This document is the Accepted Manuscript version of a Published Work that appeared in final form in ACS Appl. Mater. Interfaces 2015, 7, 1078−1086 © American Chemical Society after peer review and technical editing by the publisher. To access the final edited and published work see http://pubs.acs.org/doi/pdf/10.1021/am505197b Influence of the molecular weight and size dispersion of the electroluminescent polymer on the performance of air stable Hybrid Light Emitting Diodes Eugenia Martinez-Ferrero,*a Souren Grigorianb, James W. Ryanc,†, Werther Cambarau,c and Emilio Palomaresc,d a b Fundació CETEMMSA, Avda. Ernest Lluch 36, E-08302 Mataró (Spain). Institute of Physics, University of Siegen, Walter-Flex-Str. 3, D-57068 Siegen (Germany). c Chemical Research Institute of Catalonia (ICIQ), Avda. Païssos Catalans 16, E-43007 Tarragona (Spain). d Institució Catalana de Recerca i Estudis Avançats (ICREA), Barcelona (Spain). 1 ABSTRACT: The influence of the chain length and the molecular weight distribution of the electroluminescent polymer on the carrier transport properties and morphology of air stable Hybrid Light Emitting Diodes is reported. It is found that variations between diverse as-received commercial batches play a major role in the performance of the devices, whose maximum luminance can differ up to two orders of magnitude. Through complementary opto-electronic, structural and morphological characterization techniques we provide insights into the relationship between charge dynamics and the structure of polymeric electroluminescent materials. The carrier dynamics are found to be dominated by both the polymeric chain length and the hole transport, which in turn, is dependent on the concentration of trap states. Furthermore, the chain length is seen to affect the morphology of the active layer. KEYWORDS: electroluminescent polymers, hybrid light emitting diodes, F8BT, molecular weight, polidispersity index, charge carrier mobility INTRODUCTION The performance of optoelectronic devices relies heavily on the properties of the semiconducting polymers that make up the active layer. The packing of the molecular chains affects the charge transport in such a manner that polaron or exciton separation, charge transport and recombination depend on the resulting morphology.1-3 This has been reported in significant detail for many organic electronic devices, in particular organic solar cells and organic transistors,4 where it has 2 been demonstrated that the behaviour of the polymer depends on its molecular weight, which can be difficult to control and can lead to large differences in device performance.5 Furthermore, the chain length and its size distribution are also key factors that determine the packing of the polymer chains and their orientation towards the electrodes.6 Indeed, the distance between the aromatic groups of adjacent chains also varies with the polymeric Mw resulting in different charge mobilities.7-9 On the other hand, the polidispersity index (PDI) is ubiquitous, particularly in commercial grade samples, adding variability to the reproduction of the devices. In this work, we explore the impact of the variations of the molecular weight and the PDI of commercial electroluminescent polymer. Despite numerous investigations of polymeric properties on the performance of organic field effect transistors and solar cell applications,10-13 this topic has not been explored with the same intensity in the field of organic light emitting diodes (OLEDs).14-16 Therefore, we have chosen these types of devices due to their potential as light sources and the major influence that the electroluminescent layer has on the properties of the device. However, despite their impressive efficiencies, one significant drawback of OLEDs is the use of low work function electrodes and organic transport layers that are sensitive to moisture and oxygen and require costly protective layers.17 Therefore, it is desirable to replace the organic materials by stable transition metal oxides as charge transport layers. In this way, the robustness, transparency and electronic properties of the inorganic materials are then combined with the possibilities offered by the organic lumophores giving rise to Hybrid Light Emitting Diodes (see Scheme 1). These devices, which have shown similar performances to those of the best published polymeric OLEDS,18 are formed by sandwiching organic electroluminescent polymers between an electron transport layer (ETL) such as ZnO, TiO2 or ZrO2 and a MoO3 hole transport layer (HTL) topped with an inert high work function gold anode.19-21 In addition, the low cost and the 3 facile preparation of the metal oxides permit the reduction of production costs and the fabrication of devices in air by solution methods. F8BT (poly-9,9’-dioctylfluorene-alt-benzothiadiazole, Scheme 1b) has been used as an electroluminescent polymer since it is widely used in organic electronics and its LUMO and HOMO levels matches well with the conduction bands of the charge transport layers (Scheme 1c). Scheme 1. (a) Description of the HyLED multilayered structure, (b) the chemical structure of the F8BT polymer and (c) the energy band diagram of the device. In order to focus exclusively on the variability of the semiconducting polymer, we have acquired different commercial batches of F8BT and tested them on devices with identical configurations. Depending on the batch, the maximum luminance achieved by the diodes can differ by up to two orders of magnitude and significant differences in the current-voltage characteristics are observed. To determine the key properties controlling the device performance, a detailed study on the morphology, surface and optical properties of the active layer has been carried out. In addition, we have also probed the charge carrier dynamics of the devices under 4 working conditions to obtain