Introduction

Doping is a ubiquitous technique for enhancing the electrical properties of traditional semiconductor materials1,2. Since the pioneering reports from decades ago3,4,5,6,7,8,9, its extension to organic semiconductors has significantly boosted their conductivity by several orders of magnitude reaching as high as thousands of S/cm upon doping10,11,12. This adaptation has been critical in improving the device performance of organic light-emitting diodes, transparent conductors, electrochromic devices, bioelectronics and beyond13,14,15,16,17,18. Meanwhile, there has been tremendous progress in understanding the fundamental mechanisms of chemical doping in organic semiconductors19,20,21,22. Nevertheless, knowledge of doping in organic systems still lags that of their inorganic counterparts with fundamental mechanisms still being unraveled23.

For organic polymeric semiconductors, one of the critical challenges is to understand the complex interactions between chemical doping and morphologies such as polymer conformation, packing, and crystallinity24. For instance, the presence of significant morphological disorder is usually responsible for charge trapping and low doping efficiency24. Consider the extensively studied case of 2,3,5,6-tetra-fluoro-7,7,8,8-tetracyanoquinodimethane (F4TCNQ) doped poly(3-hexylthiophene) (P3HT), where only about 5% of the generated charge carrier pairs are found to dissociate and contribute to a free hole for electrical conduction25. Therefore, it has been proposed that films with higher crystallinity yield significantly higher doping efficiency compared to those with amorphous ones, attributed to the better delocalization of charge carriers within crystallites and the suppression of trapped charges26,27,28. On the other hand, there are proposals that low-crystallinity films can be doped more efficiently than their high-crystallinity counterparts as they can tolerate more dopants in loosely packed structures29,30. Furthermore, the microstructures can be significantly modified by introducing dopants, and both beneficial and detrimental effects on morphological ordering and charge transport have been reported22,31. We believe the key to resolving these conflicting results lies in unraveling the little known assembly pathways from solution to solid states and the resultant microstructures32, which have been rarely investigated in the context of chemical doping of conjugated polymers. Recently, some of us have discovered chiral assembly pathways of several achiral/racemic conjugated polymers33,34,35, which provides the opportunity to explore how chirality influences chemical doping. Notably, chirality has been shown to play a crucial role in various chemical reactions, including the oxygen evolution reaction and organic electrosynthesis36,37, by means of the chiral-induced spin selectivity (CISS) effect38. It remains an open question how chirality re-writes rules for redox reactions that underpin chemical doping of electronic materials – a topic not explored thus far.

In this study, we significantly modulated the chiral liquid crystal (LC) mediated assembly pathways of a racemic conjugated polymer and identified the crucial role of chiral helical assemblies in dictating the extent of redox reactions during doping. By manipulating the solution aggregate structures into networks, rigid fibers, and semiflexible fibers, we obtained three distinct assembly pathways, mediated by an isotropic phase, a continuous twist-bend nematic (Ntb) mesophase, and a striped Ntb mesophase, respectively. These pathways further resulted in the formation of three types of films with distinct multiscale morphologies: amorphous films with randomly oriented polymer chains, semicrystalline films with uniaxially-oriented helical fibers characterized by a large pitch length, and semicrystalline films with zigzag-oriented helical fibers with a small helical pitch. Also, the films exhibited varying degrees of supramolecular chirality—from achiral to weakly chiral, and to strongly chiral—corresponding to the presence or absence of Ntb mesophases. Following sequential doping, the strongly chiral film demonstrated the highest doping level, with a charge carrier concentration more than two-fold higher than that of the other two films. We discuss potential mechanisms underpinning this observation based on morphological distinctions and the chirality-induced spin selectivity (CISS) effect: The highest crystallinity and closest packing in the strongly chiral film may facilitate doping by promoting charge delocalization and reducing charge trapping; in addition, the doping process may also be facilitated by supramolecular chirality through the CISS effect by enhancing the redox reaction asymmetry. Notably, across three sets of polymer systems—a dioxythiophene copolymer (PE2-biOE2OE3), a diketopyrrolopyrrole-based polymer (DPP2T-TT), and two 1,4-(2-thienyl)−2,5-dialkoxyphenylene-based polymers (TPT-2T and TPT-T)—we observed increased conductivity in chiral structures. We present an in-depth analysis of the PE2-biOE2OE3 results and provide a summary of the findings for DPP2T-TT and TPT polymers in the final section of this manuscript.

Results

Solution-state structure and assembly pathway

A model polymer, PE2-biOE2OE3, a dioxythiophene copolymer comprised of 2, 2’-bis-(3,4-ethylenedioxy)thiophene (biEDOT) and 3,4-propylenedioxythiophene (ProDOT) substituted with branched oligo(ether) side chains39, displays distinct assembly pathways from solution to solid state in three ether solvents: 1,4-dioxane, diglyme, and 1,3-dioxane, despite the solvents being close structural analogs (Fig. 1). This is evidenced by the emergence or absence of LC mesophases beyond a critical concentration, observed both under nonequilibrium droplet-drying processes and in equilibrium solution states with progressively increased concentrations34,35,40. As shown in Fig. 2A and Movie S1, while observing a drying droplet of a 5 mg/mL 1,4-dioxane solution under a cross-polarized optical microscope (CPOM), we found the concentrated edge exhibits no birefringence, indicative of an isotropic phase-mediated polymer assembly. This is consistent with the equilibrium phases at high concentration (up to ~150 mg/mL) where the solution displays no birefringence under CPOM (Fig. 2B and S2). Furthermore, circular dichroism (CD) spectra confirm the achiral nature of the solution (Figure S2). To accurately capture the true CD signal, we used a four-scan method to eliminate linear anisotropic contributions, as previously reported41. We note that the sp³ carbon chiral centers on the side chains likely existing in a racemic form, not contributing to molecular chirality in isotropic solutions. In contrast, the droplet edge of a diglyme solution displays a continuous birefringence with a uniform texture (Fig. 2A and Movie S2). In the equilibrium solution state, we reproduced this uniform texture along with a singularity characterized by four surrounding schlieren brushes42, which resembles the texture of a nematic phase (Fig. 2B and S2)42. Additionally, CD spectra reveal the emergence of chirality in the mesophase, characterized by an anisotropic dissymmetry factor (g-factor) of 10-3 in magnitude (Fig. 2C and S2; see definition of g-factor in “Methods section”). We define the mesophase as a continuous Ntb phase. In the case of the 1,3-dioxane solution, the drying meniscus exhibits stripe-like textures (Fig. 2A and Movie S3). This resembles the morphology of a striped Ntb phase that we previously identified in the solution of several conjugated polymers adopting helical chain conformation33,35. The morphology is reproducible at solution concentrations of 90–150 mg/mL (Fig. 2B and S2), featuring twin domains that extinguish cross-polarized light at mirroring angles (Figure S3). Additionally, CD spectra uncovered the emergence of strong chirality in the mesophase, characterized by a g-factor ranging from 10-2 to 10-1 in magnitude reaching values as high as 0.141 (Fig. 2C and S2). These findings indicate that the assembly pathways of PE2-biOE2OE3 are mediated by an isotropic phase in 1,4-dioxane, a continuous Ntb phase in diglyme, and a striped Ntb phase in 1,3-dioxane, respectively. These phase assignments are supported by in-depth structural characterizations combining molecular scale X-ray diffraction, meso-scale imaging and micron-scale microscopy (shown later).

