RESEARCH ARTICLE www.advmat.de High-Hole-Mobility Fiber Organic Electrochemical Transistors for Next-Generation Adaptive Neuromorphic Bio-Hybrid Technologies Paula Alarcon-Espejo, Ruben Sarabia-Riquelme, Giovanni Maria Matrone, Maryam Shahi, Siamak Mahmoudi, Gehan S. Rupasinghe, Vianna N. Le, Antonio M. Mantica, Dali Qian, T. John Balk, Jonathan Rivnay, Matthew Weisenberger, and Alexandra F. Paterson* The latest developments in fiber design and materials science are paving the way for fibers to evolve from parts in passive components to functional parts in active fabrics. Designing conformable, organic electrochemical transistor (OECT) structures using poly(3,4-ethylenedioxythiophene):polystyrene sulfonate (PEDOT:PSS) fibers has excellent potential for low-cost wearable bioelectronics, bio-hybrid devices, and adaptive neuromorphic technologies. However, to achieve high-performance, stable devices from PEDOT:PSS fibers, approaches are required to form electrodes on fibers with small diameters and poor wettability, that leads to irregular coatings. Additionally, PEDOT:PSS-fiber fabrication needs to move away from small batch processing to roll-to-roll or continuous processing. Here, it is shown that synergistic effects from a superior electrode/organic interface, and exceptional fiber alignment from continuous processing, enable PEDOT:PSS fiber-OECTs with stable contacts, high μC* product (1570.5 F cm−1 V−1 s−1 ), and high hole mobility over 45 cm2 V−1 s−1 . Fiber-electrochemical neuromorphic organic devices (fiber-ENODes) are developed to demonstrate that the high mobility fibers are promising building blocks for future bio-hybrid technologies. The fiber-ENODes demonstrate synaptic weight update in response to dopamine, as well as a form factor closely matching the neuronal axon terminal. P. Alarcon-Espejo, R. Sarabia-Riquelme, M. Shahi, S. Mahmoudi, G. S. Rupasinghe, V. N. Le, D. Qian, M. Weisenberger, A. F. Paterson Department of Chemical and Materials Engineering Centre for Applied Energy Research University of Kentucky Lexington, KY 40506, USA E-mail: alexandra.paterson@uky.edu G. M. Matrone, J. Rivnay Department of Biomedical Engineering Northwestern University Evanston, IL 60208, USA A. M. Mantica, T. J. Balk Department of Chemical and Materials Engineering University of Kentucky Lexington, KY 40506, USA The ORCID identification number(s) for the author(s) of this article can be found under https://doi.org/10.1002/adma.202305371 DOI: 10.1002/adma.202305371 Adv. Mater. 2023, 2305371 1. Introduction Fibers are intricately woven throughout the fabric of human history. The latest developments in fiber design and materials science are paving the way for fibers to evolve from parts in passive goods to highly complex, functional components in active fabrics. Conductive materials, in particular, are being explored for a wide range of fiber-based applications, including mechanical sensors to detect audible sounds or transpose mechanical vibrations into electrical signals,[1,2] power systems,[3–6] electronic devices,[7,8] machine-learning interfaces combining multidimensional sensing,[9–11] storage, communication, logic, and processing.[12] Fibers are uniquely equipped to preserve their mechanical properties under different types of deformation and stimuli, enabling technologies that function even while being stretched, bent, and twisted in 3D.[13] Additionally, they can be sewn together to form fabrics and create customized electronic textiles that are light weight, flexible, breathable, and comfortable.[14–16] One area that would benefit greatly from these unique properties is wearable biotechnologies: the human body contains a wealth of information, which computers integrated within clothing could access, enabling machine learning and fabric artificial intelligence to process, analyze, and return digital health insights.[17] However, Loke et al. identified in their “Moore’s law in fibers”,[17] that realizing such sophisticated, wearable bioelectronic platforms requires conductive fibers that can also operate as neurons and/or transistors.[17] Although various materials, such as carbon nanotubes, graphene, metallic nanoparticles, and nanowires,[18] are being explored as fibers, organic electronic materials are competetive candidtates for mass-market, affordable wearable bioelectronics. One such material poly(3,4ethylenedioxythiophene):polystyrene sulfonate (PEDOT:PSS), is a commercially available conjugated polymer with good ambient stability, high electrical conductivity, and the 2305371 (1 of 11) © 2023 Wiley-VCH GmbH www.advmat.de Figure 1. PEDOT:PSS fiber-OECTs. a) Scanning electron microscopy image of a single, continuously spun PEDOT:PSS fiber with a diameter of 7.12 μm. b) Schematic of the PEDOT:PSS fiber-based OECTs. c) OECT transfer curves where VD = −1 V and d) transconductance for the various contact materials explored with the fibers. e) Mobilities measured for the PEDOT:PSS fiber-based OECTs with different contact materials. Statistics for graphite and silver were taken from size OECTs each, whereas μ for only one OECT is shown for Sn63:Pb37, because the contact interface was so poor that it did not produce any additional working devices. flexibility that is compatible with textiles.[19,20] In addition to promising conductive fibers,[8,20–22] PEDOT:PSS has been used in developing artificial synapses and synaptic circuits for organic neuromorphic electronics,[23–25] and is also the most widely studied material for organic bioelectronics[26–28] and organic electrochemical transistors (OECTs).[19] The latter OECTs have been identified as fundamental building blocks for advanced electronics, body–machine interfaces, drug delivery, smart textiles, bioelectronics, and adaptive healthcare.[19,29–33] Indeed, PEDOT:PSS fiber-OECTs have already demonstrated their potential for biosensors/biomonitoring devices,[34] and Kim et al. reported PEDOT:PSS fiber-OECTs with high μC* product values (≈1500 F cm−1 V−1 s−1 ) and high hole mobilities (12.9 cm2 V−1 s−1 ).[8] Employing PEDOT:PSS fibers to design conformable OECT structures may therefore offer a highly promising option for future, wearable bio-hybrid devices, at affordable costs.[7] Before scalable, bio-hybrid devices made from PEDOT:PSS fiber-OECTs can be realized, there are key challenges to overcome. First, in the context of wearable technologies, ≈4 km of thread/fiber is required to make a single T-shirt (Supporting Information). It is therefore important to move away from small batch processing (5–10 cm length for each fiber[8] ), to roll-toroll or continuous processing. Second, forming good contacts is crucial to achieve high-performance, stable OECTs,[35–37] yet creating good electrical contacts on fibers is challenging because their small radius of curvature leads to wettability issues and irregular coatings.[38–40] Here, we address these core challenges by combining continuous PEDOT:PSS fiber-processing— that can align hundreds of meters of PEDOT:PSS fiber per run— Adv. Mater. 2023, 2305371 with contact engineering. We demonstrate that the synergistic effects from exceptional fiber alignment and the superior contact/organic interface enable PEDOT:PSS fiber-OECTs with high 𝜇C* product (1570.5 F cm−1 V−1 s−1 ) and high hole mobility over 45 cm2 V−1 s−1 —the highest charge carrier mobility measured in an organic transistor to-date.[41–43] Finally, we demonstrate that the high mobility fibers are promising building blocks for future bio-hybrid integration, by designing electrochemical transistors with neurotransmitter-mediated plasticity, and a form factor that closely matches the neuronal axon terminal. 