more detailed information on the effect of polymer structure on device performance. EXPERIMENTAL SECTION F8BT samples were acquired from different suppliers and their molecular weights measured by gel permeation chromatography against polystyrene standards (Table 1). The reported purities are 99.99% and metal contents, when indicated by the provider, below 84 ppm. Zinc acetate dihydrate, methanol, xylene and MoO3 were purchased from Sigma Aldrich and used as received. Device preparation and characterization. The ZnO layers were prepared by spray pyrolysis deposition from a precursor solution of zinc acetate dihydrate in methanol (80 g/L) onto prepatterned ITO glass substrates (10 Ω/ sq, obtained from Psiotec Ltd.) at 450 °C followed by annealing at 450 ºC for 30 min. The electroluminescent F8BT layer was spin coated from a xylene solution and annealed at 155 °C for 30 min in a nitrogen rich atmosphere. The solution concentration was adjusted to get film thicknesses of 100 nm. Finally, the samples were transferred to a thermal evaporation chamber where the MoO3 (15 nm) and Au (100 nm, Kurt Lesker) layers were deposited through a shadow mask under high vacuum (1 x 10-7 mbar), which defined an active area of 0.25 cm2. The as-prepared devices were transferred out of the glove box and measured in air without any form of encapsulation. The devices were characterized with a Thornlabs silicon photodiode combined with a Keithley 2400 source controlled by in-house software. Hole only devices where prepared and characterized following the same procedure than the bipolar diodes. The ZnO layer is substituted by 50 nm of PEDOT:PSS which is deposited by spin 5 coating and annealed under nitrogen at 120 ºC for 30 min. The subsequent layers are deposited as described above. Film characterization. Thicknesses were measured by an Ambios Technology XP-1 profilometer and confirmed by X-ray reflectometry. Atomic Force Microscopy measurements were done in tapping mode on a Molecular Imaging model Pico SPM II while images were analyzed using WSxM software.22 Grazing incident X-ray diffraction (GIXD) studies were performed using a high-resolution diffractometer Seifert XRD 3003 operating at 8 keV and equipped with a point detector. Out-of-plane GIXD profiles were measured at the fixed incident angle α = 0.17º being higher than the critical angle of the polymer film and lower than the critical angle of the substrate. The surface free energy was estimated by the Owens-WendtRabel-Kaelble (OWRK) method from the contact angle measurements done on the polymer films by means of a Krüss DSA100. Differential scanning calorimetry (DSC) and thermogravimetric analysis (TGA) were performed using a Mettler-Toledo TGA-DSC 1 under nitrogen. DSC experiments were carried out in two consecutive heating cycles from 30 to 300 and cooling back to 30 ºC at a rate of 10 ºC/min while TGA consisted in a single measurement from 30 to 500 ºC at a rate of 20 ºC/min. UV-Visible absorption was measured in a Perkin Elmer Lambda 950 while fluorescence measurements were recorded using an Aminco-Bowman Series 2 luminescence spectrometer. Time correlated single photon counting (TCSPC) experiments were carried out after recording 1000 counts per second (CPS) with a Lifespec picosecond fluorescence lifetime spectrophotometer from Edinburgh Instruments© equipped with lasers as excitation sources. The instrument response was always shorter than 440 ps measured at full width half maximum (FWHM). For transient electroluminescence measurements (TEL), the devices were electrically excited by a fast pulse generator Thurlby Thandar Instruments TGP110, 6 which permitted accurate control of rise and decay times. Single square-wave voltage pulses between 4 and 7 V were applied during 100 microseconds. The electroluminescence response was detected using a PM Hamamatsu H6780-01 photomultiplier tube detector connected to an amplifier from Hamamatsu C9693. A Teknotronix TDS 2022 oscilloscope was used to record the voltage, time and electroluminescence. All the measurements were carried out at ambient conditions without using encapsulating layers. RESULTS The commercial samples employed in this study are summarized in Table 1. The Hybrid Light Emitting Diodes (HyLEDs) were structured as ITO/ZnO/F8BT/MoO3/Au and all the layers were deposited under the same conditions to permit a fair comparison of the behavior of the different polymers. Table 1. Description of the F8BT batches employed in this work. PDI accounts for the polidispersity index. Sample Mn (g/mol) Mw (g/mol) PDI Supplier F15k 5897 15686 2.7 Sigma Aldrich F23k 4107 23000 5.6 American Dye Source F27k 12273 27000 2.2 1-Material F35k 14583 35000 2.4 American Dye Source F43k 17133 43517 2.5 Sigma Aldrich F60k 31579 60000 1.9 QCR Solutions F73k 22812 73000 3.2 American Dye Source 7 F116k 34118 116000 3.4 American Dye Source The diodes were characterized by measuring the evolution of light emission with applied bias (Figure 1). The devices present turn on voltages lower than 3 V and reach the maximum luminance below 10 V (see Table 2 and Table S1 on the supplementary information). The best luminance values are achieved in devices containing the F73k sample while poorer behaviors of up to two orders of magnitude lower are observed when F15k and F60k are employed. It is interesting to note that the turn-on voltages for the samples F15k, F60k and F73k deviate from the mean estimated value of the other samples. This is rather surprising since turn on voltages are related to the electron injection barrier which is defined as the difference between the ITO workfunction and the metal oxide conduction band edge.18 Figure 1. Luminance (a) and current density (b) vs applied bias for the ITO/ZnO/F8BT/MoO3/Au devices prepared with F8BT of different Mw and PDI. 8 Figure 1b shows the evolution of the density current vs. applied