Molecular and Supramolecular Parameters Dictating the

10 Feb 2015 - MTA-SZTE “Lendület” Photoelectrochemistry Research Group, Rerrich Square 1, Szeged, H-6720, Hungary. § Department of Technical ...
0 downloads 0 Views 2MB Size
Subscriber access provided by MICHIGAN STATE UNIVERSITY | MSU LIBRARIES

Article

Molecular and Supramolecular Parameters Dictating the Thermoelectric Performance of Conducting Polymers: A Case Study Using Poly(3-alkylthiophene) Balázs Endr#di, Janos Mellar, Zoltán Gingl, Csaba Visy, and Csaba Janáky J. Phys. Chem. C, Just Accepted Manuscript • DOI: 10.1021/acs.jpcc.5b00135 • Publication Date (Web): 10 Feb 2015 Downloaded from http://pubs.acs.org on February 14, 2015

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

The Journal of Physical Chemistry C is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Molecular and Supramolecular Parameters Dictating the Thermoelectric Performance of Conducting Polymers: a Case Study Using Poly(3alkylthiophenes) Balázs Endrődi,a, b János Mellár,c Zoltán Gingl,c Csaba Visy,a and Csaba Janáky*,a, b a

Department of Physical Chemistry and Materials Science, University of Szeged, Rerrich Square 1., Szeged, H-6720, Hungary b

MTA-SZTE „Lendület” Photoelectrochemistry Research Group, Rerrich Square 1,Szeged, H6720, Hungary

c

Department of Technical Informatics, University of Szeged, Árpád Square 2., Szeged, H-6720, Hungary

KEYWORDS: doping, self-assembly, poly(3-alkylthiophene), Seebeck coefficient, alternative energy

ACS Paragon Plus Environment

1

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 28

ABSTRACT.

In this study, we investigated the impact of molecular and supramolecular structure of conducting polymers (CPs) on their thermoelectric properties. As a model system, poly(3alkylthiophene)s (P3ATs) with different side-chain length were prepared through oxidative chemical polymerization, and were recrystallized to a well-ordered lamellar structure, resulting in one dimensional self-assembled nanofibers (evidenced by TEM, XRD, and UV-Vis spectroscopic measurements). Thermoelectric characterization was performed at different doping levels (precisely tuned by doping in the redox reaction with Ag+ and Fe3+ cations) and the highly doped samples exhibited the best performance for all studied polymers. By varying the length of the alkyl side chain, the supramolecular structure and consequently the electronic properties were varied. The highest electrical conductivity was measured for poly(3-butylthiophene), rooted in its densely packed structure. The established structure-property relationships, concerning the monotonous decrease of the electrical conductivity with the alkyl side chain length, highlight the importance of the supramolecular structure (interchain distance in this case). These findings may contribute to the rational design of organic thermoelectric materials.

ACS Paragon Plus Environment

2

Page 3 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Introduction

Among other alternative energy harvesting methods, thermoelectric power generation attracts significant and continuously growing interest. Direct conversion of heat to electricity aims to reduce heat losses in industrial processes, and to recover energy from high temperature exhaust fumes. Long lifetime, offered by the lack of any moving-parts in such devices makes this approach even more attractive.1 Efficiency of thermoelectric materials is represented by their figure of merit (Z, most often used in its dimensionless form as ZT), or by the power factor (P):  =

   

and =

where T is the absolute temperature, S is the Seebeck coefficient, σ is the electrical-, and κ is the thermal conductivity. Rivaling their inorganic counterparts, conducting polymers (CPs) are promising alternatives for low temperature thermoelectric applications.2-5 Their notably high Seebeck coefficient coupled with small thermal conductivity as well as their low cost, induced significant scientific efforts in this area.3,6 Thermoelectric characterization of many different CPs have been performed so far,7 however, values of their thermoelectric figure of merit (ZT) remained inferior compared to that of their best inorganic counterpart. On the other hand, recently reported increased ZT values, (approaching 0.5 in the case of PEDOT-Tosylate and PEDOT-PSS systems) project a promising future for organic thermoelectrics.8,9 A major drawback of CPs in thermoelectrics is their usually low electrical conductivity in their neutral form. This issue can be circumvented by either doping (increasing charge carrier concentration) or by enhancing the charge carrier mobility. The effects of increased charge

