2.1. Incorporating A Nucleating Agent to Achieve Enhanced Crystallinity in Pristine PBTTT-C14 Films
As highlighted earlier, the key objective of this study is to systematically regulate the crystallinity of a model semicrystalline polymer, specifically PBTTT-C14, to elucidate the complex relationship between microstructure, morphology, charge carrier transport, and thermoelectric performance, particularly under chemical doping conditions. Prior to the sequential IEx doping process, we firstly fine-tune the crystallinity of pristine polymer films, which are spin-coated from blending solutions with varying concentrations of the nucleating agent (Fig. 1a; and see Experimental Section for details). We selected PDA, a supramolecular nucleating agent known for its remarkable ability to enhance the crystallinity of chemically similar polymer, such as Poly(3-hexylthiophene) (P3HT), with just 0.1 wt% loading20. Impressively, PDA significantly improves the crystallinity and crystalline domain size of PBTTT-C14, as corroborated by DSC, GIWAXS, and AFM. The results of these characterizations are discussed in detail below.
We conducted a series of DSC experiments on PBTTT-C14/PDA blended samples, varying additive concentrations ranging from 0 wt% to 3.4 wt%, to probe the impact of nucleating agents on the crystallization kinetics of the polymer. All samples exhibit two primary thermal transitions (Fig. 1b): a low-temperature transition (Transition I) occurring below 100 ℃, and a high-temperature transition (Transition II) centered around 220 ℃, which is characteristic of PBTTT's thermal behavior25. The effect of nucleating agents on Transition I is less pronounced, as indicated by the cooling curves in Fig. 1b. Notably, the calorimetric peak corresponding to Transition II exhibits distinct variations with increasing PDA concentrations. Therefore, our further investigations focus on film samples annealed above the Tc of Transition II, specifically the ribbon phase regime.
The addition of PDA progressively raises the crystallization temperature (Tc) of PBTTT-C14, reaching 226.33 ℃ at a 0.9 wt% concentration, an increase of 3.04 ℃. Beyond this concentration, higher PDA loadings resulted in a decline in Tc. This initial shift in Tc to higher temperatures at low additive loadings, up to 0.9 wt%, suggests heterogeneous nucleation, which enhances the polymer's crystallization capacity20. The substantial improvement in crystallinity for these blended samples is directly evidenced by the increase in the enthalpy of crystallization (ΔH), a parameter proportional to the degree of crystallinity26. Specifically, the sample with 0.9 wt% PDA exhibited a ΔH peak of 16.69 J/g at transition II (Fig. 1c), representing a 45% increase compared to pristine PBTTT films. However, at higher PDA concentrations, the degree of crystallinity declined, as evident by notable reductions in both Tc and ΔH. This reduction in crystallinity can be attributed to the overloading of the nucleating agent. Overuse of nucleating agents might lead to an excessively high nucleation rate, resulting in the formation of numerous small and imperfect crystalline grains. This high nucleation rate impedes further crystal growth, ultimately leading to a decrease in overall crystallinity27.
GIWAXS analysis was employed to investigate the evolution of crystalline order in pristine PBTTT-C14 and PBTTT-C14/PDA blended films across various nucleating agent concentrations. Both pristine and blended films exhibit high crystallinity, characterized by four orders of (h00) diffraction peaks along the qz axis, representing the lamellar stacking direction, and two additional peaks at 1.45 Å−1 and 1.71 Å−1 on the qxy axis, corresponding to the polymer backbone and π-stacking orientation, respectively (Fig. 1d-f; and Figure S1 in the Supporting Information). Overall, the 2D GIWAXS scattering patterns reveal a remarkable similarity between the samples, with enhanced diffraction intensities along the in-plane qxy directions for the film containing 0.9 wt% of the nucleating agent. To further explore the molecular stacking behavior upon the incoporation of the nucleating agent, we analyzed the 1D line-cut scattering profiles (Figure S1) and extracted crystallographic parameters, summarized in Table S1. For the film with 0.9 wt% PDA, no notable changes in paracrystallinity (gπ−π) and higher coherence length (Lc) were observed. These observations suggest that while there is a significant enhancement in crystallinity, the molecular stacking behavior of PBTTT-C14 is largely preserved up to an optimal concentration of PDA.