bias (Figure 1b) from a linear dependence that turns exponential at voltages closer to the built-in voltage (Vbi), pointing to the change from a diffusion dominated current into a space charge limited regime.23 This fact arises from the energy band alignment of the different components of the device, described in Scheme 1c. The MoO3 of the HTL forms an ohmic contact with the electroluminescent polymer allowing the injected holes to drift from the electrode to the organic-inorganic interface where, due to the energy mismatch between the valence band (VB) of the ETL and the HOMO of the polymer, accumulate leading to a built in field. The field assists the injected electrons on the other electrode to overpass the electron injection barrier between the conduction band (CB) of the ETL and the LUMO of the F8BT.18 The charge transport at voltages higher than Vbi is thus bulk limited and is best described as pseudo-hole only space charge limited current, since the predominant carriers are holes.24-26 The J-V curve can be modeled using the Mott-Gurney equation combined with field dependent mobility (equation 1): J sclc =  V 9 V2 εµ0 3 exp  β  8 L  L (1) where ε is the ε0·εr permittivity of the F8BT polymer, µo is the hole zero field mobility, V is the effective applied bias (the applied bias corrected by the built-in potential), L is the thickness of the active layer and β is the field-effect mobility coefficient (also known as Poole-Frenkel factor). The parameters obtained from the fitting of the J-V curves, listed in Table 2 and shown in the supplementary information, indicate that the room temperature hole mobility µo amounts to ~10-5 cm2/V·s for all the samples except for F15k, F43k and F60k. The β coefficient, which depends upon trap depths, is estimated to be 10-4 cm1/2/V1/2, except for F15k, F35k and F43k which is one order of magnitude lower suggesting an increased influence of traps in the charge 9 carrier mobility. Comparison with parameters obtained from hole-only devices made of ITO/PEDOT:PSS/F8BT/MoO3/Au when the hole injection takes place from the MoO3 (see Table S2 and Figure S2 in the supplementary information), indicate that the difference is less than onefold and probably arises from the beneficial influence of the ZnO layer. Independent measurements of the hole mobility under working conditions are obtained by transient electroluminescence (TEL) in which the time dependence of the rise of electroluminescence is recorded after application of square voltage pulses of different amplitude. The delay time (td) between the voltage start and the onset of light emission is a direct measurement of the transit time of the faster charge carriers (holes, vide supra) towards the recombination zone. Therefore, TEL curves can be well described by the equations used for hole-only single layer OLEDs in which holes drift faster towards the interface dominating the early stage of the rising of electroluminescence.27 Table 2. Optoelectronic data of the devices obtained from the L-V curve, the fitting of the JV curve and TEL measurements. Von Lmax µ0 β Vtrap (V)a (cd/m2)b (cm2/V·s)c (cm1/2/V1/2)c (V)d F15k 2.8 26 1.1 x10-6 4.1 x10-3 2.9 F23k 2.3 1440 4.3 x10-5 1.3 x10-4 2.7 F27k 2.1 1340 1.0 x10-5 1.0 x10-4 2.3 F35k 2.4 890 1.8 x10-5 1.2 x10-3 4.1 F43k 2.3 224 1.9 x10-7 3.9 x10-3 4.9 F60k 2.7 28 5.9 x10-7 3.9 x10-4 --- Sample 10 F73k 1.8 6400 2.8 x10-5 9.0 x10-4 3.2 F116k 2.3 530 9.7 x10-5 7.3 x10-4 2.5 a The turn-on voltage is defined as the applied bias when the luminance reaches 0.01 cd/m2; b maximum luminance; c data obtained from fitting the JV curves; d data obtained from TEL measurements. Once the holes reach the steady state, which occurs at times close to td, further rise of electroluminescence is then due to electron flow in the bulk. Even at this stage, hole density exceeds the electron concentration and is not affected by charge recombination. Consequently, we can estimate the hole mobility from the electroluminescence output using the equation of the drift mobility in a electric field (Equation 2): L2 µh = Vtd (2) where L accounts for the thickness of the active layer, V is the effective applied bias (V-Vbi) and td is the time delay. Figure 2a illustrates the linear dependence of the as-estimated µh with the electric field, except for F60k that could not be measured due to its low luminance. The variation between samples corroborates the previous observations of Figure 1b in the sense that charge dynamics are influenced by the molecular weight and PDI. 11 Figure 2. Hole mobility as a function of the square root of the electric field (a) and variation of the hole mobility with the free hole density (b). Solid lines indicate linear fitting, dashed lines indicate power law fitting. Further information can be obtained following the development made by Rihani et al. who estimated the free hole density (pe) considering the Langevin charge recombination coefficient (R) and the recombination current, Jr.27 Briefly, R can be estimated from the hole mobility through the equation 3. R= qµh ε (3) where q stands for the elementary charge, µh is the hole mobility and ε is the permittivity of F8BT. Furthermore, the recombination current is directly proportional to the light emission. Therefore, from the evolution of light emission with time we can estimate the time constant τ (equation 4), which is inversely proportional to the free hole density, pe. Equation 5 demonstrates this relationship where EL refers to the electroluminescent signal at a given time t and a is the value of the intersection at time 0: 12 t  EL(t → ∞) − EL(t)  ln  =a−  τ EL(t → ∞)   pe = τ −1 R (4) (5) The as-calculated free hole densities (equation 5) depend linearly with the applied bias so extrapolation to lower bias indicates that pe vanishes at certain voltages pointing to the presence of traps in the bulk (Vtrap, see