2. Results We span the PEDOT:PSS fibers as the active channel for an OECT (Figure 1a,b) using a previously reported continuous wetspinning process, that can continuously collect hundreds of meters of fibers per run (Figure S1, Supporting Information).[22] In short, a 2.5 wt% dispersion of PEDOT:PSS in water was extruded through a single-hole capillary spinneret into a coagulation bath containing 10 vol% dimethyl sulfoxide (DMSO) in isopropanol (IPA). The coagulated filament was then driven by rollers into a pure DMSO drawing bath followed by an in-line drying step with an air temperature of 170 °C. The fibers were stretched throughout the entire duration of the fabrication process, with a total draw ratio of 1.82, i.e., 1.82 m of fiber collected per each 1 m of filament formed in the coagulation bath.[22] Next, we identified materials to use as the contacts/electrodes for the OECTs. Contact resistance (RC ) in transistors is well-documented; the Schottky barrier formed from the difference between the 2305371 (2 of 11) © 2023 Wiley-VCH GmbH 15214095, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202305371 by Pukyong National University, Wiley Online Library on [03/03/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License www.advancedsciencenews.com www.advmat.de electrode work function (WF) and channel transport inhibits the injection and extraction of charge carriers, which reduces operating voltage, operating frequency, measured mobility, and prevents miniaturization.[44] However, the role of RC in OECTs is convoluted and under debate: on the one hand, it is reportedly so dominant that it causes the monotonic transconductance (gm )[45] and Bernards and Malliaras capacitive model to break down.[46] On the other hand, gm remains monotonic when four-point probe measurements are used to avoid RC ,[47,48] and electrochemical doping is reported to reduce Schottky barrier widths to produce the lowest ever RC values measured in organic transistors.[35,49] We therefore chose contact materials with a broad range of WF values to further explore RC in OECTs, namely: graphite paint, silver paint, a 63% tin and 37% lead alloy (Sn63:Pb37), and a 52% indium and 48% tin alloy (In52:Sn48). The measured WF values (relative to the PEDOT:PSS highest occupied molecular orbital (HOMO)) are shown in Figure S2a in the Supporting Information. The silver and graphite WF values are found to be similar, measured at −5.01 and −5.03 eV, respectively, and higher than the In52:Sn48 WF, which measures at −4.91 eV, and the Sn63:Pb37 WF which is −3.72 eV. We also note that the value for PEDOT:PSS is around 5.0 eV.[50] Based on these values, it would therefore be expected that the silver and graphite paints yield comparable device performance because they minimize RC equally, and that the silver and graphite paints surpass the performance of the alloys. We then used copper and graphite foils/tapes to complete the fabrication process by mediating between the measurement probe and aforementioned contact materials, as previously reported (Figure S2b, Supporting Information).[8,22] Further information on both fiber continuous processing and fiber-OECT fabrication are in the Experimental Section and Figures S3 and S4 in the Supporting Information. The transfer curves for representative fiber-OECTs fabricated with the different contact materials are shown in Figure 1c and Figure S5 in the Supporting Information. The peak transconductance (gm_peak ) for representative fiber-OECTs is 1162.0, 1651.7, and 2093.5 S cm−1 , for Sn63:Pb37, silver, and graphite contacts, respectively (Figure 1d). The In52:Sn48 contacts did not produce functioning fiber-OECTs. We note that gm was normalized to account for channel geometry and enable comparison with other values with the literature,[51] following the formula gmnormalized = gm AL ,[51] where L is the channel length and A denotes the fiber cross-section. The average gm (gm_avg ) over six fiber-OECTs per system is 1408.8 and 1843.9 S cm−1 for silver and graphite contacts, respectively. Statistics for the Sn63:Pb37 contacts are not presented because only one device out of six worked. In addition to showing high transconductance values, the measured charge carrier mobility (μ) values are remarkable: the time-of-flight or transient method outlined by Bernards and Malliaras[52] was used to measure the hole 𝜇, giving 21.2, 31.1, and 45.1 cm2 V−1 s−1 , for Sn63:Pb37, silver, and graphite, respectively (Figure S6, Supporting Information), with average values of 23.0 and 32.6 cm2 V−1 s−1 for silver and graphite (Figure S7, Supporting Information), respectively. Although 𝜇 extracted using this transient technique is not influenced by the channel cross-sectional area, it is essential to precisely determine L to extract the hole transient time between the source and the drain electrodes; see Figure S3 for further details.[52] Next, we explored Adv. Mater. 2023, 2305371 the impact of the high μ on the independently derived OMIEC figure-of-merit, the μC* product.[53] C* was found to be ≈34.8 F cm−3 , measured using electrochemical impedance spectroscopy (see the Experimental Section for more detail; Figure S8 in the Supporting Information shows cyclic voltammetry for a representative PEDOT:PSS fiber-OECT), which is similar to the values reported for PEDOT:PSS thin films (≈40 F cm−3 ).[54,55] The maximum μC* products for best-performing fiber-OECTs are 634.9, 1233.9, and 1570.5 F cm−1 V−1 s−1 , for the Sn63:Pb37, silver, and graphite contacts, respectively. The corresponding, average values over six fiber-OECTs per system are 914.3 and 1134.6 F cm−1 V−1 s−1 , for the silver and graphite contacts, respectively. Figure S16 compares the fibre-OECTs with 𝜇 values from a standard thin-film OECT. Overall, at time of writing this is the state-of-theart μC* product, the highest PEDOT:PSS mobility, and the highest hole mobility measured in an organic transistor to-date for the graphite contacts in the PEDOT:PSS fiber-OECTs.[41–43,56,57] 3. Discussion Interestingly, as indicated later in this Section, the measured mobility strongly depends on the contact material. Therefore, we explored the role of the contacts to begin elucidating the reasons for the extraordinarily high μ. The measured-μ has a strong relationship with the contact/organic interface because efficient carrier injection/extraction is enabled between a well-matched contact WF and organic transport level. The WF values (Figure S2a, Supporting Information) and μ (Table 1) indicate that Sn63:Pb37 has the largest energetic mismatch at the interface, and the Sn63:Pb37 fiber-OECT performance and metrics are lowest out of the systems. However, surprisingly, the silver and graphite contacts have comparable energetic interfaces (Figure S2a, Supporting Information), yet different measured-μ (31.1 and 45.1 cm2 V−1 s−1 for silver and graphite, respectively) and μC* products (1233.9 and 1570.5 F cm−1 V−1 s−1 for silver and graphite, respectively). Correspondingly, extracting resistance per length for the silver and graphite contacts using two-probe measurements gives 108.8 and 103.8 Ω mm−1 , respectively. We note that at the time of writing we were unable to extract reliable data using the transmission line method, due to the variability (device uniformity) of the fiber-OECTs across the various channel lengths.