found. The 6‐ASHTES is UV inactive at ≈l–254 nm, while the 4‐ASPTMS is active and forms gels. The difference is attributed to the aromatic nature of 4‐ASPTMS that absorb UV light at ≈l–254 nm due to π–π* transition.}, journal={MACROMOLECULAR RAPID COMMUNICATIONS}, author={Pandiyarajan, Chinnayan Kannan and Genzer, Jan}, year={2021}, month={Jun} } @article{ko_christau_klitzing_genzer_2020, title={Charge Density Gradients of Polymer Thin Film by Gaseous Phase Quaternization}, volume={9}, ISSN={["2161-1653"]}, DOI={10.1021/acsmacrolett.9b00930}, abstractNote={We report on the rapid formation of charge density gradients in polymer films by exposing poly([2-dimethylaminoethyl] methacrylate) (PDMAEMA) films resting on flat silica substrates to methyl iodide (i.e., MI, also known as iodomethane) vapors. We adjust the charge gradient by varying the MI concentration in solution and the process time. The thickness of the parent PDMAEMA film does not affect the diffusion of MI through and the reaction kinetics in the films. Instead, the diffusion of MI through the gaseous phase constitutes the limiting step in the overall process.}, number={2}, journal={ACS MACRO LETTERS}, author={Ko, Yeongun and Christau, Stephanie and Klitzing, Regine and Genzer, Jan}, year={2020}, month={Feb}, pages={158–162} } @article{schmitz_wohl_tiemsin_genzer_2020, title={Controlled heating and alignment platform enhances versatility in colloidal probe fabrication}, volume={91}, ISSN={["1089-7623"]}, DOI={10.1063/1.5111387}, abstractNote={A colloidal probe, comprising a colloidal particle attached to an atomic force microscope cantilever, is employed to measure interaction forces between the particle and a surface. It is possible to change or even destroy a particle while attaching it to a cantilever, thus limiting the types of systems to which the colloidal probe technique may be applied. Here, we present the Controlled Heating and Alignment Platform (CHAP) for fabricating colloidal probes without altering the original characteristics of the attached particle. The CHAP applies heat directly to the atomic force microscope chip to rapidly and precisely control the cantilever temperature. It minimizes particle heating and enables control over the viscosity of the thermoplastic adhesive to prevent it from contaminating the particle surface. 3D-printed components made the CHAP compatible with standard optical microscopes and streamlined the fabrication process, while increasing the platform’s versatility. To demonstrate the utility of CHAP, we conducted a case study using a thermoplastic wax adhesive to fabricate colloidal probes bearing polystyrene and silica particles between 0.7 and 40 μm in diameter. We characterized the properties and interactions of the adhesive and particles, as well as the properties of the completed probes, to demonstrate the retention of particle features throughout fabrication. Pull-off tests with CHAP’s probes measured adhesive force values in the expected ranges and demonstrated that particles were firmly attached to the cantilevers.}, number={1}, journal={REVIEW OF SCIENTIFIC INSTRUMENTS}, author={Schmitz, Russell C. and Wohl, Christopher J. and Tiemsin, Pacita I. and Genzer, Jan}, year={2020}, month={Jan} } @article{islam_kim_ko_ku_boyd_smith_nguyen_myers_baker_kim_et al._2020, title={Design of High Efficient Mid‐Wavelength Infrared Polarizer on ORMOCHALC Polymer}, volume={305}, ISSN={1438-7492 1439-2054}, url={http://dx.doi.org/10.1002/mame.202000033}, DOI={10.1002/mame.202000033}, abstractNote={AbstractWhile an organically modified chalcogenide (ORMOCHALC) can be used to fabricate a polymeric mid‐wavelength infrared (MWIR) polarizer with competitive extinction ratio compared to the commercial wire‐grid polarizers, which are composed of fragile inorganic materials, there is still a knowledge gap regarding the systematic design process to obtain high transmission efficiency and extinction ratio. To this end, a computational parameter study for design optimization is conducted with the geometric parameters of the bilayer grating ORMOCHALC polarizer. The computational study shows that the Fabry–Pérot cavity is the primary mechanism that determines the transmission behaviors and the extinction ratio. A bilayer grating design, guided by the parameter study, is realized through the thermal nanoimprint and metal deposition processes. The extinction ratios measured with the Fourier‐transform infrared are 245, 304, and 351 at the wavelength of 3, 4, and 5 µm, respectively. Compared to the state‐of‐the‐art of the polymeric MWIR linear polarizers, the extinction ratio is improved by 1.4 times, and the transmission efficiency is increased by 2.5 times. Theoretical analysis with the multiple‐layer model based on the transfer matrix method predicts a matched transmission behavior with the experiment and a full‐wave electromagnetic simulation.}, number={5}, journal={Macromolecular Materials and Engineering}, publisher={Wiley}, author={Islam, Md Didarul and Kim, Jun Oh and Ko, Yeongun and Ku, Zahyun and Boyd, Darryl A. and Smith, Evan M. and Nguyen, Vinh Q. and Myers, Jason D. and Baker, Colin C. and Kim, Woohong and et al.}, year={2020}, month={Mar}, pages={2000033} } @article{walker_genzer_santiso_2020, title={Effect of Poly(vinyl butyral) Comonomer Sequence on Adhesion to Amorphous Silica: A Coarse-Grained Molecular Dynamics Study}, volume={12}, ISSN={["1944-8252"]}, DOI={10.1021/acsami.0c10747}, abstractNote={Modulating a comonomer sequence, in addition to overall chemical composition, is the key to unlocking the true potential of many existing commercial copolymers. We employ coarse-grained molecular dynamics (MD) simulations to study the behavior of random-blocky poly(vinyl butyral-co-vinyl alcohol) (PVB) melts in contact with an amorphous silica surface, representing the interface found in laminated safety glass. Our two-pronged coarse-graining approach utilizes both macroscopic thermophysical data and all-atom molecular dynamics simulation data. Polymer-polymer nonbonded interactions are described by the fused-sphere SAFT-γ Mie equation of state, while bonded interactions are derived using Boltzmann inversion to match bond and angle distributions from all-atom PVB chains. Spatially-dependent polymer-surface interactions are mapped from a hydroxylated all-atom amorphous silica slab model and all-atom monomers to an external potential acting on the coarse-grained sites. We ran a series of interfacial coarse-grained MD simulations for PVB melts, systematically varying overall chemical composition and block length distribution. We discovered an unexpectedly complex relationship between blockiness parameter and adhesion energy. For intermediate vinyl alcohol (VA) content, adhesion strength to the silica slab was found to be maximal not for diblock copolymers, but rather random-blocky copolymers with a moderately high degree of blockiness. We attribute this to two main factors: (1) changes in morphology, which dramatically alter the number of VA beads interacting with the surface, and (2) a non-negligible contribution of vinyl butyral (VB) monomers to adhesion energy, due to their preference to adsorb to zones with low hydroxyl density on the silica surface.}, number={42}, journal={ACS APPLIED MATERIALS & INTERFACES}, author={Walker, Christopher C. and Genzer, Jan and Santiso, Erik E.}, year={2020}, month={Oct}, pages={47879–47890} } @article{islam_liu_boyd_zhong_nahid_henry_taussig_ko_nguyen_myers_et al._2020, title={Enhanced mid-wavelength infrared refractive index of organically modified chalcogenide (ORMOCHALC) polymer nanocomposites with thermomechanical stability}, volume={108}, ISSN={["1873-1252"]}, url={http://dx.doi.org/10.1016/j.optmat.2020.110197}, DOI={10.1016/j.optmat.2020.110197}, abstractNote={Abstract Organically modified chalcogenide (ORMOCHALC) polymers have proven to be alternatives to the conventional inorganic materials for mid-wavelength infrared (MWIR, λ = 3–5 μm) optical components. While the refractive index of ORMOCHALC can be reinforced by the content of chalcogenides such as sulfur (S) and selenium (Se), the increased portion of the S or Se deteriorate the thermomechanical stabilities. As a remedy, this study utilizes ZnS nanoparticles to reinforce both optical and thermomechanical properties of the sulfur-based ORMOCHALC polymer, poly(S-random-1,3-diisopropenylbenzene). The refractive index n and extinction coefficient k of the nanocomposites were characterized by Infrared Variable Angle Spectroscopic Ellipsometry (IR-VASE). The results show a significant increment in the refractive index of Δn = 6.58% at the wavelength of 4 μm by adding 20 wt% ZnS (or 7.29 vol%) in the ORMOCHALC polymer. The low extinction coefficient of the nanocomposites (}, journal={OPTICAL MATERIALS}, author={Islam, Md Didarul and Liu, Sipan and Boyd, Darryl A. and Zhong, Yaxu and Nahid, Masrur Morshed and Henry, Reece and Taussig, Laine and Ko, Yeongun and Nguyen, Vinh Q. and Myers, Jason D. and et al.}, year={2020}, month={Oct} } @article{walker_genzer_santiso_2020, title={Extending the fused-sphere SAFT-gamma Mie force field parameterization approach to poly(vinyl butyral) copolymers}, volume={152}, ISSN={["1089-7690"]}, DOI={10.1063/1.5126213}, abstractNote={SAFT-γ Mie, a molecular group-contribution equation of state with foundations in the statistical associating fluid theory framework, is a promising means for developing accurate and transferable coarse-grained force fields for complex polymer systems. We recently presented a new approach for incorporating bonded potentials derived from all-atom molecular dynamics simulations into fused-sphere SAFT-γ Mie homopolymer chains by means of a shape factor parameter, which allows for bond distances less than the tangent-sphere value required in conventional SAFT-γ Mie force fields. In this study, we explore the application of the fused-sphere SAFT-γ Mie approach to copolymers. In particular, we demonstrate its capabilities at modeling poly(vinyl alcohol-co-vinyl butyral) (PVB), an important commercial copolymer widely used as an interlayer in laminated safety glass applications. We found that shape factors determined from poly(vinyl alcohol) and poly(vinyl butyral) homopolymers do not in general correctly reproduce random copolymer densities when standard SAFT-γ Mie mixing rules are applied. However, shape factors optimized to reproduce the density of a random copolymer of intermediate composition resulted in a model that accurately represents density across a wide range of chemical compositions. Our PVB model reproduced copolymer glass transition temperature in agreement with experimental data, but heat capacity was underpredicted. Finally, we demonstrate that atomistic details may be inserted into equilibrated fused-sphere SAFT-γ Mie copolymer melts through a geometric reverse-mapping algorithm.}, number={4}, journal={JOURNAL OF CHEMICAL PHYSICS}, author={Walker, Christopher C. and Genzer, Jan and Santiso, Erik E.}, year={2020}, month={Jan} } @article{cutright_finkelstein_orlowski_mcintosh_brotherton_fabiani_khan_genzer_menegatti_2020, title={Nonwoven fiber mats with thermo-responsive permeability to inorganic and organic electrolytes}, volume={616}, ISSN={["1873-3123"]}, DOI={10.1016/j.memsci.2020.118439}, abstractNote={This study presents the development and characterization of nonwoven fiber mats (NWFs) with stimuli-controlled permeability. An ensemble of membranes was initially constructed by coating the fibers of polypropylene NWFs with a layer of poly ((N-isopropyl acrylamide)-co-(acrylic acid)) (PNIPAm-co-AA) hydrogel. Different coatings were produced by varying the PNIPAm/AA monomer ratio between 3.9 and 18.6. The thermo-responsive layer is expanded at room temperature and contracts when heated above its lower critical solution temperature (LCST). The resulting membranes were first characterized via laser scanning microscopy and fluorescence confocal microscopy to evaluate the thickness and morphology of the hydrogel layer. Microscopy shows uniform coating of the fibers, with a thickness comparable to the fiber diameter, and homogeneous filling of the pore space. The permeability of the NWFs was then evaluated using different solutes, namely an inorganic salt (sodium chloride), an organic acid (citric acid), and an amphiphilic drug (Doxorubicin). These tests consistently show that the flux of the solute is (1) higher at temperatures > LCST, where the hydrogel layer collapses and opens the pore space, and (2) decreases at room temperature ( eCD, the bulk bromination kinetics of PS in CDD is slower than that in CD because of lower solubility of PS in CDD than in CD. In addition, we demonstrate that the reaction rates for brominating PS brushes anchored to flat solid substrates are much slower than those for brominating free PS chains in bulk solution. We attribute this behavior to steric hindrance due to PS confinement on the substrate.}, number={18}, journal={Macromolecules}, author={Jhon, Y. K. and Semler, J. J. and Genzer, Jan}, year={2008}, pages={6719–6727} } @article{gorman_petrie_genzer_2008, title={Effect of substrate geometry on polymer molecular weight and polydispersity during surface-initiated polymerization}, volume={41}, ISSN={["1520-5835"]}, DOI={10.1021/ma8004857}, abstractNote={Poly(methyl methacrylate) (PMMA) anchored chains were grown on porous silicon (p-Si) and anodically etched aluminum oxide (AAO) substrates via surface-initiated atom transfer radical polymerization (ATRP). Using hydrogen fluoride, the chains could be cleaved from the substrates, as evidenced by infrared spectroscopy. The molecular weights and molecular weight distributions of PMMA could be analyzed directly on these substrates (after cleaving the chains from the support) using direct ionization mass spectrometry (DIOS-MS) and matrix-assisted laser desorption ionization mass spectrometry (MALDI-MS). Two principal conclusions were drawn from the study. First, matrix-free DIOS-MS was effective at direct analysis of the polymers up to a molecular weight of ≈6 kDa; the signal-to-noise ratio for heavier polymer chains diminished rapidly. Second, under the same polymerization conditions, PMMA grown on both p-Si and AAO substrates had a much lower molecular weight and a broader molecular weight distribution than ...}, number={13}, journal={MACROMOLECULES}, author={Gorman, Christopher B. and Petrie, Randall J. and Genzer, Jan}, year={2008}, month={Jul}, pages={4856–4865} } @article{crowe-willoughby_genzer_2009, title={Formation and Properties of Responsive Siloxane-Based Polymeric Surfaces with Tunable Surface Reconstruction Kinetics}, volume={19}, ISSN={["1616-3028"]}, DOI={10.1002/adfm.200800622}, abstractNote={AbstractHere, the formation of responsive polymeric materials with tunable response time is reported. These structures are fabricated by chemically modifying poly(vinylmethyl siloxane) (PVMS) networks with alkanethiols bearing a hydrophilic end‐group (COOH or OH). The response time is facilitated by the liquid nature of the PVMS backbone and increases with increase in length of the methylene spacer (CH2)n in the alkanethiol pendent group. While for n = 2 and 6, the surface reconstructs almost instantaneously, specimens with n = 11 resist reconstruction because of strong van der Waals forces, leading to the formation of semi‐crystalline regions. It is demonstrated that the responsive nature of PVMSS(CH2)11OH can be fine‐tuned by varying the temperature; it possesses a faster response at temperatures above the melting point of the S(CH2)11OH moiety.}, number={3}, journal={ADVANCED FUNCTIONAL MATERIALS}, author={Crowe-Willoughby, Julie Ann and Genzer, Jan}, year={2009}, month={Feb}, pages={460–469} } @article{tomlinson_genzer_2008, title={Formation and properties of multivariant assemblies of surface-tethered diblock and triblock copolymers}, volume={49}, ISSN={["0032-3861"]}, DOI={10.1016/j.polymer.2008.08.048}, abstractNote={We present methodologies for fabricating block copolymer assemblies grafted onto flat solid substrates, where each block of the copolymer possesses a systematic and gradual variation of molecular weight as a function of the position on the substrate. We demonstrate the utility of this technique on two case studies. In the first project, we generate surface-tethered poly[(2-hydroxyethyl methacrylate)-b-(methyl methacrylate)] (PHEMA-b-PMMA) diblock copolymer brushes and study systematically morphological transitions associated with collapsing either the top PMMA or the bottom PHEMA block while keeping the other block solvated. Scanning force microscopy studies of systems having the top block collapsed reveal the presence of either flat (F), or micellar (M) or bicontinuous (BC) morphologies, whose locus in the phase diagram agrees with theoretical predictions and results of computer simulations. The second case study demonstrates the extension of the deposition method to the case of surface-anchored triblock copolymer brushes. Specifically, we present results pertaining to the formation of poly[(2-hydroxyethyl methacrylate)-b-(methyl methacrylate)-b-(dimethylaminoethyl methacrylate)] brushes with independent variation of all three block lengths.}, number={22}, journal={POLYMER}, author={Tomlinson, Michael R. and Genzer, Jan}, year={2008}, month={Oct}, pages={4837–4845} } @article{sivaniah_genzer_hexemer_kramer_xiang_li_ober_magonov_2008, title={Nonplanar Surface Organization of Monodendrons in Side-Chain Modified Liquid Crystalline Block Copolymers.}, volume={41}, ISSN={0024-9297 1520-5835}, url={http://dx.doi.org/10.1021/ma8015097}, DOI={10.1021/ma8015097}, abstractNote={Scanning force microscopy (SFM) is used to investigate the surface structures of diblock copolymers of poly(styrene)-block-poly(isoprene) where a portion of the isoprene segments were modified to contain a semifluorinated alkane monodendron side group. The monodendrons consisted of either a single, double or triple attachment of the liquid crystalline mesogen group, [-(CH2)p-(CF2)qF]. Previously we reported that the surfaces of thin films of these materials contained small (∼20−30 nm) dome-like structures that were independent of the coincident block copolymer microphase architecture. We show here that these surface features can also assume a worm-like structure depending on the composition of the monodendron. Furthermore thermal studies of the block copolymer and the isolated monodendron molecule using in situ SFM studies confirm the liquid crystalline origin of the surface structures.}, number={24}, journal={Macromolecules}, publisher={American Chemical Society (ACS)}, author={Sivaniah, E. and Genzer, J. and Hexemer, A. and Kramer, E. J. and Xiang, M. L. and Li, X. F. and Ober, C. K. and Magonov, S.}, year={2008}, month={Dec}, pages={9940–9945} } @article{diamanti_arifuzzaman_elsen_genzer_vaia_2008, title={Reactive patterning via post-functionalization of polymer brushes utilizing disuccinimidyl carbonate activation to couple primary amines}, volume={49}, ISSN={["1873-2291"]}, DOI={10.1016/j.polymer.2008.06.020}, abstractNote={Polymer brushes provide an exceptional route to surface functionalization due to their chemical and mechanical robustness, lack of large-area defects, and high density of functional groups. In spite of these benefits, the synthetic difficulty and complex surface structure associated with polymer brushes have hindered their utilization for constructing multifunctional, patterned surfaces. In this contribution we describe the use of a rapid and highly efficient polymer brush post-functionalization technique as a facile method for controlling surface functionality of polymer brushes. Poly(2-hydroxyethyl methacrylate) (PHEMA) brushes are post-functionalized via activation with N,N′-disuccinimidyl carbonate (DSC) and subsequent coupling to molecules containing α-amine moieties. This post-functionalization effectively tailors surface energy resulting in water contact angles ranging from 40° to 100° using different conjugate molecules. Furthermore, the solvent tolerance, insensitivity to reactant concentration, and rapid reaction time of the aminolysis reaction enable surface energy patterning of the polymer brushes through the use of "reactive" soft lithography. Finally, these surface energy patterns could be "developed" by exposure to gold nanoparticle solutions to yield surfaces with patterned nanoparticle density.}, number={17}, journal={POLYMER}, author={Diamanti, Steve and Arifuzzaman, Shafi and Elsen, Andrea and Genzer, Jan and Vaia, Richard A.}, year={2008}, month={Aug}, pages={3770–3779} } @article{strickland_hall_genzer_2008, title={Simulation of mechanically assembled monolayers and polymers in good solvent using discontinuous molecular dynamics}, volume={41}, ISSN={["0024-9297"]}, DOI={10.1021/ma8003218}, abstractNote={We present the results of discontinuous molecular dynamics simulations of mechanically assembled monolayers in good solvent. Polymers of chain lengths 5−100 were end-grafted to surfaces at low density and then compressed laterally at varying rates. Data for brush thickness and end-monomer density were collected as a function of surface density; they were shown to correspond well with theoretical predictions and simulation results performed at constant surface density. Brush thickness for all chain lengths could be controlled by judicious choice of the compression rate. Defects in the brush layer were dependent on chain length; it was shown that quick compression for relatively short chains allowed the layer no time to relax into coil form. Quick compression on long chain systems led to entanglement in the brush layer since the longer-chained system was not being afforded the long relaxation time required to form a fully relaxed brush. Hysteresis effects were examined by allowing the brush to relax to a lo...}, number={17}, journal={MACROMOLECULES}, author={Strickland, L. Anderson and Hall, Carol K. and Genzer, Jan}, year={2008}, month={Sep}, pages={6573–6581} } @article{ritz_látalová_kříž_genzer_vlček_2008, title={Statistical copolymers of 2-(trimethylsilyloxy)ethyl methacrylate and methyl methacrylate synthesized by ATRP}, volume={46}, ISSN={0887-624X 1099-0518}, url={http://dx.doi.org/10.1002/pola.22436}, DOI={10.1002/pola.22436}, abstractNote={AbstractNo Abtract.