The morphological changes in the polymer films due to the addition of PDA were examined using AFM. The height image in Fig. 1g shows that the pristine PBTTT-C14 thin film, annealed at 260 ℃, exhibits uniform ribbons-like structures with an average width of 53.18 nm, a characteristic feature of films annealed from a smectic, liquid crystalline phase (Transition II)28. These ribbons consist of crystalline domains that maintain a smectic chain arrangement, typically comprising multiple chain-extended or chain-folded polymer chains, depending on the degree of chain alignment in the solid state29. Upon the initial incorporation of PDA, the ribbon structure was preserved, with the ribbon width increasing by 62%, from 53 nm in the pristine film to 86 nm at 0.9 wt% nucleating agent, indicating a tendency for aggregation. However, at a higher PDA concentration of 3.4 wt%, the ribbon width decreased to 42 nm. This transformation is accompanied by the gradual blurring of the nanofibrillar structure, suggesting a loss of long-range fiber periodicity (Fig. 1i; and Figure S3 in the Supporting Information). These observations imply that the addition of PDA at low concentrations (up to 0.9 wt%) enhances the degree of crystallinity and promotes the growth of individual ribbons, which is crucial for improving both structural order and charge transport. The observation of enlarged crystalline domains (ribbons) coupled with narrower grain boundaries at an optimal PDA loading suggests that the PBTTT-C14/PDA blended system may serve as a viable platform for subsequent doping and thermoelectric property investigations. Such a highly ordered microstructural morphology is anticipated to facilitate both efficient intradomain and interdomain charge transport.
2.2. Preserving Superior Structural Order in PBTTT-C14/PDA Blended Films upon Ion Exchange Doping
Chemical doping plays a pivotal role in modulating the electronic and thermoelectric properties of polymer semiconductors, effectively adjusting the energy level and optimizing the charge concentration n.11 However, achieving optimal thermoelectric conversion in a polymer system often necessitates a charge concentration n surpassing 1019 cm− 3, which may significantly disturb the molecular packing within the neat polymer matrix, potentially compromising the high carrier mobility µ derived from its high crystalline order30. To harness the enhanced crystallinity of PBTTT-C14 films induced by the nucleating agent for enhancing thermoelectric performance, we adopted a newly developed IEx doping method. This doping technique has exhibited remarkable efficacy in achieving doping concentrations exceeding 1020 cm− 3, while imparting minimal or even positive effects on the microstructure and morphology of the polymer films. This preservation of superior structural order is critical for maintaining efficient charge transport. The successful achievement of efficient doping and preservation of superior structural order in PBTTT-C14/PDA blended films was confirmed through spectroscopic, GIWAXS, and AFM characterizations (Fig. 2 and Figure S2, S4 in the Supporting Information).
The extent of doping in all polymer films, with varying PDA concentrations, both before and after IEx doping, was evaluated using Ultraviolet-Visible-Near Infrared (UV-Vis-NIR) spectroscopy. For all the blended films studied, the neutral π-π* band centered around 550 nm exhibited nearly complete bleaching (Fig. 2a; and Figure S5 in the Supporting Information), a trend consistent with the pristine PBTTT film (Figure S5). Simultaneously, two broad absorption bands, labelled P1 and P2, appeared at wavelengths > 2000 nm and around 800 nm, respectively. These absorption bands are indicative of polaron or multi-polaron states, as commonly observed in highly doped PBTTT films24,31. These UV-Vis-NIR absorption profiles suggest the blended films obtained at high doping levels comparable to those of pristine PBTTT-C14, with a doping concentration around 1021 cm− 3, corresponding to roughly one charge per monomer unit. This high carrier density, confirmed for both pristine and blended systems, is further validated by Hall effect analysis presented in Section 4 of the Supporting Information.
Fourier Transform Infrared Spectroscopy (FT-IR) was employed to investigate the degree of polaron delocalization in doped films (Fig. 2b). Polaron delocalization can be effectively assessed by examining polaron-induced absorption features in the mid-infrared (mid-IR) region32,33. Upon doping, the P1 band peaks around 1800 cm− 1 with a weak shoulder below 1500 cm− 1, corresponding to intrachain and interchain transitions, respectively32. Additionally, a number of narrow peaks are superimposed between 400 cm− 1 and 1475 cm− 1, which are typically interpreted as infrared active vibrational modes (IRAVs). The intensity of the IRAVs is an indicator of the extent of polaron delocalization along the polymer chains and between chains. The more intense the IRAVs, the greater the degree of polaron delocalization34. In the FT-IR spectra of the doped blended films, the pronounced IRAV peaks suggest that a high level of polaron delocalization was achieved via ion exchange doping, similar to that observed in pristine PBTTT-C14 films35,36.