Fig. 1: Distinct solution-state assembly pathways of PE2-biOE2OE3 and their impact on film chirality and conductivity.
figure 1

(Top) Chemical structure of PE2-biOE2OE3 and its solution-state assembly pathways mediated by an isotropic phase, a weakly chiral liquid crystal phase, and a strongly chiral liquid crystal phase. (Bottom) Distinct assembly pathways result in films with modulable electrical conductivity upon doping that scales with the absolute dissymmetric factor, |g-factor | . Error bars indicate the standard deviation based on measurements from more than ten individual films.

Fig. 2: Solution-state assembly pathways of PE2-biOE2OE3 in 1,4-dioxane, diglyme, and 1,3-dioxane.
figure 2

A Drying drops reveal that isotropic, striped twist-bend nematic, and continuous twist-bend nematic phases form in 1,4-dioxane, 1,3-dioxane, and diglyme, respectively. Arrows indicate the drop drying direction. B Direct observation of solution-state structures at 90 mg/mL under a cross-polarized optical microscope (CPOM). C Anisotropic dissymmetry factor (g-factor) from circular dichroism (CD) spectra at various concentrations (60–150 mg/mL). D Scanning electron microscopy (SEM) images of polymer samples freeze-dried from 5 mg/mL solutions. E Snapshots of separated and packed PE2-biOE2OE3 chains from molecular dynamics (MD) simulations. F Total cumulative angle ∆θtot between the conjugated units along one chain extracted from the MD simulations. Large errors in the 1-chain data indicate no inherent helicity preference in isolated chains. As chains aggregate into 4-chain assemblies, the reduced error bars and an increased ∆θtot in diglyme and 1,3-dioxane suggest an emergent tendency toward helical ordering upon aggregation. G Flory–Huggins interaction parameter (χ) for the backbone, side chains, and the entire polymer in three solvents: 1,4-dioxane, diglyme, and 1,3-dioxane at room temperature.

We next uncover that the three distinct assembly pathways arise from distinct solution-state aggregate structures. We first characterized these aggregate structures using both real-space imaging and reciprocal-space scattering techniques. As shown in Fig. 2D and S4, scanning electron microscopy (SEM) reveals three-dimensional network-like aggregates with a diameter of 251 ± 87 nm in 1,4-dioxane. In contrast, polymers in diglyme aggregate into rigid fibers with an average diameter of 133 ± 31 nm, while in 1,3-dioxane, we observe flexible fiber aggregates with an average diameter of 22 ± 15 nm. These findings suggest a trend in polymer backbone rigidity and degree of aggregation in the order of 1,4-dioxane <1,3-dioxane <diglyme. Support for this inference comes from small-angle X-ray scattering (Figure S5, Table S1), which informs an increasing polymer persistence length (lp,p) from 4.4 ± 0.2 nm in 1,4-dioxane, to 7.0 ± 0.5 nm in 1,3-dioxane, and to 9.5 ± 0.4 nm in diglyme. Further evidence is provided by UV-vis absorption spectra, which show that the absorption coefficient, the degree of red-shift in absorption peaks, and the resilience to disaggregation by heating are highest in diglyme, followed by 1,3-dioxane, and lowest in 1,4-dioxane (Figures S6 and S7). Fluorescence spectra indicate that these distinct aggregate structures are possibly different polymorphs, as detailed in the supplementary information (SI section II). Overall, these results show that formation of semiflexible fibril aggregates in solution are conducive to chiral LCs during evaporative assembly. We have found that the emergence of chirality in the lyotropic mesophases of achiral or racemic polymers is a widespread phenomenon, observed in over twenty such systems, including several previously reported examples33,34,35,43. This chiral emergence, occurring without the addition of chiral bias (such as chiral small molecules), suggests a mechanism distinct from the classic “sergeants-and-soldiers” effect44,45,46. We hypothesize that the formation of helical structures is entropy-driven, where polymers adopt a helical conformation to reduce excluded volume and maximize solution entropy. This proposed mechanism aligns with the theoretical model by Kamien et al.47, based on classic depletion interaction theory. We have also substantiated generality of chiral emergence across 20 polymer systems in a subsequent manuscript48. In this work we choose to focus on the role of chirality in chemical doping specifically.

We further explored the molecular underpinnings behind the distinct helical assemblies, identifying side-chain aggregation as a key factor through molecular dynamics (MD) simulations. To investigate this, we used MD simulations to highlight specific trends of polymer behavior in different solvents. We first calculated the cumulative angle (∆θtot) between conjugated rings along the polymer backbone to evaluate helical propensities (Figs. 2E and 2F). Ideally, in the absence of helical preference, ∆θtot should be ~0°. Our simulations indicate that isolated chains (consisting of 12-mers) maintain a ∆θtot near 0° across all solvents, showing no inherent helicity. However, upon aggregation, ∆θtot significantly diverges, reaching 132.1–142.8° in 1,3-dioxane and 89.2-91.9° in diglyme, while remaining unchanged in 1,4-dioxane. These results align with the experimental observations of g-factors (Fig. 2C), confirming that PE2-biOE2OE3 exhibits the highest extent of helicity in 1,3-dioxane, lower in diglyme, and negligible in 1,4-dioxane. It’s important to note that Δθtot is not indicative of rigidity or persistence length—e.g., both flexible coils and rigid rods can show Δθtot = 0. Instead, Δθtot reflects the consistency of helical twist along the polymer contour, correlating with helicity rather than with rigidity or aggregate size.

We then assessed the Flory–Huggins interaction parameter (χ) for the backbone, side chains, and the entire polymer in three solvents: 1,4-dioxane, diglyme, and 1,3-dioxane, which provides an established metric for evaluating solubility and aggregation tendencies (Fig. 2G). The calculated χ values for the polymer backbone in all three solvents remain below the critical χ threshold for phase separation, indicating good solubility. Among them, 1,4-dioxane displayed the lowest χ value, followed by diglyme, and the highest χ was found in 1,3-dioxane, though still within the miscible regime. A similar solvent dependence was observed for the side chains; however, in 1,3-dioxane, the χ value sharply increased to 2.2 well above the critical point, revealing a strong thermodynamic driving force for side-chain aggregation. This trend is also reflected in the χ values for the entire polymer, which follow the same order: 1,4-dioxane <diglyme <1,3-dioxane, with the polymer in 1,3-dioxane crossing the miscibility limit.

These results suggest that the enhanced helicity observed in 1,3-dioxane is primarily driven by side-chain aggregation, which overrides backbone planarity and induces a packing motif favoring side-chain interactions. Such a competition between side-chain ordering and backbone planarity in conjugated polymers has been previously reported, albeit not in the context of helical assemblies49. In contrast, diglyme promotes more balanced interactions: the χ values for side chains and backbone are comparable, reducing competition and allowing backbone-driven packing into planar conformations. In 1,4-dioxane, the polymer is the most soluble across all solvents resulting in weak driving force for either side-chain aggregation or backbone stacking, which in turn contributes to disordered chain conformation and network structures. While these results do not explicitly validate the proposed model, they offer valuable molecular-level insights that align with experimental observations, reinforcing the model’s relevance in elucidating the mechanisms driving the observed chiral assembly.