Table 2). Therefore, the low charge carrier density at lower applied bias is due to the fact that the empty traps are becoming filled. As the applied voltage increases, the number of empty traps decreases and the number of free holes increases, resulting in enhanced current densities. A close inspection reveals that Vtrap is similar to Vturn-on suggesting that, at lower bias, the dominant mechanism is non-radiative trap assisted recombination in agreement with previous observations.23 The exception is found for the samples F35k, F43K and to less extent F73k that show higher Vtrap than the rest. Again, comparison of this data with that obtained from the J-V curve reveals that these samples posses’ high β values confirming the increased influence of the traps on the mobility values. On the other hand, the variation of µh with charge carrier density follows a power law trend, in agreement with the conduct observed for other polymers at charge carrier densities higher than 1016 cm-3.28 The dependence of the field-effect mobility on charge carrier density has been related to the width of the exponential density of states; therefore, the stronger dependence of the mobility on the charge carrier density in samples F35k and F27k are indicative of larger energetic disorder.28 Moreover, the analysis of the charge dynamics suggests a major influence of traps on the performance of the diodes. The device F43k display lower charge carrier mobility reflected in reduced free hole density and current density together with high values of β and Vtrap. 13 Moreover, the devices with lower current densities (F15k, F43k and F60k) lead to higher luminance turn-on voltages regardless of the electron injection barrier. In addition, these samples give poorer luminance. However, the charge mobility does not correlate with the variations in the Mw or the PDI. Consequently and in order to gain further information, we have undertaken optical and morphological studies on polymer films. The absorption spectra, shown in Figure 3a, present the typical pattern for F8BT with the main peak centered at 465 nm. The fact that the peaks are not significantly shifted or broadened suggests that the optical band gap does not change regardless of the Mw and the PDI. Photoemission measurements after excitation at 450 nm, Figure 3b, show a main signal centered at 536 nm except for F116k whose emission maximum is red shifted to 541 nm. The intensity of the peaks varies according to their respective PDI suggesting that higher size dispersion is more beneficial to efficient light emission than higher Mw. 14 Figure 3. Optical characterization of the thin films (a) UV-Vis absorption, (b) photoluminescence, (c) TCSPC decays (IRF: instrument response function). 15 Time-Correlated Single Photon Counting (TCSPC) records the decay signals as a function of time after excitation at 405 nm. The experimental points are fitted with the equation 6: I(t) = I 0 exp [−(t / τ )] (6) where I(t) represents the time (t) dependent decay of the fluorescence intensity, Io stands for the initial intensity values and τ is the lifetime value obtained from the decay, indicating the average time that the polymer remains in the excited state. Figure 3c shows the decays for all the samples. F15k, F43k and F60k fall within the response of the instrument (IRF) and thus, the decays are shorter than 0.44 ns. The rest of the samples show monoexponential decays that can be fitted to equation 6, resulting in lifetime values that are shown in Table 3. Table 3. Photophysical parameters of the polymeric films Sample λ max Abs(nm) λ max PL(nm) τ (ns) F15k 464 537 < 0.44 F23k 464 536 1.55 F27k 462 532 1.45 F35k 465 537 1.30 F43k 462 536 < 0.44 F60k 465 536 < 0.44 F73k 467 538 1.57 F116k 466 541 1.32 Additionally, we have measured the lifetime of the devices under ambient conditions. Measurements were done every 24 hours on devices without encapsulation that were stored in air. The results, which have reached values of 600 hours, do not show significant differences between the most efficient polymers. 16 In order to gain further insight into the relationship between the molecular weight distribution and the morphology, we have proceeded with differential scanning calorimetry (DSC). Figure 4 shows the second heating scan, where the first feature above 100 ºC is an endothermic transition followed by an exothermic peak around 150 ºC. The combinations of the first cooling scan with the second heating scan indicates that it is due to the transition of the polymer from the glassy to the rubbery state that takes place alongside a crystallization step.29 The glass transition temperature (Tg) has been therefore estimated at the onset of the first change of the heat capacity defined by the point of intersection between the baseline before the thermal effect and the tangent of the curve (Table 4). A second exothermic transition located around 185 ºC is assigned to further crystallization. Finally, there is the onset of a third transition at temperatures around 200 ºC. The endothermic nature of the smooth transition points to the occurrence of the melting process to liquid crystalline phase. Figure 4. DSC data of the second heating scan for all the polymers. 