ACS Paragon Plus Environment

3

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 28

carrier density (i.e., higher doping level), however, is complex because it affects thermal conductivity and the Seebeck coefficient as well (in an undesired manner). This fact emphasizes the importance of the charge carrier mobility, which is strongly correlated to both the regioregularity of the CP, and the increased probability of interchain electron hopping. Studies on traditional CPs, most importantly on P3HT, led to the recognition that the monomer’s structure can have a decisive impact on the charge carrier mobility in the polymer. Unlike for other, first (and some second) generation CPs, where the hole mobility was mostly below 10-2 cm2 V-1s-1, this value could be either even one magnitude larger in case of P3HT.10, 11 This large improvement could be rooted in at least two factors: (i) self-ordering effect of the alkyl side-chains (ii) high regioregularity of properly synthesized P3HT. Elaborating on these results, a new era, the rational design of new monomers has been emerged within the organic electronics (more precisely the OFET) community.12 By including fused aromatic rings (e.g., pyrrolopyrrole) into the polymer backbone, extended rigidity of the polymer (compared to P3HT) leads to further enhanced charge carrier mobility. Thus, using bulky monomers, such as diketopyrrolopyrrole or cyclopentadithiophene, several orders of magnitude higher hole mobility value was reached.13, 14 Beyond the above discussed effect of the monomer molecule, precise control of the supramolecular and morphological features of the polymer is also of importance in optimizing the thermoelectric performance. Different strategies have been exploited to form CP structures with enhanced electrical conductivity recently.15-17 These methods aim the formation of ordered CP structures, in which the high electrical conductivity is ensured by 1D electron pathways. Electrospinning is a general route to prepare polymer (micro)fibers. Note however, that application of such high voltage requires complex instrumentation, making the synthesis

ACS Paragon Plus Environment

4

Page 5 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

expensive, thus diminishing the benefits of CPs versus inorganic semiconductors. Another possibility is the template synthesis of CP nanofibers. Both hard and soft templates can be employed in this manner. Infiltration of CP solutions into solid, porous templates (e.g., nanoporous Al2O3) leads to parallel standing CP fibers.18 Prominence of such structures in analytical applications is obvious due to the multiplied surface area. The lack of interaction between the individual fibers, however, limits the electrical conductivity of such architectures. Application of soft templates offers easy, one-step synthesis of highly conducting CPs. Presence of polyanions (e.g., polystyrene sulfonate (PSS)) during the synthesis leads to highly conductive products, because of the orientation of positively charged polymer chains along the (PSS) macromolecules.16,

17

Detailed investigation on the structure of PEDOT/PSS however,

revealed the presence of PSS aggregates in the formed polymer. Insulating effect of these nonconducting parts can be partially overcome by treating PEDOT/PSS with polar solvents, or acid solutions.19-23 Although most promising thermoelectric results were gathered on PEDOT/PSS systems, all the above facts emphasize the importance of template free, solution phase methods. Despite, that π-stacking interaction between the independent polymer chains has a slight ordering effect, most CPs have globular, disordered structure after synthesis. In the case of alkylsubstituted regioregular polythiophenes however, formation of highly ordered, crystalline nanofibers (NFs) was reported earlier.24-28 The self-assembly was attributed to the interaction between the alkyl-side chains, leading to an interdigitated, “zipper-like” structure. The decreased distance between the polymer chains facilitates interchain electron hopping, resulting enhanced charge carrier mobility.25, 29 Important to notice, that the above described approaches are far from being independent from each other, oppositely they are rather convoluted. For example, introducing an alkyl-chain to the

ACS Paragon Plus Environment

5

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 28

thiophene ring not only affects the molecular structure, but enables the self-assembly of the polymer to form nanofibers. Therefore, when designing thermoelectric materials, it is crucial to understand the individual role and contribution of the different parameters to the thermoelectric properties. Elaborating on our recent communication on poly(3-hexylthiophene) nanofibers,29 we present the systematic investigation on the effect of both molecular and supramolecular features as well as that of doping on the thermoelectric properties of poly(3-alkylthiophenes) (P3ATs) with different side-chain length. Our attempt to deconvolute the effect of various parameters affecting the thermoelectric performance may contribute to the rational design of new CP-materials for thermoelectric applications.