Solid-state morphology and electrical conductivity

The solution-state structures and assembly pathways directly determine the morphology of thin films prepared by meniscus-guided coating from three solvents (Figs. 3 and 4). These films, coated in the evaporation regime50, maintain a thickness of ~400 nm (Figure S10). This thickness approaches the optimal level previously identified for enhancing this polymer’s electrical conductivity39. We first examined the micron- and meso-scale morphology using CPOM, SEM and atomic force microscopy (AFM) (Fig. 3). The film coated from 1,4-dioxane solution displays no birefringence under CPOM, and its AFM phase image shows an isotropic structure, indicating the preservation of the polymer’s disordered state from the isotropic solution into the film. CD spectra further confirm the film’s achiral nature (Fig. 3D and S11). The film from diglyme exhibits an aligned, birefringent morphology under CPOM, resembling the texture of the continuous Ntb phase in solution. SEM images show the film is comprised of twisted fibers with a diameter Dfiber, WC of 171 ± 28 nm and a large helical pitch l1/2p, WC of 617 ± 117 nm, giving rise to g-factors ranging from 0 to 10-4 (Fig. 3D and S11). The film from 1,3-dioxane solution displays a twinned morphology with an 8.9 ± 2.4 µm periodicity, which resembles the texture of the striped Ntb mesophase observed in solution. AFM phase image reveals the mesoscale helically twisted fibers with an average diameter Dfiber, WC of 47 ± 13 nm and a helical pitch l1/2p, WC of 83 ± 14 nm, much smaller than in the diglyme case. Such highly twisted fibers give rise to high g-factors ranging from 10-3 to 10-2 (Figure S11). Henceforth, we have designated the films from 1,4-dioxane, diglyme, and 1,3-dioxane as achiral, weakly chiral, and strongly chiral, respectively, based on their g-factor and helical pitch across scales. It should be borne in mind that ‘achiral’ here refers to structural rather than chemical achirality in the film coated from 1,4-dioxane, as racemic side chains are attached to the polymers.

Fig. 3: Multiscale characterization of film morphology.
figure 3

A CPOM images of the films printed from different solvents. White and yellow arrows indicate the film printing direction and main axis of polarizers, respectively. B Atomic force microscopy (AFM) (left and right) phase and scanning electron microscope (middle) images of printed films. White arrows indicate film printing directions. Yellow arrows indicate fiber twists. C Grazing-incidence X-ray diffraction (GIXD) 2D images. Dashed lines indicating 0° and 90° \(\chi\) angles. Error indicates the location of 45° tilted helical backbone stacking peaks. D Circular dichroism (CD) spectra of printed films. E GIXD linecuts along \(\chi\) angle (−90° to 90°) at fixed qr = 1.35 Å. F Relative degree of crystallinity (rDoC) for lamellar, π-π and helical π-π stackings. G Lamellar dlamellar and π-π stacking dπ-π distances extracted from GIXD 1-D linecuts. H Free volume of packed chains upon solvent evaporation in MD simulations and film density calculated from XRR critical angle. Error bars for the density values indicate the standard deviation from measurements of three individual films.

Fig. 4: Schematic illustration of multiscale morphology of three printed films, with white arrows indicating film printing directions.
figure 4

The achiral film displays randomly oriented polymer chains, with a diameter (Dpolymer, AC) of 14.9 ± 0.3 nm, with π-π stacking (dπ-π) of 3.83 Å and lamellar stacking (dlamellar) of 23.38 Å, showing lowest crystallinity (rDoC(π-π) = 1.14, rDoC(lamellar) = 1.43). The weakly chiral film features uniaxially aligned helical fibers, with diameter (Dfiber, WC) of 171 ± 28 nm, pitch length (l1/2p, WC) of 617 ± 117 nm, helical backbone stacking peak at χ = ± 45° (4.65 Å), out-of-plane dπ-π of 3.75 Å, and in-plane dlamellar of 25.96 Å, and higher crystallinity (rDoC(π-π) = 2.04, rDoC(lamellar) = 4.19). The strongly chiral film exhibits twinned domains with a periodicity of 8.9 ± 2.4 μm, composed of helical fibers with diameter (Dfiber, SC) of 47 ± 13 nm, pitch length (l1/2p, SC) of 83 ± 14 nm, a backbone stacking peak also at χ = ± 45° (4.65 Å) and the closest dπ-π of 3.68 Å, largest dlamellar of 26.97 Å, and the highest crystallinity (rDoC(π-π) = 3.49, rDoC(lamellar) = 5.84).

We further analyzed the polymer packing at molecular scale using grazing-incidence X-ray diffraction (GIXD) (Fig. 3C and S12S17)51. The chiral emergence correlates with a 45° tilted peak at qr = 1.35 Å−1 attributed to the helical backbone stacking (Figs. 3C and 4). This feature is the most pronounced in strongly chiral films, less so in weakly chiral films, and absent in achiral films (Fig. 3E). Such tilted peaks are also noted in other chiral material systems52,53, and it has been suggested that the 45° tilt indicates a tightly packed helical structure54,55. We indeed observed a decrease in backbone π-π packing distances (dπ-π) from 3.80 Å in achiral films to 3.75 Å in weakly chiral films, and further to 3.68 Å in strongly chiral films, implying a tighter backbone packing in helical structures (Fig. 3G and Table S3). Conversely, the lamellar stacking distance (dlamellar) follows an opposite trend, increasing from 24.26 Å to 25.96 Å and then to 26.97 Å. This may result from the difference in side chain solubilities as suggested by χ interaction parameter shown in Fig. 2G. Poor side chain solubility in 1,3-dioxane promotes side chain aggregation and extended conformations, increasing lamellar spacing. In contrast, good solvents like 1,4-dioxane keep side chains solvated, enabling reorganization and tighter packing during film formation, resulting in smaller lamellar distances. Additionally, we observed an in-plane alkyl stacking peak at qxy = 1.417 Å−1 in the strongly chiral film, corresponding to adjacent side chain stacking distances of 4.43 Å. We attribute this to a pronounced side-chain aggregation tendency in 1,3-dioxane, which is corroborated by the solvation energy calculations shown in Fig. 2G. Furthermore, the strongly chiral film demonstrates the highest relative degree of crystallinity (rDoC) for both lamellar and (helical) π-π stacking, exceeding that of the weakly chiral film, while the achiral film displays the lowest rDoC values due to its amorphous structure, as shown in Fig. 3F. This suggests that chiral helical structures are thermodynamically favorable, driven by denser packing. MD simulations and X-ray reflectivity (XRR) experiments confirm this trend, where MD shows a decreasing free volume trend—37.1% in achiral, 36.1% in weakly chiral, and 35.4% in strongly chiral films—and XRR shows an increasing density trend—1.18 g/cm³ in achiral, 1.20 g/cm³ in weakly chiral, and 1.24 g/cm³ in strongly chiral films (Fig. 3H and S18, detailed in SI section IV).