[59] A possible reason for the lack of relationship between the WF-HOMO, measured-μ, and two-probe resistance is that the quality of the fiber/contact interface depends on the contact material fluid type, drying dynamics, thermal expansion, and commonly used nonNewtonian suspensions can lead to irregular coatings on fibers due to surface tension;[39,40] poor wettability, high curvature, and surface contamination.[38,39] In particular, the silver paint is in methyl isobutyl ketone and the graphite paint is in an isopropanol-based solution, suggesting that there may be differences in wetting between the two contact paints and the fiber channel surface. The best way to approach the wettability of the unique fiber surface is to use a dynamic contact angle measurement, such as the Wilhelmy plate method.[60,61] This technique would take into account the fiber curvature that will substantially influence the droplet spreading dynamics, and therefore also impacts the accuracy of the measured contact angles.[38,39,62] However, in this case, our systems present 2305371 (3 of 11) © 2023 Wiley-VCH GmbH 15214095, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202305371 by Pukyong National University, Wiley Online Library on [03/03/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License www.advancedsciencenews.com www.advmat.de Table 1. Table comparing published PEDOT:PSS mobility values with best-performing graphite contact mobility values. OECT Young’s modulus [GPa] Contact material μmax [cm2 V−1 s−1 ] Fiber 14.6 Graphite Fiber 8.7 Silver Thin film N/A Gold Analysis technique μC* Reference 45.1 Transient 1570.5 This work 12.9 Impedance matching 1500 [8] 11.7 (EgVDP) Linear regime 195.6 [48] gm 157.5 [58] 6.7 (OECT) Thin film N/A Gold 4.5 practical challenges from rapid paint drying times (30 s) which impact the advancing and receding contact angles, and the different particles suspended in the inks and their interaction with the fiber surface also impact the wettability phenomena and accuracy of the measured contact angles. We therefore explored the relationship between measuredμ and the quality of the mechanical contact/organic interface, using scanning electron microscopy (SEM). Figure 2a,b and Figure S9 in the Supporting Information show that there is a strong correlation between interface quality and measured-μ; we find that the In52:Sn48 does not coat the fibers, explaining the nonfunctioning fiber-OECTs. Additionally, it is clear that the sil- ver does not completely coat the fiber surface, resulting in a poor mechanical interface (Figure 2a). In stark contrast, the graphite completely covers the fiber (Figure 2b). We performed a pixel count analysis on a representative fiber OECT made with silver contacts, to better quantify the contact area, in relation to a representative graphite contact which covers 100% of the fiber (i.e., Figure 2b). Figure S10 in the Supporting Information shows the pixel count, revealing that ≈31.1% of the fiber is exposed to the silver. This can also be compared to Figure S9 in the Supporting Information, which shows that as little as 0% of the fiber is coated by the Sn63:Pb37 and In52:Sn48 alloys. Figure S11 in the Supporting Information shows the impact of the contact area Figure 2. The role of contacts in fiber-OECTs. Scanning electron microscopy images showing the morphological aspects of the a) silver and b) graphite contacts. The primary images show the electrode material coating the fiber. One inset shows how the surface of the PEDOT:PSS fiber is covered with the silver or graphite at the middle point of the fiber–electrode contact, and the other inset shows the exit point of the fiber–electrode contact. c) Real-time change in resistance, normalized to the minimum resistance, after applying the silver and the graphite paints/inks to the PEDOT:PSS fiber to fabricate the contacts. Two probe resistance measurements showing the degradation of the d) silver and e) graphite contacts over time. f) Two probe resistance measurements showing the lack of degradation in the PEDOT:PSS fiber over time, and demonstrating high electrical stability of the fibers on the spool used in this study. For each measurement, the fiber was unwound from the same spool and measured using contacts deposited on the indicated day of testing. Adv. Mater. 2023, 2305371 2305371 (4 of 11) © 2023 Wiley-VCH GmbH 15214095, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202305371 by Pukyong National University, Wiley Online Library on [03/03/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License www.advancedsciencenews.com www.advmat.de on μ for best-performing graphite contacts (at 1.5, 3, and 7 mm length) Overall, these findings indicate the importance of the contact/organic interface quality for achieving high measured-μ in transistors.[41,59,63] Considering that the organic/contact interface was recently shown to play a dominant role in OECT stability,[37] we investigated the stability of the coating quality over short and long timescales. We monitored the resistance (R) normalized to the minimum measured resistance in a two-probe configuration for silver and graphite contacts in the PEDOT:PSS fiber-OECTs. Figure 2c shows the results: in both cases, there is a sharp decrease in R when the contact material is applied to the fiber, i.e., at t = 0. After the sharp reduction in R, that corresponds to the application and drying of the paints/inks containing the contact material, it is clear that the silver contacts start to degrade In contrast, the graphite contacts are highly stable. We extended the two-probe conductivity to longer, shelf-life stability studies on the contacts (see Supporting Information). Again, the silver contacts demonstrate poor stability leading to complete loss of contact beyond the 30 days (Figure 2d), while none of the graphite contacts failed. We note that the graphite contacts showed on average a 40% increase in resistance after 20 days, with a further 3% increase after 2 months, indicating that an aging-related equilibrium may have been reached (Figure 2e). We also measured PEDOT:PSS fiber conductivity at 3 month intervals for 6 months; Figure 2f shows the fiber electrical conductivity was consistent and stable, giving an average value of 2170 ± 60 S cm−1 , indicating that the observed degradation (Figure 2d) is occurring at the contact itself. As there was no coverage change over time for the silver contacts, one possible mechanism for the degradation is the acidity of the PEDOT:PSS, which may have corroded the fiber/contact interface via corrosion of common contact metals like silver[64] or copper.[65] In other cases, this acid corrosion has been reported to degrade and limit performance of PEDOT:PSS devices,[66,67] which may suggest a key difference between the silver and graphite, and origin of the difference in the resistance changes over time (Figure 2). Namely, the PEDOT:PSS behaves as an acid and will readily accept electrons from contact materials with a high tendency to donate electrons, such as silver. If the silver donates electrons to the doped PEDOT:PSS, it equivalently removes holes and therefore effectively de-dopes the PEDOT:PSS at the contact/fiber interface region as a function of time. This in turn oxidizes the silver and increases the temporal resistance of the contact interface, where the application of current may act as a catalyst for the oxidation. On the other hand, for graphite, which has little tendency to donate electrons, we would observe a more stable contact/fiber interface. To investigate further whether oxidation increasing over time is the origin of the degradation at the silver/fiber contact interface, we used X-ray photoelectron spectroscopy (XPS) to look for shifts in the Ag5/2 peak to lower binding energy. However, the study indicated that XPS is not sensitive enough to probe for silver oxidation specifically at the silver/fiber interface. The measurement and data analysis were further complicated by the presence of oxygen in all samples (silver/fiber interface, silver paint only, the fiber only), the fact XPS only probes 5–10 nm at the sample surface, and the size of the probe relative to the fiber diameter meaning that—although the probe is incident on the contacts—it is not only investigating the interface between the silver and PE- Adv. Mater. 