Experimental

Oxidative chemical polymerization was employed to obtain polythiophene and P3ATs with different side-chain length (A = M-methyl, B-butyl, H-hexyl, O-octyl, D-dodecyl).30 3alkylthiophene and FeCl3 solutions (in chloroform) were mixed, having final reagent concentrations of 0.1 and 0.25 M, respectively. The reaction mixtures were kept in closed vessels for 6 hours on ice bath, under continuous stirring. The precipitated polymers were filtered on 1215 µm pore sized filter paper, and washed repeatedly with ethanol to remove iron-traces. The final products were dried in air, under infrared lamp. Whisker method was used to form nanofibers from the polymer powders.25 First, a larger molecular weight fraction of the respective P3AT was dissolved by tetrahydrofuran (THF). Due to the insolubility of polythiophene and P3MT, the recrystallization, and hence further studies

ACS Paragon Plus Environment

6

Page 7 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

were carried out only on the other four P3ATs. After evaporating THF, the polymers were redissolved in a mixed (9:1 ratio) anisole/chloroform solvent (final concentrations were 2.0 g dm-3 in the case of P3BT, while 2.5 g dm-3 in the case of other P3ATs). Finally, the solutions were heated up to 70 °C and then cooled down to room temperature on ice bath. P3AT nanonets were drop-casted from P3AT solutions, on plastic substrates (2 cm × 2.5 cm), previously patterned with 4 gold electrodes for 4 point probe electrical measurements. Neutral polymer films were oxidized by dipping them into AgClO4 or FeCl3 solutions (in nitrobenzene). Concentration of the oxidant was varied between 0.1 mmol dm-3 and 25 mmol dm-3 (saturated). At a selected oxidant concentration, doping level of P3ATs was controlled by the dipping time. Since the best thermoelectric results were gathered on the most highly doped polymers,29 all the characterization data (XRD, FT-IR, SEM) are presented for these samples in this paper. UV–vis–NIR spectra were recorded using an Agilent 8453 UV–visible diode array spectrophotometer in the range of 200–1100 nm. Fourier transform infrared spectra (FT-IR) were recorded using a Bio-Rad Digilab Division FTS-65A/896 instrument, equipped with a Harrick's Meridian® SplitPea single reflection diamond attenuated total reflectance (ATR) accessory. All spectra were recorded between 400 and 4000 cm-1, using 4 cm-1 optical resolution, averaging 512 interferograms. Raman spectroscopic studies were carried out on a DXR Raman Microscope with a red laser (λ=780 nm), operating at 1 mW laser power. To record transmission electron microscopic (TEM) images, P3AT solutions were drop-casted on copper mesh TEM grids covered by carbon film. Ag-doping occurred in situ on the P3AT

ACS Paragon Plus Environment

7

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 28

coated grids. A FEI Tecnai G2 20 X-Twin instrument was used, operating at an acceleration voltage of 200 kV. Scanning electron microscopic (SEM) images were recorded using a Hitachi S-4700 field emission scanning electron microscope (coupled with a Röntec EDX detector), operating at 10 kV acceleration voltage. XRD patterns were recorded between 2Θ = 3-80° at 1° minute-1 scan rate on a Rigaku Miniflex II instrument, operating with a Cu Kα,1 radiation source (λ= 0.1541 nm). A custom designed setup (Figure S1) was used to measure the Seebeck coefficient. Electrical conductivity was determined by the 4 point method, using a Keithley 2400 type general purpose source meter.