To evaluate the effect of PE2-biOE2OE3’s solution-state assembly pathway and solid-state morphology on its electronic properties, we measured the films’ electrical conductivity upon doping. After solution coating, PE2-biOE2OE3 thin films were sequentially doped with an isopropyl alcohol solution of iron (III) tosylate (Fe(Tos)3) or nitrosonium tetrafluoroborate (NOBF4) placed on top of the films, allowing the dopant molecules to penetrate the film (Fig. 5A). We then obtained the electrical conductivity values using van der Pauw method (Fig. 5B). Upon doping with Fe(Tos)3, the achiral film exhibits a conductivity of 41 ± 24 S/cm. In comparison, the weakly chiral film shows a 3.3 times higher conductivity at 136 ± 26 S/cm, while the strongly chiral film’s conductivity is 7.4 times greater than the achiral case, reaching 305 ± 42 S/cm. Films doped with NOBF4 also exhibit a similar trend as shown in Figure S19. To explain the differences in the electrical conductivity across the three films, we quantified the extent of doping and charge carrier density using a combination of X-ray photoelectron spectroscopy (XPS) and optical absorption spectroscopy. As shown in Figure S20, the XPS S2p features from doped films were deconvoluted into two contributions: polymer sulfur peaks occur between ca.163.5 and 166.0 eV, and tosylate S = O peaks occur between 167.0 to 170.0 eV (Table S6)56. The polymer sulfur peaks were further deconvoluted into neutral sulfur (S) and charged sulfur (S*) following previously reported procedures (detailed in SI section IV)56. We then calculated the carrier ratio (c) of the S* peak area to the total S* + S peak area, which is assumed to be equal to the number of holes per thiophene unit56. The carrier ratios are comparable in achiral and weakly chiral films, reaching 0.14 ± 0.02 and 0.12 ± 0.03, respectively. In contrast, the strongly chiral film demonstrates a roughly two-fold increase, achieving a ratio of 0.26 ± 0.03. The carrier concentration can be estimated from the carrier ratio, using a previously reported method described in SI Section IV. The drift mobility is then calculated by dividing the measured electrical conductivity by the estimated carrier concentration56. As shown in Fig. 5C, the strongly chiral film shows the highest level of doping with the highest carrier concentration of (4.9 ± 0.6) × 1020 cm-3. In contrast, the weakly chiral and achiral films show relatively lower carrier concentrations of (2.2 ± 0.6) × 1020 and (2.5 ± 0.4) × 1020 cm-3, respectively. In addition, compared to the polymer sulfur peaks before doping, the S2p peak shifts to lower binding energy with increased doping density (Figure S20)57. The strongly chiral film shows the largest extent of shift, 0.41 eV compared to 0.26 eV in the other two films, further confirming its highest level of doping. Furthermore, the degree of oxidation can also be quantified by determining the abundance ratio of tosylate counterions (using the sulfoxide doublet between 167.0 to 170. eV) with respect to thiophene58. As shown in Fig. 5E, this trend also aligns closely with the calculated carrier concentration and the shifts in the S2p peaks. Due to the surface-sensitive nature of XPS, which exclusively probes a limited volume close to the surface, we employ the optical absorbance of the charge carriers (i.e., polaron and bipolarons) to assess the representativeness of the results for the bulk material. As illustrated in Fig. 5D, the chiral film exhibits notably higher charge carrier absorbance; quantitative comparisons were performed by integrating the area under the curve. As shown in Fig. 5E, the relative carrier ratios obtained from S*/(S + S*) in XPS align well with those from charge carrier absorbance. In sum, all these four pieces of evidence support that the strongly chiral films exhibit the highest doping level. As for the nature of the charge carriers after doping, we believe that in all three cases, the charge carriers are predominantly di-cation fragments based on the absence of the characteristic PE2 polaron absorption peak near 1000 nm and the presence of a broad absorbance extending deep into the near-infrared (NIR) region39,59.

Fig. 5: Characterization of doped films.
figure 5

A Schematic of the doping reaction of PE2-biOE2OE3 with the molecular structures of the dopants Fe(Tos)3 and NOBF4 used in this study. B Conductivity measured from three films upon doping with Fe(Tos)3. Conductivity values represent the average of measurements from more than ten independently prepared films doped with Fe(Tos)3. Error bars indicate the standard deviation across these independent samples. C Calculated mobility and charge carrier concentration. Standard deviation in carrier concentration arises from propagation of errors in X-ray photoelectron spectroscopy (XPS) peak fitting of one set XPS data and X-ray reflectivity (XRR) density measurements on three independent films. Mobility standard deviation is obtained from propagation of conductivity and carrier concentration standard deviations. D Optical absorption spectra of doped films. E Quantification of doping level from XPS and absorption. Error bars for the S*/S and S(Tos)/S(P) peak area represent the standard error derived from XPS peak fitting. Error bars for absorption represent the standard deviation based on measurements from three individual films. F Relative degree of crystallinity (rDoC) before and after doping. G Schematic illustrations of microstructural changes upon doping in strongly chiral and achiral films.

We further analyzed the films’ morphological changes upon doping. As shown in Figure S21, AFM images suggest that the mesoscale morphologies closely resemble those before doping. The molecular scale packing, however, is significantly modified as revealed by GIXD and CD. First, the dπ-π reduces and the dlamellar expands in all three films (Figure S22 and Table S4), which is commonly observed and suggests that the dopants are mostly located between the side chains22. Second, the variation in rDoC values across the three films is significantly distinct (Fig. 5F). In the strongly chiral film, the lamellar stacking rDoC increases significantly from 5.84 to 10.32, indicating a dopant-induced enhancement in lamellar order, aligning with earlier findings56. In contrast, the change in the weakly chiral film is minor, moving from 4.20 to 5.52. The achiral film shows minimal change, with rDoC values shifting only slightly from 1.43 to 1.61, implying that the dopant preferentially resides in the amorphous regions of this low-crystallinity film22. The π-π stacking rDoC drops in both chiral films, indicating a disruption of ordering in the backbone stacking region due to enhanced side-chain ordering49. Moreover, the characteristic 45° tilted peak in the amorphous ring region at qr = 1.35 Å−1 is significantly reduced in the strongly chiral film (Figure S23), leading to a substantial decrease in the rDoC of helical backbone stacking. In contrast, the helical backbone stacking rDoC in the weakly chiral film shows only a slight alteration, consistent with the largely unchanged tilted peak signature. This observation further aligns with the CD spectra after doping (Figure S23). The g-factors for weakly chiral films remain within the range of 0 to 10-4, while for strongly chiral films, they decrease from 10−3-10−2 before doping to predominantly 10-4 after doping (Figure S23). Moreover, the in-plane alkyl stacking peak diminishes upon doping (Figure S22), implying dopants enter the lamellar region and disrupt the lateral packing between side chains. These inferred morphological changes upon doping are summarized in Fig. 5G. All these analyses indicate that a higher quantity of dopants has entered the strongly chiral film, resulting in more pronounced changes in the crystalline region of the films compared to the other two cases, which is consistent with the XPS elemental analysis shown in Fig. 5E.