}, number={5}, journal={Journal of Polymer Science Part A: Polymer Chemistry}, publisher={Wiley}, author={Ritz, Pavel and Látalová, Petra and Kříž, Jaroslav and Genzer, Jan and Vlček, Petr}, year={2008}, pages={1919–1923} } @misc{genzer_bhat_2008, title={Surface-bound soft matter gradients}, volume={24}, ISSN={["0743-7463"]}, DOI={10.1021/la7033164}, abstractNote={This feature article describes the progress realized over the past half century in the field of surface-bound gradient structures created on or from soft materials (oligomers and/or polymers), or those enabling the study of the behavior of soft materials. By highlighting our work in the field and accounting for the contribution of other groups, we emphasize the exceptional versatility of gradient assemblies in facilitating fast screening of physicochemical phenomena, acting as "recording media" for monitoring a process, and playing a key role in the design and fabrication of surface-bound molecular and macromolecular motors capable of directing a transport phenomenon.}, number={6}, journal={LANGMUIR}, author={Genzer, Jan and Bhat, Rajendra R.}, year={2008}, month={Mar}, pages={2294–2317} } @article{guo_resnick_efimenko_genzer_desimone_2008, title={Alternative fluoropolymers to avoid the challenges associated with perfluorooctanoic acid}, volume={47}, ISSN={["0888-5885"]}, DOI={10.1021/ie0703179}, abstractNote={The degradation of stain-resistant coating materials leads to the release of biopersistent perfluorooctanoic acid (PFOA) to the environment. In order to find the environmentally friendly substitutes, we have designed and synthesized a series of nonbiopersistant fluorinated polymers containing perfluorobutyl groups in the side chains. The surface properties of the new coating materials were characterized by static and dynamic contact angle measurements. The new coating materials demonstrate promising hydrophobic and oleophobic properties with low surfaces tensions. The wetting properties and surface structure of the polymers were tuned by varying the “spacer” structures between the polymer backbones and the perfluorinated groups of the side chains. The relationship between orientations of the fluorinated side chains and performances of polymer surfaces were further investigated by near-edge X-ray fine absorption structure (NEXAFS) experiments and differential scanning calorimetry (DSC).}, number={3}, journal={INDUSTRIAL & ENGINEERING CHEMISTRY RESEARCH}, author={Guo, Ji and Resnick, Paul and Efimenko, Kirill and Genzer, Jan and DeSimone, Joseph M.}, year={2008}, month={Feb}, pages={502–508} } @article{wu_gong_szleifer_vlcek_subr_genzer_2007, title={Behavior of surface-anchored poly(acrylic acid) brushes with grafting density gradients on solid substrates: 1. Experiment}, volume={40}, ISSN={["1520-5835"]}, DOI={10.1021/ma0710176}, abstractNote={We describe experiments pertaining to the formation of surface-anchored poly(acrylic acid) (PAA) brushes with a gradual variation of the PAA grafting densities on flat surfaces and provide detailed analysis of their properties. The PAA brush gradients are generated by first covering the substrate with a molecular gradient of the polymerization initiator, followed by the “grafting from” polymerization of tert-butyl acrylate (tBA) from these substrate-bound initiator centers, and finally converting the PtBA into PAA. We use spectroscopic ellipsometry to measure the wet thickness of the grafted PAA chains in aqueous solutions at three different pH values (4, 5.8, and 10) and a series of ionic strengths (IS). Our measurements reveal that at low grafting densities, σ, the wet thickness of the PAA brush (H) remains relatively constant, the polymers are in the mushroom regime. Beyond a certain value of σ, the macromolecules enter the brush regime, where H increases with increasing σ. For a given σ, H exhibits a ...}, number={24}, journal={MACROMOLECULES}, author={Wu, Tao and Gong, Peng and Szleifer, Igal and Vlcek, Petr and Subr, Vladimir and Genzer, Jan}, year={2007}, month={Nov}, pages={8756–8764} } @article{gong_wu_genzer_szleifer_2007, title={Behavior of surface-anchored poly(acrylic acid) brushes with grafting density gradients on solid substrates: 2. Theory}, volume={40}, ISSN={["0024-9297"]}, DOI={10.1021/ma071018y}, abstractNote={We apply a molecular theory to predict the structural properties of poly(acrylic acid) macromolecules grafted via one of their ends to solid surfaces. The theory explicitly incorporates the acid-base equilibrium responsible for the charge regulation of the acrylic groups, as well as the conformations, size, shape, and charge distributions of all the molecular species present. We compare the predictions of the theory with experimental observations presented in the preceding article for the height of the layer as a function of ionic strength for different polymer molecular weight and polymer surface coverage. The calculated heights are found to be in good agreement with the experimental observations. The theory predicts the distribution of charges within the layer. We find that the counterions adsorb to the grafting surface, overcompensating the charge of the polymer. The charge regulation within the polymer layer is determined by the interplay between the bulk pH, the ionic strength, and the density of polymer. The system tends to become uncharged with decreasing ionic strength of the solution and increasing polymer density. In all cases the charge regulation acts in order to minimize the electrostatic repulsions in the system. The local distribution of protons within the polymer layer is predicted to be very different from that of the bulk solution. The local pH within the polymer layer can be tuned by varying the solution ionic strength and the polymer surface coverage; the variation can be large as two pH units, relative to the bulk pH. This large variation of the local pH within a couple of nanometers within the brush can be used in the design of biosensors.}, number={24}, journal={MACROMOLECULES}, author={Gong, Peng and Wu, Tao and Genzer, Jan and Szleifer, Igal}, year={2007}, month={Nov}, pages={8765–8773} } @article{rhodes_lappi_fischer_sambasivan_genzer_franzen_2008, title={Characterization of monolayer formation on aluminum-doped zinc oxide thin films}, volume={24}, ISSN={["0743-7463"]}, DOI={10.1021/la701741m}, abstractNote={The optical and electronic properties of aluminum-doped zinc oxide (AZO) thin films on a glass substrate are investigated experimentally and theoretically. Optical studies with coupling in the Kretschmann configuration reveal an angle-dependent plasma frequency in the mid-IR for p-polarized radiation, suggestive of the detection of a Drude plasma frequency. These studies are complemented by oxygen depletion density functional theory studies for the calculation of the charge carrier concentration and plasma frequency for bulk AZO. In addition, we report on the optical and physical properties of thin film adlayers of n-hexadecanethiol (HDT) and n-octadecanethiol (ODT) self-assembled monolayers (SAMs) on AZO surfaces using reflectance FTIR spectroscopy, X-ray photoelectron spectroscopy (XPS), contact angle, and near-edge X-ray absorption fine structure (NEXAFS) spectroscopy. Our characterization of the SAM deposition onto the AZO thin film reveals a range of possible applications for this conducting metal oxide.}, number={2}, journal={LANGMUIR}, author={Rhodes, Crissy L. and Lappi, Simon and Fischer, Daniel and Sambasivan, Sharadha and Genzer, Jan and Franzen, Stefan}, year={2008}, month={Jan}, pages={433–440} } @article{jayaraman_hall_genzer_2007, title={Computer simulation study of probe-target hybridization in model DNA microarrays: Effect of probe surface density and target concentration}, volume={127}, ISSN={["1089-7690"]}, DOI={10.1063/1.2787618}, abstractNote={We use lattice Monte Carlo simulations to study the thermodynamics of hybridization of single-stranded “target” genes in solution with complementary “probe” DNA molecules immobilized on a microarray surface. The target molecules in our system contain 48 segments and the probes tethered on a hard surface contain 8–24 segments. The segments on the probe and target are distinct, with each segment representing a sequence of nucleotides that interacts exclusively with its unique complementary target segment with a single hybridization energy; all other interactions are zero. We examine how surface density (number of probes per unit surface area) and concentration of target molecules affect the extent of hybridization. For short probe lengths, as the surface density increases, the probability of binding long stretches of target segments increases at low surface density, reaches a maximum at an intermediate surface density, and then decreases at high surface density. Furthermore, as the surface density increases, the target is less likely to bind completely to one probe; instead, it binds simultaneously to multiple probes. At short probe lengths, as the target concentration increases, the fraction of targets binding completely to the probes (specificity) decreases. At long probe lengths, varying the target concentration does not affect the specificity. At all target concentrations as the probe length increases, the fraction of target molecules bound to the probes by at least one segment (sensitivity) increases while the fraction of target molecules completely bound to the probes (specificity) decreases. This work provides general guidelines to maximizing microarray sensitivity and specificity. Our results suggest that the sensitivity and specificity can be maximized by using probes 130–180 nucleotides long at a surface density in the range of 7×10−5–3×10−4probemoleculespernm2.}, number={14}, journal={JOURNAL OF CHEMICAL PHYSICS}, author={Jayaraman, Arthi and Hall, Carol K. and Genzer, Jan}, year={2007}, month={Oct} } @article{semler_jhon_tonelli_beevers_krishnamoorti_genzer_2007, title={Facile method of controlling monomer sequence distributions in random copolymers}, volume={19}, DOI={10.1002/adma.200602359}, abstractNote={Inthis report, we present a simple methodology facilitating theformation of A-B random copolymers with tunable sequencedistributions. We demonstrate that varying the degree ofblockiness in the sequence distribution of A and B monomershas a profound impact on the partition of random copolymersat interfaces.Random copolymers (RCPs) are long chain moleculesmade of covalently bound monomers comprising at least twodifferent chemical moieties (say, A and B). In addition to theoverall molecular weight, RCPs are characterized by theircomposition and monomer sequence distribution. The abilityof A-B RCPs to act as “homopolymers with tunable composi-tion”, ranging between A and B homopolymers, has recentlyattracted considerable attention in controlling polymer misci-bility}, number={19}, journal={Advanced Materials}, author={Semler, J. J. and Jhon, Y. K. and Tonelli, A. and Beevers, M. and Krishnamoorti, R. and Genzer, Jan}, year={2007}, pages={2877-} } @article{cerruti_rhodes_losego_efremenko_maria_fischer_franzen_genzer_2007, title={Influence of indium-tin oxide surface structure on the ordering and coverage of carboxylic acid and thiol monolayers}, volume={40}, ISSN={["1361-6463"]}, DOI={10.1088/0022-3727/40/14/016}, abstractNote={This paper analyses the variability of self-assembled monolayers (SAMs) formation on ITO depending on the substrate surface features. In particular, we report on the formation of carboxylic acid- and thiol-based SAMs on two lots of commercially prepared indium–tin oxide (ITO) thin films. Contact angle measurements, electrochemical experiments, and near-edge x-ray absorption fine structure (NEXAFS) spectroscopy showed that the quality of monolayers formed differed substantially between the two ITO batches. Only one of the two ITO substrates was capable of forming well-organized thiol- and carboxylic acid-based SAMs. In order to rationalize these observations, atomic force microscopy and x-ray diffraction analyses were carried out, and SAMs were prepared on ITO substrates fabricated by sputtering in our laboratories. An attempt was made to influence the film microstructure and surface morphology by varying substrate temperatures during ITO deposition. Good-quality thiol and carboxylic acid SAMs were obtained on one of the ITO substrates prepared in-house. While our characterization could not single out conclusively one specific parameter in ITO surface structure that could be responsible for good SAMs formation, we could point out homogeneous surface morphology as a relevant factor for the quality of the SAMs. Evidence was also found for ITO crystallographic orientation to be a parameter influencing SAMs organization.}, number={14}, journal={JOURNAL OF PHYSICS D-APPLIED PHYSICS}, author={Cerruti, Marta and Rhodes, Crissy and Losego, Mark and Efremenko, Alina and Maria, Jon-Paul and Fischer, Daniel and Franzen, Stefan and Genzer, Jan}, year={2007}, month={Jul}, pages={4212–4221} } @article{loo_genzer_2007, title={Introduction}, volume={45}, ISSN={["0887-6266"]}, DOI={10.1002/polb.21356}, abstractNote={Fifteen years in its tradition, the December issue of the Journal of Polymer Science, Polymer Physics features a series of papers by members of the Division of Polymers (DPOLY) who presented their work at the annual meeting of the American Physical Society (APS). Two years ago, this special APS DPOLY issue received a significant “face-lift.” Kari Dalnoki-Veress (recipient of the 2008 Dillon Medal) from McMaster University in Canada, who served as the Editor of the special issue, introduced a new and exciting layout which incorporated a series of feature articles by invited authors1 in addition to regular contributed papers. In these feature articles, experts in their respective fields shared their points-of-views on current and future developments in polymer physic related topics; these articles were meant to provide perspective to non-experts in the field. Additionally, the winners of the Polymer Physics Prize and the Dillon Medal wrote review articles pertaining to their research. This fresh blueprint introduced by Kari made a very favorable impression on the readers of the journal and a similar format was chosen by the co-editors of the special issue last year, Chang Y. Ryu and Rahmi Ozisik from the Rensselaer Polytechnic Institute, the following year.2 Inspired by our predecessors and by the success of the new format of the special APS issue, the decision for us to replicate the existing layout was thus very easy. The 2007 DPOLY special issue of Journal of Polymer Science, Polymer Physics highlights a paper co-authored by Glenn Fredrickson, the winner of the 2007 DPOLY Polymer Physics Prize. Concentrating on recent advances in the simulation of charged polymeric systems, Glenn's review article provides an excellent overview of the field, which we hope will inspire theorists and experimentalists alike. The subsequent 6 articles represent viewpoints/opinions by experts in the diverse fields of polymer physics. The section commences with a paper authored by Buck Christ, who commemorates the 50th anniversary of chain folding. Steve Granick and Janet Wong share their perspectives on friction in polymer systems. Ron Larson's article examines entanglements in polymer dynamics by computer simulations. Marc Hillmyer highlights the influence of molecular weight distribution on microstructure development in block copolymers. New insights into polymer nanocomposites are presented by Ramanan Krishnamoorti and Rich Vaia. And if you ever wondered how polymer physics is related to foodstuffs, you should not miss Athene Donald's article. The second part of the journal features 13 contributed papers, including several from finalists of the Frank J. Padden, Jr. Award, highlighting selected works presented during the March 2007 APS meeting. It has been our pleasure to co-edit the 2007 special issue of Journal of Polymer Science, Polymer Physics. We would like to take this opportunity to thank the authors for their contributions and the reviewers for their assistance in assessing these manuscripts. And last but not least, we would like to express our sincere gratitude to Ms. Hope Inman, the journal's managing editor, for her help and superb organization and managing skills. Enjoy! YUEH-LIN (LYNN) LOO Professor Loo recently joined Princeton University as an associate professor in the Chemical Engineering Department. She received bachelor degrees in Chemical Engineering as well as Materials Science and Engineering from the University of Pennsylvania in 1996 and a doctoral degree in Chemical Engineering from Princeton University in 2001. In 2002, Lynn joined the University of Texas at Austin as Assistant Professor in Chemical Engineering and General Dynamics Endowed Faculty Fellow in Engineering. Lynn's research focuses on developing patterning and processing tools for making organic electronic devices and understanding the structure-property relationships in organic electrically-active materials, especially solution-processable organic molecular semiconductors and polymer conductors. Her honors include a Camille and Henry Dreyfus New Faculty Award (2002), a DuPont Young Professor Grant (2003), an NSF-CAREER Award (2004), and an Arnold and Mabel Beckman Young Investigator Award (2005). In 2004, Lynn was also selected as one of Top 100 Young Innovators under 35 by MIT's Technology Review. More recently, Lynn was the inaugural recipient of the Peter and Edith O'Donnell Award in Engineering awarded by the Academy of Medicine, Science, and Engineering of Texas and the 2006 recipient of the Alan P. Colburn Award of the AIChE. Lynn will be presenting the 2007 Thiele Lecture in the Chemical Engineering Department at the University of Notre Dame this fall. JAN GENZER Professor Genzer is a professor of Chemical & Biomolecular Engineering at North Carolina State University. He received the "Diploma-engineer" degree (Dipl.-Ing.) in Chemical & Materials Engineering from the Institute of Chemical Technology in Prague, Czech Republic in 1989 and Ph.D. in Materials Science & Engineering from the University of Pennsylvania in 1996. Between 1996 and 1998 he was a post-doctoral fellow at Cornell University and the University of California at Santa Barbara, working with Professor Ed Kramer. In the fall 1998 he joined the faculty of chemical engineering at the North Carolina State University as an assistant professor and was promoted to the associate and full professor in 2004 and 2006, respectively. He also holds a position of an adjunct professor at the Norwegian University of Science & Technology (NTNU) in Trondheim, Norway. His honors include: Camille Dreyfus Teacher-Scholar Award, Sigma Xi research award, NSF CAREER award, John H. Dillon Award of the American Physical Society, NSF Award for Special Creativity, and NCSU's Outstanding Teacher award. He has published over 100 peer-reviewed journal articles and delivered more than 100 invited lectures. His group at North Carolina State University is actively involved in research related to the behavior of polymers at interfaces and in confined geometries, with particular emphasis on self-assembly and forced assembly and combinatorial methods.}, number={24}, journal={JOURNAL OF POLYMER SCIENCE PART B-POLYMER PHYSICS}, author={Loo, Yueh-Lin and Genzer, Jan}, year={2007}, month={Dec}, pages={3221–3222} } @inproceedings{crowe_efimenko_genzer_2007, title={Manipulating siloxane surfaces: Obtaining the desired surface function via engineering design}, volume={964}, DOI={10.1021/bk-2007-0964.ch015}, abstractNote={We present a synopsis of recent accomplishment in our group in the area of surface-functionalized silicone elastomer networks. Specifically, we show that by combining mechanical manipulation of poly(dimethylsiloxane) (PDMS) networks with activation via ultraviolet/ozone (UVO) treatment and subsequent chemical modification of the preactivated surfaces, one can generate so-called mechanically assembled monolayers (MAMs). This technology can be successfully applied to create a variety of surfaces, including dense polymer brushes, long-lived superhydrophobic surfaces, molecular gradients comprising tunable length scales and two-dimensional chemical gradients. In addition, the UVO modification of mechanically strained PDMS sheets provides a convenient route for creating "buckled" elastomeric sheets. These surfaces have a multitude of applications; ranging from anti-fouling surfaces to directed particle assembly. Finally we demonstrate control of surface wettability and responsiveness using poly(vinylmethylsiloxane) (PVMS) networks. We have shown that PVMS surfaces can be chemically tailored by reacting with thiols. The degree of response (including response rate) of such surfaces to hydrophobic-hydrophilic interactions can be adjusted by varying the chemical and structural properties of the side-group modifiers.}, booktitle={Science and technology of silicones and silicone-modified materials}, author={Crowe, J. A. and Efimenko, Kirill and Genzer, Jan}, year={2007}, pages={222–255} } @misc{liebmann-vinson_chaney_efimenko_genzer_2007, title={Methods of surface modification of a flexible substrate to enhance cell adhesion}, volume={7,198,855}, number={2007 Apr. 7}, publisher={Washington, DC: U.S. Patent and Trademark Office}, author={Liebmann-Vinson, A. and Chaney, B. and Efimenko, K. and Genzer, J.}, year={2007} } @article{schultz_hall_genzer_2007, title={Obtaining concentration profiles from computer simulation structure factors}, volume={40}, ISSN={["1520-5835"]}, DOI={10.1021/ma062836d}, abstractNote={The structure factor, S(q), is an important analysis tool for analyzing the structure of macromolecular crystals, 1,2 block copolymers, 3 micelles, 4 and the glass transition. 5 As the Fourier transform of the density profile, S(q) gives a quantitative description of the concentration fluctuations in a system as a function of the fluctuation frequency and direction. S(q) can be obtained experimentally from scattering experiments or can be extracted from simulation data on particle positions. In the past,S(q) has been used to determine the size and morphology of micelles, the degree of ordering, the identity of an ordered structure, and even the interaction strength between components of a copolymer. 