17 Structural analysis to probe ordering was performed by GIXD measurements at the incident angle of αi = 0.17º. This incident angle is above the critical angle of the polymer (αc = 0.15º) and below that of the glass substrate (αc = 0.21º) to permit the penetration of the X-beam through the depth of the whole film. Figure 5 shows the scattering profiles, which present a peak at small angles suggesting the presence of similar structural organization of crystalline domains regardless of the Mw or PDI of the samples (see supporting information Figure S3). The peaks at higher angles could not be resolved despite long counting time (200s) due to the weak signals arising from the low thickness of the samples. Therefore, X-ray studies were focused on the most intense scattering peaks centred at a 2θ position of ≈ 5.2º (see Figure 5). According to the literature, polyfluorenes usually form lamellas that stack parallel to the substrate in a configuration described as the β-phase and defined by the π-π stacking, the size of the F8-BT repeating unit and the interchain spacing.30,9 We have focused on the peak corresponding to the interchain spacing that lies between 1.66 to 1.69 nm, depending on the sample, which agrees with the 1.6 nm reported in the literature.30 Moreover, since the estimated length of the alkyl side chains is 8.75 Å we can deduce that the observed interchain distance corresponds to two polyfluorene backbones separated by alkyl chains. Similar edge-on orientation with the strongest (100) peak associated with side chain distances along the out-of-plane direction has been reported for a large family of thiophenes.31,32 All samples here show these characteristic features but for three exceptions, the F60k, F73k and F43k samples. The first, F60k, shows a rather broad distribution shifted to higher 2θ angles with two peak positions suggesting the existence of two polymorphs where the chains are separated by 1.6 or 1.4 nm. In addition, it confirms the higher amount of disordered domains for the F60k in comparison to all other samples in the series that provide narrower peaks. Interestingly, F73k and F43k samples in turn show just one peak (very 18 intense for F73k) located at even higher 2θ angles (see supplementary information) that correspond to interchain distances of about 1.2 nm. The size of the crystallites is obtained from the width of the peak in combination with the Scherrer equation (Table 4). The as-calculated size of the domains ranges from 10 to 17 nm with marginal changes of crystalline domains as a function of the molecular weight, except for F43k that measures 30 nm. Figure 5. Out-of-plane GIXD in the vicinity of the (100) peak of F8BT films. Further information has been obtained by AFM studies (Figure 6). It is interesting to point out the low roughness of the samples ranging between 1.8 and 3.4 nm (see Table 4). The films are all formed by domains of aggregated crystallites that agree with literature reports about the polycrystalline nature of films forming well-ordered domains when annealed above Tg.29,33 However, there are clear differences in the local morphology. The pattern in F43k is formed by long chains while the aggregates are much closer in F15K and F60k. Samples F23k, F35k, F73k and F116k are arranged in a similar fashion than F43k, whereas the F27k film is rather smooth but contains aggregates of 200 x 140 nm. It is worthy to note that F60k, which has the lowest PDI among the samples, forms small aggregates while larger domains are observed in samples with PDI values above 2.5. 19 Figure 6. AFM topography images (2 µm x 2 µm) of the F8BT films. The formation of molecular aggregates is known to be strongly affected by the surface energy of the molecules.34 Consequently, we have estimated the surface free energy of the different samples through the measurement of the contact angle of water, ethylene glycol and dimethyl sulfoxide on F8BT films. These values are applied in equation 7 to calculate the dispersive and polar interactions between F8BT and the liquids. σ L (1+ cosθ ) 2 σ D L P = σS P σL σ D L D + σS (7) 20 where θ is the contact angle and σ is the surface free energy of the F8BT (S) or the liquid (L). The total surface energy is calculated from the sum of the dispersive and the polar components and the results are summarized in Table 4. The results indicate that surface free energy (SFE) increases with the Mw especially due to the decrease of the polar component. These observations agree with Wu who defined a linear variation between the surface energy and Mn-2/3 so the variation of the Mw provides a reliable method to control the surface energy.34,35 However, the increase of Mw alone cannot explain the differences found in the structural characterization. The SFE of the chain ends is usually lower than that of the chain; therefore the increment of SFE values with higher Mw is assigned to the decreased ratio of chain ends to the repeat unit. The heterogeneity caused by the chain ends in the surface energy can affect the grain boundaries, the size of the domains and the order of the molecular arrangement.34 Moreover, it could be the cause of the irregular change in the mobility of semiconducting layers.36 In addition, we also need to consider the value of the PDI of each sample since it affects the heterogeneity of the surface as well. For example, on one hand, F60k has the lowest PDI, 1.9, shows the lowest mobility and forms films containing a broad dispersion of crystallites with a mean size of 9 nm arranged in a closed pattern resulting in low surface roughness. On the other hand, F23k has the highest PDI, 5.6, the charge mobility and luminance values are among the highest of the samples and it forms large and homogenous domains (15.2 nm) resulting in relatively rougher films. In fact, the SFE values are higher than expected for this Mw and the polar component is comparable to F43k. So we can deduce that the polymer samples with higher PDI (F23k, F35k, F43k, F73k and F116k) produce films with bigger domains that form large patterns as seen in the AFM images. The SFE values correlate with the Mw although some subtle deviations are observed due mainly to the influence of the PDI on the heterogeneity of the surface. Finally, and 21 although the films present certain surface roughness, it is worthy noticing that the average values are lower than 4 nm, and as reported by Wang et al., in these cases, the influence of the topography can be considered negligible.37 On the other hand, the