ACS Paragon Plus Environment

8

Page 9 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Results and discussion

Recrystallization of P3ATs to nanofibers The recrystallization procedure is accompanied by the color change of the polymer solutions from orange to ruby-purple. On the UV-Visible spectra (Fig. 1.) significant changes could be detected when the solvent was changed from pure THF to 9/1 ratio anisole/chloroform mixture. The first important difference is the red shift of the absorption maximum in all cases. Such decrease in the transition energy can be related to the extension of the conjugation length. Furthermore, appearance of the fine vibronic structure (π-π* absorption bands at higher wavelengths of about 510, 550, and 600 nm) is a clear indication of the nanofiber formation.26-28 Figure 1. Supramolecular structure of the bulk and nanofibrillar P3AT samples was investigated by XRD, after drop-casting them on glass substrates. As seen in Fig. 2A, bulk P3ATs show both sharp reflections and a notable, broad, hill-type reflection too. While the first reflection (100) can be attributed to the lamellar ordering of the polymer chains, the latter is typically observed for amorphous, regiorandom P3ATs, obtained by simple oxidative chemical polymerization.25 Contrarily, the presence of the crystalline phase is almost exclusive after the recrystallization: beyond the nearly complete disappearance of the hill type reflection, this is further supported by the appearance of the (200) and (300) reflections (Fig. 2B). This alteration can be attributed to a highly-ordered structure, in which the individual P3AT chains are stacked together due to the interaction of the overlapping aromatic rings (π-stacking) and the “zipper-like” connection of the alkyl side chains.25, 31, 32 Figure 2.

ACS Paragon Plus Environment

9

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 28

Direct evidence on the nanofiber formation was furnished by TEM measurements. As seen in Fig. 3A, the recrystallization process results in a randomly oriented network of CP nanofibers. Interestingly, the average diameter of the individual nanofibers is almost identical for all P3ATs (around 18 nm). At the same time, the length of the nanofibers is highly dependent on the alkyl side-chain: while it is about 1 µm for P3BT and P3HT, for polymers with longer side side-chains it decreases drastically (in the case of P3DT larger aggregates of very short nanofibers were detected) (Fig. 3B). Figure 3. Effect of silver perchlorate doping To tune the oxidation level of the P3AT samples, the nanonets were gradually doped in their reaction with silver cations: 3  +  = 3   +  Raman spectra (Fig. 4A for P3BT and Fig. S2 for all other polymers) of the neutral and the doped P3ATs nanonets are consistent with that of other P3ATs studied earlier.33 For example, as shown for P3BT, changes in the relative intensities of the Cα-S-Cα’ deformation (721 cm-1) and Cα-Cα’ stretching vibration (1212 cm-1, characteristic to the head-to-tail structure), and the shift of the band at 1446 cm-1 (related to the Cα=Cβ stretching) to a significantly lower wavenumber (1407 cm-1) confirm the formation of a heavily doped form of the polymer. This fact was further supported by FTIR measurements. As the doping reaction proceeds, characteristic tenor of the spectrum ceases (Fig. 4B for P3BT and Fig. S3 for all other polymers), indicating the formation of the oxidized, conducting form of P3ATs.34 Figure 4.

ACS Paragon Plus Environment

10

Page 11 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Modifications in the crystal structure were monitored by XRD. Upon oxidation, XRD pattern of the P3ATs changed significantly. The sharp reflection at lower 2Θ values – corresponding to the distance between the interdigitated chains – gradually shifted to lower values (larger distances) (Fig. 5 and Table S1). This phenomenon is a side-effect of the oxidation: as the polymer chains get positively charged, Coulomb repulsion leads to larger interchain distances. Furthermore, appearance of the new reflections at 2Θ = 27.7, 32.0, 38.1, 44.4, and 46.1° confirms the formation of metallic silver particles on the P3AT nanonets. Although XRD measurements are not quantitative, the high relative intensity of the latter reflections indicates the formation of large amount of silver, and indirectly the high doping level of the polymers.

Figure 5.