GIWAXS characterization was employed to examine the microstructural changes in both pristine PBTTT-C14 and PBTTT-C14/PDA blended films after IEx doping. As depicted in the 2D diffraction patterns in Fig. 2d-f, all doped films retain their lamellar packing, featuring an edge-on orientation of the polymer backbone and considerable long-range crystallinity. Both the well-ordered out-of-plane peaks (h00) and in-plane stacking peaks remain clearly visible, resembling their undoped counterparts. Upon IEx doping, the lamellar spacing notably expands for both the pristine and blended films, for instance, from 20.937 Å to 26.449 Å for the 0.9 wt% blended sample. This expansion suggests that the counterions predominantly integrate into the alkyl side-chain regions, consistent with previous findings on semicrystalline polythiophenes36–38. Interestingly, a reduction in the π-π stacking distance is observed across all samples: in the pristine film, this distance decreases from 3.601 Å to 3.526 Å (Fig. 2d; and Table S2 in the Supporting Information), while in the 0.9 wt% and 3.4 wt% blended films, it reduces from 3.646 Å to 3.516 Å and from 3.654 Å to 3.524 Å, respectively (Fig. 2e-f; and Table S2 in the Supporting Information). This reduction can be attributed to the strengthening of polaronic coupling between chains, facilitated by the IEx doping process36. Furthermore, all three films display a ~ 10% decrease in paracrystallinity parameter (gπ−π) compared to their undoped states, suggesting that ion exchange doping effectively enhances structural order without causing discernible disruption, even in the presence of nucleating agent additives (Table S2 in the Supporting Information). The high electrical conductivities of both the pristine and blended systems, with σmax ranging from 878 to 1800 S cm− 1, are strongly correlated with reduced paracrystalline disorder, in line with previous findings in high-mobility semicrystalline polymers and donor-acceptor copolymers36.
It is important to note the differences in charge transport and thermoelectric properties between the pristine and blended systems. In particular, the doped 0.9 wt% blended system achieves a conductivity of approximately 1800 S cm− 1 and an exceptional thermoelectric power factor of about 150 µW m− 1 K− 2 (as disscussed in Section 2.3 and 2.4). These remarable enhancements cannot be solely attributed to molecular packing, as there are no significant differences in paracrystallinity and coherence length between doped pristine PBTTT film and 0.9 wt% blended film (Table S2 in Supporting Information). Instead, the enhanced thermoelectric performance in the blended systems can be primarily linked to a notable increase in crystallinity, as demonstrated through detailed microstructural and morphological analyses of the undoped counterparts.
In conclusion, the GIWAXS analysis reveals that the counterions primarily occupy the side-chain regions, while IEx doping enhances backbone planarization, promoting efficient charge transport in all doped films—especially in the 0.9 wt% PBTTT-C14/PDA blended system. Overall, the introduction of the nucleating agent into PBTTT-C14 does not significantly interfere with the IEx doping process, enabling highly effective doping while maintaining or even improving the superior crystalline order.
2.3. Thermoelectric Properties of Ion Exchange Doped PBTTT-C14/PDA Blended Films
We now turn our attention to the thermoelectric performance of the ion exchange (IEx) doped PBTTT-C14/PDA blended films. As shown in Figs. 3a and 3b, the electrical conductivity (σ) and Seebeck coefficient (S) of these doped polymers exhibit distinct trends as a function of PDA concentration, ranging from 0 wt% to 3.4 wt%.
Specifically, Fig. 3a shows a significant increase in electrical conductivity, reaching a peak value of 1178 S cm− 1 at 0.9 wt% PDA loading with a doping concentration of 3 mM. This represents an 85% enhancement compared to the conductivity of the doped pristine PBTTT-C14 film under the same doping conditions. This improvement is consistent with the crystallinity evolution with PDA concentration, as discussed in Section 2.1. Notably, the Seebeck coefficient follows a similar trend, decoupling from the electrical conductivity. The doped 0.9 wt% blended film exhibits a maximum Seebeck coefficient of approximately 36 µV K− 1, which is 148% higher than that of the doped pristine film. As a result, the power factor (PF) of the doped 0.9 wt% blended film attains an impressive 150 µW m− 1 K− 2, representing a remarkable 1011% improvement over the pristine film at an equivalent doping concentration, as shown in Fig. 3c.