measurements of the contact angle have been done using the advancing angle, which has been demonstrated to be relatively less sensitive to surface roughness.38 Therefore, the influence of the surface roughness on the measurements of the contact angle, and therefore, on the determination of the SFE have not been taken into account. Table 4. Temperature of the glass transition (Tg), wettability properties described by the surface energy (γe) and the dispersive (σD) and polar components (σP), the crystalline size (S) calculated by GIXS and surface average roughness (Rav) of the F8BT films. Film Tg γe (mJ/m2) σD (mJ/m2) σP (mJ/m2) S (nm) RAV (nm) (ºC) F15k --- 27.5 27.2 0.3 --- 2.2 F23k 104 29.9 29.7 0.2 15.2 2.3 F27k 107 28.3 27.9 0.4 16.9 3.1 F35k 109 28.1 27.8 0.3 14.2 3.0 F43k 104 27.9 27.7 0.2 30.6 2.8 F60k 101 29.7 29.5 0.2 9.7 1.8 F73k 109 29.7 29.6 0.1 11.2 2.6 F116k 112 30.4 30.3 0.1 13.3 3.8 DISCUSSION We have prepared devices with commercial F8BT samples ranging from 15686 to 116000 g/mol that present a broad dispersion of sizes. The performance of the devices depends on the 22 morphology of the active layer, which in turns depends on the polidispersity index and chain length of the polymer, as summarized in Table 5. We consider first the samples F43k and F60k. The estimated charge carrier mobilities and current densities are lower than the rest of the samples, and consequently, the devices offer poor results with low luminance values and high turn-on voltages. This trend is also observed in the optical properties where the emission rates and the lifetime of the photogenerated excitons are also lower. But despite the similar optoelectronic properties of these two devices, their morphologies differ and so do the reasons for their poor behavior. F60k is composed of two polymorphs that form small crystalline domains. The features of the polymorphism have been detected by GIXD and by the width of the endothermic peak at DSC measurements, which despite its lower PDI displays a broad feature. The increased number of domains implies the presence of more grain boundaries that act as traps, which hinder charge transport and photoemission. On the other hand, F43k possess shorter spacing between polymer chains that can induce the formation of well-ordered nanodomains. The presence of these crystalline domains has not been detected by optical measurements but the structural analysis reveals the closer interchain spacing that gives rise to homogeneous domains ranging up to 30 nm. The decrease of spacing between polymer chains causes a decrease of the charge mobility. This effect is also seen for F60k where one of the polymorphs shares similar structural characteristics with F43k. Interestingly, the sample F73k with shorter interlammelae distance due to interdigitation of the side chains provides relatively high field effect mobility and the highest luminance. Despite that in samples F43k and F60k such interdigitations cause a dramatic decrease of the optoelectronics characteristics, the sample F73k reveals the best performance. The origin of this enhancement can be related to the strong improvement of the orientation features confirmed by the highest intensity of the out-of-plane GIXD profile. 23 Thermogravimetric analysis of F60k reveals an unusual loss of weight between 100 and 360 ºC that have not seen in the rest of the samples (see supplementary information). Samples F15K, F23k, F27k, F35k, F73k and F116k display similar morphological features with homogeneous domains made of crystallites of similar sizes giving rise to moderate surface roughness. The performance of the devices improves with increasing length of the polymer chains, and therefore, F15k has low luminance values coincident with lower mobility values. This trend is usually assigned to the concentration of chain ends that can act as trapping centres. However, F116k does not produce the best device, as the trend would suggest. Reasons for the lower than expected performance can be deduced from the red shift of the photoluminescence peak and a moderate increase in the roughness of the film surface that suggests the formation of non-emissive aggregates. These aggregates could originate from partial chain entanglements that limit molecular motion and crystallization despite the annealing treatment.29 Finally, it is worth mentioning F23k, which behaves better than expected for its Mw. A priori, the chain ends act as exciton dissociation sites so shorter polymer chains with more chain ends are expected to have lower emission efficiency, but this effect is mitigated by the broad dispersion of chain lengths in the sample that contribute to increase the photoluminescent efficiency. As has been reported before, polydispersity enhances the ability of polymers to self-organize during the annealing process at optimized temperatures. In fact, devices made of P3HT with low Mw and high polidispersity have been observed to be more efficient than devices with higher Mw due to the enhanced crystallinity of the polymer chains.39 On the other hand, the presence of small fractions of high molecular weight polymers can drastically affect film formation in blends with polymers of lower Mw, their morphology and carrier dynamics.40 Finally, if we are to conclude on any 24 general trend from the experimental results, high PDI values favor good optical properties and the formation of large domains, which result in enhanced performance of the devices. Table 5. Summary of the experimental results observed in each sample. Sample F15k Mw (g/mol) 15686 PDI Experimental observations 2.7 Shortest polymer chain length and moderate PDI Poor charge mobility and optical properties Poor device performance F23k 23000 5.6 Broad dispersion of chain lengths Large crystalline domains Excellent device performance F27k 27000 2.2 Moderate polidispersity Formation of large