6-8}, number={8}, journal={MACROMOLECULES}, author={Schultz, Andrew J. and Hall, Carol K. and Genzer, Jan}, year={2007}, month={Apr}, pages={2629–2632} } @article{gong_genzer_szleifer_2007, title={Phase behavior and charge regulation of weak polyelectrolyte grafted layers}, volume={98}, ISSN={["1079-7114"]}, DOI={10.1103/physrevlett.98.018302}, abstractNote={The stability of weak polyelectrolytes end grafted to a planar surface has been studied with a molecular theory. The effective quality of the solvent is found to depend on the interplay between polymer grafting density, acid-base equilibrium, and salt concentration. Our results reveal that increasing salt concentration results in a thermodynamically more stable layer. This reverse salt effect is due to the competition between the solvent quality and the dual role of the ionic strength in screening the electrostatic interactions (reducing stability with increasing salt concentration), and regulating the charge on the polymer (increasing charge with increasing salt concentration). Grafted weak polyelectrolyte layers are found to be thermodynamically unstable at intermediate surface coverages. Additionally, it is established that the increased solubility of the layer at low surface coverage is due to the relatively large charge of the grafted polymers. The range of stability of the film with regard to polymer surface coverage, temperature, bulk pH and salt concentration is demonstrated.}, number={1}, journal={PHYSICAL REVIEW LETTERS}, author={Gong, Peng and Genzer, Jan and Szleifer, I.}, year={2007}, month={Jan} } @article{douglas_efimenko_fischer_phelan_genzer_2007, title={Propagating waves of self-assembly in organosilane monolayers}, volume={104}, ISSN={["0027-8424"]}, DOI={10.1073/pnas.0703620104}, abstractNote={ Wavefronts associated with reaction–diffusion and self-assembly processes are ubiquitous in the natural world. For example, propagating fronts arise in crystallization and diverse other thermodynamic ordering processes, in polymerization fronts involved in cell movement and division, as well as in the competitive social interactions and population dynamics of animals at much larger scales. Although it is often claimed that self-sustaining or autocatalytic front propagation is well described by mean-field “reaction–diffusion” or “phase field” ordering models, it has recently become appreciated from simulations and theoretical arguments that fluctuation effects in lower spatial dimensions can lead to appreciable deviations from the classical mean-field theory (MFT) of this type of front propagation. The present work explores these fluctuation effects in a real physical system. In particular, we consider a high-resolution near-edge x-ray absorption fine structure spectroscopy (NEXAFS) study of the spontaneous frontal self-assembly of organosilane (OS) molecules into self-assembled monolayer (SAM) surface-energy gradients on oxidized silicon wafers. We find that these layers organize from the wafer edge as propagating wavefronts having well defined velocities. In accordance with two-dimensional simulations of this type of front propagation that take fluctuation effects into account, we find that the interfacial widths w ( t ) of these SAM self-assembly fronts exhibit a power-law broadening in time, w ( t ) ≈ t β , rather than the constant width predicted by MFT. Moreover, the observed exponent values accord rather well with previous simulation and theoretical estimates. These observations have significant implications for diverse types of ordering fronts that occur under confinement conditions in biological or materials-processing contexts. }, number={25}, journal={PROCEEDINGS OF THE NATIONAL ACADEMY OF SCIENCES OF THE UNITED STATES OF AMERICA}, author={Douglas, Jack F. and Efimenko, Kirill and Fischer, Daniel A. and Phelan, Fredrick R. and Genzer, Jan}, year={2007}, month={Jun}, pages={10324–10329} } @article{kim_efimenko_genzer_carbonell_2007, title={Surface properties of poly[2-(perfluorooctyl)ethyl acrylate] deposited from liquid CO2 high-pressure free meniscus coating}, volume={40}, ISSN={["1520-5835"]}, DOI={10.1021/ma0623791}, abstractNote={The surface characteristics of poly[2-(perfluorooctyl)ethyl acrylate] (PFOEA) films deposited using a high-pressure free meniscus coating (hFMC) process with liquid CO{sub 2} (l-CO{sub 2}) as the coating solvent on 12.5 cm diameter silicon wafer substrates were investigated using contact angle measurements, atomic force microscopy (AFM), X-ray photoelectron spectroscopy (XPS), and near-edge X-ray adsorption fine structure (NEXAFS) spectroscopy. The results were compared with surface property measurements of PFOEA films deposited from 1,1,2-trichlorotrifluoroethane (Freon 113) under normal dip coating conditions at atmospheric pressure. NEXAFS measurements showed that perfluoroalkyl groups in the films from l-CO{sub 2} and Freon 113 were well-organized and oriented normal to the substrate at the air/polymer interface. AFM images and XPS measurements revealed that a terrace-like structure of the PFOEA film from l-CO{sub 2} resulted in carbonyl group exposure at the air/polymer interface. This leads to smaller contact angles on the films cast from l-CO{sub 2} relative to the specimens deposited from Freon 113. Annealing the films deposited from the solvents resulted in droplet formation on the surface due to dewetting. The critical surface tension ({gamma}{sub c}) after annealing the film prepared from Freon 113 increased from 6.5 to 8.5 mJ/m{sup 2}, whereas {gamma}{sub c} of the film deposited frommore » l-CO{sub 2} decreased slightly from 9.7 to 8.9 mJ/m{sup 2}. We discuss how surface morphology changes before and after annealing play a role in the variation of {gamma}{sub c}.« less}, number={3}, journal={MACROMOLECULES}, author={Kim, Jaehoon and Efimenko, Kirill and Genzer, Jan and Carbonell, Ruben G.}, year={2007}, month={Feb}, pages={588–597} } @article{ritz_latalova_janata_toman_kriz_genzer_vlcek_2007, title={Synthesis of amphiphilic copolymers by ATRP initiated with a bifunctional initiator containing trichloromethyl groups}, volume={67}, ISSN={["1873-166X"]}, DOI={10.1016/j.reactfunctpolym.2007.06.007}, abstractNote={Bifunctional polystyrene macroinitiators, having various molecular weights, were prepared by atom transfer radical polymerization (ATRP), initiated with bifunctional initiator 1,3-bis{1-methyl-1[(2,2,2-trichloroethoxy) carbonylamino]ethyl}benzene in conjunction with CuCl catalyst and polyamine ligands. These macroinitiators were subsequently used for ATRP of tert-butyl acrylate (t-BuA), giving BAB triblocks poly[(t-BuA)-b-(Sty)-b-(t-BuA)] as precursors of amphiphilic copolymers. Both the polymerization steps proceeded as controlled processes with linear semi-logarithmic conversion plots and lengths of the blocks following theoretical predictions. Hydrolysis of outer poly(t-BuA) blocks led to triblock copolymers with the central polystyrene block and outer blocks of poly(acrylic acid), the molecular weights of which ranged from ca. 5 × 103 to almost 1 × 105 Da.}, number={10}, journal={REACTIVE & FUNCTIONAL POLYMERS}, author={Ritz, P. and Latalova, P. and Janata, M. and Toman, L. and Kriz, J. and Genzer, J. and Vlcek, P.}, year={2007}, month={Oct}, pages={1027–1039} } @article{jayaraman_santiso_hall_genzer_2007, title={Theoretical study of kinetics of zipping phenomena in biomimetic polymers}, volume={76}, ISSN={["1550-2376"]}, DOI={10.1103/physreve.76.011915}, abstractNote={In this work we use theory to obtain a mathematical expression for a time correlation function c(l,t) that provides insight into the zipping phenomena along a polymer going through a conformational transition. The polymer is modeled as an Ising-like chain with each segment being in one of two states: bound (+1) or unbound (-1). The time correlation function c(l,t) predicts the correlation between the state of the jth polymer segment at time 0 and the state of the (j+/-l)th polymer segment at time t . The expressions for c(0,t) , c(1,t), and c(2,t) obtained from our theory are dependent on the values of k0 and k1, where 2k0 is the rate coefficient for one segment changing from an unbound state to a bound state when both the neighboring segments are in an unbound state, and 2k1 is the rate coefficient for one segment changing from an unbound state to a bound state when both the neighboring segments are in a bound state. The ratio k1/k0 is an indication of the extent of cooperativity of binding adjacent segments on the polymer. We observe that c(0,t), c(1,t), and c(2,t) decay to 0 (no correlation) more slowly and the maximum values of c(1,t) and c(2,t) are lower for low values of k1/k0 as compared to high values of k1/k0. This is because at low values of k1/k0 the consecutive binding of adjacent segments along the polymer occurs slowly, while at high values of k1/k0 the cooperativity of binding adjacent segments is high and the segments along the polymer bind in a fast zipping mechanism.}, number={1}, journal={PHYSICAL REVIEW E}, author={Jayaraman, Arthi and Santiso, Erik E. and Hall, Carol K. and Genzer, Jan}, year={2007}, month={Jul} } @article{jayaraman_hall_genzer_2006, title={Computer Simulation Study of Molecular Recognition in Model DNA Microarrays}, volume={91}, ISSN={0006-3495}, url={http://dx.doi.org/10.1529/biophysj.106.086173}, DOI={10.1529/biophysj.106.086173}, abstractNote={DNA microarrays have been widely adopted by the scientific community for a variety of applications. To improve the performance of microarrays there is a need for a fundamental understanding of the interplay between the various factors that affect microarray sensitivity and specificity. We use lattice Monte Carlo simulations to study the thermodynamics and kinetics of hybridization of single-stranded target genes in solution with complementary probe DNA molecules immobilized on a microarray surface. The target molecules in our system contain 48 segments and the probes tethered on a hard surface contain 8-24 segments. The segments on the probe and target are distinct and each segment represents a sequence of nucleotides ( approximately 11 nucleotides). Each probe segment interacts exclusively with its unique complementary target segment with a single hybridization energy; all other interactions are zero. We examine how the probe length, temperature, or hybridization energy, and the stretch along the target that the probe segments complement, affect the extent of hybridization. For systems containing single probe and single target molecules, we observe that as the probe length increases, the probability of binding all probe segments to the target decreases, implying that the specificity decreases. We observe that probes 12-16 segments ( approximately 132-176 nucleotides) long gave the highest specificity and sensitivity. This agrees with the experimental results obtained by another research group, who found an optimal probe length of 150 nucleotides. As the hybridization energy increases, the longer probes are able to bind all their segments to the target, thus improving their specificity. The hybridization kinetics reveals that the segments at the ends of the probe are most likely to start the hybridization. The segments toward the center of the probe remain bound to the target for a longer time than the segments at the ends of the probe.}, number={6}, journal={Biophysical Journal}, publisher={Elsevier BV}, author={Jayaraman, Arthi and Hall, Carol K. and Genzer, Jan}, year={2006}, month={Sep}, pages={2227–2236} } @article{jhon_bhat_jeong_rojas_szleifer_genzer_2006, title={Cover Picture: Macromol. Rapid Commun. 9/2006}, volume={27}, ISSN={1022-1336 1521-3927}, url={http://dx.doi.org/10.1002/marc.200690014}, DOI={10.1002/marc.200690014}, abstractNote={Cover: Quartz crystal microbalance with dissipation monitoring is used to probe the effect of salt concentration on the phase transition behavior of thermoresponsive surface‐grafted poly(N‐isopropylacryl‐amide) (PNIPAM) brushes brushes. The phase transition temperature of grafted polymer is found to decrease as the concentration of salt is increased. This depression in phase transition temperature is similar to that observed for free polymer and is attributed to unbinding of the water molecules from the PNIPAM by the presence of ions. Further details can be found in the article by Y. K. Jhon, R. R. Bhat, C. Jeong, O. J. Rojas, I. Szleifer, and J. Genzer* on page 697.}, number={9}, journal={Macromolecular Rapid Communications}, publisher={Wiley}, author={Jhon, Young K. and Bhat, Rajendra R. and Jeong, Changwoo and Rojas, Orlando J. and Szleifer, Igal and Genzer, Jan}, year={2006}, month={May}, pages={645–645} } @article{semler_genzer_2006, title={Design of random copolymers with statistically controlled monomer sequence distributions via Monte Carlo simulations}, volume={125}, ISSN={["1089-7690"]}, DOI={10.1063/1.2210011}, abstractNote={We use Monte Carlo simulations to model the formation of random copolymers with tunable monomer sequence distributions. Our scheme is based on the original idea proposed a few years ago by Khokhlov and Khalatur [Physica A 249, 253 (1998); Phys. Rev. Lett. 82, 3456 (1999)], who showed that the distribution of species B in A-B random copolymers can be regulated by (a) adjusting the coil size of a homopolymer A and (b) chemically modifying (“coloring”) monomers that reside at (or close to) the periphery of the coil with species B. In contrast to Khokhlov and Khalatur’s work, who modeled the polymer modification by performing the coloring instantaneously, we let the chemical coloring reaction progress over time using computer simulations. We show that similar to Khokhlov and Khalatur’s work, the blockiness (i.e., number of consecutive monomers) of the B species along the A-B copolymer increases with increasing degree of collapse of the parent homopolymer A. A simple analysis of the A-B monomer sequences in the copolymers reveals that monomer sequence distributions in homopolymers “colored” under collapsed conformations possess certain degrees of self-similarity, while there is no correlation found among the monomer sequence distributions formed by coloring homopolymers with expanded conformations.}, number={1}, journal={JOURNAL OF CHEMICAL PHYSICS}, author={Semler, James J. and Genzer, Jan}, year={2006}, month={Jul} } @article{smith_tong_genzer_fischer_kilpatrick_2006, title={Effects of synthetic amphiphilic alpha-helical peptides on the electrochemical and structural properties of supported hybrid bilayers on gold}, volume={22}, ISSN={["0743-7463"]}, DOI={10.1021/la051104w}, abstractNote={Amphiphilic alpha-helices were formed from designed synthetic peptides comprising alanine, phenylalanine, and lysine residues. The insertion of the alpha-helical peptides into hybrid bilayers assembled on gold was studied by a variety of methods to assess the resulting structural characteristics, such as electrical resistance and molecular orientation. Self-assembled monolayers (SAMs) of dodecanethiol (DDT); octadecanethiol (ODT); and 1,2-dipalmitoyl-sn-glycero-3-phosphothioethanol (DPPTE) were formed on gold substrates with and without incorporated peptide. Supported hybrid bilayers and multilayers of 1,2-dimyristoyl-sn-glycero-3-phosphocholine (DMPC) were formed on SAMs by the "paint-freeze" method of bilayer formation. Modeling of electrochemical impedance spectroscopy data using equivalent electrochemical circuits revealed that the addition of peptide decreased dramatically the resistive element of the bilayer films while maintaining the value of the capacitive element, indicating successful incorporation of peptide into a well-formed bilayer. Near-edge X-ray absorption fine structure spectroscopy data provided evidence that the molecules in the SAMs and hybrid multilayers were ordered even in the presence of peptide. The peptide insertion into the SAM was confirmed by observing the pi* resonance peak correlating with phenylalanine and a peak in the nitrogen K-edge regime attributable to the peptide bond.}, number={4}, journal={LANGMUIR}, author={Smith, MB and Tong, JH and Genzer, J and Fischer, D and Kilpatrick, PK}, year={2006}, month={Feb}, pages={1919–1927} } @article{genzer_efimenko_fischer_2006, title={Formation mechanisms and properties of semifluorinated molecular gradients on silica surfaces}, volume={22}, ISSN={["0743-7463"]}, DOI={10.1021/la061016r}, abstractNote={The goal of this study is to elucidate the formation of molecular gradients made of semifluorinated organosilanes (SFOs) on flat substrates by using a methodology developed by Chaudhury and Whitesides (Science 1992, 256, 1539). We use surface-sensitive combinatorial near-edge X-ray absorption fine structure (combi-NEXAFS) spectroscopy to measure the position-dependent concentration and orientation of SFO molecules in SFO molecular gradients on flat silica surfaces. Using the combi-NEXAFS data, we establish the correlation between the fraction of the F(CF(2))(8)(CH(2))(2)- species on the substrate and the average tilt angle of the -(CF(2))(8)F group in the SFO as a function of the deposition gas medium (air vs nitrogen) and the end group around the silicon atom (monofunctional vs trifunctional). In addition, we utilize the gradient geometry to comprehend the mechanism of formation of SFO self-assembled monolayers (SAMs). Specifically, we provide evidence that depending on the nature of the end group in the SFO and the vapor phase the SFO molecules add themselves into the existing SAMs either as individual molecules or as multimolecular complexes.}, number={20}, journal={LANGMUIR}, author={Genzer, Jan and Efimenko, Kirill and Fischer, Daniel A.}, year={2006}, month={Sep}, pages={8532–8541} } @article{genzer_2006, title={In silico polymerization: Computer simulation of controlled radical polymerization in bulk and on flat surfaces}, volume={39}, ISSN={["1520-5835"]}, DOI={10.1021/ma061155f}, abstractNote={We use Monte Carlo computer simulation to study the effect of several molecular parameters on controlled/“living” radical polymerization in bulk and on flat substrates. Specifically, we investigate how the molecular weight and molecular weight distribution of grown polymers depend on the initial number of initiators, the initial number of monomers, the initiator activation probability, the initial probability of addition of a new monomer to a growing chain, the probability of terminating two “living” polymers, and the numbers of “living” polymers and their lifetime. Some observations reported here are common to both bulk- and surface-initiated polymerizations. Specifically, we demonstrate that increasing the termination probability and/or decreasing the initial probability of addition of a new monomer to a growing chain broadens the molecular weight distribution. In addition, decreasing the lifetime of the “living” radicals results in polymers with narrower molecular weight distributions. One of the goals...}, number={20}, journal={MACROMOLECULES}, author={Genzer, Jan}, year={2006}, month={Oct}, pages={7157–7169} } @article{clare_efimenko_fischer_genzer_abbott_2006, title={Orientations of liquid crystals in contact with surfaces that present continuous gradients of chemical functionality}, volume={18}, ISSN={["1520-5002"]}, DOI={10.1021/cm052537n}, abstractNote={We report the formation of continuous spatial gradients in the density of grafted semifluorinated chains on silicon oxide surfaces by vapor-phase diffusion of semifluorinated silanes. We quantify the orientations of the nematic liquid crystal (LC) 4-cyano-4‘-pentylbiphenyl on these surfaces as a function of local surface composition obtained by using NEXAFS. These measurements demonstrate that it is possible to obtain the full range of tilt angles of a LC on these surfaces. We also use the data provided by these gradient surfaces to test hypotheses regarding the nature of the interaction between the LC and surfaces that give rise to the range of tilted orientations of the LC. We conclude that the orientations of the LC are not determined solely by the density of grafted semifluorinated chains or by the density of residual hydroxyl groups presented at these surfaces following reactions with the silanes. Instead, our results raise the possibility that the tilt angles of the semifluorinated chains on these s...}, number={9}, journal={CHEMISTRY OF MATERIALS}, author={Clare, BH and Efimenko, K and Fischer, DA and Genzer, J and Abbott, NL}, year={2006}, month={May}, pages={2357–2363} } @misc{genzer_efimenko_2006, title={Recent developments in superhydrophobic surfaces and their relevance to marine fouling: a review}, volume={22}, ISSN={["1029-2454"]}, DOI={10.1080/08927010600980223}, abstractNote={Abstract In this review, a brief synopsis of superhydrophobicity (i.e. extreme non-wettability) and its implications on marine fouling are presented. A short overview of wettability and recent experimental developments aimed at fabricating superhydrophobic surfaces by tailoring their chemical nature and physical appearance (i.e. substratum texture) are reviewed. The formation of responsive/“smart” surfaces, which adjust their physico-chemical properties to variations in some outside physical stimulus, including light, temperature, electric field, or solvent, is also described. Finally, implications of tailoring the surface chemistry, texture, and responsiveness of surfaces on the design of effective marine fouling coatings are considered and discussed.}, number={5}, journal={BIOFOULING}, author={Genzer, Jan and Efimenko, Kirill}, year={2006}, pages={339–360} } @article{jhon_bhat_jeong_rojas_szleifer_genzer_2006, title={Salt-induced depression of lower critical solution temperature in a surface-grafted neutral thermoresponsive polymer}, volume={27}, ISSN={["1521-3927"]}, DOI={10.1002/marc.200600031}, abstractNote={AbstractSummary: Quartz crystal microbalance with dissipation monitoring (QCM‐D) is employed to determine the effect of salt on the volume phase transition of thermoresponsive polymer brushes. Changes in mass and viscoelasticity of poly(N‐isopropylacrylamide) (PNIPAM) layers grafted from a QCM‐D crystal are measured as a function of temperature, upon contact with aqueous solutions of varying salt concentrations. The phase‐transition temperature of PNIPAM brushes, TC,graft, quantified from the QCM‐D measurements is found to decrease as the concentration of salt is increased. This phenomenon is explained by the tendency of salt ions to affect the structure of water molecules (Hofmeister effect). However, in contrast to the linear decrease in phase‐transition temperature upon increasing salt concentration observed for free PNIPAM, the trend in TC,graft for PNIPAM brushes is distinctively non‐linear.Schematic representation of the effect of salt concentration on the phase transition behavior of thermoresponsive polymer brushes.magnified imageSchematic representation of the effect of salt concentration on the phase transition behavior of thermoresponsive polymer brushes. }, number={9}, journal={MACROMOLECULAR RAPID COMMUNICATIONS}, author={Jhon, Young K. and Bhat, Rajendra R. and Jeong, Changwoo and Rojas, Orlando J. and Szleifer, Igal and Genzer, Jan}, year={2006}, month={May}, pages={697–701} } @misc{genzer_groenewold_2006, title={Soft matter with hard skin: From skin wrinkles to templating and material characterization}, volume={2}, ISSN={["1744-6848"]}, DOI={10.1039/b516741h}, abstractNote={The English-language dictionary defines wrinkles as "small furrows, ridges, or creases on a normally smooth surface, caused by crumpling, folding, or shrinking". In this paper we review the scientific aspects of wrinkling and the related phenomenon of buckling. Specifically, we discuss how and why wrinkles/buckles form in various materials. We also describe several examples from everyday life, which demonstrate that wrinkling or buckling is indeed a commonplace phenomenon that spans a multitude of length scales. We will emphasize that wrinkling is not always a frustrating feature (, wrinkles in human skin), as it can help to assemble new structures, understand important physical phenomena, and even assist in characterizing chief material properties.