2023, 2305371 DOT:PSS fibers. Although beyond the scope of the current paper, going forward more sensitive techniques such as high-resolution transmission electron microscopy and selected area electron diffraction could be used to probe the interface between PEDOT and both the silver and the graphite contacts. Finally, Figure S12 in the Supporting Information further demonstrates superior mechanical resilience and electrical contact for the graphite contacts compared to silver. The contacts were subjected to a brief mechanical distress that clearly unveiled the mechanical deficiencies at the silver/fiber interface, which are then recovered by applying pressure. On the other hand, the graphite contacts were unaffected, maintaining noise free, initial resistance values. Overall, the data indicate that the contact/organic interface is the source of instability rather than the fiber/channel itself. Next, we explored the role of the fiber channel on the high OECT mobility. Table 2 shows the hole μ measured in the PEDOT:PSS fiber-OECTs is significantly higher than μ reported for thin-film OECTs (6.7 cm2 V−1 s−1 ),[48] and higher than μ reported for PEDOT:PSS thin films measured using a fourprobe technique (11.7 cm2 V−1 s−1 ).[48] Furthermore, μ is 2.5 × greater than previously reported for a PEDOT:PSS fiber OECT (silver contacts: 12.9 cm2 V−1 s−1[8] compared to 31.1 cm2 V−1 s−1 measured here). A key difference between the previously reported values and the fibers used in this work is the drawing in the continuous processing. On the one hand, drawing in continuous processing applies a permanent extensional deformation along the axis of the fiber,[68] which, in turn, induces preferential orientation of the polymer chains by aligning them with the fiber-axis direction. On the other hand, the previously reported high mobility fibers were fabricated discontinuously, i.e., without applied draw. In the discontinuous case, the polymer chains were orientated by drying the sulfuric acid-treated fibers and using hanging weights, limiting the amount of force applied and therefore the level of polymer chain orientation achieved.[8] Indeed, improved chain orientation in continuously processed fibers is reported to give exceptional mechanical properties and enhanced thermal conductivity, where thermal conductivity is also considered a good indicator of alignment. For example, oriented polyethylene nanofibers have reported values up to 104 W m−1 K−1 ,[69] and PEDOT:PSS-fibers with values 3–4 W m−1 K−1 , over one order of magnitude higher than typical PEDOT:PSS thin films.[22] Additionally, increased electrical conductivity with increasing draw has been reported for conducting polymer fibers made from poly(3-alkylthiophene),[70,71] polyaniline,[72,73] and PEDOT:PSS,[22] as a result of polymer chains aligned along the fiber axis, and charge carrier transport happens fastest along the backbone of the polymer chains in conjugated polymers.[74] We therefore hypothesize that the drawing in continuous processing is the reason for the higher μ compared to discontinuously processed fibers. To investigate this further, we first measured the Young’s modulus of the continuously processed PEDOT:PSS fibers, where the Young’s modulus is a key indicator that can be used as a proxy to compare polymer chain alignment and orientation in fibers made from the same polymer. Figure S13 in the Supporting Information shows the Young’s modulus of the continuously processed fibers is 14.2 GPa—almost twice the value of the optimized (i.e., dried under the maximum tensile stress) discontinuously processed fibers (8.7 GPa[8] ).[21,22] We measured the electrical conductivity and break stress to be 2305371 (5 of 11) © 2023 Wiley-VCH GmbH 15214095, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202305371 by Pukyong National University, Wiley Online Library on [03/03/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License www.advancedsciencenews.com www.advmat.de Table 2. Comparative data for maximum charge carrier mobilities for different contact materials in PEDOT:PSS-fiber OECTs. Silver and graphite contact statistics are taken over N = 6 devices. Statistics for the Sn63:Pb37 contacts are not presented because only one device worked. Work function values and interface quality are compared to demonstrate their collective impact on μ. Material μmax μavg Work Function [eV] Interface quality −4.9 Extremely poor In52:Sn48 - Sn63:Pb37 21.2 - −3.7 Poor Silver 31.1 23.0 −5.0 Poor to moderate Graphite 45.1 32.6 −5.0 Excellent 2130 S cm−1 and 357 MPa, respectively. We note that changing the solvent that the PEDOT:PSS is drawn from, i.e., changing the DMSO to concentrated sulfuric acid, has been reported to further enhance the Young’s modulus (22 GPa[21,22] ) and therefore may offer an approach to further enhance the mobility when combined with the optimized graphite contacts. However, we also note that sulfuric acid may degrade the channel material over time.[8] Wide-angle X-ray scattering (WAXS) was then used to study the alignment in the continuously drawn fibers (Figure 3a–c). The 2D-WAXS spectrum of the fibers is shown in Figure 3a, where the intense arcs observed along the horizontal line of the spectrum are an indication of the anisotropy of the fibers. Since the fibers were placed vertically and normal to the transmitted incident X-ray beam, diffracted horizontal arcs are an indication of preferential orientation of crystal planes in the fiber axis direction. Figure 3b shows the 2D-WAXS 2𝜃 integrated intensities for the fibers. The first peak at 3.6° corresponds to the (100) lamella stacking of PEDOT and PSS chains with a d-spacing of 24.2 Å.[75,76] The peak appearing at 26.3° is attributed to the (020) re- flection of the 𝜋–𝜋 stacking of PEDOT chains. The 𝜋–𝜋 stacking distance was 3.4 Å, in agreement with previous reports.[22,75,76] Lastly, the peak observed at 17.8° corresponds to amorphous PSS.[75,76] The azimuthal intensity profiles for the three reflections are shown in Figure 3c. For PSS, the azimuthal profile reflection is flat, indicating random orientation. However, the peaks observed at 0° and 180° for the (100) and (020) reflections indicate a preferential orientation of those planes parallel to the axial direction of the fiber, meaning that the PEDOT polymer backbones are also oriented in that direction. Finally, we studied the surface morphology of the fibers with SEM and atomic force microscopy (AFM). Figure 3d,e shows longitudinal ridges and striations running parallel to the fiber axis along the length of the fibers, and Figure 3f shows that these aligned fibrils are also visible in AFM surface topography. We note that, based on previous reports, the ridges and striations in the SEM images are characteristic of wetspun fibers that undergo a solvent diffusion-induced coagulation process,[77,78] and are associated with polymer chain alignment and orientation of polymer fibrils on the fiber surface in the fiber axis direction. Additionally, AFM was used to analyze the surface Figure 3. Continuous processing enables highly aligned PEDOT:PSS fibers. a) 2D WAXS pattern of the fibers. b) 2 theta scan integrated intensity with the main peaks for PEDOT:PSS noted. c) Azimuthal intensity profiles of (100) lamella stacking, PSS, and (020) 𝜋–𝜋 stacking reflections. d,e) SEM image showing the ridges and striations running along the length of the fibers indicative of fibril orientation. f) AFM topography image showing fibril orientation in the fiber axis direction. Adv. Mater. 