SEM images together with the XRD data proved the formation of silver nanoparticles in the doping reaction. Using concentrated oxidant solution, doping level of the polymers was about 0.4 in all cases (as derived from elemental ratio of S to Ag, measured by EDX measurements). This value indicates the presence of highly doped form of polymers and the formation of large amount of silver. According to the SEM images, the Ag nanoparticles form large aggregates, with a size of about 420 ± 44 nm in case of P3BT and about 320 ± 50 nm in case of the other P3ATs (Fig 6 and Fig. S4). Careful analysis of TEM images (e.g., Fig. S5) also confirmed the above average particle sizes. Clearly, the deposited Ag particles do not form a percolation pathway through the bulk nanonet structure.35 This observation indicates an intimate electronic contact among the individual P3AT nanofibers in the self-assembled network. Figure 6.

ACS Paragon Plus Environment

11

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 28

Thermoelectric properties of silver doped P3AT nanonets Changes in the thermoelectric properties of P3ATs upon the doping reaction were monitored at different oxidant concentrations. Most importantly, the oxidant concentration had a large impact on both the reaction rate and on the maximum doping level, reached in the reaction. Higher electrical conductivities, registered with higher concentrations indicate higher doping level (see data in Fig. S6 for P3BT). Similarly to P3HT nanofibers29 the best thermoelectric performance was achieved at the highest oxidant concentration for all polymers in this study. As presented in Fig. 7A-D, results obtained for all the polymers follow a similar pattern: electrical conductivity increases with the doping level, while the Seebeck coefficient ceases to a minimum value. This trend is entirely consistent to earlier reports on P3HT. 29, 36 The maximum power factor, registered for the studied P3ATs however, was highly dependent on the length of the alkyl side-chain. While the best power factor was 10 µWm-1K-2 in the case of P3BT, it was about 40 times lower for P3DT. Figure 7. Differences in the thermoelectric performance can be further highlighted by comparing and contrasting the electrical conductivity and the Seebeck coefficient values for the best performing P3AT samples. Similarly to earlier findings on different CPs, the value of the Seebeck coefficient is totally independent from the length of the alkyl side chain (Fig. 8A).37-39 This means, that the large difference in the power factors is mostly dictated by the electrical conductivity values: as the length of the alkyl side chain increased, the reasonable electrical conductivity measured for P3BT (35 S cm-1) decreased to 0.77 S cm-1 (about 45 times lower value) in the case of P3DT. Figure 8.

ACS Paragon Plus Environment

12

Page 13 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

To link the above observations with structural attributes, and thus obtain structure-property relationships- electrical conductivity values were plotted as a reciprocal function of the lamellar interchain distance of the highly doped polymers (derived from XRD data, see Table S1) in the self-assembled architecture (Fig. 8B). The more or less linear trend indicates that besides the interchain electron hopping facilitated by π-stack interactions,40 the closely packed structure (observed for P3ATs with shorter side-chains) is also a major contributor to the electrical conductivity.

Thermoelectric properties of FeCl3 doped P3ATs To evaluate the role of the dopants’ nature

41-42

and the possible contribution of the in situ

formed silver particles to the enhanced electrical conductivity, doping was also performed using FeCl3, an oxidant, forming no metal precipitate on the polymer upon the redox reaction (Fe3+  Fe2+). As coined from Raman and FTIR measurements, there is no significant difference between the effects of the two substantially different dopants. Similarly to the above discussed case (using AgClO4 as oxidant), both vibrational spectroscopic techniques confirmed the presence of heavily doped P3ATs (Fig. 9A,B). As it was shown above, XRD pattern of P3BT changes strongly upon oxidation (Fig. 2B and Fig. 5). Beyond the obvious disappearance of the Ag-related sharp reflections, the shift of the reflection at lower 2Θ values upon oxidation is almost identical in the Fe3+ case (Fig. 9C). This fact reveals that the larger distance between the polymer chains in the oxidized state originates only from the charge repulsion, and not from the presence of Ag particles. Figure 9.