The inset of Fig. 3c further emphasizes the competitive advantage of our IEx doped 0.9 wt% blended system when compared to other unaligned benchmark systems, achieved through morphology control39, doping engineering40,41. Remarkably, this optimal performance is on par with a recently developed highly oriented PBTTT system42, which, despite requiring complex processing techniques—such as controlled tie chain incorporation through polymer blending coupled with large-scale alignment—results in notable charge transport anisotropy. The simultaneous increase in both σ and S, leading to a substantial enhancement in the PF with the appropriate PDA additive loading, is unprecedented. The detailed mechanisms underlying this phenomenon are elaborated in Section 2.4.
To further fine-tune the thermoelectric performance of the 0.9 wt% blended system, we optimized the S and σ as a function of doping concentration. For comparison, the same experiment was performed on a pristine PBTTT-C14 system to better understand the impact of PDA additives on thermoelectric properties. As shown in Fig. 3d, the electrical conductivity initially increases with rising doping concentration, peaking at approximately 7 mM for the doped 0.9 wt% blended system. At this concentration, the blended film achieves an electrical conductivity of 1800 S cm− 1, more than double the maximum conductivity of 878 S cm− 1 observed in the doped pristine system at a doping concentration of 12.5 mM. In terms of the Seebeck coefficient, both the blended and pristine systems exhibit a comparable downward trend with increasing doping concentration. Consequently, the calculated power factors are presented in Fig. 3f. Notably, the ion exchange doped PBTTT-C14/PDA blended films reach a significantly higher maximum power factor of 150 µW m− 1 K− 2, far surpassing the 31 µW m− 1 K− 2 achieved by the doped pristine counterpart.
These findings underscore the promising potential of nucleating agent additive engineering in significantly enhancing the thermoelectric performance of high-performance semicrystalline conjugated polymers. A comparative analysis of the thermoelectric properties between blended and pristine systems reveals that the remarkable PF achieved by the IEx doped PBTTT-C14/PDA blend (0.9 wt%) is primarily driven by its exceptional charge transport properties. To further investigate, we explored the charge transport physics to uncover the fundamental mechanisms by which the nucleating agent facilitates improved charge transport in the doped PBTTT systems.
2.4. Mechanisms Underlying the Remarkable Enhancement in Charge Transport and Thermoelectric Performance of IEx Doped PBTTT-C14/PDA Blended Films
Charge transport measurements are essential for characterizing transport parameters in semiconducting materials, allowing for the correlation between microscopic spatial and energetic distributions through microstructural characterization and charge transport models43. For doped polymer semiconductors, temperature-dependent measurements of electrical conductivity, the Seebeck coefficient, and the Hall coefficient provide insight into the dominant transport mechanisms—whether they involve localization (hopping-like) or delocalization (metal-like)—that govern electronic transport44–46. In this study, we applied this experimental framework to investigate how the incorporation of a nucleating agent additive affects charge transport physics and the microstructural characteristics of PBTTT-C14 films subjected to sequential ion exchange doping.
Figure 4a depicts the temperature (T) dependence of electrical conductivity for both pristine and blended films doped at a concentration of 5 mM, highlighting the significant effect of PDA concentration. Across all systems the temperature dependence of conductivity reflects that of high conductivity PBTTT, characterized by a plateau or slight decrease above 250 K, particularly prominent in the 0.9 wt% blended system. The appearance of a conductivity peak suggests a shift toward metallic conduction rather than thermally activated hopping conduction, especially at grain boundaries42. To quantify the thermal activation energy (Ea) governing each sample's conductivity, we applied the Arrhenius equation (see Supporting Information Section 5). The Ea, which reflects energetic disorder within integer charge transfer complex (ICTC) states47, decreases from 3.25 meV in pristine PBTTT to 2.67 meV at 0.9 wt% PDA loading, where conductivity peaks (Fig. 4b, upper panel). This minimal Ea in the 0.9 wt% blended sample signifies the lowest energetic disorder, which can be attributed to optimal crystallinity and structural ordering (Sections 2.1 & 2.3). As a result, the 0.9 wt% blended film achieves a peak charge carrier mobility of 2.92 cm2 V− 1 s− 1, confirmed by Hall effect measurements (Fig. 4e).