2023, 2305371 2305371 (6 of 11) © 2023 Wiley-VCH GmbH 15214095, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202305371 by Pukyong National University, Wiley Online Library on [03/03/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License www.advancedsciencenews.com www.advmat.de Figure 4. An artificial bio-hybrid synapse with dopamine-dependent plasticity. a) Schematic showing the double PEDOT:PSS fiber-ENODe configuration. b) Measurements of conductance modulation in the channel terminal while periodically pulsing VG . In the presence of dopamine, the conductance modulation is permanent and depends directly on the concentration of the neurotransmitter in the electrolyte solution. roughness of the PEDOT:PSS fiber, where a root mean square (RMS) value was found to be 17.5 nm with a standard deviation of 3.8 nm. This is considerably smaller than the particle size distribution of the graphite and the silver paints, as reported by the manufacturer, which ranges from 0.4 to 1 μm. Overall, the collective Young’s modulus, WAXS, SEM, and AFM evidence that continuously processed PEDOT:PSS fibers are highly aligned and orientated, which contributes to the high μ measured in the fiberOECTs, and may play a role in increasing the intrinsic mobility of the PEDOT:PSS fiber. Finally, active elements based on building blocks, such as these high-performance fibers and their associated OECTs, could offer unprecedented fiber-based sensors, circuits, and neuromorphic components. Artificial bio-hybrid synapses are one emergent bioelectronic platform that show synaptic conditioning from neuronal physiological biochemical signaling activity. For example, electrochemical neuromorphic organic devices (ENODes) have been reported to work as artificial synapses by emulating neurotransmitter synaptic weight modulation and recycling machinery at the synaptic cleft,[79] where the gate represents the Adv. Mater. 2023, 2305371 pre-synapse and the channel acts as the post-synapse. Although ENODes made from PEDOT:PSS exploit the reversible dedopingdoping (depletion mode) for conductance modulation,[80,81] the oxidaton electro-active neurotransmitters can produce protons and electrons that cause problematic, nonreversible decreases in conductance. The conductance modulation is important in bio-hybrid interfacing platforms because it represents the weight of the artificial synapse, that is directly updated in response to the biological element’s neurochemical signal activity (i.e., the neuronal pre-synapse). Additionally, unavoidable molecular crosstalk can occur in some cell-interfacing scenarios, e.g., Keene et al. translated chemical activity into post-synaptic weight modulation[79] by seeding neuronal model cells (PC12) on a gate electrode. However, the neuron-to-neuron signal transmission was lost because the neuon ensembles expressed multiple neurotransmitters. Overall, it is of interest to design neuromorphic platforms with the sensitivity and physical dimension and to match biological pre-to-post synaptic terminals, so that pre-to-post synaptic chemical activity can be combined with neuromorphic processors for in-sensor computing. 2305371 (7 of 11) © 2023 Wiley-VCH GmbH 15214095, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202305371 by Pukyong National University, Wiley Online Library on [03/03/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License www.advancedsciencenews.com www.advmat.de Inspired by the fact that the fibers approach the aspect ratio of neuron axons, we developed neuromorphic fiber-ENODes, finding that the fiber-ENODes display synaptic weight update (conductance modulation) in response to a key biological cue, dopamine (DA) (Figure 4a).[82] The results are shown in Figure 4b and Figures S14 and S15 in the Supporting Information. First, in presence of NaCl electrolyte only, periodically pulsing 0.3 V for 0.85 s, followed by an 8 s delay, causes cations to penetrate and decrease the conductance (conditioning). When the gate bias is removed, the cations return to the electrolyte (extinction), and the initial conductance level of the transistor channel is restored. On the other hand, when DA is present, a permanent conductance modulation is recorded, which we find increases with the number of pulses. Additionally, we demonstrate that this permanent conductance modulation depends on the concentration of the neurotransmitter in the electrolyte. For 0.025 × 10−3 and 0.05 × 10−3 m DA solutions, conductance values of 2.5 × 10−2 and 4.9 × 10−2 mS are recorded, respectively. Overall, the remarkably stable conductance modulation emulates synaptic conditioning mechanisms from neural signaling activity on the pre-synaptic terminal, and the biological electrical-to-chemical signal transduction. 4. Conclusion We have fabricated organic electrochemical transistors using PEDOT:PSS fibers as the active channel material, and measured a hole mobility of 45.1 cm2 V−1 s−1 —the highest reported mobility measured in an organic transistor to date. Additionally, the independently derived μC* product, 1570.5 F cm−1 V−1 s−1 . We find that the underlying fiber-OECT performance arises from a synergistic combination of the drawing step in continuously processing the fibers and optimizing the mechanical interface between the source/drain electrodes and the channel. On the one hand, Young’s modulus measurements, X-ray scattering, and microscopy imaging show the former leads to highly aligned, polymer chains that are preferentially orientated along the fiber axis, i.e., in the charge transport direction, increasing the intrinsic mobility. On the other hand, SEM, mechanical distress, and stability interface testing indicate that the latter, superior mechanical interface improves the measured mobility, and we find that it is equally or more important than the good electronic/energetic match (Kelvin probe and two-probe resistance measurements). Finally, we show that the fiber-OECTs could enable neuromorphic devices, such as artificial synapses, with reduced molecular crosstalk effects. Overall, the collective evidence indicates synergistic effects from contact engineering and processing optimization produce organic electrochemical transistors that are compatible with both large area processing and neuromorphic computing, with high μC* product and high charge carrier mobility. 5. Experimental Section PEDOT:PSS Fiber Fabrication: PEDOT:PSS fibers were fabricated following a procedure previously described by Sarabia-Riquelme, et al.[22] Aqueous PEDOT:PSS dispersion was purchased from Heraeus (PH1000; solid content 1.3 wt%). The dispersion was concentrated to 2.5 wt% by placing it on a hot plate at 90 °C while magnetically stirring. Then, 5 wt% Adv. Mater. 