Niklas J. Herrmanna,
Oleksandra Korychenskab,
Ngoc Phi Taa,
Guy E. Mayneordc,
Xabier Rodríguez-Martínez
ad,
Daniel T. W. Toolan
e,
Craig C. Robertsonb,
Ahmed Iraqib,
Jenny Clark
*b and
Jana Zaumseil
*a
aInstitute for Physical Chemistry, Heidelberg University, 69120 Heidelberg, Germany. E-mail: zaumseil@pci.uni-heidelberg.de
bSchool of Mathematical and Physical Sciences, Hounsfield Road, Sheffield S3 7RH, UK. E-mail: jenny.clark@sheffield.ac.uk
cSchool of Biosciences, University of Sheffield, Firth Court, Western Bank, Sheffield, S10 2TN, UK
dUniversidade da Coruña, Centro de Investigación en Tecnoloxías Navais e Industriais (CITENI), Campus de Esteiro, 15403 Ferrol, Spain
eDepartment of Materials, The University of Manchester, Engineering Building A, Booth Street East, Manchester M13 9PL, UK
First published on 25th July 2025
Many small-molecule organic semiconductors can crystallize in different polymorphs, which influences their electronic (charge transport) and optical (absorption and fluorescence) properties. Understanding polymorph formation and transformation is crucial to control these properties for potential applications. Here we explore different thin films of a newly synthesized derivative of 5,10-dihydroindolo[3,2-b]indole, which shows high photoluminescence quantum yields even in the solid state as well as hole transport. Deposition of this indoloindole-based p-type semiconductor by thermal evaporation and zone-casting from solution results in distinct film morphologies and a metastable polymorph that can transform into more stable polymorphs at room temperature over timescales from hours to months. Their conversion and changing structural, optical and electronic properties are characterized by grazing incidence wide-angle X-ray scattering (GIWAXS), atomic force microscopy (AFM), absorption, fluorescence, and Raman spectroscopy as well as charge transport measurements in field-effect transistors (FET). Rough and polycrystalline films result in very fast polymorph transformation, while defect-free and smooth zone-cast films are stable for several months. Annealing these films does not lead to faster conversion but instead to thermodynamic stabilisation of the metastable polymorph and thus even a reversal of aged thin films. For this and potentially other organic polymorphic systems, long-term retention of a metastable polymorph and its electronic and optical properties can be achieved by controlling the initial film morphology, while annealing can indeed induce the formation of metastable polymorphs.
The formation of a specific polymorph can be controlled by the type of film deposition or crystal growth methods (e.g., physical vapor transport, thermal evaporation, or solution processing) and their specific conditions.19,20 However, the initially formed polymorphs are often not the most thermodynamically stable polymorph,21 allowing for polymorph transformations and potential device degradation to take place over time. Long-term stability or transformation into other metastable or stable polymorphs depend on intrinsic as well as extrinsic factors. Intrinsic factors are, for example, the intermolecular interaction strength and film morphology, while temperature, humidity, and the presence of additional compounds (e.g., solvent vapours) are important external parameters.3,10,20,22–25 These parameters mainly affect the kinetic barriers to molecular reorientation and/or the relative thermodynamic stability (i.e., Gibbs free energy) of the different polymorphs.14 The latter determines the thermodynamic driving force behind polymorph transformations.
Here we investigate the polymorph transformations of a newly synthesized derivative of 5,10-dihydroindolo[3,2-b]indole (ININ, also known as dibenzopyrolo[3,2-b]pyrrole) with methylated nitrogen atoms and an octyl chain at each of the benzene rings (3,8-dioctyl-5,10-dimethylindolo[3,2-b]indole; C8-C-inin, see Fig. 1a). C8-C-inin can be viewed as nitrogen-substituted near-analogue of the benchmark p-type organic semiconductor C8-BTBT.26 ININ derivatives possess an electron-rich and rigid planar core structure,27 which has led to their use as p-type semiconductors in organic field-effect transistors (OFETs)28,29 and as donor molecules in solar cells (both as small molecules and donor–acceptor copolymers).30–34 In addition, many ININ derivatives have high photoluminescence quantum yields (PLQYs), even in the solid state, and can be functionalized to exhibit thermally activated delayed fluorescence (TADF).35,36 Consequently, ININ-based materials have been used for high-efficiency organic light-emitting diodes (OLEDs).35,37,38 The small molecular core and easy functionalization with additional side-chains enable thin film deposition via thermal evaporation as well as solution-based methods. This versatility of film formation and functionalization also facilitates variations of morphology and molecular packing over a wide range.29 However, unlike the sulfur-substituted analogues of ININs (such as BTBTs: benzothieno[3,2-b]benzothiophene),13,24,39 and despite the intriguing properties of the ININ class molecules, their polymorphism and its impact on optoelectronic properties have not yet been explored.