}, number={4}, journal={SOFT MATTER}, author={Genzer, J and Groenewold, J}, year={2006}, month={Apr}, pages={310–323} } @article{tomlinson_efimenko_genzer_2006, title={Study of kinetics and macroinitiator efficiency in surface-initiated atom-transfer radical polymerization}, volume={39}, ISSN={["0024-9297"]}, DOI={10.1021/ma061885n}, abstractNote={Generation of surface-tethered block copolymer brushes with well-defined physicochemical characteristics requires achieving good control over the degree of polymerization of each block of the copolymer. In order to precisely form these block copolymer layers, one must (1) utilize a polymerization scheme that is capable of generating nearly monodisperse polymers, (2) fully characterize the kinetics of surface-initiated polymerization, and (3) produce macroinitiators with living characteristics capable of reinitiating the growth of each subsequent block. In this work, we describe technological steps leading to the controlled growth of surface-tethered homopolymers and multiblock copolymers via surface-initiated atom transfer radical polymerization (ATRP) from flat substrates. We first report on investigating the ability of a macroinitiator to reinitiate a homopolymer brush. We use computer simulations to illustrate the advantages of an “added deactivator” type ATRP over the traditional “sacrificial initiato...}, number={26}, journal={MACROMOLECULES}, author={Tomlinson, Michael R. and Efimenko, Kirill and Genzer, Jan}, year={2006}, month={Dec}, pages={9049–9056} } @article{smith_efimenko_fischer_lappi_kilpatrick_genzer_2007, title={Study of the packing density and molecular orientation of bimolecular self-assembled monolayers of aromatic and aliphatic organosilanes on silica}, volume={23}, ISSN={["0743-7463"]}, DOI={10.1021/la062475v}, abstractNote={Bimolecular self-assembled monolayers (SAMs) of aromatic and aliphatic chlorosilanes were self-assembled onto silica, and their characteristics were established by contact angle measurement, near-edge X-ray absorption fine structure spectroscopy, and Fourier transform infrared spectroscopy. Three aromatic constituents (phenyltrichlorosilane, benzyltrichlorosilane, and phenethyltrichlorosilane) were studied in combination with four aliphatic coadsorbates (butyltrichlorosilane, butyldimethylchlorosilane, octadecyltrichlorosilane, and octadecyldimethylchlorosilane). Our results demonstrate that whereas SAMs made of trichlorinated organosilanes are densely packed, SAMs prepared from monochlorinated species are less dense and poorly ordered. In mixed systems, trichlorinated aromatics and trichlorinated aliphatics formed SAMs with highly tunable compositions; their surfaces were compositionally homogeneous with no large-scale domain separation. The homogeneous nature of the resulting SAM was a consequence of the formation of in-plane siloxane linkages among neighboring molecules. In contrast, when mixing monochlorinated aliphatics with trichlorinated aromatics, molecular segregation occurred. Although the two shortest aromatic species did not display significant changes in orientation upon mixing with aliphatics, the aromatic species with the longest polymethylene spacer, phenethyltrichlorosilane, displayed markedly different orientation behavior in mixtures of short- and long-chain aliphatics.}, number={2}, journal={LANGMUIR}, author={Smith, Matthew B. and Efimenko, Kirill and Fischer, Daniel A. and Lappi, Simon E. and Kilpatrick, Peter K. and Genzer, Jan}, year={2007}, month={Jan}, pages={673–683} } @article{bhat_tomlinson_wu_genzer_2006, title={Surface-grafted polymer gradients: Formation, characterization, and applications}, volume={198}, ISBN={["3-540-30251-4"]}, ISSN={["1436-5030"]}, DOI={10.1007/12_060}, journal={SURFACE- INITIATED POLYMERIZATION II}, publisher={Berlin; New York: Springer}, author={Bhat, Rajendra R. and Tomlinson, Michael R. and Wu, Tao and Genzer, Jan}, year={2006}, pages={51–124} } @article{wei_lam_braunfeld_agard_genzer_spontak_2006, title={Tunable instability mechanisms of polymer thin films by molecular self-assembly}, volume={22}, ISSN={["0743-7463"]}, DOI={10.1021/la061391j}, abstractNote={Incorporation of a block copolymer into a thin polymer film is observed to alter both the rate and mechanism by which the film dewets from an immiscible polymer substrate. Films with little or no copolymer dewet by classical nucleation and growth of circular holes, and the dewetting rate decreases with increasing copolymer concentration. Increasing the copolymer content at constant film thickness generates copolymer micelles that adsorb/aggregate along the polymer/polymer interface and promote nonclassical dewetting fluctuations similar in appearance to spinodal dewetting. At higher copolymer concentrations, dewetting proceeds after a lengthy induction period by the nucleation and growth of flower-shaped holes suggestive of film pinning or viscous fingering. Atomic force microscopy of the polymer/polymer interface after removal of the top film by selective dissolution reveals substantial structural development due to copolymer self-assembly.}, number={21}, journal={LANGMUIR}, author={Wei, Bin and Lam, Peter G. and Braunfeld, Michael B. and Agard, David A. and Genzer, Jan and Spontak, Richard J.}, year={2006}, month={Oct}, pages={8642–8645} } @article{bhat_genzer_2007, title={Tuning the number density of nanoparticles by multivariant tailoring of attachment points on flat substrates}, volume={18}, ISSN={["0957-4484"]}, DOI={10.1088/0957-4484/18/2/025301}, abstractNote={We report on the organization of nanoparticles on a flat surface when there is strong yet tunable interaction between the particles and the surface. Specifically, we tailor the number density of citrate-stabilized gold nanoparticles on flat substrates by varying the concentration of the grafted amino groups on the surfaces and their degree of ionization. While the former effect is accomplished by decorating silica-based substrates with a molecular gradient of (3-aminopropyl)triethoxysilane (APTES), the latter effect is achieved by varying the degree of ionization of the −NH2 groups in APTES by varying the pH of the gold sol. We show that the measurement of particle number density on an APTES concentration gradient substrate at different pH values provides a simple, non-spectroscopic means to deduce the relative molecular concentration profile of APTES on the substrate.}, number={2}, journal={NANOTECHNOLOGY}, author={Bhat, Rajendra R. and Genzer, Jan}, year={2007}, month={Jan} } @article{striolo_jayaraman_genzer_hall_2005, title={Adsorption of comb copolymers on weakly attractive solid surfaces}, volume={123}, ISSN={["1089-7690"]}, DOI={10.1063/1.1993557}, abstractNote={In this work continuum and lattice Monte Carlo simulation methods are used to study the adsorption of linear and comb polymers on flat surfaces. Selected polymer segments, located at the tips of the side chains in comb polymers or equally spaced along the linear polymers, are attracted to each other and to the surface via square-well potentials. The rest of the polymer segments are modeled as tangent hard spheres in the continuum model and as self-avoiding random walks in the lattice model. Results are presented in terms of segment-density profiles, distribution functions, and radii of gyration of the adsorbed polymers. At infinite dilution the presence of short side chains promotes the adsorption of polymers favoring both a decrease in the depletion-layer thickness and a spreading of the polymer molecule on the surface. The presence of long side chains favors the adsorption of polymers on the surface, but does not permit the spreading of the polymers. At finite concentration linear polymers and comb polymers with long side chains readily adsorb on the solid surface, while comb polymers with short side chains are unlikely to adsorb. The simple models of comb copolymers with short side chains used here show properties similar to those of associating polymers and of globular proteins in aqueous solutions, and can be used as a first approximation to investigate the mechanism of adsorption of proteins onto hydrophobic surfaces.}, number={6}, journal={JOURNAL OF CHEMICAL PHYSICS}, author={Striolo, A and Jayaraman, A and Genzer, J and Hall, CK}, year={2005}, month={Aug} } @article{bhat_genzer_2006, title={Combinatorial study of nanoparticle dispersion in surface-grafted macromolecular gradients}, volume={252}, ISSN={["0169-4332"]}, DOI={10.1016/j.apsusc.2005.03.237}, abstractNote={Surface-tethered assemblies of polymers with gradually varying molecular weight (MW) and/or grafting density are utilized to control the dispersion of nanosized particles. Using several case studies we show that these gradient polymer specimens represent ideal systems for combinatorial exploration of the parameters that control the distribution of the particles in surface-grafted layers. We demonstrate that the particle distribution is governed by the interplay between the particle size and the grafting density and molecular weight of the polymer brush.}, number={7}, journal={APPLIED SURFACE SCIENCE}, author={Bhat, RR and Genzer, J}, year={2006}, month={Jan}, pages={2549–2554} } @article{schultz_hall_genzer_2005, title={Computer simulation of block copolymer/nanoparticle composites}, volume={38}, ISSN={["1520-5835"]}, DOI={10.1021/ma0496910}, abstractNote={Discontinuous molecular dynamics simulation is used to study the phase behavior of diblock copolymer/nanoparticle composites. The copolymers are modeled as chains of tangent hard-spheres with square shoulder repulsions between unlike species, while the nanoparticles are modeled as hard-spheres with a square shoulder repulsion with one of the copolymer blocks. The resulting phase diagrams are presented for composites containing nanoparticles of various sizes and interaction strengths and include lamellae, perforated lamellae, cylinders, and disordered phases. Composites containing large nanoparticles also exhibit two-phase coexistence between different copolymer phases or between a copolymer phase and a nanoparticle phase, depending on the nanoparticle interaction strength. We also present concentration profiles perpendicular to the lamellar interface for nanoparticles of different sizes and interaction strengths. Neutral nanoparticles concentrate at the interface between copolymer domains, while interacti...}, number={7}, journal={MACROMOLECULES}, author={Schultz, AJ and Hall, CK and Genzer, J}, year={2005}, month={Apr}, pages={3007–3016} } @article{jayaraman_hall_genzer_2005, title={Computer simulation study of pattern transfer in AB diblock copolymer film adsorbed on a heterogeneous surface}, volume={123}, ISSN={["1089-7690"]}, DOI={10.1063/1.2043048}, abstractNote={In this work we investigate how a pattern imposed in a copolymer film at a certain distance from the surface propagates through the film onto an adsorbing heterogeneous surface. We bias the copolymer film to adopt a specified target pattern and then use simulation to design a surface pattern that helps the adsorbed film to maintain that target pattern. We examine the effect of varying the copolymer chain length, the size of the target pattern, and the distance from the surface where the target pattern is applied, z′, on the extent of pattern transfer. For each chain length, target pattern, and z′ we compare the energy of the system when a pattern is applied in the bulk to the energy when no pattern is applied in order to understand why a certain pattern size is transferred to the surface with higher fidelity than the others. At constant chain length, pattern transfer is best when the pattern size brings the energy of the system close to the energy when no pattern is applied. At constant pattern size, pattern transfer is best in the systems with longer chains. This is because longer chains are more likely to adsorb as brushes and loops which then helps transfer the pattern through the adsorbed film down to the surface.}, number={12}, journal={JOURNAL OF CHEMICAL PHYSICS}, author={Jayaraman, A and Hall, CK and Genzer, J}, year={2005}, month={Sep} } @article{crowe_genzer_2005, title={Creating responsive surfaces with tailored wettability switching kinetics and reconstruction reversibility}, volume={127}, ISSN={["0002-7863"]}, DOI={10.1021/ja056926u}, abstractNote={We report on the formation of responsive surfaces with tailorable surface reconstruction kinetics and switching hysteresis by thiolene radical addition of mercaptoalkanols with variable lengths to poly(vinylmethylsiloxane) networks. Exposing these elastomer surfaces to water results in rearrangement of the hydrophilic alkanes at the surface. The rearrangement kinetics decreases with increasing number of the methylene spacers (n) in the mercaptoalkanol. The response kinetics is found to be very fast for n = 2 and 6. For instance, upon exposing to water, the water contact angle on 3-mercaptopropanol-based surfaces decreases by approximately 35 degrees at the rate of 2 degrees /s. The high flexibility of the siloxane backbone endows these materials with switching longevity; the materials were able to switch their wettability over 10 cycles with minimum hysteresis. Increasing the number of methylene spacers to n = 11 decreases the surface reorganization dramatically. Formation of semicrystalline regions in such materials (detected via IR) is responsible for initial "sluggish" kinetics and eventual surface "freezing".}, number={50}, journal={JOURNAL OF THE AMERICAN CHEMICAL SOCIETY}, author={Crowe, JA and Genzer, J}, year={2005}, month={Dec}, pages={17610–17611} } @article{jayaraman_hall_genzer_2005, title={Designing pattern-recognition surfaces for selective adsorption of copolymer sequences using lattice Monte Carlo simulation}, volume={94}, ISSN={["1079-7114"]}, DOI={10.1103/physrevlett.94.078103}, abstractNote={We describe a simulation method to design surfaces for recognizing specific monomer sequences in copolymers. We fix the monomer sequence statistics of the AB copolymers on a surface containing two types of sites and allow the simulation to iterate towards an optimal surface pattern that can recognize and selectively adsorb the sequence in the copolymer. During the simulation the surface pattern is designed by switching identities of two randomly picked sites. For copolymers with less blocky sequences the designed surfaces recognize the correct sequence well when the segment-surface interactions dominate over the intersegment interactions. For copolymers with more blocky sequences recognition is good when the segment-surface interactions are only slightly stronger than the intersegment interactions.}, number={7}, journal={PHYSICAL REVIEW LETTERS}, author={Jayaraman, A and Hall, CK and Genzer, J}, year={2005}, month={Feb} } @article{tomlinson_genzer_2005, title={Evolution of surface morphologies in multivariant assemblies of surface-tethered diblock copolymers after selective solvent treatment}, volume={21}, ISSN={["0743-7463"]}, DOI={10.1021/la051523t}, abstractNote={We study systematically the topography behavior of PHEMA-b-PMMA block as a function of the PHEMA and PMMA block lengths after selectively collapsing the top (PMMA) block by using surface-anchored assemblies of poly(2-hydroxyethyl methacrylate-b-methyl methacrylate), PHEMA-b-PMMA, block copolymer with orthogonally varying lengths of each block. Our experimental results are in excellent qualitative agreement with topology diagrams predicted by self-consistent field calculations of Zhulina and co-workers.}, number={25}, journal={LANGMUIR}, author={Tomlinson, MR and Genzer, J}, year={2005}, month={Dec}, pages={11552–11555} } @article{bhat_chaney_rowley_liebmann-vinson_genzer_2005, title={Inside Front Cover: Tailoring Cell Adhesion Using Surface-Grafted Polymer Gradient Assemblies (Adv. Mater. 23/2005)}, volume={17}, ISSN={0935-9648 1521-4095}, url={http://dx.doi.org/10.1002/adma.200590122}, DOI={10.1002/adma.200590122}, abstractNote={AbstractA three‐step approach to tailor cell adhesion via surface‐grafted polymer gradients is shown in this image from the work of Genzer and co‐workers on p. 2802. Surface‐anchored polymer assemblies with gradients in polymer molecular weight and/or grafting density are first employed to tailor adsorption of the protein, which in turn governs the number density as well as the extent of spreading of osteoblastic cells. Increasing the surface coverage of the polymer results in a decrease in the amount of protein adsorbed, which causes a decrease in the number of cells adhered and a change in cell morphology.}, number={23}, journal={Advanced Materials}, publisher={Wiley}, author={Bhat, R. R. and Chaney, B. N. and Rowley, J. and Liebmann-Vinson, A. and Genzer, J.}, year={2005}, month={Dec}, pages={NA-NA} } @article{efimenko_rackaitis_manias_vaziri_mahadevan_genzer_2005, title={Nested self-similar wrinkling patterns in skins}, volume={4}, ISSN={["1476-4660"]}, DOI={10.1038/nmat1342}, abstractNote={Stiff thin films on soft substrates are both ancient and commonplace in nature; for instance, animal skin comprises a stiff epidermis attached to a soft dermis. Although more recent and rare, artificial skins are increasingly used in a broad range of applications, including flexible electronics, tunable diffraction gratings, force spectroscopy in cells, modern metrology methods, and other devices. Here we show that model elastomeric artificial skins wrinkle in a hierarchical pattern consisting of self-similar buckles extending over five orders of magnitude in length scale, ranging from a few nanometres to a few millimetres. We provide a mechanism for the formation of this hierarchical wrinkling pattern, and quantify our experimental findings with both computations and a simple scaling theory. This allows us to harness the substrates for applications. In particular, we show how to use the multigeneration-wrinkled substrate for separating particles based on their size, while simultaneously forming linear chains of monodisperse particles.}, number={4}, journal={NATURE MATERIALS}, author={Efimenko, K and Rackaitis, M and Manias, E and Vaziri, A and Mahadevan, L and Genzer, J}, year={2005}, month={Apr}, pages={293–297} } @article{carla_wang_hussain_efimenko_genzer_grant_sarti_carbonell_doghieri_2005, title={Nonequilibrium model for sorption and swelling of bulk glassy polymer films with Supercritical carbon dioxide}, volume={38}, ISSN={["1520-5835"]}, DOI={10.1021/ma0506684}, abstractNote={A new procedure is introduced for the calculation of solubility isotherms of plasticizing agents in glassy polymer matrices with particular application to the case of absorption of supercritical gases in bulk glassy polymer films. The model presented is an extension of the nonequilibrium thermodynamics for glassy polymers (NET-GP) approach, modified to allow for the calculation of the effects of pressure, temperature, and gas concentration on the glass transition. Mass sorption and one- dimensional swelling behavior are analyzed for the carbon dioxide (CO2)-poly(methyl methacrylate) (PMMA) system at high pressure. A quantitative comparison is presented between the model performance and experimental data measured using quartz crystal microbalance (QCM) and high-pressure ellipsometry (HPE).}, number={24}, journal={MACROMOLECULES}, author={Carla, V and Wang, K and Hussain, Y and Efimenko, K and Genzer, J and Grant, C and Sarti, GC and Carbonell, RG and Doghieri, F}, year={2005}, month={Nov}, pages={10299–10313} } @article{bhat_tomlinson_genzer_2005, title={Orthogonal surface-grafted polymer gradients: A versatile combinatorial platform}, volume={43}, ISSN={["1099-0488"]}, DOI={10.1002/polb.20640}, abstractNote={AbstractOrthogonal polymer brush gradients are assemblies of surface‐anchored macromolecules, in which two material properties of the grafted chains (e.g., grafting density, molecular weight) vary independently in orthogonal directions. Here, we describe the formation and applications of two such orthogonal assemblies, involving: (1) molecular weight and grafting density (MW/σ) gradients of a given polymer and (2) molecular weight gradients (MW1/MW2), of two different polymers. Each point on orthogonal gradient substrate represents a unique combination of the two surface properties being varied, thus facilitating systematic investigation of a phenomenon that depends on the two said properties. We illustrate this point by employing orthogonal structures to study systematically: (1) formation of polymer brush‐nanoparticle composite assemblies, (2) protein adsorption and cell adhesion, and (3) chain conformations in tethered diblock copolymers exposed to selective solvents. © 2005 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 43: 3384–3394, 2005}, number={23}, journal={JOURNAL OF POLYMER SCIENCE PART B-POLYMER PHYSICS}, author={Bhat, RR and Tomlinson, MR and Genzer, J}, year={2005}, month={Dec}, pages={3384–3394} } @article{efimenko_crowe_manias_schwark_fischer_genzer_2005, title={Rapid formation of soft hydrophilic silicone elastomer surfaces}, volume={46}, ISSN={["1873-2291"]}, DOI={10.1016/j.polymer.2005.07.046}, abstractNote={We report on the rapid formation of hydrophilic silicone elastomer surfaces by ultraviolet/ozone (UVO) irradiation of poly(vinylmethyl siloxane) (PVMS) network films. Our results reveal that the PVMS network surfaces render hydrophilic upon only a short UVO exposure time (seconds to a few minutes). We also provide evidence that the brief UVO irradiation treatment does not cause dramatic changes in the surface modulus of the PVMS network. We compare the rate of formation of hydrophilic silicone elastomer surfaces made of PVMS to those of model poly(dimethyl siloxane) (PDMS) and commercial-grade PDMS (Sylgard-184). We find that relative to PVMS, 20 times longer UVO treatment times are needed to oxidize the PDMS network surfaces in order to achieve a comparable density of surface-bound hydrophilic moieties. The longer UVO treatment times for PDMS are in turn responsible for the dramatic increase in surface modulus of UVO treated PDMS, relative to PVMS. We also study the formation of self-assembled monolayers (SAMs) made of semifluorinated organosilane precursors on the PVMS-UVO and PDMS-UVO network surfaces. By tuning the UVO treatment times and by utilizing mono- and tri-functional organosilanes we find that while mono-functionalized organosilanes attach directly to the substrate, SAMs of tri-functionalized organosilanes form in-plane networks on the underlying UVO-modified silicone elastomer surface, even with only short UVO exposure times.