2023, 2305371 of DMSO was added and the dope was stirred for 2 more hours at room temperature. Lastly, the dope was sonicated for 30 min and degassed in a vacuum oven for 5 min at room temperature. Once the dope was finished, it was transferred to a 5 mL syringe and placed in a syringe pump (KD Scientific). The dope was extruded at a constant rate of 0.25 mL h−1 through a nylon syringe filter with average pore size of 5 μm and a capillary spinneret with 100 μm in diameter into a coagulation bath of 10 vol% DMSO in IPA. After coagulation, fibers were drawn in a DMSO bath and then dried passing through a keyhole-cylinder-shaped oven kept at 170 °C before being taken up on a spool. Both baths were kept at room temperature. The total draw ratio applied during processing was 1.82, which corresponds to the ratio of velocities at which the fiber was being collected on a spool and the velocity at which the dispersion was being extruded, i.e., per each 1 m of nascent filament being extruded, 1.82 m of fiber was being continuously collected. A spool of a single-filament PEDOT:PSS fibers was characterized before implementing the fibers in devices. PEDOT:PSS Fiber Mechanical Characterization: The mechanical properties of the fiber were extracted using an automatic single-fiber test system, FAVIMAT+ from Textechno. Pretension was 0.50 cN tex−1 , test speed 5.0 mm min−1 over a gauge length of 25.4 mm. Each value obtained was the average for five fibers. Young’s modulus was then extracted from the slope of the linear section of the stress–stress curve. PEDOT:PSS Fiber Electrical and Contact Stability Characterization: The fibers were characterized electrically by placing a single fiber between two graphite foil electrodes and contacted using graphite paint. The resistance was then measured using a two-probe method with a Keithley 2100 microvoltmeter. By conducting resistance measurements on specimens of different lengths, the resistance of the fibers and the contact resistance could be extracted from the slope and intercept, respectively, by plotting the total resistance measured versus the length, as shown in Figure 2d–f, where Rmeasured = Rspecimen + Rcontact . This method was commonly used to determine the conductivity of the specL . The cross-sectional area was thereimen, where Rspecimen = 𝜎 1 A specimen fore not required for the two-probe measurements to obtain values of total resistance per length fibers from the same spool and fabricated during the same run. However, to obtain the electrical conductivity of the fibers, the cross-sectional area became necessary. To determine the crosssectional area, a bundle of fibers was cut from the spool and placed with their cross-sections perpendicular to the electron beam in the SEM. The diameter of ten fibers was measured to calculate an average, and to obtain the electrical conductivity from the length, resistance, and cross-section of the fiber. Organic Electrochemical Transistor Fabrication and Characterization: Six OECT devices per each pair of electrode–contact material were fabricated using the continuously processed highly aligned PEDOT:PSS fibers as the channel. Figure 1b shows a scheme of the PEDOT:PSS fiber-based OECTs and Figure S4 in the Supporting Information shows a photograph of actual devices using graphite foil–graphite paint as contacts. To fabricate the devices, the chosen electrode (copper tape or graphite foil) was first placed on a 3D-printed support structure (see Figure 1b). Next, a PEDOT:PSS fiber was placed on top of the electrodes and, then, the contact material (silver paint (Silver Conductive adhesive 503, Electron Microscopy Sciences) or graphite paint (Graphite Conductive Adhesive 154, Electron Microscopy Sciences)) was applied to make electrical contact between the fiber and the electrodes. The fiber was left suspended above a well that was then filled with the electrolyte solution in which the Ag/AgCl gate was immersed. The curvature of the electrolyte solution at the top edges of the well, formed due to surface tension, prevented the electrical contacts from being wetted while ensuring an almost complete coverage of the fiber (see the Experimental Section and Supporting Information for more details). A Keysight B2912A Precision Source/Measure Unit was used to obtain current–voltage characteristics. The devices were then placed on a microprobe station. With two microprobes, contacts were made to the electrodes, called source and drain. OECTs were measured with 0.1 m NaCl electrolyte, where de-doping was caused by the application of a positive gate bias at the gate, so that cations from the NaCl electrolyte were injected into the channel terminal and compensated the anions on PSS chains decreasing the number of holes and so the drain current.[55] 2305371 (8 of 11) © 2023 Wiley-VCH GmbH 15214095, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202305371 by Pukyong National University, Wiley Online Library on [03/03/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License www.advancedsciencenews.com www.advmat.de OECTs were measured in depletion mode with a drain–source voltage from 0 to −1.2 V with a gate voltage from −0.57 to 0.7 V in 0.212 V steps for output characteristics, and a gate voltage from −0.57 to 0.7 V at −0.1 V drain voltage for the linear region and at −1 V for the saturation region for transfer characteristics. To obtain carrier mobilities, a gate pulse current with a specific drain voltage (0.15 V) was applied, then the source–drain current was measured. Extracting the slopes of each plot, called the transient slopes of source–drain current, and plotting them versus the gate current applied in each case, the electronic transient time could be obtained from the slope of this plot which could be correlated directly to mobility ( 𝜏 e = L2 /μVD ).[52] Finally, to measure μ using the transient method in a control sample, thin-film PEDOT:PSS OECTs were prepared using a PEDOT:PSS dispersion, by mixing Clevios PH 1000, 5 vol% ethylene glycol (EG), 1 vol% 3-glycidoxypropyltrimethoxysilane, and 0.002 vol% 4-dodecylbenzenesulfonic acid.[55,80,83,84] OECTs were fabricated in a cleanroom environment and patterned with photolithography.[85] The PEDOT:PSS dispersion was then spin-coated onto OECT substrates at 500 RPM for 15 s, followed by 2500 rpm for 45 s. The coated substrates were then baked at 140 °C for 45 min. The device dimensions for the control OECT were 500 × 500 μm. Electrochemical Impedance Spectroscopy and Cyclic Voltammetry: Electrochemical impedance spectroscopy was used to determine the PEDOT:PSS fiber capacitance. The impedance spectra were extracted from each set of electrodes with the contact material having the fiber submerged in 0.1 m NaCl aqueous electrolyte solution to function as the working electrode. An Autolab potentiostat was used to take the measurements in ambient conditions using a Pt wire as the counter electrode and an Ag/AgCl pellet as the reference electrode, applying a sine wave of 10 mV at frequencies from 105 to 0.1 Hz, and a 0.32 V DC offset potential. The data analysis was performed using Metrohm Autolab NOVA software. Capacitance was then normalized by the fiber volume to give volumetric capacitance (C*). An average C* was calculated for every contact system. Cyclic voltammetry data were obtained at a scan rate of 0.05 V s−1 for three cycles. Scanning Electron Microscopy: Images were obtained using a Hitachi S-4800 field emission scanning electron microscope at 10 keV accelerating voltage and 10 μA beam current. PEDOT:PSS fibers and contacts were imaged. Gold sputtering was not required for any of the samples due to their conductive nature. SEM was also used to obtain an average diameter of the fiber. 