Here we employ thermal evaporation and zone-casting from solution to create thin films of C8-C-inin with markedly different macro- and microscopic morphologies and study their influence on polymorph formation, transformation, and stability. By combining atomic force microscopy (AFM), GIWAXS, Raman and absorption/fluorescence spectroscopy we explore the different polymorphs and the molecular rearrangements during polymorph transformation. Variations in crystallinity and morphology directly affect the transformation kinetics (fast or slow) in different thin films as large molecular reorientations occur. Furthermore, we identify conditions for which polymorph transformation processes can be reversed.
The absorption spectrum of C8-C-inin in hexane solution (Fig. 1b) shows two main absorption bands that are common for ININ-derivatives and are associated with the S0 → S1 (376 nm) and the S0 → S2 (333.5 nm) transition.40 Absorption and emission band positions were determined from the second derivatives of the smoothed spectra to correct for peak shifts due to band overlap. The S0 → S1 transition shows a typical vibronic progression with the 0–0 band at 376 nm and the 0–1 band at 357.5 nm. The photoluminescence spectrum (Fig. 1c) mirrors the vibrational progression of the absorbance (0–0 band at 379 nm and 0–1 band at 403 nm). The relatively rigid molecular core structure is reflected in a small Stokes shift of only 3 nm. High photoluminescence quantum yields (PLQY) of 87 ± 11% were observed for deoxygenated toluene solutions of C8-C-inin.
Thin films of C8-C-inin were deposited in two different ways, by slow thermal evaporation (0.2 Å s−1) in vacuum and by zone-casting from toluene solutions (Tstage = 80 °C, Tsolution = 60 °C, casting speed 0.1 mm s−1). Both types of films show nearly identical absorption spectra (see Fig. 1b) indicating very similar molecular packing orders. An additional peak not detectable in the solution spectrum appears at ∼354 nm in the fresh thin-film spectra. Its exact identity remains uncertain, but it might correspond to the 0–2 peak, which is too weak to be observed for the isolated molecules in solution as it would be covered by the strong S0 → S2 transition. For both thin films, the 0–0/0–1 band intensity ratio is lower than in the solution spectrum. This change in ratio indicates the formation of H-aggregates as a result of close molecular packing with direct overlap of the π-electron systems of the neighbouring molecules.41 However, the observed redshift of the 0–0 (solution: 376 nm, therm. evap.: 395 nm, zone-cast: 394 nm) and 0–1 absorption bands (solution: 357.5 nm, therm. evap.: 373 nm, zone-cast: 372 nm) is in contrast to the expected blueshift for H-aggregates. Due to the rigidity of the molecular core (the π-electron system) we can rule out conformational changes in the solid state as the reason behind the changed vibronic coupling and the spectral redshift. However, this discrepancy might be attributed to a relatively small excitonic coupling, with the blueshift due to H-aggregation being smaller than the stabilization by dispersive forces in the solid state resulting in a ‘solution-to-crystal’ redshift. Similar behaviour has been observed for H-aggregates of other molecules.41 Thermally evaporated C8-C-inin films still showed fairly high PLQYs of 34 ± 4% despite H-aggregates usually leading to suppressed PL.41 Some solid state emitters with vibrational coupling and low excitonic coupling have been shown to retain high PLQYs if the non-radiative decay rates are slow.41–43 However, this is relatively rare for polycrystalline films, in contrast to single-crystals, as large numbers of grain boundaries often lead to fast non-radiative decay.