aggregates Excellent device performance F35k 35000 2.4 Moderate polidispersity High values for Vtrap Good device performance F43k 43517 2.5 Short interchain distances generating large domains Poor charge mobility and optical properties Moderate device performance 1.9 Presence of polymorphs and impurities Small domains originating grain boundaries Poor mobility and optical properties Poor device performance F60k 60000 F73k 73000 3.2 High molecular weight and polidispersity Short interchain distances, strong orientation of the chains Excellent device performance F116k 116000 3.4 High molecular weight and polidispersity Formation of optical non emissive aggregates Moderate device performance CONCLUSIONS In conclusion, this study determines the influence of the molecular weight and polidispersity index of commercial F8BT in the performance of Hybrid Light Emitting Diodes. Contrary to 25 previous work on this topic made on purified polymers, the commercial batches have broad size dispersion that affects the performance of the devices. So, despite higher Mw favors charge mobility, the presence of long polymer chains does not always guarantee the formation of adequate morphologies that lead to efficient devices. Consequently, it is necessary to consider the molecular weight, the size distribution and the purity of the batches because of their influence on the morphology and the final properties of the devices. We believe that these conclusions will impact several important research fields where polymers are utilized such as organic solar cells, light emitting diodes and transistors promoting the industrial development of optoelectronic devices. ASSOCIATED CONTENT Supporting Information. Fittings of the JV curve of bipolar and hole-only devices, long range out-of-plane GIXD data and thermogravimetric analysis of the polymers. This material is available free of charge via the Internet at http://pubs.acs.org. AUTHOR INFORMATION Corresponding Author *Fax: +34 93 741 9228; Tel: +34 93 741 9100; E-mail: emartinez@cetemmsa.com Present Addresses † Present address: Department of Chemistry, The University of Tokyo, 7-3-1 Hongo, Bunkyoku, Tokyo 113-0033 (Japan) Author Contributions 26 The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript. Notes The authors declare no competing financial interest. ACKNOWLEDGMENT E.M-F. thanks the Spanish MINECO for the Ramon y Cajal fellowship RYC-2010-06787 and funding through the project MAT2012-31570. C. Carenas help for the acquisition of TEL data is acknowledged. S.G. is grateful to L. Grodd and C. Kwamen for assistance during X-ray measurements and to BMBF for financial support through the project Nr05K13PS4. Financial support from CETEMMSA is appreciated as well. REFERENCES (1) Ballantyne, A. M.; Chen, L.; Dane, J.; Hammant, T.; Braun, F. M.; Heeney, M.; Duffy, W.; McCulloch, I.; Bradley, D. D. C.; Nelson, J. The Effect of Poly(3-hexylthiophene) Molecular Weight on Charge Transport and the Performance of Polymer:Fullerene Solar Cells. Advanced Functional Materials 2008, 18, 2373-2380. (2) Salleo, A.; Kline, R. J.; DeLongchamp, D. M.; Chabinyc, M. L. Microstructural Characterization and Charge Transport in Thin Films of Conjugated Polymers. Advanced Materials 2010, 22, 3812-3838. 27 (3) Cornil, J.; Beljonne, D.; Calbert, J. P.; Brédas, J. L. Interchain Interactions in Organic π-Conjugated Materials: Impact on Electronic Structure, Optical Response, and Charge Transport. Advanced Materials 2001, 13, 1053-1067. (4) Chu, T.-Y.; Lu, J.; Beaupré, S.; Zhang, Y.; Pouliot, J.-R.; Zhou, J.; Najari, A.; Leclerc, M.; Tao, Y. Effects of the Molecular Weight and the Side-Chain Length on the Photovoltaic Performance of Dithienosilole/Thienopyrrolodione Copolymers. Advanced Functional Materials 2012, 22, 2345-2351. (5) Kline, R. J.; McGehee, M. D.; Kadnikova, E. N.; Liu, J.; Fréchet, J. M. J.; Toney, M. F. Dependence of Regioregular Poly(3-hexylthiophene) Film Morphology and Field-Effect Mobility on Molecular Weight. Macromolecules 2005, 38, 3312-3319. (6) Hiorns, R. C.; de Bettignies, R.; Leroy, J.; Bailly, S.; Firon, M.; Sentein, C.; Khoukh, A.; Preud'homme, H.; Dagron-Lartigau, C. High Molecular Weights, Polydispersities, and Annealing Temperatures in the Optimization of Bulk-Heterojunction Photovoltaic Cells Based on Poly(3-hexylthiophene) or Poly(3-butylthiophene). Advanced Functional Materials 2006, 16, 2263-2273. (7) Joshi, S.; Pingel, P.; Grigorian, S.; Panzner, T.; Pietsch, U.; Neher, D.; Forster, M.; Scherf, U. Bimodal Temperature Behavior of Structure and Mobility in High Molecular Weight P3HT Thin Films. Macromolecules 2009, 42, 4651-4660. (8) Joshi, S.; Grigorian, S.; Pietsch, U.; Pingel, P.; Zen, A.; Neher, D.; Scherf, U. Thickness Dependence of the Crystalline Structure and Hole Mobility in Thin Films of Low Molecular Weight Poly(3-hexylthiophene). Macromolecules 2008, 41, 6800-6808. (9) Donley, C. L.; Zaumseil, J.; Andreasen, J. W.; Nielsen, M. M.; Sirringhaus, H.; Friend, R. H.; Kim, J.-S. Effects of Packing Structure on the Optoelectronic and Charge 28 Transport Properties in Poly(9,9-di-n-octylfluorene-alt-benzothiadiazole). J. Am. Chem. Soc. 2005, 127, 12890-12899. (10) Tumbleston, J. R.; Collins, B. A.; Yang, L.; Stuart, A. C.; Gann, E.; Ma, W.; You, W.; Ade, H. The Influence of Molecular Orientation on Organic Bulk Heterojunction Solar Cells. Nature Photonics 2014, 8, 385-391. (11) Merari Masillamani, A.; Orgiu, E.; Samori, P. Effect of the Molecular Weight of the Polymer Gate Dielectric on the Performances of Solution-processed Ambipolar OTFTs. Journal of Materials Chemistry C 2013, 1, 7725-7730. (12) Intemann, J. J.; Yao, K.; Yip, H.-L.; Xu, Y.-X.; Li, Y.-X.; Liang, P.-W.; Ding, F.