10 diameters were measured at different points of the sample and then they were averaged to obtain an average cross-section. Since the fibers had a circular cross-section, the area was calculated by 𝜋4 diameter2 . To image the cross-section, fibers were cut by a razor blade in liquid nitrogen to avoid deformation, then the bundle was positioned perpendicular to the electron beam. Energy-Dispersive Spectroscopy: Energy-dispersive spectroscopy (EDS) spectra were acquired using a Thermo Scientific Quanta 250 microscope, operating at an accelerating voltage of 20 kV for EDS analysis. Data from the EDS measurements were collected and analyzed using Oxford Instruments’ Aztec software. Wide-Angle X-ray Scattering: Experiments were performed in transmission mode using the Xenocs Xeuss 2.0 SAXS/WAXS system located at the Electron Microscopy Center at the University of Kentucky. The system uses a source of GeniX3D Cu ULD of 8 keV and a wavelength of 1.54189 Å. Several fibers were grouped together and placed on an aperture card. The sample was located in a vacuum chamber to obtain the 2D diffraction pattern. The sample-to-detector distance was 108.070 mm and the exposure time was 600 s. Data analysis was processed using the software Foxtrot from Xenocs. Atomic Force Microscopy: Topographical information was obtained using a Cypher S atomic force microscope operating in tapping mode. Igor Pro was used for image processing. PEDOT:PSS fibers were fixed using conductive carbon adhesive tab on a glass substrate. Kelvin Probe Microscopy: The WF of the materials was quantified by measuring contact potential difference (CPD) with an ambient Kelvin probe system, manufactured by KP Technology, LTD. The WF of the Kelvin probe tip was calibrated by making a CPD measurement on a sputtered thin film of gold with largely Au (111) orientation, and has a WF of 5.16 eV Adv. Mater. 2023, 2305371 according to the Materials Project.[86] The confidence of the WF measurements was within less than 0.01 eV.[87] X-ray Photoelectron Spectroscop: Measurements were performed on Thermo Scientific K-Alpha, which uses an aluminum mono-chromatic Xray source (1486.69 eV) and an electron flood gun for charge neutralization. XPS data were collected on three points of each sample, with X-ray spot size of 30 μm. At each point, wide survey scans were performed at a pass energy of 160 eV, and high-resolution scans were performed at a pass energy of 20 eV. Thermo Scientific software “Avantage” was used to transfer raw XPS data into excel format for further peak fittings. ENODes and Artificial Synapse Measurements: For this series of experiments, a 3D-printed stage containing two fibers was used. With the use of a Keithley, one of the two fibers was used as a gate terminal, instead of the Ag/AgCl pellet, while the other was used as a channel in a transistor configuration. Periodic voltage pulses were applied (pulses of 0.85 s, delay 8 s, 0.3 V amplitude) and the drain current was recorded, then converted into conductance. NaCl solutions with different concentrations of DA were prepared (0.025 × 10−3 and 0.05 × 10−3 m). For each experiment, the stage well was filled with the target solution in order to completely cover the fibers while avoiding contact wetting. Before each DA experiments, the fibers were rinsed in a pristine NaCl solution at least three times. Supporting Information Supporting Information is available from the Wiley Online Library or from the author. Acknowledgements The authors gratefully acknowledge the support by the National Science Foundation under Cooperative Agreement no. 1849213. Parts of the work were supported by Excet contract ID-08180074, Advanced Fiber Technologies, Inc. P.O. 22205UKY, and US Army Combat Capabilities Development Command, Chemical Biological Center. Any opinions, findings, and conclusions or recommendations expressed in this material are those of the authors and do not necessarily reflect the views of the National Science Foundation. G.M. and J.R. acknowledge the support of the Alfred P. Sloan Foundation under award FG-2019-12046. Conflict of Interest The authors declare no conflict of interest. Data Availability Statement The data that support the findings of this study are available from the corresponding author upon reasonable request. Keywords bio-hybrid technologies, contact engineering, hole mobility, organic electrochemical transistors, organic electronics Received: June 5, 2023 Revised: September 29, 2023 Published online: [1] W. Yan, G. Noel, G. Loke, E. Meiklejohn, T. Khudiyev, J. Marion, G. Rui, J. Lin, J. Cherston, A. Sahasrabudhe, J. Wilbert, I. Wicaksono, R. W. Hoyt, A. Missakian, L. Zhu, C. Ma, J. Joannopoulos, Y. Fink, Nature 2022, 603, 616. 2305371 (9 of 11) © 2023 Wiley-VCH GmbH 15214095, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202305371 by Pukyong National University, Wiley Online Library on [03/03/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License www.advancedsciencenews.com www.advmat.de [2] J. Marion, N. Gupta, H. Cheung, K. Monir, P. Anikeeva, Y. Fink, Adv. Mater. 2022, 34, 2201081. [3] T. Khudiyev, B. Grena, G. Loke, C. Hou, H. Jang, J. Lee, G. H. Noel, J. Alain, J. Joannopoulos, K. Xu, J. Li, Y. Fink, J. T. Lee, Mater. Today 2022, 52, 80. [4] W. Seung, M. K. Gupta, K. Y. Lee, K.-S. Shin, J.-H. Lee, T. Y. Kim, S. Kim, J. Lin, J. H. Kim, S.-W. Kim, ACS Nano 2015, 9, 3501. [5] Z. Zhang, J. Deng, X. Li, Z. Yang, S. He, X. Chen, G. Guan, J. Ren, H. Peng, Adv. Mater. 2015, 27, 356. [6] J. Zhong, Y. Zhang, Q. Zhong, Q. Hu, B. Hu, Z. L. Wang, J. Zhou, ACS Nano 2014, 8, 6273. [7] M. Hamedi, R. Forchheimer, O. Inganäs, Nat. Mater. 2007, 6, 357. [8] Y. Kim, H. Noh, B. D. Paulsen, J. Kim, I.-Y. Jo, H. Ahn, J. Rivnay, M.-H. Yoon, Adv. Mater. 2021, 33, 2007550. [9] Y.-C. Lai, J. Deng, S. L. Zhang, S. Niu, H. Guo, Z. L. Wang, Adv. Funct. Mater. 2017, 27, 1604462. [10] A. Leber, A. G. Page, D. Yan, Y. Qu, S. Shadman, P. Reis, F. Sorin, Adv. Funct. Mater. 2020, 30, 1904274. [11] S. Seyedin, J. M. Razal, P. C. Innis, A. Jeiranikhameneh, S. Beirne, G. G. Wallace, ACS Appl. Mater. Interfaces 2015, 7, 21150. [12] G. Loke, T. Khudiyev, B. Wang, S. Fu, S. Payra, Y. Shaoul, J. Fung, I. Chatziveroglou, P. W. Chou, I. Chinn, W. Yan, A. Gitelson-Kahn, J. Joannopoulos, Y. Fink, Nat. Commun. 2021, 12, 3317. [13] M. Kang, T.-W. Kim, Appl. Sci. 2021, 11, 6131. [14] J. Lee, S. Jeon, H. Seo, J. T. Lee, S. Park, Appl. Sci. 2021, 11, 531. [15] J. S. Heo, J. Eom, Y.-H. Kim, S. K. Park, Small 2018, 14, 1703034. [16] M. Kang, S.-A. Lee, S. Jang, S. Hwang, S.-K. Lee, S. Bae, J.-M. Hong, S. H. Lee, K.-U. Jeong, J. A. Lim, T.-W. Kim, ACS Appl. Mater. Interfaces 2019, 11, 22575. [17] G. Loke, T. Khudiyev, B. Wang, S. Fu, S. Payra, Y. Shaoul, J. Fung, I. Chatziveroglou, P.-W. Chou, I. Chinn, W. Yan, A. Gitelson-Kahn, J. Joannopoulos, Y. Fink, Nat. Commun. 2021, 12, 3317. [18] H. Zhang, Z. Wang, Z. Wang, B. He, M. Chen, M. Qi, Y. Liu, J. Xin, L. Wei, Front. Optoelectron. 2022, 15, 2. [19] J. Rivnay, S. Inal, A. Salleo, R. M. Owens, M. Berggren, G. G. Malliaras, Nat. Rev. Mater. 2018, 3, 17086. [20] R. Sarabia Riquelme, Thes. Dissert–Chem Mater Eng. 2020, 114. [21] R. Sarabia-Riquelme, R. Andrews, J. E. Anthony, M. C. Weisenberger, J. Mater. Chem. C 2020, 8, 11618. [22] R. Sarabia-Riquelme, M. Shahi, J. W. Brill, M. C. Weisenberger, ACS Appl. Polym. Mater. 2019, 1, 2157. [23] Y. Van De Burgt, E. Lubberman, E. J. Fuller, S. T. Keene, G. C. Faria, S. Agarwal, M. J. Marinella, A. Alec-Talin, A. Salleo, Nat. Mater. 2017, 16, 414. [24] X. Ji, B. D. Paulsen, G. K. K. Chik, R. Wu, Y. Yin, P. K. L. Chan, J. Rivnay, Nat. Commun. 