Proposed mechanisms for enhanced doping by supramolecular chirality

What enables the strongly chiral film to incorporate substantially more dopants at the expense of pronounced structural changes? We first note that all three films exhibit similar ionization energies despite their structural differences: 4.64 eV in both the achiral and weakly chiral films, and 4.67 eV in the strongly chiral film (Figure S24). We now turn to possible morphological impact on doping efficiency. The strongly chiral film demonstrates the highest crystallinity in both backbone and side-chain stacking regions and has the highest density, factors that could typically hinder dopant penetration due to reduced ease of swelling29,30. While it possesses slightly larger lamella spacing (by < 1 Å) and slightly larger volume of the side chain region where the dopants reside (Figures S25-S27), such small differences can hardly account for a nearly doubled doping level in this film compared to the achiral and weakly chiral cases. However, despite the potential for impeded dopant diffusion, previous works have shown or implied in a number of instances, that high crystallinity can facilitate chemical doping. For example, this phenomenon has been observed in chemical doping of graphite, polyacetylene, poly(phenylene sulfide), or poly(3-hexylthiophene)26,27,28,60,61. It was proposed that the ordered structure of polymer chains in backbone crystalline domains can serve as an effective precursor for doped species26,27. The high planarity of these chains also facilitates the delocalization of hole carriers, thereby improving doping efficiency. In this work, the close-packing nature and high crystallinity of the strongly chiral films may impose molecular-level constraints and reduce dihedral angle distribution. This could lead to fewer charge traps and lower activation energy62, enhancing charge delocalization and further improving doping efficiency. This suggests that the superior doping levels observed in strongly chiral films can in part be attributed to their higher degree of crystallinity and reduced structural disorder, alongside the reduced π-π stacking distance dπ-π, allowing for improved interchain charge delocalization (Fig. 6A). Nonetheless, the trend in doping levels also reflects other important factors beyond differences in crystallinity. This is evidenced by the achiral and weakly chiral films, which, despite having significant differences in their degree of crystallinity, show similar charge carrier concentrations. Further, while there is only moderate gain in crystallinity from weakly to strongly chiral films, the conductivity gain is substantial.

Fig. 6: Hypothesis on the mechanism underpinning chirality-modulated chemical doping.
figure 6

A Illustration of the hypotheses for enhanced doping efficiency in the strongly chiral films. Arrows indicate charge carrier transfer direction. I: The high crystallinity in chiral films enhances charge delocalization and reduces trapped charges compared to achiral films, resulting in a higher doping efficiency. II: The (chiral-induced spin selectivity) CISS effect improves doping efficiency by affecting the reaction asymmetry: while the oxidation reaction is promoted, the reduction reaction is reduced. B Forward (oxidation) and reverse (reduction) reaction rate constants \({k}_{{ox}}^{\circ }\) and \({k}_{{red}}^{\circ }\) measured using scanning electrochemical microscopy (SECM). The log \(k\) vs. potential (η) plots were fitted with the Butler-Volmer model to determine reaction rate constants (\({k}^{o}\)) from intercepts. The fitting region is -0.05 to -0.20 V for oxidation reaction and 0.05 to 0.20 for reduction reaction. The cyan area around the curve represents the error bar for log \(k\). C The redox reaction asymmetry quantified by the normalized ratio of \({k}_{{ox}}^{\circ }\)/\({k}_{{red}}^{\circ }\) across three films, and the extent of oxidation quantified by the normalized ratio of charged S*/neutral S peak area from XPS. The error bars for the reaction rate constants reflect the standard deviation from three individual films, whereas the error bars for the S*/S peak area represent the standard error derived from XPS peak fitting from one set of XPS data.

Intriguingly, another potential contributing factor could be the impact of chirality on the redox reaction during doping. Indeed, recent studies have highlighted the presence of an unconventional spin-orbit coupling (SOC) in helical structures of chiral conjugated polymers63,64, which underpins the spin-selective electron transport known as the CISS effect38. CISS has proven effective in manipulating the electron spin, which is crucial in chemical bonding and reactive processes. For example, CISS has been shown to enhance the efficiency of oxygen evolution and reduction reactions65,66,67, facilitate organic electrosynthesis and polymerization37,68, and influence electron transfer kinetics in biological redox processes69. An earlier study has indicated that the CISS effect causes an asymmetry in the oxidation and reduction rate constants during electron transfer processes through oligopeptides. This asymmetry, observed in both forward and backward electron transfers, is attributed to spin-selective electron transfer mechanisms within these chiral molecules70. Thus, the supramolecular chirality in the strongly chiral PE2-biOE2OE3 film along with the higher crystallinity could enhance the polymer oxidation rate relative to the reduction rate by enhancing redox reaction asymmetry during doping (Fig. 6A). To probe this hypothesis, we performed kinetic measurements of the Fe3+/Fe2+ redox couple via scanning electrochemical microscopy (SECM) on PE2-biOE2OE3 films (Figures S30 and S31). The results reveal that the forward (oxidation) reaction is enhanced, while the reverse (reduction) reaction is reduced in the strongly chiral films (Figure S6B and Table S8). The findings reveal a significant enhancement in the oxidation reaction rates in strongly chiral films, where \({k}_{{ox}}^{\circ }\) = (9.49 ± 2.09) × 10-6 m/s, which is more than twice the rates observed in weakly chiral films (\({k}_{{ox}}^{\circ }\) = (4.36 ± 0.72) × 10-6 m/s) and achiral films (\({k}_{{ox}}^{\circ }\) = (4.57 ± 1.11) × 10-6 m/s). Additionally, the reduction reaction rates increase by a factor of 33% when transitioning from \({k}_{{red}}^{\circ }\,\)= (6.39 ± 0.48) × 10-6 m/s in strongly chiral films to \({k}_{{red}}^{\circ }\,\)= (8.48 ± 0.71) × 10-6 m/s in achiral films. In comparison, only a 6% increase is observed when moving from \({k}_{{red}}^{\circ }\,\)= (7.97 ± 0.27) × 10-6 m/s in weakly chiral films to achiral films, as shown in Fig. 6B and Table S8. The extent of redox asymmetry measured by SECM is highly correlated with the extent of oxidation across the three films measured by XPS (Fig. 6C). It must be noted that, as the oxidation and reduction reactions were carried out in distinct reaction media (iron (II) and iron (III) tosylate solutions, respectively) (Figures S30 and S31), the absolute values of \({k}_{{ox}}^{\circ }\) and \({k}_{{red}}^{\circ }\) cannot be directly compared; only the oxidation-to-reduction rate comparisons are meaningful.

The possible manifestation of the CISS effect on the observed reaction asymmetry is supported by the exceptionally high SOC in the strongly chiral films inferred from the damping factors \(\alpha\) measured by the ferromagnetic resonance (FMR) technique63. As shown in Fig. 7, the strongly chiral film showed a larger damping factor \(\alpha\) = 0.033 compared to that of weakly chiral (\(\alpha\) = 0.010) and achiral (\(\alpha\) = 0.022) films, reaching a value on par with that of a heavy metal such as platinum71. We further validated that the damping factor is measured in the saturation regime and thus intrinsic to the polymer material (Figure S32). This large damping factor suggests that the strongly chiral film possesses a surprisingly large chirality-induced SOC42 underpinning a possible CISS effect, corroborating the observed redox reaction asymmetry. The fact that the achiral film exhibits a moderately high \(\alpha\) value of 0.022 can be attributed to the highly torsional chain conformation rather than chirality (Figure S28)63. This trend in the magnitude of the damping factor \(\alpha\) aligns with the oxidation trend, where weakly chiral films exhibit the lowest degree of oxidation, followed by an increase in achiral films, and a significant increase in strongly chiral films (Fig. 7E).