}, number={22}, journal={POLYMER}, author={Efimenko, K and Crowe, JA and Manias, E and Schwark, DW and Fischer, DA and Genzer, J}, year={2005}, month={Oct}, pages={9329–9341} } @article{bhat_chaney_rowley_liebmann-vinson_genzer_2005, title={Tailoring cell adhesion using surface-grafted polymer gradient assemblies}, volume={17}, ISSN={["1521-4095"]}, DOI={10.1002/adma.200500858}, abstractNote={higher-energy emissions due to impurity phases. In conclusion, well-defined [ZnSe](DETA)0.5 nanobelts have been synthesized successfully by tuning the composition of a ternary solution made of DETA, hydrazine hydrate, and deionized water. We find that an appropriate amount of hydrazine hydrate is essential for the formation of elegant [ZnSe](DETA)0.5 nanobelts. While the optical properties of the nanobelts are not changed significantly, the ability to make these II–VI-based hybrid semiconductor nanostructure microparticles into nanocrystals with uniform shape and size is one step further towards the miniaturization of devices. In addition, surface modification or combination with other materials may introduce new phenomena and properties into this system with remarkable quantum size effects and expand their potential for applications in advanced semiconductor devices.}, number={23}, journal={ADVANCED MATERIALS}, author={Bhat, RR and Chaney, BN and Rowley, J and Liebmann-Vinson, A and Genzer, J}, year={2005}, month={Dec}, pages={2802-+} } @article{genzer_2005, title={Templating surfaces with gradient assemblies}, volume={81}, ISSN={["1545-5823"]}, DOI={10.1080/00218460590944855}, abstractNote={ABSTRACT One of the most versatile and widely used methods of forming surfaces with position-dependent wettability is that conceived by Chaudhury and Whitesides more than a decade ago [Science 256, 1539 (1992)]. In this paper we review several projects that utilize this gradient-forming methodology for: controlled of deposition of self-assembled monolayers on surfaces, generating arrays of nanoparticles with number density gradients, probing the mushroom-to-brush transition in surface-anchored polymers, and controlling the speed of moving liquid droplets on surfaces. *This paper is dedicated to Professor Manoj K. Chaudhury for his contributions to the field of self-assembly and surface chemistry. The author is grateful for countless fruitful discussions he has had with Professor Chaudhury over the past several years, his friendship, encouragement, and overwhelming support. The author wants to acknowledge all past and current post-doctoral and doctoral students from his group who have contributed to the various projects outlined in this paper: Dr. Kirill Efimenko, Dr. Tao Wu, Rajendra R. Bhat, Michael R. Tomlinson, Tiffani N. Bailey, and Randal J. Petrie. The author also thanks Dr. Daniel A. Fischer (NIST) and Professor Christopher S. Gorman (NCSU) for their collaboration during the various stages of the research described in this paper. We are grateful to the National Science Foundation, The Camille & Henry Dreyfus Foundation, and 3M for their generous financial support. NEXAFS spectroscopy experiments were carried out at the National Synchrotron Light Source, Brookhaven National Laboratory, which is supported by the U. S. Department of Energy, Division of Materials Sciences and Division of Chemical Sciences.}, number={3-4}, journal={JOURNAL OF ADHESION}, author={Genzer, J}, year={2005}, pages={417–435} } @article{bhat_genzer_2005, title={Using spectroscopic ellipsometry for quick prediction of number density of nanoparticles bound to non-transparent solid surfaces}, volume={596}, ISSN={["1879-2758"]}, DOI={10.1016/j.susc.2005.09.014}, abstractNote={We report on the use of spectroscopic ellipsometry (SE) in predicting number density of nanoparticles bound to the surfaces decorated with either organic monolayers or surface-grafted polymers. Two systems are considered that comprise citrate-stabilized gold nanoparticles adsorbed on: (1) 3-aminopropyltriethoxysilane (APTES) self-assembled monolayer (SAM), and (2) surface-tethered polyacrylamide (PAAm). Number density of gold nanoparticles on the surface is varied systematically by gradually increasing either the concentration of APTES molecules in the SAM or molecular weight of grafted PAAm. The adsorption of gold nanoparticles on APTES gradient surfaces is monitored via atomic force microscopy (AFM), near-edge X-ray absorption fine structure (NEXAFS) spectroscopy, and SE. The partition of gold nanoparticles on PAAm gradient assemblies is characterized by AFM, ultraviolet–visible (UV–vis) spectroscopy, and SE. By correlating the results obtained from the various techniques on nanoparticle coatings, we derive an empirical linear relationship between the number density of nanoparticles on surfaces and cos (Δ) parameter measured in SE. Excellent agreement between nanoparticle number density determined experimentally from AFM scans and that predicted by SE proves the potential of SE as a quick, predictive technique to estimate number density of nanoparticles bound to solid, non-transparent substrates.}, number={1-3}, journal={SURFACE SCIENCE}, author={Bhat, RR and Genzer, J}, year={2005}, month={Dec}, pages={187–196} } @article{schultz_hall_genzer_2004, title={Box length search algorithm for molecular simulation of systems containing periodic structures}, volume={120}, ISSN={["1089-7690"]}, DOI={10.1063/1.1636156}, abstractNote={We have developed a box length search algorithm to efficiently find the appropriate box dimensions for constant-volume molecular simulation of periodic structures. The algorithm works by finding the box lengths that equalize the pressure in each direction while maintaining constant total volume. Maintaining the volume at a fixed value ensures that quantitative comparisons can be made between simulation and experimental, theoretical or other simulation results for systems that are incompressible or nearly incompressible. We test the algorithm on a system of phase-separated block copolymers that has a preferred box length in one dimension. We also describe and test a Monte Carlo algorithm that allows the box lengths to change while maintaining constant volume. We find that the box length search algorithm converges at least two orders of magnitude more quickly than the variable box length Monte Carlo method. Although the box length search algorithm is not ergodic, it successfully finds the box length that minimizes the free energy of the system. We verify this by examining the free energy as determined by the Monte Carlo simulation.}, number={4}, journal={JOURNAL OF CHEMICAL PHYSICS}, author={Schultz, AJ and Hall, CK and Genzer, J}, year={2004}, month={Jan}, pages={2049–2055} } @article{semler_genzer_2004, title={Cover Picture: Macromol. Theory Simul. 3/2004}, volume={13}, ISSN={1022-1344 1521-3919}, url={http://dx.doi.org/10.1002/mats.200490004}, DOI={10.1002/mats.200490004}, abstractNote={Cover: The picture on the cover represents Monte Carlo simulation generated conformations of copolymers with alternating and diblock sequences on chemically patterned substrates with strong and week attractions between the surface active sites and the adsorbing monomers along the copolymer backbone.Further details can be found in the Full Paper by J. J. Semler and J. Genzer* on page 219.}, number={3}, journal={Macromolecular Theory and Simulations}, publisher={Wiley}, author={Semler, James J. and Genzer, Jan}, year={2004}, month={Apr}, pages={193–193} } @article{wei_genzer_spontak_2004, title={Dewetting behavior of a block copolymer/homopolymer thin film on an immiscible homopolymer substrate}, volume={20}, ISSN={["0743-7463"]}, DOI={10.1021/la049562c}, abstractNote={Numerous previous studies have established that the addition of a microphase-ordered AB diblock copolymer to a thin homopolymer A (hA) film can slow, if not altogether prevent, film rupture and subsequent film dewetting on a hard substrate such as silica. However, only a few reports have examined comparable phenomena when the hA/AB blend resides on a soft B-selective surface, such as homopolymer B (hB). In this work, the dewetting kinetics of thin films composed of polystyrene (PS) and a symmetric poly(styrene-b-methyl methacrylate) (SM) diblock copolymer on a poly(methyl methacrylate) substrate is investigated by hot-stage light microscopy. Without the SM copolymer, the dewetting rate of the PS layer is constant under isothermal conditions and exhibits Arrhenius behavior with an apparent activation energy of approximately 180 kJ/mol. Addition of the copolymer promotes a crossover from early- to late-stage dewetting kinetics, as evidenced by measurably different dewetting rates. Transmission electron microscopy reveals the morphological characteristics of dewetted PS/SM films as functions of film thickness and SM concentration.}, number={20}, journal={LANGMUIR}, author={Wei, B and Genzer, J and Spontak, RJ}, year={2004}, month={Sep}, pages={8659–8667} } @article{wei_gurr_genzer_qiao_solomon_spontak_2004, title={Dewetting of star nanogel/homopolymer blends from an immiscible homopolymer substrate}, volume={37}, ISSN={["0024-9297"]}, DOI={10.1021/ma048636o}, abstractNote={Departments of Chemical & BiomolecularEngineering and Materials Science & Engineering,North Carolina State University,Raleigh, North Carolina 27695, and Polymer Science Group,Department of Chemical & Biomolecular Engineering,University of Melbourne, Victoria 3010, AustraliaReceived July 6, 2004Revised Manuscript Received August 26, 2004}, number={21}, journal={MACROMOLECULES}, author={Wei, B and Gurr, PA and Genzer, J and Qiao, GG and Solomon, DH and Spontak, RJ}, year={2004}, month={Oct}, pages={7857–7860} } @article{petrie_bailey_gorman_genzer_2004, title={Fast directed motion of "Fakir" droplets}, volume={20}, ISSN={["0743-7463"]}, DOI={10.1021/la048612a}, abstractNote={In this Letter, we report on the motion of water droplets on surfaces decorated with molecular gradients comprising semifluorinated (SF) organosilanes. SF molecular gradients deposited on flat silica substrates facilitate faster motion of water droplets relative to the specimens covered with an analogous hydrocarbon gradient. Further increase in the drop speed is achieved by advancing it along porous substrates coated with the SF wettability gradients. The results of our experiments are in quantitative agreement with a simple scaling theory that describes the faster liquid motion in terms of reduced friction at the liquid/substrate interface.}, number={23}, journal={LANGMUIR}, author={Petrie, RJ and Bailey, T and Gorman, CB and Genzer, J}, year={2004}, month={Nov}, pages={9893–9896} } @misc{genzer_efimenko_2004, title={Methods for forming tunable molecular gradients on substrates}, volume={6,770,323}, number={2004 Aug. 03}, publisher={Washington, DC: U.S. Patent and Trademark Office}, author={Genzer, J. and Efimenko, K.}, year={2004} } @article{semler_tonelli_beevers_genzer_2004, title={Monomer sequence mediated mobility in random copolymers under confinement}, volume={3}, journal={Nature Materials}, author={Semler, J. J. and Tonelli, A. E. and Beevers, M. and Genzer, J.}, year={2004} } @article{semler_genzer_2004, title={Monte Carlo simulations of copolymer adsorption at planar chemically patterned surfaces: Effect of interfacial interaction}, volume={13}, ISSN={["1521-3919"]}, DOI={10.1002/mats.200300039}, abstractNote={AbstractSummary: Monte Carlo simulation utilizing the bond fluctuation model in conjunction with single and configurational biased Monte Carlo moves is used to study the adsorption of diblock (A‐block‐B) and alternating (A‐alt‐B) copolymers at flat, chemically heterogeneous surfaces comprising C and D domains. The main objective of this work is to address the effect of the strength of attraction between the adsorbing surface domains, D, and the copolymer adsorbing segments, B, on the copolymer's ability to recognize the chemical pattern on the surface. The results of our simulations reveal that both block and alternating copolymers have the ability to recognize the surface motif and transcribe it into the bulk material. The extent to which diblock copolymers transfer the chemical pattern from the surface to the bulk is relatively unaffected when the attractive B‐D potential is increased beyond a certain critical value. This behavior stems from the brush‐like conformation adopted by the diblock copolymer at the substrate. In contrast to the diblock copolymer, the adsorption of the alternating copolymer is influenced by the strength of the attraction between the copolymer's adsorbing segments and the adsorbing domains on the surface. Since the B segments are distributed evenly along the backbone, the alternating copolymers are more likely to adopt conformations in which the whole chain is “zipped” to the surface. The resultant entropic frustration is then alleviated through an increased formation of loops with little change to their length. Such conformational changes endow the alternating copolymer with the ability to invert the substrate pattern as the distance away from the surface is increased.image}, number={3}, journal={MACROMOLECULAR THEORY AND SIMULATIONS}, author={Semler, JJ and Genzer, J}, year={2004}, month={Apr}, pages={219–229} } @article{seog_dean_rolauffs_wu_genzer_plaas_grodzinsky_ortiz_2005, title={Nanomechanics of opposing glycosaminoglycan macromolecules}, volume={38}, ISSN={["1873-2380"]}, DOI={10.1016/j.jbiomech.2004.09.010}, abstractNote={In this study, the net intermolecular interaction force between a chondroitin sulfate glycosaminoglycan (GAG)-functionalized probe tip and an opposing GAG-functionalized planar substrate was measured as a function of probe tip-substrate separation distance in aqueous electrolyte solutions using the technique of high resolution force spectroscopy. A range of GAG grafting densities as near as possible to native cartilage was used. A long-range repulsive force between GAGs on the probe tip and substrate was observed, which increased nonlinearly with decreasing separation distance between probe tip and substrate. Data obtained in 0.1 M NaCl was well predicted by a recently developed Poisson-Boltzmann-based theoretical model that describes normal electrostatic double layer interaction forces between two opposing surfaces of end-grafted, cylindrical rods of constant volume charge density and finite length, which interdigitate upon compression. Based on these results, the nanomechanical data and interdigitated rod model were used together to estimate the electrostatic component of the equilibrium modulus of cartilage tissue, which was then compared to that of normal adult human ankle cartilage measured in uniaxial confined compression.}, number={9}, journal={JOURNAL OF BIOMECHANICS}, author={Seog, J and Dean, D and Rolauffs, B and Wu, T and Genzer, J and Plaas, AHK and Grodzinsky, AJ and Ortiz, C}, year={2005}, month={Sep}, pages={1789–1797} } @article{fleming_ulrich_efimenko_genzer_chan_madey_oh_zhou_rowe_2004, title={Near-edge absorption fine structure and UV photoemission spectroscopy studies of aligned single-walled carbon nanotubes on Si(100) substrates}, volume={22}, ISSN={["2166-2746"]}, DOI={10.1116/1.1775190}, abstractNote={We report near-edge absorption fine structure (NEXAFS) and UV photoemission spectroscopy (UPS) studies of aligned single-walled carbon nanotube films on Si(100) substrates. Orientation of the films was detected in the NEXAFS spectra, with the intensity of the π* core exciton at 284.4 eV showing a strong dependence on nanotube alignment with respect to the polarization of the incident radiation. At lower angles of incidence, the intensity of the π* peak was higher for all orientations, which we attribute to the greater accessibility of the π* orbitals. UPS spectra of the films showed little angular dependence and included features consistent with the total density of states of graphite. As a result of the nanotube curvature and the distribution of nanotube chiralities, the UPS spectra are similar to angle-integrated graphite spectra.}, number={4}, journal={JOURNAL OF VACUUM SCIENCE & TECHNOLOGY B}, author={Fleming, L and Ulrich, MD and Efimenko, K and Genzer, J and Chan, ASY and Madey, TE and Oh, SJ and Zhou, O and Rowe, JE}, year={2004}, pages={2000–2004} } @article{wu_tomlinson_efimenko_genzer_2003, title={A combinatorial approach to surface anchored polymers}, volume={38}, ISSN={["1573-4803"]}, DOI={10.1023/A:1027373216389}, number={22}, journal={JOURNAL OF MATERIALS SCIENCE}, author={Wu, T and Tomlinson, M and Efimenko, K and Genzer, J}, year={2003}, month={Nov}, pages={4471–4477} } @article{wu_sambasivan_wang_wallace_genzer_fischer_2003, title={A direct comparison of surface and bulk chain-relaxation in polystyrene}, volume={12}, ISSN={["1292-8941"]}, DOI={10.1140/epje/i2003-10050-6}, abstractNote={Near-edge X-ray absorption fine-structure (NEXAFS) spectroscopy was used to measure simultaneously the relaxation rates of polystyrene (PS) molecules at the free surface and in the bulk. The samples were uniaxially stretched and annealed at temperatures below the bulk glass transition temperature of PS. The surface and bulk chain relaxation was monitored by measuring the partial-electron and the fluorescence NEXAFS yields, respectively, both parallel and perpendicular to the stretching direction. The decay of the optical birefringence was also measured to provide an independent measure of the bulk relaxation. Relaxation of PS chains was found to occur faster on the surface relative to the bulk. The magnitude of the surface glass transition temperature suppression over the bulk was estimated based on the information on the temperature dependence of the rates.}, number={1}, journal={EUROPEAN PHYSICAL JOURNAL E}, author={Wu, WL and Sambasivan, S and Wang, CY and Wallace, WE and Genzer, J and Fischer, DA}, year={2003}, month={Sep}, pages={127–132} } @article{bhat_tomlinson_genzer_2004, title={Assembly of nanoparticles using surface-grafted orthogonal polymer gradients}, volume={25}, ISSN={["1521-3927"]}, DOI={10.1002/marc.200300163}, abstractNote={AbstractSummary: We report on preparing poly(2‐(dimethylamino)ethyl methacrylate) (PDMAEMA) gradient substrate, wherein molecular weight (MW) and grafting density (σ) of the surface‐anchored PDMAEMA chains vary continuously in two orthogonal directions. Such a specimen is used to control the assembly of charged gold nanoparticles. Increasing MW and σ of the grafted PDMAEMA cause an enhanced binding of the nanoparticles to PDMAEMA, thus leading to an orthogonal number density gradient of surface‐bound gold nanoparticles.magnified image}, number={1}, journal={MACROMOLECULAR RAPID COMMUNICATIONS}, author={Bhat, RR and Tomlinson, MR and Genzer, J}, year={2004}, month={Jan}, pages={270–274} } @article{genzer_fischer_efimenko_2003, title={Combinatorial near-edge x-ray absorption fine structure: Simultaneous determination of molecular orientation and bond concentration on chemically heterogeneous surfaces}, volume={82}, ISSN={["1077-3118"]}, DOI={10.1063/1.1535271}, abstractNote={We show that simultaneous molecular orientation and bond chemistry of planar chemically heterogeneous surfaces can be obtained by combining near-edge x-ray absorption fine structure (NEXAFS) spectroscopy and rastering the incident x-ray beam on the specimen. This rastering produces serially two-dimensional NEXAFS images in space and energy, revealing information about the chemistry (including bond concentration) and orientation of the surface-bound molecules with submillimeter planar spatial resolution and submonolayer molecular sensitivity. We illustrate the power of the combinatorial NEXAFS method by simultaneously probing the concentration and molecular orientation of semifluorinated (SF) molecules in double-SF molecular gradients on flat silica substrates.}, number={2}, journal={APPLIED PHYSICS LETTERS}, author={Genzer, J and Fischer, DA and Efimenko, K}, year={2003}, month={Jan}, pages={266–268} } @article{bhat_genzer_chaney_sugg_liebmann-vinson_2003, title={Controlling the assembly of nanoparticles using surface grafted molecular and macromolecular gradients}, volume={14}, ISSN={["1361-6528"]}, DOI={10.1088/0957-4484/14/10/313}, abstractNote={We report on the generation of assemblies comprising number density gradients of nanoparticles in two (2D) and three (3D) dimensions. These structures are fabricated by creating a surface-bound organic template which directs the spatial arrangement of gold nanoparticles. The 2D template is made of amine-terminated organosilane with a concentration gradient along the solid substrate. The 3D matrix comprises surface-anchored poly(acryl amide), whose molecular weight changes gradually on the specimen. In both cases, the composite is assembled at low pH, where the positively charged –NH3+ groups within the organic scaffold attract negatively charged gold nanoparticles. We use a battery of experimental tools to determine the number density of particles along the gradient substrate and in the case of 3D structures also their spatial distribution. For 2D gradient assemblies, we show that gold nanoparticle coverage on the surface decreases gradually as the concentration of substrate-bound aminosilane decreases. The number of particles in the polymer brush/particle hybrid is found to increase with increasing polymer molecular weight. We show that for a given grafting density of polymer brush, larger particles predominantly stay near the brush–air interface. In contrast, smaller nanoparticles penetrate deeper into the polymer brush, thus forming a 3D structure. Finally, we discuss possible applications of these nanoparticle gradient assemblies.