2021, 12, 2480. [25] Y. van de Burgt, A. Melianas, S. T. Keene, G. Malliaras, A. Salleo, Nat. Electron. 2018, 1, 386. [26] M. J. Donahue, A. Sanchez-Sanchez, S. Inal, J. Qu, R. M. Owens, D. Mecerreyes, G. G. Malliaras, D. C. Martin, Mater. Sci. Eng., R 2020, 140, 100546. [27] S. Inal, J. Rivnay, A.-O. Suiu, G. G. Malliaras, I. McCulloch, Acc. Chem. Res. 2018, 51, 1368. [28] A. J. Petty, R. L. Keate, B. Jiang, G. A. Ameer, J. Rivnay, Chem. Mater. 2020, 32, 4095. [29] Y. Lee, T.-W. Lee, Acc. Chem. Res. 2019, 52, 964. [30] B. D. Paulsen, K. Tybrandt, E. Stavrinidou, J. Rivnay, Nat. Mater. 2020, 19, 13. [31] A. Marks, S. Griggs, N. Gasparini, M. Moser, Adv. Mater. Interfaces 2022, 9, 2102039. [32] A. Gumyusenge, Acc. Mater. Res. 2022, 3, 669. [33] F. Torricelli, D. Z. Adrahtas, Z. Bao, M. Berggren, F. Biscarini, A. Bonfiglio, C. A. Bortolotti, C. D. Frisbie, E. Macchia, G. G. Malliaras, Adv. Mater. 2023, 2305371 [34] [35] [36] [37] [38] [39] [40] [41] [42] [43] [44] [45] [46] [47] [48] [49] [50] [51] [52] [53] [54] [55] [56] [57] [58] [59] [60] [61] [62] [63] [64] I. Mcculloch, M. Moser, T.-Q. Nguyen, R. M. Owens, A. Salleo, A. Spanu, L. Torsi, Nat. Rev. Methods Primers 2021, 1, 66. X. Qing, Y. Wang, Y. Zhang, X. Ding, W. Zhong, D. Wang, W. Wang, Q. Liu, K. Liu, M. Li, Z. Lu, ACS Appl. Mater. Interfaces 2019, 11, 13105. A. F. Paterson, H. Faber, A. Savva, G. Nikiforidis, M. Gedda, T. C. Hidalgo, X. Chen, I. McCulloch, T. D. Anthopoulos, S. Inal, Adv. Mater. 2019, 31, 1902291. M. Berggren, G. G. Malliaras, Science 2019, 364, 233. E. A. Schafer, R. Wu, D. Meli, J. Tropp, M. Moser, I. McCulloch, B. D. Paulsen, J. Rivnay, ACS Appl. Electron. Mater. 2022, 4, 1391. J. Ashmore, A. Q. Shen, H. P. Kavehpour, H. A. Stone, G. H. Mckinley, J. Eng. Math. 2008, 60, 17. D. Quéré, Annu. Rev. Fluid Mech. 1999, 31, 347. J. C. Maxwell, Nature 1874, 10, 119. Z. A. Lamport, K. J. Barth, H. Lee, E. Gann, S. Engmann, H. Chen, M. Guthold, I. Mcculloch, J. E. Anthony, L. J. Richter, D. M. Delongchamp, O. D. Jurchescu, Nat. Commun. 2018, 9, 5130. A. F. Paterson, L. Tsetseris, R. Li, A. Basu, H. Faber, A.-H. Emwas, J. Panidi, Z. Fei, M. R. Niazi, D. H. Anjum, M. Heeney, T. D. Anthopoulos, Adv. Mater. 2019, 31, 1900871. A. F. Paterson, S. Singh, K. J. Fallon, T. Hodsden, Y. Han, B. C. Schroeder, H. Bronstein, M. Heeney, I. Mcculloch, T. D. Anthopoulos, Adv. Mater. 2018, 30, 1801079. C. Liu, Y. Xu, Y.-Y. Noh, Mater. Today 2015, 18, 79. V. Kaphle, S. Liu, A. Al-Shadeedi, C.-M. Keum, B. Lüssem, Adv. Mater. 2016, 28, 8766. V. Kaphle, P. R. Paudel, D. Dahal, R. K. Radha Krishnan, B. Lüssem, Nat. Commun. 2020, 11, 2515. J. T. Friedlein, J. Rivnay, D. H. Dunlap, I. McCulloch, S. E. Shaheen, R. R. McLeod, G. G. Malliaras7, Appl. Phys. Lett. 2017, 111, 023301. F. Bonafè, F. Decataldo, B. Fraboni, T. Cramer, Adv. Electron. Mater. 2021, 7, 2100086. H. Klauk, Adv. Electron. Mater. 2018, 4, 1700474. Z. Li, Y. Liang, Z. Zhong, J. Qian, G. Liang, K. Zhao, H. Shi, S. Zhong, Y. Yin, W. Tian, Synth. Met. 2015, 210, 363. D. Ohayon, V. Druet, S. Inal, Chem. Soc. Rev. 2023, 52, 1001. D. A. Bernards, G. G. Malliaras, Adv. Funct. Mater. 2007, 17, 3538. S. Inal, G. G. Malliaras, J. Rivnay, Nat. Commun. 2017, 8, 1767. J. Rivnay, P. Leleux, M. Ferro, M. Sessolo, A. Williamson, D. A. Koutsouras, D. Khodagholy, M. Ramuz, X. Strakosas, R. M. Owens, C. Benar, J.-M. Badier, C. Bernard, G. G. Malliaras, Sci. Adv. 2015, 1, 1400251. J. Rivnay, S. Inal, B. A. Collins, M. Sessolo, E. Stavrinidou, X. Strakosas, C. Tassone, D. M. Delongchamp, G. G. Malliaras, Nat. Commun. 2016, 7, 11287. A. F. Paterson, N. D. Treat, W. Zhang, Z. Fei, G. Wyatt-Moon, H. Faber, G. Vourlias, P. A. Patsalas, O. Solomeshch, N. Tessler, M. Heeney, T. D. Anthopoulos, Adv. Mater. 2016, 28, 7791. P. Li, T. Lei, J. Polym. Sci. 2022, 60, 377. C.-Y. Lo, Y. Wu, E. Awuyah, D. Meli, D. M. Nguyen, R. Wu, B. Xu, J. Strzalka, J. Rivnay, D. C. Martin, L. V. Kayser, Polym. Chem. 2022, 13, 2764. M. Waldrip, O. D. Jurchescu, D. J. Gundlach, E. G. Bittle, Adv. Funct. Mater. 2020, 30, 1904576. T. Langston, J. Reinf. Plast. Compos. 2009, 29, 2156. W. Garat, M. F. Pucci, R. Leger, Q. Govignon, F. Berthet, D. Perrin, P. Ienny, P.-J. Liotier, Colloids Surf., A 2021, 611, 125787. E. Lorenceau, T. Senden, D. Quéré, Molecular Gels: Materials with SelfAssembled Fibrillar Networks (Eds: R. G. Weiss, P. Terech), Springer, Dordrecht, The Netherlands 2006. H. Sirringhaus, Adv. Mater. 2005, 17, 2411. S. Chen, L. Song, Z. Tao, X. Shao, Y. Huang, Q. Cui, X. Guo, Org. Electron. 2014, 15, 3654. 2305371 (10 of 11) © 2023 Wiley-VCH GmbH 15214095, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202305371 by Pukyong National University, Wiley Online Library on [03/03/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License www.advancedsciencenews.com www.advmat.de [65] A. Cho, S. Kim, S. Kim, W. Cho, C. Park, F. S. Kim, J. H. Kim, J. Polym. Sci., Part B: Polym. Phys. 2016, 54, 1530. [66] J. Cameron, P. J. Skabara, Mater. Horiz. 2020, 7, 1759. [67] Y. Xia, G. Yan, J. Lin, Nanomaterials 2021, 11, 3119. [68] W. D. Callister, D. G. Rethwisch, Fundamentals of Materials Science and Engineering: An Integrated Approach, John Wiley & Sons, Incorporated, Hoboken, NJ 2015. [69] S. Shen, A. Henry, J. Tong, R. Zheng, G. Chen, Nat. Nanotechnol. 2010, 5, 251. [70] J. Moulton, P. Smith, Synth. Met. 1991, 40, 13. [71] J. Moulton, P. Smith, Polymer 1992, 33, 2340. [72] B. R. Mattes, H. L. Wang, D. Yang, Y. T. Zhua, W. R. Blumenthala, M. F. Hundleya, Synth. Met. 1997, 84, 45. [73] J. E. Fischer, X. Tang, E. M. Scherr, V. B. Cajipe, A. G. Macdiarmid, Synth. Met. 1991, 41, 661. [74] K. E. Aasmundtveit, E. J. Samuelsen, L. A. A. Pettersson, O. Inganäs, T. Johansson, R. Feidenhans’l, Synth. Met. 1999, 101, 561. [75] N. Massonnet, A. Carella, A. De Geyer, J. Faure-Vincent, J.-P. Simonato, Chem. Sci. 2015, 6, 412. [76] T. Takano, H. Masunaga, A. Fujiwara, H. Okuzaki, T. Sasaki, Macromolecules 2012, 45, 3859. Adv. Mater. 2023, 2305371 [77] E. A. Morris, M. C. Weisenberger, S. B. Bradley, M. G. Abdallah, S. J. Mecham, P. Pisipati, J. E. Mcgrath, Polymer 2014, 55, 6471. [78] E. Morris, M. Weisenberger, G. Rice, Fibers 2015, 3, 560. [79] S. T. Keene, C. Lubrano, S. Kazemzadeh, A. Melianas, Y. Tuchman, G. Polino, P. Scognamiglio, L. Cinà, A. Salleo, Y. Van De Burgt, F. Santoro, Nat. Mater. 2020, 19, 969. [80] A. Savva, S. Wustoni, S. Inal, J. Mater. Chem. C 2018, 6, 12023. [81] R. Wu, B. D. Paulsen, Q. Ma, J. Rivnay, Chem. Mater. 2022, 34, 9699. [82] G. M. Matrone, U. Bruno, C. Forró, C. Lubrano, S. Cinti, Y. Van De Burgt, F. Santoro, Adv. Mater. Technol. 2023, 8, 2201911. [83] Y. Zhang, S. Inal, C.-Y. Hsia, M. Ferro, M. Ferro, S. Daniel, R. M. Owens, Adv. Funct. Mater. 2016, 26, 7304. [84] M. ElMahmoudy, S. Inal, A. Charrier, I. Uguz, G. G. Malliaras, S. Sanaur, Macromol. Mater. Eng. 2017, 302, 1600497. [85] V. N. Le, J. H. Bombile, G. S. Rupasinghe, K. N. Baustert, R. Li, I. P. Maria, M. Shahi, P. Alarcon Espejo, I. McCulloch, K. R. Graham, C. Risko, A. F. Paterson, Adv. Sci. 2023, 10, 2207694. [86] A. M. Mántica, T. J. Balk, in 2020 IEEE 21st Int. Conf. Vacuum Electronics (IVEC), IEEE, Piscataway, NJ, USA, 2020, pp. 391–392. [87] A. M. Mantica, M. J. Detisch, T. J. Balk, IEEE Trans. Electron Devices 2023, 70, 2864. 2305371 (11 of 11) © 2023 Wiley-VCH GmbH 15214095, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202305371 by Pukyong National University, Wiley Online Library on [03/03/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License www.advancedsciencenews.com