Zone-cast films of C8-C-inin clearly show fewer grain boundaries, however, their PL spectra also differ substantially from the evaporated polycrystalline films. The PL spectrum (see Fig. 1c) shows a strong and broad emission band around 500 nm and a reduced 0–0/0–1 intensity ratio compared to the thermally evaporated film. The spectral evolution of the PL spectra of thermally evaporated C8-C-inin films at 77 K from 0 to 10 ns and time-resolved PL decay measurements reveal a long-lived emission at 500 nm (see ESI,† Fig. S14). Together with its occurrence only in the solid state, we attribute this broad emission feature to excimer species that are more pronounced in the zone-cast films due to increased long-range order. The increased excimer intensity and its overlap with the 0–1 band but not the 0–0 band explains the smaller 0–0/0–1 PL band ratio of the zone-cast film compared to the thermally evaporated film despite their very similar 0–0/0–1 absorption band ratios.
Large single-crystals of C8-C-inin were grown from a THF:hexane solvent mixture (slow solvent diffusion) and also by recrystallization from toluene. The obtained single-crystal structures are different for crystals grown from the THF:hexane compared to crystals from toluene solution (see ESI,† Fig. S16–S19). However, both crystal types show nearly the same peaks in differential scanning calorimetry (DSC) (see ESI,† Fig. S20) with an endothermic peak on first heating at 103.5 °C/104.6 °C (potential polymorph transformation) and a second endothermic peak at 151.6 °C/152.6 °C (melting). We attribute the similarity of the DSC data for the two crystal types to their similar intermolecular interactions. The two bulk-phase single-crystal structures share an interlocking pattern of the octyl chains between layers and small overlaps between the π-electron systems of neighbouring molecules. Both single-crystal structures exhibit unit cell parameters that are different from those of the thin films (see ESI,† Table S3). A much larger c-axis length of 3.5–3.6 nm compared to the 2.8 nm of the thin films is most obvious. The smaller out-of-plane packing distance of the thin films excludes the possibility of interlocking octyl chains between layers. Overall, C8-C-inin layers created by thermal evaporation and zone-casting yield a unique thin-film polymorph that is different from bulk single-crystals. The existence of multiple crystal structures poses the question of packing stability and the possibility of polymorph transformations.
For a polycrystalline thermally evaporated C8-C-inin film, the first morphology changes occurred within a few hours after deposition (see ESI,† Fig. S21). Consecutive AFM images collected over 8 hours at room temperature (ESI,† Video S1) show significant diffusion of material and drastically changing surface topology. In contrast to that, zone-cast films only showed significant morphology changes after 1–2 months. Fig. 3 shows the morphology changes of the thermally evaporated and zone-cast films after more than 4 months of storage in dry nitrogen at room temperature. A clear transformation to needle-like layers and a strong reduction of the minimal step height from 2.1 ± 0.2 nm to 0.7 ± 0.2 nm occurred for most zone-cast films (Fig. 3a and b) and all of the thermally evaporated films (Fig. 3d and e). However, for a few zone-cast crystalline ribbons (Fig. 3c) with very smooth surfaces (homogeneous height, very low number of molecular steps, edges, and defects), no morphology or significant step height changes (2.1 ± 0.2 nm to 2.3 ± 0.2 nm) were observed even after over 4 months. These different transformation kinetics for thermally evaporated films and zone-cast films of different roughness clearly indicate different kinetic barriers for the transformation in these films.
The decreased out-of-plane packing distance of the transformed films suggests that the C8-C-inin molecules not only diffuse over the substrate but also tilt to a more parallel orientation to the substrate plane (see schematic of molecular orientations in Fig. 3f). GIWAXS data of the aged thermally evaporated films provide further evidence for the reorientation of the molecules (see ESI,† Fig. S22). They show sharp reflections instead of diffuse rings, thus indicating increased in-plane order. The unit cell parameters of the aged thermally evaporated film were extracted from GIWAXS data via iterative diffraction peak fitting and simulation (see ESI,† Table S4). These unit cell parameters differ from all other polymorphs that were discussed so far, thus indicating a polymorph transformation during film aging. The extracted out-of-plane distance of ∼0.7 nm corresponds well to the minimal AFM step height. The longest in-plane distance of ∼2.8 nm is in good agreement with a mostly parallel orientation of the molecular long axis relative to the substrate plane.