Fig. 7: Mechanism of chirality-induced chemical doping and characterization of damping factors.
figure 7

A Schematic of the CISS. When electrons move through the chiral system, one direction of spin is selected along its moving direction, and the selected spin direction is altered by switching the direction of electron flow and structural chirality. B Experimental setup for the ferromagnetic resonance on the polymer/NiFe bilayer. C Schematics of spin-current dissipation across the interface between FM and adjacent material. When the material has a large spin-orbit coupling (SOC), the spin-current from the FM dissipates strongly, leading to the enhancement of damping. D Ferromagnetic resonance (FMR) linewidth versus microwave frequency. Insets show the representative FMR spectra measured at 10 GHz for each film. E Comparison of SOC trends quantified by damping factors with relative doping levels derived from XPS and absorption. Error bars for the S*/S peak area represent the standard error derived from XPS peak fitting. Error bars for absorption represent the standard deviation based on measurements from three individual films.

To summarize the doping mechanism discussion above, we propose that the chiral structure increases doping efficiency through inherently higher crystallinity with a possible contribution from the CISS effect, which increases the apparent oxidation rate of the chiral film. An enhanced oxidation rate builds up a steeper dopant concentration gradient in the film, thereby driving faster dopant diffusion into the strongly chiral film to boost charge carrier concentration and doping efficiency, resulting in improved conductivity upon doping. We rule out the possibility that faster dopant diffusion may instead arise from a higher diffusion coefficient, as the relative degrees of crystallinity of both the side chain and backbone regions are the highest in the strongly chiral film, leading to the densest packing in the strongly chiral film (Figs. 3F and 3H). Such morphology characteristics can only reduce the dopant diffusion coefficient in the chiral film. Despite a lower diffusion coefficient, the more than two-fold higher oxidation rate “forces” the dopant into the strongly chiral film as to cause the most substantial morphological changes upon doping, evidenced by diminished in-plane alkyl stacking and marked changes in rDoC values after doping (Figs. 5F and 5G). Besides CISS and the crystallinity effect, variations in dopant distribution near the polymer backbone may contribute to the differences in apparent oxidation rates; however, the similar extent of changes observed in the stacking distances of both backbone and side chains upon doping across the three films suggest similar dopant distributions within the crystalline regions (Tables S3 and S4). It must be stressed that definitively proving the CISS effect requires studying the magnetic field dependence of the reaction rate asymmetries. Unfortunately, bulk doping is not compatible with such measurements given the non-directional character of the electron transfer processes. Indeed, the CISS effect is usually studied in model systems involving an oriented monolayer of chiral molecules with directional electron transfer, distinct from a practical chemical doping process shown in this work.

Generality of the role of chirality in chemical doping

To assess whether supramolecular chirality influences chemical doping in other polymer systems, we tested two additional sets of systems: DPP2T-TT and TPT-2T/TPT-T (Fig. 8). As shown in Fig. 8A-D, using chlorobenzene (CB) or anisole as solvents allowed control over supramolecular chirality in blade coated DPP2T-TT films, yielding chiral films from CB and achiral films from anisole. Films were subsequently doped with a 5:1 molar ratio of F4TCNQ (2, 3, 5, 6-tetrafluoro-7, 7, 8, 8-tetracyanoquinodimethane) to BCF (tris(pentafluorophenyl)borane) mixture which forms Lewis-paired dopant complexes72. Doped chiral films showed a much higher conductivity (0.55 ± 0.05 S/cm) than achiral films (0.02 ± 0.01 S/cm). In the second system (Fig. 8E-H), TPT-2T exhibited chiral structure upon annealing due to its thermotropic liquid crystalline phases, whereas TPT-T is absent from chiral assembly upon annealing73. After doped with iron (III) tosylate, chiral TPT-2T films reached a conductivity of 24.7 ± 4.0 S/cm, outperforming the achiral TPT-T films (3.3 ± 1.8 S/cm).

Fig. 8: Supramolecular chirality enhances chemical doping in two additional sets of polymer systems (DPP2T-TT and TPT Derivatives).
figure 8

A Molecular structures of DPP2T-TT and dopants (BCF and F4TCNQ). B CPOM images of DPP2T-TT films coated from chlorobenzene (CB) and anisole, with arrows indicating the coating direction. C CD spectra showing chirality in CB-coated films and lack thereof in anisole-coated films. D Conductivity measurements comparing chiral and achiral DPP2T-TT films. Bars represent average conductivity, with error bars indicating the standard deviation from measurements of three individual films. E Molecular structures of TPT-T and TPT-2T. F CPOM images of annealed TPT films. G CD spectra indicating chirality in TPT-2T films and achirality in TPT-T films. H Conductivity measurements contrasting chiral TPT-2T and achiral TPT-T films doped with iron (III) tosylate post film deposition. Bars represent average conductivity, with error bars indicating the standard deviation from measurements of three individual films.

Further analysis of the TPT-2T/TPT-T system suggests that chirality, rather than high crystallinity, is the key factor enhancing doping efficiency. As shown in Figure S33, TPT-T achiral films exhibit higher crystallinity, evidenced by sharper GIXD peaks and lower paracrystalline disorder parameters (g), with values of 12.27% and 7.00% for lamellar and π-π stacking peaks, respectively, compared to the higher disorder in TPT-2T chiral films (24.46% and 17.65%, respectively). This higher crystallinity in TPT-T contradicts the expected trend if crystallinity were the dominant factor, as it would predict higher conductivity for TPT-T rather than TPT-2T. Furthermore, we found TPT-T exhibited a higher mobility than TPT-2T, which further reinforce this discrepancy. Despite its lower mobility and crystallinity, TPT-2T exhibits significantly higher conductivity, which can be attributed to the role of supramolecular chirality in enhancing doping efficiency. These findings highlight the potential of chirality in overcoming low charge-carrier mobility by facilitating effective oxidation and boosting conductivity.

Discussion

To conclude, independent of the actual extent of contribution of the CISS effect, our study points to the unprecedented observation that introducing supramolecular chirality via helical assemblies can serve as a surprisingly effective approach to significantly enhance doping efficiency and conductivity. This discovery points to previously unrecognized principles governing redox reactions. Given the pivotal role of chemical doping in a broad range of polymer materials, we anticipate that our finding will have broad impact across many technological areas spanning renewable energy and bioelectronics.