- Z.; Li, X.; Jen, A. K. Y. Molecular Weight Effect on the Absorption, Charge Carrier Mobility, and Photovoltaic Performance of an Indacenodiselenophene-Based Ladder-Type Polymer. Chemistry of Materials 2013, 25, 3188-3195. (13) Fitzner, R.; Mena-Osteritz, E.; Mishra, A.; Schulz, G.; Reinold, E.; Weil, M.; Körner, C.; Ziehlke, H.; Elschner, C.; Leo, K.; Riede, M.; Pfeiffer, M.; Uhrich, C.; Bäuerle, P. Correlation of π-Conjugated Oligomer Structure with Film Morphology and Organic Solar Cell Performance. Journal of the American Chemical Society 2012, 134, 11064-11067. (14) Haldi, A.; Kimyonok, A.; Domercq, B.; Hayden, L. E.; Jones, S. C.; Marder, S. R.; Weck, M.; Kippelen, B. Optimization of Orange-Emitting Electrophosphorescent Copolymers for Organic Light-Emitting Diodes. Advanced Functional Materials 2008, 18, 30563062. (15) Yim, K.-H.; Doherty, W. J.; Salaneck, W. R.; Murphy, C. E.; Friend, R. H.; Kim, J.-S. Phase-Separated Thin Film Structures for Efficient Polymer Blend Light-Emitting Diodes. Nano Letters 2010, 10, 385-392. 29 (16) Weinfurtner, K.-H.; Fujikawa, H.; Tokito, S.; Taga, Y. Highly Efficient Pure Blue Electroluminescence from Polyfluorene: Influence of the Molecular Weight Distribution on the Aggregation Tendency. Applied Physics Letters 2000, 76, 2502-2504. (17) So, F.; Kondakov, D. Degradation Mechanisms in Small-Molecule and Polymer Organic Light-Emitting Diodes. Advanced Materials 2010, 22, 3762-3777. (18) Sessolo, M.; Bolink, H. J. Hybrid Organic–Inorganic Light-Emitting Diodes. Advanced Materials 2011, 23, 1829-1845. (19) Bolink, H. J.; Coronado, E.; Sessolo, M. White Hybrid Organic-Inorganic Light- Emitting Diode Using ZnO as the Air-Stable Cathode. Chemistry of Materials 2009, 21, 439441. (20) Tokmoldin, N.; Griffiths, N.; Bradley, D. D. C.; Haque, S. A. A Hybrid Inorganic-Organic Semiconductor Light-Emitting Diode Using ZrO2 as an Electron-Injection Layer. Advanced Materials 2009, 21, 3475-3478. (21) Morii, K.; Kawase, T.; Inoue, S. High Efficiency and Stability in Air of the Encapsulation-free Hybrid Organic-inorganic Light-emitting Diode. Applied Physics Letters 2008, 92, 213304/213301-213304/213303. (22) Horcas, I.; Fernández, R.; Gómez-Rodríguez, J. M.; Colchero, J.; Gómez-Herrero, J.; Baro, A. M. WSXM: A Software for Scanning Probe Microscopy and a Tool for Nanotechnology. Review of Scientific Instruments 2007, 78, 013705. (23) Wetzelaer, G. A. H.; Kuik, M.; Nicolai, H. T.; Blom, P. W. M. Trap-assisted and Langevin-type Recombination in Organic Light-emitting Diodes. Physical Review B 2011, 83, 165204. 30 (24) Kabra, D.; Lu, L. P.; Song, M. H.; Snaith, H. J.; Friend, R. H. Efficient Single- Layer Polymer Light-Emitting Diodes. Advanced Materials 2010, 22, 3194-3198. (25) Lu, L. P.; Kabra, D.; Friend, R. H. Barium Hydroxide as an Interlayer Between Zinc Oxide and a Luminescent Conjugated Polymer for Light-Emitting Diodes. Advanced Functional Materials 2012, 22, 4165-4171. (26) Lu, L.-P.; Kabra, D.; Johnson, K.; Friend, R. H. Charge-Carrier Balance and Color Purity in Polyfluorene Polymer Blends for Blue Light-Emitting Diodes. Advanced Functional Materials 2012, 22, 144-150. (27) Rihani, A.; Hassine, L.; Fave, J. L.; Bouchriha, H. Study of the Transient EL Slow Rise in Single Layer OLEDs. Organic Electronics 2006, 7, 1-7. (28) Tanase, C.; Meijer, E. J.; Blom, P. W. M.; de Leeuw, D. M. Unification of the Hole Transport in Polymeric Field-Effect Transistors and Light-Emitting Diodes. Physical Review Letters 2003, 91, 216601. (29) Banach, M. J.; Friend, R. H.; Sirringhaus, H. Influence of the Molecular Weight on the Thermotropic Alignment of Thin Liquid Crystalline Polyfluorene Copolymer Films. Macromolecules 2003, 36, 2838-2844. (30) Faria, G. C.; Plivelic, T. S.; Cossiello, R. F.; Souza, A. A.; Atvars, T. D. Z.; Torriani, I. L.; deAzevedo, E. R. A Multitechnique Study of Structure and Dynamics of Polyfluorene Cast Films and the Influence on Their Photoluminescence. The Journal of Physical Chemistry B 2009, 113, 11403-11413. (31) Hukic-Markosian, G.; Basel, T.; Singh, S.; Valy Vardeny, Z.; Li, S.; Laird, D. Study of Photoexcitations in Poly(3-hexylthiophene) for Photovoltaic Applications. Applied Physics Letters 2012, 100, 213903. 31 (32) Bao, Z.; Lovinger, A. J. Soluble Regioregular Polythiophene Derivatives as Semiconducting Materials for Field-Effect Transistors. Chemistry of Materials 1999, 11, 26072612. (33) Watts, B.; Schuettfort, T.; McNeill, C. R. Mapping of Domain Orientation and Molecular Order in Polycrystalline Semiconducting Polymer Films with Soft X-Ray Microscopy. Advanced Functional Materials 2011, 21, 1122-1131. (34) Sun, X.; Liu, Y.; Di, C.-A.; Wen, Y.; Guo, Y.; Zhang, L.; Zhao, Y.; Yu, G. Interfacial Heterogeneity of Surface Energy in Organic Field-Effect Transistors. Advanced Materials 2011, 23, 1009-1014. (35) Wu, S. Polymer Interface and Adhesion, 1 ed.; Taylor & Francis: New York, United States of America, 1982; Vol. 1. (36) Umeda, T.; Kumaki, D.; Tokito, S. Surface-energy-dependent Field-effect Mobilities up to 1 cm2/V s for Polymer Thin-film Transistor. Journal of Applied Physics 2009, 105, 024516. (37) Wang, C.-F.; Su, Y.-C.; Kuo, S.-W.; Huang, C.-F.; Sheen, Y.-C.; Chang, F.-C. Low-Surface-Free-Energy Materials Based on Polybenzoxazines. Angewandte Chemie International Edition 2006, 45, 2248-2251. (38) Drelich, J.; Miller, J. D.; Good, R. J. The Effect of Drop (Bubble) Size on Advancing and Receding Contact Angles for Heterogeneous and Rough Solid Surfaces as Observed with Sessile-Drop and Captive-Bubble Techniques. Journal of Colloid and Interface Science 1996, 179, 37-50. 32 (39) Kim, Y. S.; Lee, Y.; Lee, W.; Park, H.; Han, S.-H.; Lee, S.-H. Effects of Molecular Weight and Polydispersity of Poly (3-hexylthiophene) in Bulk Heterojunction Polymer Solar Cells. Current Applied Physics 2010, 10, 329-332. (40) Gasperini, A.; Sivula, K. Effects of Molecular Weight on Microstructure and Carrier Transport in a Semicrystalline Poly(thieno)thiophene. Macromolecules 2013, 46, 93499358. 33 TABLE OF CONTENTS GRAPHIC 34