The aging/polymorph transformation processes of C8-C-inin thin films are also reflected in their absorption and PL spectra. Fig. 4a and b show the corresponding spectra of a fresh and aged thermally evaporated film and bulk-phase single-crystals (recrystallized from hexane). The relative absorption band intensities changed significantly with aging while transition energies remained mostly the same with only a slight blueshift of 9 nm, indicating a modified intermolecular orientation but no chemical degradation. The 0–0/0–1 absorption band ratio is sensitive to changes in the intermolecular excitonic coupling, which in turn is related to changes in the molecular packing.41 The similarity between the 0–0/0–1 absorption band ratios of the aged film and the single-crystal powder suggests a similar packing order in both. The 0–0/0–1 ratio is near unity and thus equal to the ratio observed for the molecules in solution (see Fig. 1b), suggesting very small intermolecular excitonic coupling in contrast to the small 0–0/0–1 ratio of the fresh films indicative of H-aggregates.
The main difference between the PL spectra of aged thermally evaporated film and the bulk-phase single-crystals (Fig. 4b) is the very low emission intensity at the 0–0 band for the single-crystals. This can be explained by the higher self-absorption of the latter caused by an increased optical path length due to the thickness of the crystallites as well as waveguiding. Both single-crystals and aged films show relatively low excimer emission intensities despite the clearly higher long-range order compared to the freshly thermally evaporated film. For single-crystals, this could be a consequence of the small overlap of the π–π-electron systems between neighbouring molecules (ESI,† Fig. S16–S19) and represents additional evidence for a similar molecular packing order in aged films.
To summarize, while the unit cell parameters of the polymorphs of the aged films and single-crystals are clearly different, the next-neighbour orientations and interactions appear to be quite similar for the two samples. Only the longest unit cell axes are significantly different (∼2.8 nm vs. 3.5 nm) which might be explained by a slightly higher symmetry in the packing along these axes. Nearest neighbour interactions along these axes should only consist of interactions between octyl-chains. Hence, the intermolecular excitonic coupling along these axes is negligible due to the large distances between the molecular cores. They should have no significant influence on the absorption and emission properties. These observations explain the seemingly contradictory results of similar optical properties but different unit cell parameters of the aged thin film polymorph and the bulk-phase single-crystals.
To further explore the similarities and distinctions of the different molecular packing orders, we applied low-frequency Raman microscopy. As Raman modes below ∼200 cm−1 are mostly attributed to intermolecular vibrations and rotations they depend on the molecular packing and can be used to distinguish between different polymorphs.4,16,17 Fig. 4c shows the low-frequency Raman spectra of a freshly zone-cast film of two different aged zone-cast films and of bulk-phase single-crystals. While the single-crystals showed Raman bands at 80, 90 and 158 cm−1, the freshly zone-cast films exhibited one main band at 107 cm−1 and smaller features at 87, 96, and 116 cm−1, again indicating different polymorphs in agreement with AFM, GIWAXS, and absorption data. Aged zone-cast films exhibited two distinct types of Raman spectra (referred to as type 1 and type 2 in the following) associated with slightly different morphologies (ESI,† Fig. S23). Here, type 1 was the predominant form of aged zone-cast films and had a mostly needle-like morphology as already shown in Fig. 3b (see also ESI,† Fig. S23). This morphology is similar to that of aged evaporated films, suggesting that the type 1 polymorph observed in aged zone-cast films is probably the same as that of aged evaporated films. Type 1 films display Raman signals identical to those of single-crystal powder (80, 90 and 158 cm−1), indicating similar intermolecular vibrations and rotations, thereby confirming the presence of the same or at least very similar polymorphs in the aged zone-cast film and the single-crystal powder.
In contrast to that, type 2 was only observed for a small minority of aged zone-cast C8-C-inin ribbons and exhibited a more island-like surface morphology (ESI,† Fig. S23) but still a similar minimum step height of 0.7 ± 0.2 nm. Type 2 films feature a unique Raman signal at 75 cm−1, alongside additional peaks with small relative intensities close to those observed in fresh zone-cast films (85, 97 and 110 cm−1) and single-crystal powder (153 cm−1). We postulate a co-existence of the fresh thin-film, the bulk-phase single-crystal-like, and a potential third polymorph in different areas of the type 2 films. This third polymorph may represent a transitional phase from the metastable thin-film polymorph to the stable bulk-phase single-crystal-like polymorph. Overall, low-frequency Raman spectroscopy can be used to identify and track different polymorph transformations of C8-C-inin films by probing the whole volume of a measurement spot and not just the surface (as in AFM).