}, number={10}, journal={NANOTECHNOLOGY}, author={Bhat, RR and Genzer, J and Chaney, BN and Sugg, HW and Liebmann-Vinson, A}, year={2003}, month={Oct}, pages={1145–1152} } @article{genzer_fischer_efimenko_2003, title={Fabricating two-dimensional molecular gradients via asymmetric deformation of uniformly-coated elastomer sheets}, volume={15}, ISSN={["0935-9648"]}, DOI={10.1002/adma.200305151}, abstractNote={Molecular gradients comprising two‐dimensional in‐plane variation of density of substrate‐grafted molecules (see Figure, darker regions in the plot denote regions with higher substrate density) have been fabricated. This is achieved by utilizing the technology of “mechanically assembled monolayers” in conjunction with in‐plane asymmetric stretching of elastomeric substrates uniformly‐coated with organosilane precursors.}, number={18}, journal={ADVANCED MATERIALS}, author={Genzer, J and Fischer, DA and Efimenko, K}, year={2003}, month={Sep}, pages={1545-+} } @article{wu_efimenko_vlcek_subr_genzer_2003, title={Formation and properties of anchored polymers with a gradual variation of grafting densities on flat substrates}, volume={36}, ISSN={["1520-5835"]}, DOI={10.1021/ma0257189}, abstractNote={We show that assemblies comprising anchored polymers with a gradual variation of grafting densities on solid substrates can be generated by first covering the substrate with a molecular gradient of the polymerization initiator, followed by polymerization from the substrate-bound initiator centers (“grafting from”). We apply this technique to prepare grafting density gradients of poly(acrylamide) (PAAm) on flat silica substrates. We demonstrate that using the grafting density gradient geometry, the mushroom-to-brush transition can be accessed on a single sample. This transition is detected by monitoring the dependence of the thickness of the grafted PAAm in a good solvent using variable angle spectroscopic ellipsometry. Wettability experiments performed on the gradient PAAm substrate provide complementary information about the nature of the mushroom-to-brush transition.}, number={7}, journal={MACROMOLECULES}, author={Wu, T and Efimenko, K and Vlcek, P and Subr, V and Genzer, J}, year={2003}, month={Apr}, pages={2448–2453} } @article{tomlinson_genzer_2003, title={Formation of grafted macromolecular assemblies with a gradual variation of molecular weight on solid substrates}, volume={36}, ISSN={["1520-5835"]}, DOI={10.1021/ma025937u}, abstractNote={The development of new polymeric materials and structures relies primarily on controlled polymerizations, chain growth polymerizations that proceed without irreversible chain transfer, and chain termination. Recent advances in living radical polymerizations have led to the advancement of robust and simple polymerization methodologies. In particular, atom transfer radical polymerization (ATRP), pioneered by Sawamoto1 and Matyjaszewski,2 provides a convenient means of synthesizing end-functionalized and body-functionalized polymers, thereby enabling the synthesis of a variety of polymer architectures, including block copolymers, multiarmed stars, hyperbranched polymers, and polymer combs with well-defined composition and relatively low molecular weight polydispersities.3 Many types of monomers including acrylates, methacrylates, styrenes, vinylpyridines, acrylonitrile, and acrylamides have been polymerized successfully via ATRP. Continuous molecular gradients represent chief tools for combinatorial chemistry and materials science.4,5 These multivariant methods enable systematic variation of one or more physicochemical properties, thus enabling systematic exploration of the broad parameter space, improved efficiency, and lower cost.6 Surface-tethered polymer structures represent an effective means of tuning the physicochemical properties of substrates. Recently, techniques, involving the patterning of polymer layers grafted to the substrate, have been developed that utilize ATRP.7-14 The latter group of technologies is based on selectively decorating the material surfaces with polymerization initiators and then performing the polymerization directly on the surface (“grafting from”). In this Communication we report on creating surfacegrafted layers with a gradual variation of polymer molecular weight (or length). We use ATRP of methyl methacrylate (MMA) from water/methanol solution. Previous reports indicated that such polymerization proceeds rapidly even at a room temperature.11,15 Surface-bound macromolecular assemblies with position-dependent molecular weight were prepared in a custom-designed apparatus shown schematically in Figure 1. Silicon wafer (≈1 × 5 cm) was first covered with (11-(2-bromo-2-methyl)propionyloxy)undecyltri* Corresponding author: e-mail Jan_Genzer@ncsu.edu; Tel +1-919-515-2069. Figure 1. Schematic of the apparatus for creating surfacegrafted polymer assemblies with a gradient in molecular weight. The chamber is loaded with a solution comprising a monomer, bipyridine, CuCl2, and the solvent. The chamber is then purged with nitrogen in order to remove oxygen. CuCl is added, and the silicon wafer, covered with a chemisorbed ATRP initiator, is lowered into the solution. During the reaction, the micropump removes the solution from the chamber causing a steady decrease in the vertical position of the three-phase (substrate/solution/inert) contact line. The profile (including the “steepness”) of the polymer brush length gradient on the substrate can be controlled by varying the removal rate. Volume 36, Number 10 May 20, 2003}, number={10}, journal={MACROMOLECULES}, author={Tomlinson, MR and Genzer, J}, year={2003}, month={May}, pages={3449–3451} } @article{tomlinson_genzer_2003, title={Formation of surface-grafted copolymer brushes with continuous composition gradients}, ISSN={["1364-548X"]}, DOI={10.1039/b303823h}, abstractNote={We report on a simple methodology that facilitates the generation of surface-grafted assemblies comprising block copolymers with tunable composition and molecular weight gradients along flat solid substrates.}, number={12}, journal={CHEMICAL COMMUNICATIONS}, author={Tomlinson, MR and Genzer, J}, year={2003}, pages={1350–1351} } @article{fischer_efimenko_bhat_sambasivan_genzer_2004, title={Mapping surface chemistry and molecular orientation with combinatorial near-edge X-ray absorption fine structure spectroscopy}, volume={25}, ISSN={["1521-3927"]}, DOI={10.1002/marc.200300178}, abstractNote={AbstractSummary: Mapping the bond chemistry and molecular orientation of self‐assembled monolayer gradients on flat surfaces and reaction intermediates in catalyst arrays is made possible using combinatorial near‐edge X‐ray absorption fine structure (NEXAFS) spectroscopy. These spatially resolved NEXAFS maps have been made by utilizing synchrotron‐based NEXAFS spectroscopy in conjunction with a computer controlled precision sample manipulator. The NEXAFS maps reveal bond concentration, rehybridization, and orientation of the surface‐bound molecules with sub‐millimeter planar spatial resolution and sub‐monolayer molecular sensitivity. The wide applicability of the combinatorial NEXAFS method is illustrated by mapping: (1) the concentration and molecular orientation of semifluorinated molecules in molecular gradients; (2) the concentration of amino groups in molecular gradients used for nanoparticle templating, and (3) the rehybridization of propylene intermediates on zeolite catalyst arrays used for measuring solid‐state acidity and catalyst activity.The principle of “combinatorial” NEXAFS. The molecular gradient is indicated pictorially by the shading on the specimen.magnified imageThe principle of “combinatorial” NEXAFS. The molecular gradient is indicated pictorially by the shading on the specimen. }, number={1}, journal={MACROMOLECULAR RAPID COMMUNICATIONS}, author={Fischer, DA and Efimenko, K and Bhat, RR and Sambasivan, S and Genzer, J}, year={2004}, month={Jan}, pages={141–149} } @article{semler_genzer_2003, title={Monte Carlo simulations of copolymer adsorption at planar chemically patterned surfaces: Effect of surface domain sizes}, volume={119}, ISSN={["0021-9606"]}, DOI={10.1063/1.1597872}, abstractNote={We present results of Monte Carlo simulation studies utilizing the bond fluctuation model in conjunction with single and configurational biased Monte Carlo moves to investigate the adsorption of diblock (A–b–B) and alternating (A–alt–B) copolymers at physically flat surfaces made of an equal number of two chemically different sites, C and D. The adsorption of the copolymer to the surface is driven by the repulsion between the A and B segments along the copolymer and the attraction between the B segments and the D sites on the surface. We address the critical role of the commensurability between the copolymer’s monomer sequence distribution and the size and spatial distribution of the surface adsorbing sites on the copolymer adsorption. We show that both copolymer architectures have the ability to recognize the surface motif and transcribe it into the bulk material. Diblock copolymers can transfer the pattern once the heterogeneous domain sizes match the size of the parallel component to the radius of gyration, which is constituted primarily of the adsorbing species. This behavior results from the ability of the diblock copolymer to adopt a brush type conformation. In contrast to the diblocks, copolymers with the alternating sequence distribution are more likely to “zip to” the surface since the adsorbing species are evenly distributed along the copolymer. This chain conformation creates an entropic penalty, which must be alleviated by the formation of loops and tails. These conformational changes endow the alternating copolymer with the ability to recognize patterns with periodicities much less than the parallel component to the radius of gyration, and to invert the pattern as the distance away from the surface is increased.}, number={10}, journal={JOURNAL OF CHEMICAL PHYSICS}, author={Semler, JJ and Genzer, J}, year={2003}, month={Sep}, pages={5274–5280} } @article{genzer_kramer_fischer_2002, title={Accounting for Auger yield energy loss for improved determination of molecular orientation using soft x-ray absorption spectroscopy}, volume={92}, ISSN={["0021-8979"]}, DOI={10.1063/1.1516258}, abstractNote={Partial (Auger) yield near edge x-ray absorption fine structure (NEXAFS) is a structural analytical technique that has been primarily used to measure the spatial orientation and chemical bonding of small molecules on solid (i.e., inorganic or semiconductor) surfaces. In this article we demonstrate that the building block (BB) scheme proposed by Outka and co-workers [Phys. Rev. Lett. 59, 1321 (1987)] for analyzing NEXAFS spectra can be applied to model the molecular orientation of larger molecules, provided one accounts properly for kinetic energy losses of the Auger electrons traversing through the sample and hence the attenuation in measured Auger yield. We test the applicability of the proposed “modified” BB (MBB) model by measuring the orientation of a self-assembled monolayer (SAM) of –O1.5Si–(CH2)2–(CF2)8F, SF–SAM (SiOx), deposited on top of SiOx-covered silicon wafer as a function of the entrance grid bias (EGB) of the channeltron photoelectron detector. Our measurements of the EGB-dependent electron escape depth reveal that a crude depth profiling within the top ≈5 nm of the sample is possible by increasing the negative EGB on the channeltron detector, at the highest bias thus selecting only the Auger electrons, which have suffered negligible energy loss. In addition, we discuss how the order parameter method introduced recently by Stöhr and Samant [J. Electron Spectrosc. Relat. Phenom. 98–99, 189 (1989)] can be used to determine the molecular orientation of large organic molecules on surfaces. We also show that by accounting for energy losses of the NEXAFS Auger electrons (attenuation of measured Auger yield), the corrected order parameter (COP) approach gives good estimates of the orientation of molecules. We present a comparison between the MBB and COP models using experimental data collected from NEXAFS experiments from semifluorinated (SF) mesogens, –(CH2)x(CF2)yF, which are attached to: (1) the isoprene backbone of polyisoprene or a styrene–isoprene diblock copolymer and (2) a SiOx-covered solid substrate. We show that on both surfaces, the SF groups are oriented and on average are tilted by an angle 〈τF-helix〉 from the sample normal. We show that at higher 〈τF-helix〉 the results from the COP approach agree almost quantitatively with those extracted using the MBB model.}, number={12}, journal={JOURNAL OF APPLIED PHYSICS}, author={Genzer, J and Kramer, EJ and Fischer, DA}, year={2002}, month={Dec}, pages={7070–7079} } @misc{composto_walters_genzer_2002, title={Application of ion scattering techniques to characterize polymer surfaces and interfaces}, volume={38}, ISSN={["1879-212X"]}, DOI={10.1016/S0927-796X(02)00009-8}, abstractNote={Ion beam analysis techniques, particularly elastic recoil detection (ERD) also known as forward recoil spectrometry (Frcs) has proven to be a value tool to investigate polymer surfaces and interfaces. A review of ERD, related techniques and their impact on the field of polymer science is presented. The physics of the technique is described as well as the underlying principles of the interaction of ions with matter. Methods for optimization of ERD for polymer systems are also introduced, specifically techniques to improve the depth resolution and sensitivity. Details of the experimental setup and requirements are also laid out. After a discussion of ERD, strategies for the subsequent data analysis are described. The review ends with the breakthroughs in polymer science that ERD enabled in polymer diffusion, surfaces, interfaces, critical phenomena, and polymer modification.}, number={3-4}, journal={MATERIALS SCIENCE & ENGINEERING R-REPORTS}, author={Composto, RJ and Walters, RM and Genzer, J}, year={2002}, month={Jul}, pages={107–180} } @article{wu_efimenko_genzer_2002, title={Combinatorial study of the mushroom-to-brush crossover in surface anchored polyacrylamide}, volume={124}, ISSN={["0002-7863"]}, DOI={10.1021/ja027412n}, abstractNote={We present a method for fabricating anchored polymers with a gradual variation of grafting densities on solid substrates. The technique for generating such structures comprises (i) formation of a molecular gradient of polymerization initiator on the solid substrate and (ii) polymerization from the substrate-bound initiator centers ("grafting from"). We measure the mushroom-to-brush transition in grafted polyacrylamides and show that the mushroom and brush behavior can be described using existing scaling theories.}, number={32}, journal={JOURNAL OF THE AMERICAN CHEMICAL SOCIETY}, author={Wu, T and Efimenko, K and Genzer, J}, year={2002}, month={Aug}, pages={9394–9395} } @article{schultz_hall_genzer_2002, title={Computer simulation of copolymer phase behavior}, volume={117}, ISSN={["1089-7690"]}, DOI={10.1063/1.1519839}, abstractNote={Discontinuous molecular dynamics simulation is used to study the phase behavior of diblock copolymers modeled as chains of tangent hard spheres with square shoulder repulsions between unlike species as a function of chain length, volume fraction and interaction strength (χ). The location of the order–disorder transition for a symmetric copolymer is close to the predictions of Fredrickson and Helfand. Our simulation results for packing fractions of 0.35, 0.40, and 0.45 and chain lengths 10 and 20 are summarized in phase diagrams which display disordered, lamellae, perforated lamellae (PL), cylindrical, and BCC spherical (S) phases in the χN versus f plane. These phase diagrams are consistent with phase diagrams from other simulation studies. Contrary to theoretical predictions we observe the PL phase near regions of predicted gyroid stability, and the S phase only in the systems with high packing fraction and long chain length. These discrepancies may be due to the short chain lengths considered, as they are less evident in the 20-bead chains than the 10-bead chains. We examine the structural spacing of the microphases and the variation of that spacing with χN. We also examine the internal energy and entropy and their variation with χN. Our results are consistent with self-consistent field theory results for the strong segregation limit.}, number={22}, journal={JOURNAL OF CHEMICAL PHYSICS}, author={Schultz, AJ and Hall, CK and Genzer, J}, year={2002}, month={Dec}, pages={10329–10338} } @article{genzer_2002, title={Copolymer adsorption on planar chemically heterogeneous substrates: The interplay between the monomer sequence distribution and interaction energies}, volume={11}, ISSN={["1521-3919"]}, DOI={10.1002/1521-3919(20020601)11:5<481::AID-MATS481>3.0.CO;2-G}, abstractNote={We use three-dimensional self-consistent field model to study the adsorption of A-B copolymers from A-B copolymer/A homopolymer blends on planar substrates comprising two chemically distinct regions C and D. The interplay between the spatial distribution of the surface chemical heterogeneities and the monomer sequence distribution in the copolymer is examined for diblock (A-B), triblock (A-B-A), inverted triblock (B-A-B), and altermating (A-a/t-B) copolymer. Our results demonstrate that when the chemically heterogeneous motifs on the substrate are detected by the copolymer adsorbing segments, the copolymers can transcribe them with high fidelity into three dimensions. The way the surface pattern gets transferred is dictated by the monomer sequence distribution. We show that relative to alternating copolymers, block copolymers are generally better at capturing the chemical pattern shape and transcribing it into the polymer mixture. Moreover, block copolymers with shorter adsorbing blocks are capable of better recognizing the substrate motifs. In order to address the interplay between the monomer sequence distribution in the copolymer and the interaction energies, we systematically vary the repulsion between A and B, and the attraction between B and D. Our calculations reveal that reveal that increasing i) the interaction between the copolymer I adsorbing segments. (B) and the sticky points at the substrate (D), and/or ii) the repuilsion between the copolymer segments (A and B) increases the total amount of the copolymer adsorbed at the mixture/substrate interface, and decreases (increases) the fidelity of the substrate chemical pattern recognition by compositionally symmetric (asymmetric) copolymers.}, number={5}, journal={MACROMOLECULAR THEORY AND SIMULATIONS}, author={Genzer, J}, year={2002}, month={Jun}, pages={481–493} } @misc{bhat_fischer_genzer_2002, title={Fabricating planar nanoparticle assemblies with number density gradients}, volume={18}, ISSN={["0743-7463"]}, DOI={10.1021/la025524m}, abstractNote={We report on preparing assemblies of gold nanoparticles with continuous gradients in number density on flat silica-covered substrates. The methodology consists of (i) first forming a one-dimensional molecular gradient of amino groups (−NH2) on the substrate by vapor deposition of amine-terminated silane molecules, followed by (ii) attachment of gold nanoparticles to −NH2 functional groups by immersing the substrate in a colloidal gold solution. Experiments using atomic force microscopy reveal that the number density of nanoparticles on the substrate varies continuously as a function of the position on the substrate. Near-edge X-ray absorption fine structure studies confirm that the nanoparticle number density gradient is closely correlated with the concentration gradient of −NH2 groups anchored to the substrate. We demonstrate that the number density of nanoparticles within the gradient and the length of the gradient can be tuned by controlling the vapor diffusion of silane molecules. In addition we show ...}, number={15}, journal={LANGMUIR}, author={Bhat, RR and Fischer, DA and Genzer, J}, year={2002}, month={Jul}, pages={5640–5643} } @article{efimenko_novick_carbonell_desimone_genzer_2002, title={Formation of self-assembled monolayers of semifluorinated and hydrocarbon chlorosilane precursors on silica surfaces from liquid carbon dioxide}, volume={18}, ISSN={["0743-7463"]}, DOI={10.1021/la011813j}, abstractNote={We report on the formation and properties of self-assembled monolayers (SAMs) prepared by depositing semifluorinated and hydrocarbon trichlorosilane precursors, F(CF2)8(CH2)2SiCl3 (F8H2) and H(CH2)18SiCl3 (H18), respectively, from vapor, organic solvent, and liquid CO2 (l-CO2). Contact angle measurements of the SAM deposition kinetics reveal that regardless of the molecule type, the deposition rates from l-CO2 exceed those from vapor or organic solvents by several orders of magnitude. We derive two different transport models describing the formation of SAMs. We show that while the diffusion-limited model is not capable of describing the experimental data, the adsorption-limited model captures the major features of the adsorption kinetics quite well. We apply the results of the adsorption-limited model to conclude that the observed behavior is a consequence of (i) a relatively high bulk concentration (l-CO2 vs vapor) and (ii) higher solution diffusivity (l-CO2 vs organic solvent) of the silanes in l-CO2. N...}, number={16}, journal={LANGMUIR}, author={Efimenko, K and Novick, B and Carbonell, RG and DeSimone, JM and Genzer, J}, year={2002}, month={Aug}, pages={6170–6179} } @article{genzer_efimenko_fischer_2002, title={Molecular orientation and grafting density in semifluorinated self-assembled monolayers of mono-, di-, and trichloro silanes on silica substrates}, volume={18}, ISSN={["0743-7463"]}, DOI={10.1021/la025921x}, abstractNote={Near-edge X-ray absorption fine structure (NEXAFS) spectroscopy is used to measure the molecular orientation in semifluorinated self-assembled monolayers (SAMs) prepared by vapor deposition of mono- (F3C(CF2)8(CH2)2Si(CH3)2Cl, m-F8H2), di- (F3C(CF2)8(CH2)2Si(CH3)Cl2, d-F8H2), and trichloroorganosilanes (F3C(CF2)8(CH2)2SiCl3, t-F8H2) on flat silica-covered substrates. The average tilt angles (from the sample normal) of the fluorocarbon part, F(CF2)8-, of t-F8H2, d-F8H2, and m-F8H2 measured by carbon K-edge NEXAFS are 10 ± 2°, 35 ± 2°, and 45 ± 3°, respectively. We show that the increase of the tilt angle is associated with the steric hindrance of the methyl groups attached to silicon close to the bonding substrate. We also show that the molecular orientation obtained from the NEXAFS measurements can be used to estimate the grafting densities of the F8H2 molecules on the substrates. We present a simple one-dimensional geometric model to show that the grafting density of m-F8H2 is approximately one-half of t...