Methods

Materials

The dioxythiophene copolymer, comprised of 2, 2’-bis-(3,4-ethylenedioxy)thiophene (biEDOT) and 3,4-propylenedioxythiophene (ProDOT) substituted with branched oligo(ether) side chains (PE2-biOE2OE3, Mn = 20.0 kg/mol, PDI = 1.7), was synthesized using direct hereto(arylation) polymerization (DHAP) as previously described (the synthetic protocol is given in the SI)39. All starting materials used for DHAP were purchased from a commercial supplier and were used without further purification. The number average molecular weight was determined by gel permeation chromatography with a Tosoh EcoSEC (calibrated versus polystyrene) equipped and a refractive index detector and using stabilized chloroform at 40 °C as the eluent. The polymer solutions were prepared by dissolving the polymer (5 mg mL−1) in 1,4-dioxane (anhydrous, 99.8%, Sigma-Aldrich), diglyme (anhydrous, 99.5%, Sigma-Aldrich), and 1,3-dioxane (97.0%, Sigma-Aldrich). Corning glass was used as substrates. The substrates were cleaned with toluene, acetone and isopropyl alcohol, and then dried under a stream of nitrogen to remove contaminants. The substrates were then plasma treated for 20 minutes before printing to improve the wetting of the solution. For SECM and differential pulse voltammetry measurements the substrates were ITO-coated glass substrates. Dopant iron(III) p-toluenesulfonate hexahydrate (technical grade, Sigma-Aldrich) was dissolved in isopropyl alcohol (IPA) (40 mM). The diketopyrrolopyrrole-based polymer DPP2T-TT (Mn = 60.7 kg/mol, PDI = 3.2) was synthesized following a previously published method74. The polymer was dissolved in anhydrous chlorobenzene (99.8%, Sigma-Aldrich) and anisole (99.7%, Sigma-Aldrich) at a concentration of 5 mg/mL, and films were prepared using a meniscus-guided coating technique in the evaporation regime on glass substrates. The coating was performed at room temperature with a blade angle of 7° relative to the substrate, using coating speeds of 0.002 mm/s and 0.004 mm/s for CB and anisole solutions, yielding film thicknesses of 648 ± 42 nm and 705 ± 73 nm, respectively. The dopant, consisting of a 5:1 molar ratio of F4TCNQ (2, 3, 5, 6-tetrafluoro-7, 7, 8, 8-tetracyanoquinodimethane) (97%, Sigma-Aldrich) to BCF (tris(pentafluorophenyl)borane) (95%, Sigma-Aldrich) mixture, was dissolved in acetonitrile ( > 99.9%, Sigma-Aldrich) at a concentration of 10 wt%72. Two additional polymers, TPT-2T (Mn = 15.7 kg/mol, PDI = 1.6) and TPT-T (Mn = 17.2 kg/mol, PDI = 1.6), were synthesized as previously reported75 and dissolved in anhydrous chlorobenzene at 5 mg/mL. Films of these polymers were spin-coated at 500 rpm onto glass substrates, heated to 180 °C (above the melting temperature), and cooled to room temperature at a rate of 10 °C/min. The film thicknesses were 103 ± 8 and 110 ± 13 nm for TPT-2T and TPT-T before annealing, respectively. This annealing process induces the thermotropic liquid crystal phase of TPT-2T and promotes chiral emergence. The dopant solution was prepared with iron(III) p-toluenesulfonate hexahydrate (technical grade, Sigma-Aldrich) in isopropyl alcohol at a concentration of 40 mM.

Solution-state structures characterization

UV–visible (Agilent Cary 60 UV-Vis) and fluorescence (SLM-ISS fluorescence Spectrometer) spectroscopy was used to measure the absorption and emission of PE2-biOE2OE3 solutions. The solution SAXS measurement was performed at beamline 12ID-E within the Advanced Photon Source at Argonne National Laboratory. Scattering profiles were acquired using an incident X-ray beam with an energy of 13.3 keV and a sample-to-detector distance of 3622.78 mm. The polymer scattering data were correction for background scattering, which was determined using solvent-only measurements. Subsequent data analysis was carried out using SAXSLee, a tool developed for MATLAB, and fitting was conducted employing models developed in our previous work76 (Supplementary Information section II).

Freeze-drying experiments were conducted to image the structure of aggregates in solution using SEM (JEOL 7000 F). This process began with the rapid freezing of polymer solutions in a liquid ethane and propane mixture (63% propane and 37% ethane), immediately followed by storage in liquid nitrogen. The chosen mixture facilitated swift freezing owing to its high heat capacity, effectively minimizing polymer aggregation during the cooling phase. Subsequently, the frozen polymer solution was swiftly moved to a Linkam stage set at −100 °C under vacuum. There, the temperature was incrementally raised at a rate of 0.5 °C/min up to -80 °C, remaining decidedly below the solvents’ melting points (with melting points at 11.8 °C for 1,4-dioxane, -64 °C for diglyme, and -42 °C for 1,3-dioxane). The process concluded with 8–10 h of continuous observation using cross-polarized microscopy, ensuring complete solvent sublimation.

Solution characterization under CPOM (Nikon Eclipse Ci-POL) was conducted by encapsulating the polymer solutions between two microscope glass coverslips, which were then placed on a Si wafer for the CPOM measurements in reflection mode. High-concentration solutions were prepared using the drop and dry method, a process in which the concentration of the pristine solution was enhanced by evaporating a specified volume and subsequently redissolving the residue with 2 μL of the solution. This concentrated solution was then homogenized by shearing with the microscope cover glasses to further increase its concentration. The solution samples then underwent thermal annealing cycles to remove the shearing history and reach an equilibrium state before imaging under CPOM.

Film printing and doping

The PE2-biOE2OE3 thin films were deposited onto substrates using a meniscus-guided blade coating technique, as detailed in our earlier work53 In brief, an OTS-treated Si blade set at a 7° printing angle and a 100 µm gap was set from the blade to the substrate. The blade moved linearly across the stationary substrate, maintaining the ink solution within the specified gap. The PE2-biOE2OE3 films were coated at a speed of 5.0 µm s−1 at 25 °C for 1,4-dioxane and 1,3-dioxane solutions, and at 1.0 µm s−1 at 55 °C for diglyme solutions. The elevated temperature for diglyme aimed to align its vapor pressure with that of the other solvents, while the reduced speed ensured a consistent film thickness of 400 nm across different solvent conditions. Printing began with a low solution concentration (5 mg/mL); as the solvent evaporated during coating, the local concentration increased, promoting the in-situ formation of chiral mesophases typically observed at higher equilibrium concentrations. To preserve these structures in the final film, coating was performed at low speeds (1–5 µm/s) within the evaporation regime, allowing near-equilibrium formation of chiral liquid crystalline phases.

When doping the film using Fe(Tos)3, 40 μL of Fe(Tos)3 solution was distributed across the film using a pipette. The film was then shielded with a petri dish for 2 minutes. Then around 100 μL of clean IPA was applied over the film to rinse away residual dopant solution, which was then carefully pipetted off. This washing step was repeated 5 times and the whole doping process was repeated 3 times for one film. Subsequently, the films are placed in a vacuum oven, set at room temperature under vacuum, for a duration of 10 minutes before conductivity measurement. The doping procedure is the same when using NOBF4/IPA solution. TPT-T and TPT-2T were doped using same dopant solutions and procedures. DPP2T-TT was doped using the same procedure but with F4TCNQ/BCF dopant solutions in acetonitrile. The dopant solution is composed of a 5:1 molar ratio of F4TCNQ to BCF, dissolved in acetonitrile at a concentration of 10 wt%. This solution forms Lewis-paired dopant complexes that effectively dope the polymers, as previously reported72.