Interestingly, smooth zone-cast films did not show any morphology changes (AFM) or changes of the low-frequency Raman modes compared to the fresh zone-cast films even after nine months (ESI,† Fig. S24). Hence, we hypothesize that the differences in polymorph stability and the transformation speed of rough and smooth zone-cast films and thermally evaporated films are the result of a higher kinetic barrier for molecular diffusion and rotation/reorientation in closely packed crystalline ribbons. Molecular reorientation requires free volume, and such space would only be available at grain boundaries, crystal edges, and packing defects. At these positions, which are numerous in the polycrystalline thermally evaporated films and rougher zone-cast ribbons, the kinetic barrier should be low enough to permit polymorph transformation at room temperature.
We tested this hypothesis by observing the birefringence of different C8-C-inin films with cross-polarized microscopy at different temperatures. The transformed polymorphs of aged thin films (both type 1 and type 2) were strongly birefringent but nearly no birefringence was observed for the metastable polymorph of freshly deposited films (see ESI,† Fig. S25a and b). Bright-field optical microscope images did not show enough contrast to locate areas with polymorph transformation (see ESI,† Fig. S25c and d). However, differences in birefringence allowed for tracking the polymorph transformations over much larger areas than AFM or Raman measurements.
Films consisting mainly of very smooth, low-defect ribbons were deposited by lowering the C8-C-inin concentration of the toluene solutions used for zone-casting to 0.25 mg mL−1 (compared to 2.5 mg mL−1). The smooth zone-cast films showed no transformation of the metastable thin-film polymorph even after prolonged heating (up to 4 hours) at 60 °C, 80 °C, 110 °C, or 130 °C. Heating to 150–160 °C resulted in melting of the thin film (melting temperature of bulk-phase single-crystals ∼152 °C). The stability observed at low temperatures (close to room temperature) might be attributed to a high kinetic barrier for polymorph transformation in highly ordered crystal ribbons. However, the absence of transformations at elevated temperatures, such as 130 °C, indicates the presence of another contributing factor. In addition to kinetics, thermodynamic aspects must also be considered.
Polymorphic systems exhibit a transformation temperature at which the relative thermodynamic stability (Gibbs free energy) of the polymorphs switch order (see schematic for a dimorphic system in Fig. 5a).21 Hence, while the kinetic barrier of molecular reorientation in a system might be overcome at higher temperatures, the thermodynamic driving force for the transformation is lost. Consequently, no conversion from thin-film polymorph to bulk-phase or aged thin-film polymorph can occur. However, at temperatures above the transformation point the reverse process should occur, i.e., conversion of the formerly thermodynamically stable polymorph into the metastable polymorph. The only additional conditions are that the transformation temperature must be lower than the melting temperature of both polymorphs (enantiotropic system)23 and the kinetic barrier can be overcome.
To test for such a reverse polymorph transformation, an aged zone-cast C8-C-inin film with clear birefringence was heated from 40 °C to 80 °C in steps of 10 °C for 10 min each. The birefringent polymorph remained stable up to 70 °C but around 80 °C a fast transformation towards the non-birefringent polymorph occurred (see Fig. 5b and ESI,† Fig. S26). AFM images show that the film surface changed from a needle-like morphology to an island-like morphology (see ESI,† Fig. S27). The minimum step-height of the film changed from 0.7 ± 0.2 nm to 2.1 ± 0.2 nm. In addition, the low-frequency Raman spectrum of the annealed film showed no significant differences to the spectra of freshly zone-cast films (see ESI,† Fig. S28). Thus, AFM and Raman data confirm that annealing indeed induced a reverse polymorph transformation back to the thin-film polymorph of freshly zone-cast C8-C-inin films.