}, number={24}, journal={LANGMUIR}, author={Genzer, J and Efimenko, K and Fischer, DA}, year={2002}, month={Nov}, pages={9307–9311} } @article{efimenko_wallace_genzer_2002, title={Surface modification of Sylgard-184 poly(dimethyl siloxane) networks by ultraviolet and ultraviolet/ozone treatment}, volume={254}, DOI={10.1006/jcis.202.8594}, number={2}, journal={Journal of Colloid and Interface Science}, author={Efimenko, Kirill and Wallace, W. E. and Genzer, Jan}, year={2002}, pages={306–315} } @article{genzer_2001, title={Copolymer adsorption on planar substrates with a random distribution of chemical heterogeneities}, volume={115}, ISSN={["1089-7690"]}, DOI={10.1063/1.1391254}, abstractNote={We use a 3D SCF model of polymer adsorption to investigate the adsorption of A-B copolymers from A-B/homopolymer A mixtures onto planar substrates composed of two chemically distinct randomly distributed sites, one of which has a preferential affinity for the B segments of the copolymer. Our results show that when the chemically heterogeneous substrate motifs are recognized by the copolymer, the copolymers can transcript them with a relatively high fidelity into three dimensions. The way the surface motif is transferred is strongly dictated by the copolymer sequence. We show that block copolymers are capable of detecting small clusters of the substrate adsorption sites. The fidelity of the pattern shape and the distance from the substrate to which the pattern gets transferred increases with decreasing the length of the adsorbing block of the copolymer. Our results also indicate that increasing (i) the interactions between the copolymer adsorbing segments and the “sticky” points at the substrate, and/or (ii) the repulsion between the copolymer segments increases the total adsorbed amount of the copolymer at the mixture/substrate interface but it decreases the fidelity of the substrate chemical pattern transfer into the mixture. We show that, in contrast to the block copolymers, macromolecules with alternating sequence distributions adopt different conformations on random substrates in that they tend to localize at the boundaries between the C/D surface sites, where the substrate chemical pattern more closely matches the sequence distribution of the B stickers along the copolymer. We claim that this feature allows us to use alternating copolymers in situations where one needs to suppress the chemical pattern transfer on such random substrates.}, number={10}, journal={JOURNAL OF CHEMICAL PHYSICS}, author={Genzer, J}, year={2001}, month={Sep}, pages={4873–4882} } @article{genzer_2001, title={Copolymer-assisted generation of three-dimensional patterns by replicating two-dimensional substrate motifs - art. no. 022601}, volume={6302}, number={2}, journal={Physical Review. E, Statistical Physics, Plasmas, Fluids, and Related Interdisciplinary Topics}, author={Genzer, J.}, year={2001}, pages={2601} } @article{efimenko_genzer_2001, title={How to prepare tunable planar molecular chemical gradients}, volume={13}, ISSN={["0935-9648"]}, DOI={10.1002/1521-4095(200110)13:20<1560::AID-ADMA1560>3.0.CO;2-Z}, abstractNote={Mechanically assembled monolayers (MAMs) can be employed to adjust the surface density of the initiators and consequently that of molecular gradients prepared on flexible supports. The steepness and the position of the gradient of molecules on a substrate are fine-tuned by a combination of grafting and mechanical manipulation of the films. Planar molecular gradients ranging from several millimeters to several centimeters can thus be fabricated. In addition, the wetting properties of the hydrophobic part of the substrate can be adjusted by altering the chemical nature of the films.}, number={20}, journal={ADVANCED MATERIALS}, author={Efimenko, K and Genzer, J}, year={2001}, month={Oct}, pages={1560-+} } @article{wallace_fischer_efimenko_wu_genzer_2001, title={Polymer chain relaxation: Surface outpaces bulk}, volume={34}, ISSN={["1520-5835"]}, DOI={10.1021/ma002075t}, abstractNote={Introduction: The time scale that controls polymer chain relaxation profoundly influences the economics of industrial plastics processing. However, not much is known about how fast the surface relaxes relative to the bulk in amorphous polymeric materials. This is in part because no study has reported simultaneous measurement of both surface and bulk chain relaxation dynamics on a single macroscopic sample. In this work we show how carbon near-edge X-ray absorption fine structure (NEXAFS) can be applied to detect both surface and bulk segmental relaxation in uniaxially deformed polystyrene samples. We show that by simultaneously monitoring the partial electron yield (PEY, kinetic energy > 150 eV) and the fluorescence yield (FY, carbon Kα 277eV) NEXAFS signals, with probing depths of approximately 2 and 200 nm, respectively, one can examine segmental motion throughout a polymer sample. Using this methodology, chain relaxation is found to occur approximately 30% faster at the surface than in the bulk. Methods and Materials: Rectangular samples (12.5 x 12.5 x 6 mm), prepared by vacuum hot-pressing at 150 °C monodisperse polystyrene (M = 228,000 g/mole), were placed into a steel channel die and uniaxially elongated at room temperature along one of the long dimensions to about 130% of their original length. This sample preparation allowed us to examine true bulk samples as encountered in industrial practice and to avoid thin film effects that may alter the free surface behavior. The NEXAFS experiments were conducted on the U7A NIST/Dow Materials Characterization end-station at the National Synchrotron Light Source at Brookhaven National Laboratory. For each sample the PEY and FY carbon K-edge NEXAFS intensities were measured with the incident polarized X-ray beam normal to the sample surface and at two azimuthal sample orientations:with the electric field vector, E, parallel (φ = 0°) and perpendicular (φ = 90°) to the elongation direction as shown in Fig. 1. Results: Orientation of the chain backbone was determined by monitoring the C=C phenyl ring 1s → π1* NEXAFS resonance intensity at 285.5 eV, which involves the excitation of carbon 1s electrons to the unfilled π* antibonding orbitals of the phenyl ring. Enhancement of the 1s → π1* resonance intensity was observed when E was parallel to the elongation direction. Since the phenyl π* orbitals are oriented normal to the phenyl rings, and the phenyl rings, free to rotate around the pendant bond, will have a component normal to the chain axis, the intensity of the π* signal has been shown to be an unambiguous signature of backbone orientation. This orientation, seen both in the PEY and the FY NEXAFS signals of the elongated samples, provides clear evidence of chain orientation at the outset of the experiment. A direct measure of the chain relaxation rates at the surface and in the bulk can be obtained by defining an orientation factor, OF (see Fig. 1 caption) that is evaluated from the time dependence of the 1s → π1* resonance intensity in the PEY and FY NEXAFS spectra, respectively, during annealing. Figure 1 shows that while the OF for both surface and bulk chains decays as a function of increasing annealing time at 60 °C, the surface orientation is initially greater and decays faster than for the bulk. Fitting the decay rates to exponential functions gives characteristic time constants of approximately 33 and 43 minutes for the surface and the bulk, respectively. Conclusions: These results show conclusively that polystyrene surface chain relaxation dynamics are significantly faster than the bulk. Recalling that chain relaxation rates in amorphous polymers increase with decreasing number of entanglements, our results are in agreement with recent theory that predicts the number of entanglements at the free surface to be significantly reduced compared to the bulk. Acknowledgments: This research was supported by NSF CAREER (JG) and NCSU FR&PD Grant (JG). Figure 1. Time evolution of the orientation factor, OF, from an elongated (oriented) polystyrene sample reveals that when annealed at 60 °C the surface chains relax to an equilibrium (unoriented) configuration faster than the bulk chains. OF is calculated from (I|| I⊥)/(I|| + I⊥), where I|| and I⊥ are the 1s → π1* resonance NEXAFS intensities collected with the sample elongation direction parallel (φ = 0°) and perpendicular (φ = 90°) to the electric vector of the soft Xray beam, E, respectively. 0 40 80 120 10 10 10}, number={15}, journal={MACROMOLECULES}, author={Wallace, WE and Fischer, DA and Efimenko, K and Wu, WL and Genzer, J}, year={2001}, month={Jul}, pages={5081–5082} } @article{wu_efimenko_genzer_2001, title={Preparing high-density polymer brushes by mechanically assisted polymer assembly}, volume={34}, ISSN={["1520-5835"]}, DOI={10.1021/ma001750w}, abstractNote={It is well-known that when polymers are endanchored in a sufficient concentration to a substrate, they form a so-called polymer “brush” whereby the chains are stretched with respect to their preferred configuration away from the interface. The characteristics of polymer brushes have been analyzed using a variety of theoretical methods and experimental probes and are now fairly well established.1-4 For example, the thickness of the brush layer, H, is known to depend linearly on the number of repeat units of the polymer, N, and on the power law of the brush grafting density at the substrate, σ, with the exponent of /3 or 1, depending on the surface coverage.1,2 To fine-tune the polymer brush properties, one needs to have a good control over H and σ. While H can be adjusted by simply varying the polymerization time, monomer concentration, and monomer conversion, σ depends on the methods by which polymer brushes are formed. Previous reports established that polymer brushes with moderately high grafting densities can be prepared by harnessing the “grafting from” principle in which the polymer chains are synthesized using radical initiators that are covalently bound to the substrate. A vast majority of experiments involved classical radical growth methods with either azo(e.g., AIBN) or peroxide-based initiators that either were created directly on the substrate or were attached to the substrate via selfassembly.5-11 While the advantage of such methods is their ease of use and the ability to prepare brushes from a variety of monomers, a major disadvantage is a hardto-control polymerization process, which usually leads to brushes with broad molecular weight distributions. Recently several reports appeared that described the formation of polymer brushes using surface-initiated “living” radical polymerization,12 such as the atom transfer radical polymerization (ATRP).13,14 Because of its simplicity, robustness, and the ability to synthesize polymers with narrow molecular weight distributions, ATRP has been the method of choice for most surfaceinitiated “living” radical polymerization processes.15-21 As mentioned previously, one of the crucial parameters governing the behavior of polymer brushes is their grafting density at the polymer/substrate interface. The “grafting from” techniques offer a fairly good control over the grafting density of the polymer brushsin the ideal case σ is simply equal to the surface density of the polymerization initiators. While a relatively high density of initiators can be achieved by assembling the molecules on the surface by means of LangmuirBlodgett (LB) techniques17,20 or by forming organized self-assembled monolayers (SAMs),19 tailoring the grafting density of the SAM chains is not an easy task. SAMs are usually formed through self-assembly processes that are governed by the chemical and structural nature of the SAM molecules and the means of their attachment to the substrate. To overcome this limitation, one would need to seek another way of controlling the grafting density of the surface initiators that is independent of the system thermodynamics. Recently, we have demonstrated that one can finetune the grafting density of molecules on surfaces by fabricating MAMs (“mechanically assembled monolayers”), structures that are based on the combination of natural self-assembly and mechanical manipulation of the grafted molecules on surfaces.22 In this work we utilize the MAMs to adjust the surface density of the initiators and consequently that of polymer brushes grown by surface-initiated ATRP. Following the steps involved in the preparation procedure (see below), we call this technique MAPA (“mechanically assisted polymer assembly”). In this communication we illustrate the principles of MAPA on preparing dense polymer brushes of polyacrylamide (PAAm).15 Figure 1 shows schematically how PAAm-MAPA works. First, a pristine PDMS network film was prepared (cf. Figure 1a) by casting a mixture of PDMS and a cross-linker24 into a thin (≈1 mm) film and curing it at 70 °C for about an hour. The film was then cut into small strips (≈1 × 5 cm2) and mechanically uniaxially elongated by ∆x (cf. Figure 1b). Subsequent exposure to UV/ozone (UVO) treatment produced hydrophilic PDMS surfaces (PDMS-UVO) composed mainly of hydroxyl groups (HO-[Si]surface) that served as attachments points for chlorosilane-based ATRP initiators. Following previous work on PAAm brushes, we used 1-trichlorosilyl-2-(m-p-chloromethylphenyl)ethane (CMPE) (United Chemical Technologies, Inc.) as the initiator. The CMPE molecules were deposited from vapor25 onto this stretched substrate (cf. Figure 1d) and formed an organized CMEP-SAM following the well-known set of reactions:}, number={4}, journal={MACROMOLECULES}, author={Wu, T and Efimenko, K and Genzer, J}, year={2001}, month={Feb}, pages={684–686} } @article{martinache_royer_siripurapu_henon_genzer_khan_carbonell_2001, title={Processing of polyamide 11 with supercritical carbon dioxide}, volume={40}, ISSN={["0888-5885"]}, DOI={10.1021/ie010410b}, abstractNote={The supercritical carbon dioxide induced swelling and plasticization of polyamide 11 were investigated. The swelling kinetics exhibit an initial region of large swelling, in which the diffusion of CO2 into the polymer follows Fickian behavior, and a subsequent region of small volume increase that asymptotically approaches an equilibrium swelling value. The diffusion coefficient of CO2 in polyamide 11 was calculated from the initial slope of the swelling kinetics data. CO2, injected up to 3 wt % using an extrusion rheometer, is shown to be an effective plasticizer for polyamide 11, lowering the viscosity of the polymer melt by as much as 50%. The use of CO2 as a blowing agent was also investigated, and we report preliminary foaming attempts using a batch process. We obtained homogeneously distributed foams, featuring well-defined closed cells with an average diameter of 30 μm that had an unfoamed skin layer.}, number={23}, journal={INDUSTRIAL & ENGINEERING CHEMISTRY RESEARCH}, author={Martinache, JD and Royer, JR and Siripurapu, S and Henon, FE and Genzer, J and Khan, SA and Carbonell, RG}, year={2001}, month={Nov}, pages={5570–5577} } @article{genzer_2001, title={Self-consistent field study of copolymer adsorption at planar chemically 'rough' surfaces: an interplay between the substrate chemical pattern and copolymer sequence distribution}, volume={94}, ISSN={["0001-8686"]}, DOI={10.1016/S0001-8686(01)00057-4}, abstractNote={We extend the one-dimensional self-consistent field (SCF) scheme of Scheutjens-Fleer to three dimensions (3D) and use this three-dimensional SCF model to investigate the adsorption of A–B copolymers from A homopolymer matrices onto planar substrates composed of two chemically distinct sites (C and D), one of which has a preferential affinity for the B segments of the copolymer. To address the role of the substrate chemical heterogeneity on copolymer adsorption, we keep the fraction of the C and D sites constant (50% of each site) and vary their spatial distribution on the substrate. The interplay between the surface chemical heterogeneity and the chain microstructure is examined for A–B diblock, A–B–A, B–A–B triblock, and A–alt–B alternating copolymers. Our results indicate that regardless of the type of the surface chemical heterogeneity, the A–B diblock and triblock copolymers adopt ‘brush’-like and ‘bridge’-like structures, respectively, with the B block being anchored to the substrate. In contrast, the A–alt–B macromolecule is found to be ‘zipped’ to the substrate. For a fixed chemical potential of the copolymer in the A–B/A mixture and the surface adsorption energy of B, the amount of the adsorbed copolymer depends on: (1) the number of the B segments; (2) the copolymer microstructure; and (3) the distribution of the C and D regions on the substrate. Three-dimensional maps of the spatial density of copolymer segments provide insight into copolymer conformation at the mixture/substrate interface and also the ability of the copolymer to recognize and mimic the substrate pattern. In addition, we discuss the circumstances under which the substrate pattern is transferred deep inside the A–B/A mixture and those, which lead to strong damping of the substrate motif as one moves away from the substrate/mixture interface.}, number={1-3}, journal={ADVANCES IN COLLOID AND INTERFACE SCIENCE}, author={Genzer, J}, year={2001}, month={Nov}, pages={105–134} } @article{genzer_efimenko_2000, title={Creating long-lived superhydrophobic polymer surfaces through mechanically assembled monolayers}, volume={290}, ISSN={["1095-9203"]}, DOI={10.1126/science.290.5499.2130}, abstractNote={We show that elastomeric surfaces can be tailored using “mechanically assembled monolayers” (MAMs), structures that are fabricated by combining self-assembly of surface grafting molecules with mechanical manipulation of the grafting points in the underlying elastic surface. The versatility of this surface modification method is demonstrated by fabricating MAMs with semifluorinated (SF) molecules. These SF-MAMs have superior nonwetting and barrier properties in that they are “superhydrophobic” and nonpermeable. We also establish that these material characteristics do not deteriorate even after prolonged exposure to water, which usually causes surface reconstruction in conventionally prepared SF self-assembled monolayers.}, number={5499}, journal={SCIENCE}, author={Genzer, J and Efimenko, K}, year={2000}, month={Dec}, pages={2130–2133} } @article{manias_chen_krishnamoorti_genzer_kramer_giannelis_2000, title={Intercalation kinetics of long polymers in 2 nm confinements}, volume={33}, ISSN={["0024-9297"]}, DOI={10.1021/ma0009552}, abstractNote={The motion of confined polymers is measured experimentally between parallel, atomically smooth solid surfaces, separated by 2 nm. In particular, the kinetics of intercalation of monodispersed polystyrene in alkylammonium modified mica-type silicates were studied using X-ray diffraction. The kinetics of the neat polymers and their functionalized derivatives were measured as functions of molecular weight, extent of functionalization, and silicate surface organic modification, at various temperatures. Selective corroborative studies were also performed using in-situ small-angle neutron scattering (SANS). The kinetic data are interpreted in terms of an effective diffusion coefficient (Deff) of the polymer, which undergoes a dramatic decrease with stronger silicate surface-polymer interaction. This interaction is varied by either changing the silicate surface modification or by increasing the extent of functionalization of the polystyrene chains. Furthermore, the diffusion coefficient exhibits an inverse dependence on chain length (N), i.e., Deff ∝ N -1 , for chains up to 900 000 molecular weight.}, number={21}, journal={MACROMOLECULES}, author={Manias, E and Chen, H and Krishnamoorti, R and Genzer, J and Kramer, EJ and Giannelis, EP}, year={2000}, month={Oct}, pages={7955–7966} } @article{genzer_sivaniah_kramer_wang_xiang_char_ober_bubeck_fischer_graupe_et al._2000, title={Molecular orientation of single and two-armed monodendron semifluorinated chains on "soft" and "hard" surfaces studied using NEXAFS}, volume={33}, ISSN={["0024-9297"]}, DOI={10.1021/ma991710w}, abstractNote={Near-edge absorption fine structure (NEXAFS) measurements are used to probe the molecular orientation of semifluorinated (SF) mesogens, −(CH2)x(CF2)yF, which are attached to (i) the isoprene backbone of polyisoprene or a styrene-isoprene diblock copolymer (“soft” substrate), and (ii) a Au-covered solid substrate via a thiol link (“hard” substrate). The SF groups on both surfaces are oriented and on average are tilted from the sample normal. The tilt angle, 〈τF-helix〉, of the fluorinated part of the SF group on each substrate is determined exclusively by the combination of x and y, increasing with increasing x and with decreasing y. Moreover, 〈τF-helix〉 is found to be independent of the surface topology (flat surfaces vs surfaces covered with holes or islands of the copolymer), casting solvent, and the architecture of the SF group (single vs 2-armed monodendron). Comparing the orientation of the SF groups on both substrates reveals that 〈τF-helix〉 is approximately 14° higher on the “soft” substrate.}, number={16}, journal={MACROMOLECULES}, author={Genzer, J and Sivaniah, E and Kramer, EJ and Wang, JG and Xiang, ML and Char, K and Ober, CK and Bubeck, RA and Fischer, DA and Graupe, M and et al.}, year={2000}, month={Aug}, pages={6068–6077} } @misc{genzer_efimenko_2000, title={Tailoring the grafting density of organic modifiers at solid/liquid interfaces}, volume={6,423,372}, number={2000 Dec 13}, publisher={Washington, DC: U.S. Patent and Trademark Office}, author={Genzer, J. and Efimenko, K.