Printed film morphology and electrical property characterization

The birefringence of the PE2-biOE2OE3 thin films was characterized using cross-polarized optical microscopy (Nikon Eclipse Ci-POL). The mesoscale morphology was characterized using tapping-mode atomic force microscopy (Asylum Research Cypher). UV–visible (Cary 60 UV-Vis and Varian Carry 5 G, Agilent) spectroscopy was used to measure the absorption of films before and after doping. GIWAXS measurements were performed at the 7.3.3 beamline of the Advanced Light Source at Lawrence Berkeley National Laboratory, using incident angles of 0.14° with an X-ray energy of 10 keV and a beam size of 30 μm × 50 μm. All samples were scanned 10 s in a helium chamber for measurement. XRR measurement was done using Bruker D8 Advance XRR system with an X-ray energy of 8.04 keV.

CD spectra were recorded using a JASCO 1500 spectrophotometer. To accurately capture the true CD signal, we used a four-scan method, as previously reported41, to eliminate artificial CD contributions from linear dichroism and linear birefringence (CDLDLB), as well as residual static birefringence from the photoelastic modulator (PEM) (CDα). Further details are provided in the Supplementary Information and Figure S28.

The CD g-factor quantifies the differential absorption of left- and right-circularly polarized light relative to the total absorption of unpolarized light77. It normalizes circular dichroism (ellipticity) to eliminate dependencies on sample concentration, path length, or film thickness. Mathematically, it is given by:

$${g-{factor}}_{{abs}}=\frac{\triangle A}{A}=\frac{\theta }{32,980\,\times \,A}=\frac{2\left({A}_{L}\,-{A}_{R}\right)}{{A}_{L}\,+{A}_{R}}$$
(1)

where \(\triangle A={A}_{L}-{A}_{R}\) is the difference in absorption between left- and right-circularly polarized light, \(A=\frac{{A}_{L}\,+{{A}}_{R}}{2}\) is the absorption of unpolarized light, \(\theta\) is the ellipticity in millidegrees,

and the constant 32,980 is used to convert ellipticity to absorbance units.

To quantify crystallinity, we extracted the relative degree of crystallinity (rDoC) from GIXD patterns. The angle χ is defined such that χ = 0° corresponds to a “face-on” π–π orientation and χ = 90° to an “edge-on” π–π orientation, as illustrated in Fig. 3C. The rDoC values for the lamellar, π–π, and helical π–π stacking peaks were calculated by integrating the corresponding peak intensities over χ according to the expression78:

$${{rDoC}}\propto {\int }_{0}^{\frac{\pi }{2}}sin (\chi ){{{\rm{I}}}}({{{\rm{\chi }}}}){{{\rm{d}}}}\chi$$
(2)

The film sheet resistivity (\({R}_{s}\)) was measured using van der Pauw method in Accent HL5500PC Measurement system. The \({R}_{s}\) was then converted to conductivity \(\sigma\) using equation:

$$\sigma=\frac{1}{{R}_{s}\times t}$$
(3)

where \(t\) is the film thickness. To estimate the charge carrier concentration, XPS spectra were measured using a Kratos Axis Supra photoelectron spectrometer and a phi-Versa probe hemispherical analyzer system. For photoelectron excitation, a monochromatized X-ray source (Al Kα, 1486.6 eV) was used. The base pressure of the analysis chamber was maintained below 10–8 Torr) The energy reference was calibrated using clean Au (111).

Ionization energies and charge density were determined under argon atmosphere by differential pulse voltammetry and chronocoulometry in degassed 0.5 M tetraethylammonium tosylate (TEAtos) in propylene carbonate. The potential was controlled with an EG&G PAR 273 A potentiostat/galvanostat under CorrWare control. The three-electrode measurements incorporated a Ag/Ag+ pseudoreference electrode (calibrated vs ferrocene/ferrocenium, E1/2 = 90 mV) and a platinum flag counter electrode.

FMR and SECM measurements

For ferromagnetic resonance (FMR) measurement, the NiFe (6 nm) thin films were deposited on three assembled polymers by e-beam evaporation. The sample was attached to the coplanar wave guide with the Kapton tape. The FMR spectra were obtained using a commercial NanOsc Phase FMR spectrometer connected with the wave guide in the Physical Property Measurement System (PPMS, Quantum Design). The FMR spectra were fitted with the derivative of the asymmetric Lorentzian equation19,

$$S= {K}_{1}\cdot \frac{\left(\frac{\triangle H}{2}\right)\left(H-{H}_{{res}}\right)}{{\left({\left(H-{H}_{{res}}\right)}^{2}+{\left(\frac{\triangle H}{2}\right)}^{2}\right)}^{2}}+{K}_{2}\cdot \frac{{\left(\frac{\triangle H}{2}\right)}^{2}-{\left(H-{H}_{{res}}\right)}^{2}}{{\left({\left(H-{H}_{{res}}\right)}^{2}+{\left(\frac{\triangle H}{2}\right)}^{2}\right)}^{2}} \\ +{Offset}+H\cdot {Slope}$$
(4)

where \(S,\,{K}_{1},\,{K}_{2},{H},\,{H}_{{res}},\) and \(\triangle H\) are absorption derivative, fitting constant for the first term, constant for the second term, applied magnetic field, resonant field, and FMR linewidth (FWHM).

In SECM, an ultramicroelectrode is positioned over a substrate, and the current response at the tip and substrate depends on the potential applied to the tip and substrate. Following Tafel fitting gives the standard rate constant (ko) and charge transfer coefficient (\({\alpha }_{{SECM}}\)) of the redox reaction. A similar work can be found elsewhere79.

MD simulations

All-atom (AA) molecular dynamics (MD) simulations were performed using the LAMMPS package80, and the Optimized Potentials for Liquid Simulations−All Atom (OPLS-AA) force field81. Atom types and parameters were defined using the LigParGen web server82. Long-range electrostatic interactions of charged particles were computed using the particle-particle-particle-mesh (PPPM) algorithm with a root-mean-accuracy of 10-583. The Verlet integrator with a time step of 1 fs was used, and the Nose−Hoover thermostat/barostat was employed for temperature/pressure control. All related simulations were performed using NPT (constant number of molecules, pressure, and temperature) at pressure of 1 atm. A cutoff of 12 Å was applied for the summation of van der Waals interactions.

MD simulations were used to investigate the conformational dynamics of PE2-biOE2OE3 chains, with an emphasis on the behavior of single chains and their collective organization in various solvents. These simulations were conducted at 300 K within the NPT ensemble, featuring 12-mer polymer chains within a solvent-filled simulation box measuring 18.0 × 10.5 × 10.5 nm3. Each system underwent an equilibration phase of 20 ns, succeeded by a production phase lasting 30 ns for the purposes of data collection and averaging. For structures involving multiple chains, an additional metadynamics simulation was implemented to expedite the packing process through the application of a biased potential.