Cross-polarized microscope images during annealing of the aged zone-cast film show that the polymorph transformation occurred via a nucleation and growth mechanism (see Video S2 and Fig. S29, ESI†). Nucleation and growth transitions are the most common transition types and occur generally for polymorph transformations with significant changes in molecular packing, where molecules move individually across phase boundaries (in contrast to martensitic transitions).3,45 Note, that all nucleation events occur at positions where lower kinetic barriers towards molecular reorientation can be expected (i.e., crystal edges, grain boundaries). For nucleation and growth transitions, defects in molecular packing and grain boundaries increase the rate of polymorph transformation.45–48
Overall, these experiments support our hypothesis that the polymorph metastability of C8-C-inin is strongly influenced by the film morphology, especially the presence or absence of grain boundaries and defects. They are the origin of the vastly different polymorph transformation kinetics in thermally evaporated films and different zone-cast films.
Both fresh films showed low hole mobilities of up to μsat = 0.01 cm2 V−1 s−1 in the saturation regime. Determining hole mobilities in the linear regime was not possible due to large bias-stress instabilities when only small drain voltages were applied. Large and mostly irreversible negative shifts of the threshold voltage occurred during those measurements (see ESI,† Fig. S30). The irreversible shift of the threshold voltage might be caused by the irreversible secondary oxidation step of C8-C-inin molecules (to C8-C-inin2+) as indicated by cyclic voltammery (see ESI,† Fig. S13). Permanently trapped positive charges in the semiconducting layer could lead to such large threshold voltage shifts.
The charge transport properties of the zone-cast C8-C-inin thin films changed over the course of 9 months depending on their morphology and polymorph stability. For rough films with polymorph transformation, a strong reduction in the drain current on/off ratio was observed, the hole mobility dropped below 0.001 cm2 V−1 s−1, and the onset voltage shifted to more negative voltages. This degradation might be explained by differences in the transfer integrals for the changed molecular packing orders. Furthermore, the presence of many different crystal orientations in the transformed polymorph as well as phase boundaries between transformed and metastable polymorphs (see ESI,† Fig. S23) are likely to contribute to charge trapping and overall reduced charge carrier mobilities.
In stark contrast to these devices, the hole mobilities of OFETs with thin smooth films of zone-cast C8-C-inin, which showed no polymorph transformation, actually increased by an order of magnitude (up to μsat = 0.1 cm2 V−1 s−1) after aging for 9 months. A steeper subthreshold swing also indicates a lower density of trap states. This enhanced performance might be due to the high molecular mobility of C8-C-inin molecules at crystal edges and defects as discussed above, which may facilitate healing of trap-states through equilibration of the molecular orientations at grain boundaries. It also confirms that the observed device degradation for rough films was not the result of any chemical changes of the C8-C-inin molecules.
In summary, the stabilization of the C8-C-inin thin-film polymorph in highly ordered and smooth zone-cast films not only prevents device degradation due to polymorph transformation but even enhances charge transport after long-term storage in contrast to many other organic semiconductors.
GIWAXS measurements of thermally evaporated films were carried out on a Xeuss 2.0 instrument equipped with an Excillium MetalJet liquid gallium X-ray source. Films prepared on glass were collected for 900 s using collimating slits of 0.5 × 0.6 mm (“high flux” mode). Alignment was performed via three iterative height (z) and rocking curve (Ω) scans, with the final grazing incidence angle set to Ω = 0.3°. Scattering patterns were recorded on a vertically-offset Pilatus 1M detector with a sample to detector distance of 385 mm, calibrated using a silver behenate standard to achieve a q-range of 0.035–2.0 Å−1. Data reduction was performed using the instrument-specific Foxtrot software.
Unit cell sizes and lattice parameters were obtained from GIWAXS data using the MatLab script GIWAXS-SIIRkit provided by Savikhin et al.50 GIWAXS-SIIRkit allows calculating lattice parameters from a GIWAXS pattern by fitting well-defined and oriented diffraction peaks.
Crystallographic data for C8-C-inin single-crystal polymorphs 1 and 2 have been deposited at the CCDC under CCDC 2448535 and 2448536.
Footnote |
† Electronic supplementary information (ESI) available. CCDC 2448535 and 2448536. For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d5tc01928a |
This journal is © The Royal Society of Chemistry 2025 |