}, year={2000} } @article{genzer_sivaniah_kramer_wang_korner_xiang_char_ober_dekoven_bubeck_et al._2000, title={The orientation of semifluorinated alkanes attached to polymers at the surface of polymer films}, volume={33}, ISSN={["0024-9297"]}, DOI={10.1021/ma991182o}, abstractNote={The surface molecular orientation of a liquid crystalline (LC) layer made up of semifluorinated (SF) single side groups [−CO−(CH2)x-1−(CF2)yF] (single SF groups) attached to polyisoprene homopolymer or the isoprene block of a styrene−isoprene diblock copolymer was determined by analyzing the partial electron yield C-edge NEXAFS signal. The results show that the surfaces of thin SF polymer films are covered with a uniform layer, consisting of the SF−LC groups whose average −CF2− tilt angle with the surface normal lies in the range 29−46°. This is in direct contrast to the bulk, where the directors of the SF−LC mesogens are aligned parallel to the polystyrene/SF−polyisoprene interface of the block copolymers. This average tilt angle increases with increasing the length of the −(CH2)x-1− group (x increases) but decreases with increasing the length of the −(CF2)y− part of the molecule (y increases) at constant x.}, number={5}, journal={MACROMOLECULES}, author={Genzer, J and Sivaniah, E and Kramer, EJ and Wang, JG and Korner, H and Xiang, ML and Char, K and Ober, CK and DeKoven, BM and Bubeck, RA and et al.}, year={2000}, month={Mar}, pages={1882–1887} } @article{oslanec_genzer_faldi_composto_garrett_1999, title={Surface enrichment in a miscible random copolymer blend: Influence of polydispersity and architecture}, volume={32}, ISSN={["0024-9297"]}, DOI={10.1021/ma980601z}, abstractNote={Neutron reflectivity and low-energy forward-recoil spectrometry were used simultaneously to monitor surface segregation in blends of poly(styrene-ran-acrylonitrile), dSAN23 and SAN27, having 23 and 27 wt % AN, respectively. Because of its lower AN content, dSAN23 was found to partition at the polymer/air interface, in agreement with the work of Mansfield et al. (Physica B 1991, 173, 207) and Kim et al. (Polymer 1995, 36, 2427). In contrast to the previous studies, this work explores surface segregation at high dSAN23 bulk volume fractions, φ∞ (i.e., φ∞ > 0.50). New results include the observation of a maximum dSAN23 surface excess (∼75 A) near φ∞ ≈ 0.35 and a dSAN23 surface volume fraction that approaches 1, rather than 0.75, as φ∞increases to 1. The dSAN23 profile deviates from the exponential profile predicted by Schmidt−Binder and self-consistent mean-field (SCMF) models. For φ∞ > 0.50, the profile displays a surface flattening not observed at low φ∞. The SCMF model is extended to account for polymer p...}, number={12}, journal={MACROMOLECULES}, author={Oslanec, R and Genzer, J and Faldi, A and Composto, RJ and Garrett, PD}, year={1999}, month={Jun}, pages={4098–4105} } @article{genzer_sivaniah_kramer_wang_korner_char_ober_dekoven_bubeck_fischer_et al._2000, title={Temperature dependence of molecular orientation on the surfaces of semifluorinated polymer thin films}, volume={16}, ISSN={["0743-7463"]}, DOI={10.1021/la9910327}, abstractNote={Near-edge X-ray absorption fine structure is used to investigate the temperature dependence of molecular orientation of semifluorinated liquid crystalline (SF-LC) mesogens, which are attached to the modified isoprene backbone of (i) a poly(1,2-isoprene) homopolymer and (ii) a diblock copolymer consisting of polystyrene and poly(1,2-isoprene) blocks. Our experiments reveal the existence of two temperature regions in which the surface orientation of the SF-LC mesogens changes abruptly, but even 30 K above the highest such temperature region the surface orientation does not become isotropic. The lower temperature surface transition for both homopolymer and block copolymer occurs close to the temperature of the bulk homopolymer smectic-B to smectic-A transition and well above the bulk smectic-B to smectic-A transition in the block copolymer. It seems to be controlled exclusively by the ordering phenomena originating from the surface. In contrast, the change in the surface organization of the SF-LC mesogens at...}, number={4}, journal={LANGMUIR}, author={Genzer, J and Sivaniah, E and Kramer, EJ and Wang, JG and Korner, H and Char, K and Ober, CK and DeKoven, BM and Bubeck, RA and Fischer, DA and et al.}, year={2000}, month={Feb}, pages={1993–1997} } @article{genzer_composto_1999, title={The interface between immiscible polymers studied by low-energy forward recoil spectrometry and neutron reflectivity}, volume={40}, ISSN={["0032-3861"]}, DOI={10.1016/s0032-3861(98)00664-8}, abstractNote={The superb depth resolution of low-energy, forward recoil, spectrometry, (LE-FRES), is demonstrated by measuring the interfacial profile between 1,4-polybutadiene (PB), and deuterated polystyrene (dPS), at 175°C. To enhance depth resolution, the top dPS layer is ‘thinned’ by ion sputtering, prior to LE-FRES analysis, to achieve a resolution of 13 nm at the interface. An interfacial width of 6.0±3.5 nm is measured. To compliment this study, neutron reflectivity (NR) is used to determine the interfacial volume fraction profile between PB and an isotopic mixture of polystyrene (dPS:PS) having a dPS volume fraction of 0.30. After annealing at 175°C, the interface broadens to 3.3 nm. However, no interfacial segregation of dPS or PS is observed suggesting that the difference between the PB–dPS and PB–PS interaction parameters (χ) is insufficient to drive interfacial segregation. Using a self-consistent field model (SCF) and known values of χ, the width is calculated to be 2.2 nm. By adding capillary wave broadening, this width increases to 2.8 nm, in better agreement with the NR value. In agreement with the NR measurement, the SCF model predicts no segregation of dPS or PS.}, number={15}, journal={POLYMER}, author={Genzer, J and Composto, RJ}, year={1999}, month={Jul}, pages={4223–4228} } @article{heier_genzer_kramer_bates_walheim_krausch_1999, title={Transfer of a chemical substrate pattern into an island-forming diblock copolymer film}, volume={111}, ISSN={["0021-9606"]}, DOI={10.1063/1.480469}, abstractNote={We investigate the transfer of a chemical pattern on a substrate into a symmetric diblock copolymer thin film of poly(styrene-2-vinylpyridine) (PS-PVP). The substrates have patterns of self-assembled monolayers (SAMs) produced by microcontact printing H3C-terminated (H3C-) SAM stripes alternating with HO-terminated (HO-) SAM stripes. The PS-PVP lamellae over the H3C-SAM have a defect structure that attracts excess PS-PVP that would normally form islands on a uniform HO-SAM stripe. We seek to understand the process that limits our ability to accommodate all excess polymers on top of the H3C-SAM. In the early stages of annealing, waves of thickness develop from the H3C/HO-SAM boundary and propagate into the film over the HO-SAM. For very short annealing times, the wavelength λ of these thickness waves is constant at any given time for all grating periodicities. Large amplitude patterns develop when λ=2d/(2n−1), where d is the width of the HO-SAM stripe and n is an integer ⩾1. Such patterns suggest constructive interference of the thickness waves and indeed much lower amplitudes over the HO-SAM stripes are observed when λ=d/n (destructive interference). This behavior seems close to that seen for surface-directed spinodal decomposition waves in thin films of binary polymer mixtures. We achieve more complete transfer of excess copolymers from the HO-SAM stripe to the H3C-SAM ones if the film is preordered under a confining layer that does not permit the formation of surface features.}, number={24}, journal={JOURNAL OF CHEMICAL PHYSICS}, author={Heier, J and Genzer, J and Kramer, EJ and Bates, FS and Walheim, S and Krausch, G}, year={1999}, month={Dec}, pages={11101–11110} } @article{genzer_heier_kramer_1999, title={Wetting reversal transition in phase-separated polymer mixtures}, volume={139}, ISSN={["1521-3900"]}, DOI={10.1002/masy.19991390109}, abstractNote={AbstractWe investigate a wetting reversal transition in thin films of two‐phase mixtures of poly(ethylene‐propylene) (PEP) and its deuterated analogue (dPEP) on a substrate covered by a self‐assembled monolayer (SAM) whose surface energy, γSAM, is tuned by varying the SAM composition. As γSAM increases from 21 to 24 mJ/m2, a transition from a three‐layer (air/dPEP/PEP/dPEP/SAM) to a two‐layer (air/dPEP/PEP/SAM) structure occurs at increasing Tc ‐ T, where Tc and T are the critical and transition temperatures, respectively. As the system structure changes from three‐layer to two‐layer, the thicknesses of the dPEP‐rich wetting layers at the air/mixture and mixture/SAM interfaces are found to smoothly increase and decrease, respectively, while the thickness of the PEP‐rich layer (ca. one half of the total film thickness) does not change. The dependence of the transition temperature on γSAM is predicted by a simple model using the experimental data on the surface energies of PEP/dPEP and estimates of the interfacial energy between PEP and dPEP.}, number={1999 Apr.}, journal={MACROMOLECULAR SYMPOSIA}, author={Genzer, J and Heier, J and Kramer, EJ}, year={1999}, month={Apr}, pages={77–85} } @article{genzer_composto_1998, title={Effect of molecular weight on the interfacial excess, tension, and width in a homopolymer binary polymer blend system}, volume={31}, ISSN={["0024-9297"]}, DOI={10.1021/ma970993u}, abstractNote={The interfacial properties of a three-component, two-phase A/B:C blend are investigated to understand the effect of the A and B chain lengths on the interfacial excess of B, zB*, the interfacial tension, γABC, and the interfacial width, wABC. The A/B:C components are polystyrene/poly(d8-styrene-co-4-bromostyrene):poly(styrene-co-4-bromostyrene), where B and C have 4-bromostyrene mole fractions of 0.154 and 0.177, respectively. Low-energy forward recoil spectrometry (LE-FRES) is used to measure zB* as a function of the B volume fraction in the B:C blend, φB∞. The experimental zB*'s are found to be in excellent agreement with those calculated using the self-consistent field (SCF) model of the A/B:C interface. In addition, the SCF model is used to evaluate γABC and the widths for the A/B:C, A/B, and A/C interfaces (i.e., wABC, wAB, and wAC, respectively). Our results demonstrate that increasing the number of B segments, NB, greatly increases the magnitude of zB*, particularly at low φB∞. On the other hand, v...}, number={3}, journal={MACROMOLECULES}, author={Genzer, J and Composto, RJ}, year={1998}, month={Feb}, pages={870–878} } @article{genzer_kramer_1998, title={Pretransitional thinning of a polymer wetting layer}, volume={44}, ISSN={["0295-5075"]}, DOI={10.1209/epl/i1998-00454-4}, abstractNote={Mixtures of two immiscible polymers A and B phase-separate into A-rich (α) and B-rich (β) phases. Thin films of such mixtures exhibit a transition from an equilibrium 3-layer structure (substrate/α/β/α/air) to a 2-layer structure (substrate/β/α/air) as the surface energy of the substrate is increased. In this letter we show that as an (equilibrium) precursor to this transition, the thickness of the α layer in the 3-layer structure at the substrate/polymer interface decreases, while that of the α layer against air increases until the α layer against the substrate vanishes leaving the 2-layer structure. This pretransitional behavior originates from the long-range nature of the van der Waals interaction between layers and is predicted by a simple model that considers the dependence of the free energies of the two α layers on their thicknesses.}, number={2}, journal={EUROPHYSICS LETTERS}, author={Genzer, J and Kramer, EJ}, year={1998}, month={Oct}, pages={180–185} } @article{genzer_composto_1997, title={A self-consistent field study of the wetting transition in binary polymer blends}, volume={106}, ISSN={["0021-9606"]}, DOI={10.1063/1.473222}, abstractNote={A self-consistent field approach is used to investigate the partial to complete wetting transition for an A:B polymer blend at coexistence where polymers A and B have equal numbers of segments, N. The surface free energy, Fs, is modeled using the quadratic form suggested by Schmidt and Binder [J. Phys. II (France) 46, 1631 (1985)], namely, Fs=−μφ1−0.5gφ12, where μ and g are the surface equivalents of the bulk chemical potential and interaction energy, respectively, and φ1 is the surface volume fraction of the surface preferred component (A). For selected values of g and the bulk volume fraction of A, φ∞, the volume fraction profile and A surface excess, z*, are calculated as a function of increasing μ. The first and second order wetting transitions are indicated by a discontinuity and divergence, respectively, of z* and φ1. In our simulations, at high values of φ∞ only first order transitions are detected for both N=100 and N=1000. However, both first and second order wetting transitions are observed for low values of φ∞ depending on the value of g. The latter results contrast with those of Carmesin and Noolandi [Macromolecules 22, 1689 (1989)], who found that only first order wetting transitions are possible polymer mixtures. However, our results are in agreement with recent Monte Carlo simulations carried out by Wang and Binder [J. Chem. Phys. 94, 8537 (1991)] and Pereira and Wang [J. Chem. Phys. 104, 5294 (1996)].}, number={3}, journal={JOURNAL OF CHEMICAL PHYSICS}, author={Genzer, J and Composto, RJ}, year={1997}, month={Jan}, pages={1257–1263} } @article{genzer_composto_1997, title={Surface segregation amplification in miscible polymer blends near criticality}, volume={38}, ISSN={["0295-5075"]}, DOI={10.1209/epl/i1997-00221-7}, abstractNote={In atomic, small molecule or polymeric multicomponent materials, surface compositions naturally differ from the bulk because one component (or phase) will generally favor the surface region. Binary polymer blends represent a model system to investigate surface enrichment because segregation is enhanced by the small combinatorial entropy of mixing and amplified by chain connectivity (relative to small-molecule systems). Therefore, polymers are advantageous systems for probing the thermodynamic complexities underlying surface enrichment in mixtures. In this work, the surface excess of binary polymer blends is studied as a function of composition and temperature in the vicinity of the critical point. Although the surface excess away from criticality behaves as anticipated, it is found to grow slower than expected as criticality is approached from the one-phase region. These results suggest that the surface and bulk thermodynamics are coupled.}, number={3}, journal={EUROPHYSICS LETTERS}, author={Genzer, J and Composto, RJ}, year={1997}, month={Apr}, pages={171–176} } @article{genzer_kramer_1997, title={Wetting of substrates with phase-separated binary polymer mixtures}, volume={78}, ISSN={["0031-9007"]}, DOI={10.1103/physrevlett.78.4946}, abstractNote={We investigate a wetting reversal transition in thin films of two-phase mixtures of poly(ethylenepropylene) (PEP) and its deuterated analog (dPEP) on substrates covered by self-assembled monolayers (SAM) whose surface energy, gSAM, is tuned by varying the SAM composition. As gSAM increases from 21 to 24 mJ m2, a transition from a dPEP PEP dPEP SAM to a dPEP PEP SAM structure occurs at increasing TC 2 T , where TC and T are the critical and transition temperatures, respectively. The dependence of T on gSAM is predicted by a simple model from surface and interfacial energies of PEP dPEP. [S0031-9007(97)03445-5]}, number={26}, journal={PHYSICAL REVIEW LETTERS}, author={Genzer, J and Kramer, EJ}, year={1997}, month={Jun}, pages={4946–4949} } @article{genzer_faldi_composto_1996, title={Mean-field theory of the interface between a homopolymer and a binary-polymer mixture}, volume={105}, ISSN={["0021-9606"]}, DOI={10.1063/1.472936}, abstractNote={In this paper we investigate the properties of the interface between a homopolymer (A) and a binary polymer mixture (B:C). We have extended the self-consistent field (SCF) model of Helfand [Macromolecules 25, 1676 (1992)] by including the effects of the numbers of segments of polymers on the polymer volume fractions, the interfacial excess of B, zB*, the interfacial tension, γABC, the interfacial width, wABC, and the A–B interfacial overlap, wAB. Other parameters include the polymer interaction parameters, χAB, χAC, and χBC, and the B volume fraction in the B:C mixture, φB∞. As expected, the B component segregates to the A/B:C interface when A–B interactions are more favorable than the A–C ones (χAB<χAC). This interfacial adsorption lowers the interfacial tension and width compared to the pure A/C case, demonstrating the compatibilizing effect of B. As χBC varies from favorable to unfavorable, B segregation is found to increase. Similarly, at constant interaction parameters, B segregation increases as the number of the B segments increases. Furthermore, the addition of a small amount of high molecular weight B to an immiscible A/C blend is found to reduce greatly γABC and thus stabilize the system. Calculations for the case of an athermal mixture B:C with χAB=χAC reveals that the shorter chains are ‘‘entropically driven’’ to the A/B:C interface.}, number={22}, journal={JOURNAL OF CHEMICAL PHYSICS}, author={Genzer, J and Faldi, A and Composto, RJ}, year={1996}, month={Dec}, pages={10134–10144} } @article{genzer_faldi_oslanec_composto_1996, title={Surface enrichment in a miscible polymer blend: An experimental test of self-consistent field and long-wavelength approximation models}, volume={29}, ISSN={["0024-9297"]}, DOI={10.1021/ma951108f}, abstractNote={Neutron reflectivity (NR) and low-energy forward recoil spectrometry (LE-FRES) were used to study surface enrichment in miscible blends of deuterated polystyrene, d-PS, and poly(styrene-co-4-bromostyrene), PBr0.049S, having a 0.049 mole fraction of 4-bromostyrene units. The d-PS component was found to segregate preferentially to the polymer blend/air interface, whereas no enrichment of either component was detected at the polymer blend/silicon interface. The experimental values of the surface concentration, φ1, and the surface excess, z*, of d-PS were interpreted using both the theory of Schmidt and Binder (SB) (J. Phys. II (Paris) 1985, 46, 1631) and the self-consistent field (SCF) approach of Genzer et al. (Phys. Rev. E 1994, 50, 2373). Although both SB and SCF models were found to be in good qualitative agreement with the experimentally measured values of φ1 and z*, we demonstrate that the latter proves to be in better quantitative agreement with the experimental results. Moreover, a comparison of the ...}, number={16}, journal={MACROMOLECULES}, author={Genzer, J and Faldi, A and Oslanec, R and Composto, RJ}, year={1996}, month={Jul}, pages={5438–5445} } @article{faldi_genzer_composto_dozier_1995, title={SEGREGATION AT THE INTERFACE BETWEEN A HOMOPOLYMER AND A BINARY POLYMER BLEND}, volume={74}, ISSN={["0031-9007"]}, DOI={10.1103/physrevlett.74.3388}, abstractNote={Low-energy forward-recoil spectrometry studies show that the interface between polystyrene ({ital A}) and a miscible blend of two poly(styrene-{ital co}-4-bromostyrene)s ({ital B}:{ital C}) has an interfacial excess of {ital B}, {ital z}{sub {ital B}}{sup *}, where {ital B} is the component with the lower bromostyrene level. At 170 {degree}C, {ital z}{sub {ital B}}{sup *} increases with the bulk volume fraction of {ital B}, {phi}{sub {ital B}{infinity}}, has a maximum near 0.25, and then decreases. These results are in qualitative agreement with self-consistent field (SCF) calculations. SCF calculations of the interfacial tension and interfacial width predict that {ital B} has an optimum compatibilizing effect for {phi}{sub {ital B}{infinity}}{similar_to}0.25.}, number={17}, journal={PHYSICAL REVIEW LETTERS}, author={FALDI, A and GENZER, J and COMPOSTO, RJ and DOZIER, WD}, year={1995}, month={Apr}, pages={3388–3391} } @article{genzer_rothman_composto_1994, title={IMPROVED HYDROGEN AND DEUTERIUM DEPTH PROFILING IN POLYMERS USING LOW-ENERGY FORWARD RECOIL SPECTROMETRY}, volume={86}, ISSN={["0168-583X"]}, DOI={10.1016/0168-583x(94)95300-7}, abstractNote={We demonstrate that the hydrogen and deuterium depth resolution of forward recoil spectrometry (FRES) is greatly improved by utilizing 1.3 MeV 4He ions and sample tilting. The near-surface depth resolution is measured on samples of alternating deuterated and protonated polystyrene layers, each ca 250 Å thick. The depth resolution at the surface is dramatically improved (from 800 Å to 250 Å) as the incident beam energy decreases from 3.0 to 1.3 MeV. We call this technique low-energy FRES (LE-FRES) to differentiate it from standard FRES, which uses 3 MeV 4He ions incident at θ = 75°. Furthermore, by using a glancing incident or exit geometry, the LE-FRES depth resolution is improved to ca 125 Å, which is comparable to the resolution of secondary ion mass spectrometry and nuclear reaction analysis. Measurements of the energy and glancing angle dependence of the depth resolution are in quantitative agreement with theoretical predictions. The probing depth of LE-FRES is found to have a maximum at θ ≈ 71° and decreases at glancing exit and incident angles. A contour plot illustrates the relationship between depth resolution and probing depth and provides guidance for future users of LE-FRES.}, number={3-4}, journal={NUCLEAR INSTRUMENTS & METHODS IN PHYSICS RESEARCH SECTION B-BEAM INTERACTIONS WITH MATERIALS AND ATOMS}, author={GENZER, J and ROTHMAN, JB and COMPOSTO, RJ}, year={1994}, month={Apr}, pages={345–354} } @article{wallace_zhong_genzer_composto_bonnell_1993, title={ON THE USE OF ION-SCATTERING TO EXAMINE THE ROLE OF HYDROGEN IN THE REDUCTION OF TIO2}, volume={8}, ISSN={["0884-2914"]}, DOI={10.1557/jmr.1993.1629}, abstractNote={Rutherford backscattering spectrometry (RBS) was used to measure the titanium concentration profile for hydrogen-reduced, vacuum-reduced, and as-received, stoichiometric rutile. These profiles give the degree of reduction, specifically, the extent of oxygen deficiency, as a function of depth below the sample surface. Using forward-recoil spectrometry (FRES), the hydrogen-reduced rutile was found to contain more bulk and near-surface hydrogen than the as-received, stoichiometric rutile. This observation provides additional evidence for a hydrogen-diffusion model for the reduction of rutile in a hydrogen environment.}, number={7}, journal={JOURNAL OF MATERIALS RESEARCH}, author={WALLACE, WE and ZHONG, Q and GENZER, J and COMPOSTO, RJ and BONNELL, DA}, year={1993}, month={Jul}, pages={1629–1634} } @article{genzer_svorcik_rybka_1990, title={Survey of the reactions of some metals with gallium arsenide applicable in semiconductor technology}, volume={84}, number={11}, journal={Chemicke Listy}, author={Genzer, J. and Svorcik, V. and Rybka, V.}, year={1990}, pages={1146–1160} }