Additionally, by fitting the Seebeck coefficient-conductivity (S-σ) data using the Kang-Snyder model48 (see the Supporting Information Section 6 for more details), we derived the energy barrier for charge percolation between ordered domains, denoted as Wγ. Unlike Ea, which provides insights into the overall transport mechanism encompassing both intra- and inter-domain transport, Wγ specifically offers insights into inter-domain transport, providing a macroscopic perspective on conductivity across crystalline regions48,49. Wγ follows a trend similar to Ea (Fig. 4b, bottom panel), reaching its minimum at intermediate PDA concentrations before slightly increasing at higher concentrations, likely due to decreased overall crystallinity from excessive PDA loadings, as observed in DSC, GIWAXS, and AFM data discussed earlier. Notably, the decreasing ratio of Wγ to Ea with increasing PDA concentration (Fig. 4c) suggests that inter-domain energetic disorder plays a diminished role in limiting conductivity and carrier mobility.
To further investigate the synergistic impact of nucleating agent additive engineering and IEx doping on the charge transport properties of PBTTT-C14, we conducted comparative analyses of the S-σ relationship (Fig. 4d) across varying doping concentrations for both pristine and 0.9 wt% blended systems. By applying the Kang-Snyder model to the S-σ relationship, we observe that a single parameter set (s and σE0) inadequately describes the entire conductivity range (See the Supporting Information Section 6 for more details). Notably, both systems exhibit a distinct change in charge transport characteristics around 200 S cm− 1, transitioning from s = 3 to s = 1 with increasing conductivity. This transition is generally indicative of an insulator-to-metal transition occurring in semicrystalline polymer systems, particularly when carrier densities exceed 1021 cm− 3 40,50.
In both the 's = 3' and 's = 1' regimes, the 0.9 wt% blended system demonstrates notably higher σE0 values, specifically 0.03 S cm− 1 and 120 S cm− 1, respectively, which are approximately more than 2 times greater than those of the pristine system. Since σE0 acts as a weighted mobility factor tied to intrinsic carrier mobility49,51, this observation highlight the enhanced charge transport capability of the doped 0.9 wt% blended system. This superiority is further evidenced by the consistently higher Hall mobility measurements across the entire conductivity spectrum (Fig. 4e). Consequently, the 0.9 wt% blended system exhibits significantly improved electrical conductivity for a given Seebeck coefficient, resulting in a substantially higher PF.
A comprehensive correlation analysis between transport measurements and microstructural characterization data allowed us to formulate a physical model (Fig. 4f) that elucidates the mechanism behind the enhanced charge transport properties induced by the nucleating agent additive. Upon incorporating an optimal concentration (0.9 wt%) of PDA, PBTTT-C14 films experience a significant increase in crystallinity, as demonstrated by DSC analysis. This increase is accompanied by improved structural order at the grain boundaries without disrupting the crystalline lattices within the domains. The polymer chain extension, supported by AFM and GIWAXS characterizations, contributes to this enhanced structural order. Such improvements are preserved, and in some cases intensified, across a broad range of carrier concentrations through ion exchange doping (Section 2.2).
Recent experimental and theoretical investigations have underscored structural disorder, rather than Coulombic trapping, as the primary determinant limiting charge transport in doped polymers, especially at high doping levels24,52. This explains the superior carrier mobility of the doped 0.9 wt% blended system, which achieves substantially higher electrical conductivities compared to the unblended system when sufficient charge densities are introduced. Specifically, the chain extension reduces grain boundary sizes and promotes the formation of tie chains that bridge adjacent crystalline domains by clustering together. This enables charge carriers to traverse domains without relying on interchain hopping through disordered regions. This dual effect—grain boundary size reduction and tie chain formation—significantly alleviates energetic disorder at grain boundaries, as evidenced by the substantial decreases in the activation energies Ea and the transport energy barriers Wγ. The macroscopic resistance of the 0.9 wt% system is therefore reduced, with grain boundary resistance often being a major limiting factor for conductivity in many polymer semiconductors42,53. Additionally, the enhanced crystalline order within domains after IEx doping, indicated by reduced gπ−π values, fosters intradomain charge transport, promoting two-dimensional charge carrier delocalization. This further reduces the overall energetic disorder and lowers Ea.
This synergistic improvement in both intra- and inter-domain charge transport, stemming from nucleating agent additive engineering and ion-exchange doping, results in a record-high electrical conductivity of approximately 1800 S cm− 1 and an outstanding thermoelectric performance, with a PF of 150 µW m− 1 K− 2 for PBTTT-C14 films.