ACS Paragon Plus Environment

13

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 28

As shown in Fig. 9D, changes in both the electrical conductivity and Seebeck are very similar to those observed in the case of AgClO4 doping (Fig. 7). Although the kinetics of the reaction differs using a different oxidant, both the Seebeck coefficient and the electrical conductivity (and hence the power factor) values are almost identical for the heavily doped polymers. Since no nanoparticle formation occurred during Fe3+ doping, the identical value of the power factor in the two cases suggests that the in situ formed Ag particles do not contribute to the thermoelectric performance.

Conclusions As a result of our systematic study on the influence of both molecular/supramolecular structure and doping on the thermoelectric properties of P3ATs we may conclude that the observed effects of the above factors are complex and convoluted. Both shorter alkyl side chains (molecular structure) and the ordered nanofibrillar structure (supramolecular structure) manifested in an enhanced electrical conductivity. At the same time, the value of the Seebeck coefficient was independent from these parameters. As a result, the best thermoelectric power factor was registered for P3BT, the polymer with the shortest alkyl side chain among the studied P3ATs. Selecting the most appropriate parameters from our studies, the highest achieved thermoelectric power factor (10 µWK-2m-1 for heavily doped P3BT nanofibers) showed a notable improvement compared to the values reported earlier for this class of polymers. Effect of the dopant’s nature was also investigated, by comparing the thermoelectric power factor of Fe3+ and Ag+ doped P3BT nanonets. Even though in the latter case the formation of silver particles was evidenced, the best power factor reached with the two very different oxidants

ACS Paragon Plus Environment

14

Page 15 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

is almost identical, indicating that the enhanced electrical conductivity is rooted in molecular changes upon doping. Finally we note that by evaluating the individual (yet convoluted) contribution of the above parameters on the thermoelectric performance, new avenues may be paved for the rational design of CP thermoelectrics. Supramolecular and morphological engineering of CPs with intrinsically high charge carrier mobility (e.g., CPs containing diketopyrrolopyrrole moieties) may contribute to unprecedently high ZT values, and such studies are in progress in our laboratory.

ACS Paragon Plus Environment

15

The Journal of Physical Chemistry

1.0 0.8 0.6 0.4 0.2 0.0

Norm. Absorbance

FIGURES AND CAPTIONS

P3DT P3OT P3HT P3BT 400

500

600

700

800

900

λ / nm

Figure 1. UV-Visible spectra of P3AT solutions. Spectra recorded in THF and anisole/chloroform solutions (after cooling) are overlaid for each P3AT.

P3DT P3OT P3HT P3BT

0

10

20

30

40

B

P3DT P3OT P3HT P3BT

Normalized intensity

A Normalized intensity

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 28

0

10

2 Θ / Deg

20

30

40

2 Θ / Deg

Figure 2. XRD pattern of (A) bulk and (B) nanofibrillar P3ATs.

ACS Paragon Plus Environment

16

B

1200

35

Length Diameter

Diameter / nm

30

1000

25

800

20

600

15

400

10

200

5

0 4

6

8

10

12

Alkyl side-chain length

Figure 3. (A) TEM image of P3BT nanofiber network (B) Average diameter and length of the formed nanofibers for the different P3ATs.

B

1.0

P3BT NF P3BT NF + Ag

0.8 0.6 0.4 0.2 0.0 1600

1400

1200

1000

800

600

P3BT NF + Ag P3BT NF

Normalized absorbance

A Normalized intensity

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Length / nm

Page 17 of 28

3500

3000

Raman shift / cm-1

2500

1500

1000

500

Wavenumber / cm-1

Figure 4. (A) Raman- and (B) absorption FTIR-spectra of the neutral and the doped P3BT nanofiber network.

ACS Paragon Plus Environment

17

The Journal of Physical Chemistry

P3DT/Ag P3OT/Ag P3HT/Ag P3BT/Ag

Normalized intensity

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 28

0

10

20

30

40

50

2 Θ / Deg

Figure 5. XRD pattern of the doped P3ATs.

Figure 6. SEM images the Ag-doped P3ATs (A: P3BT, B: P3HT, C:P3OT, D:P3DT).

ACS Paragon Plus Environment

18

B

A 100

10 8

100

65

55

S / µV K-1

60

σ / S cm-1

10

70

P / µW m-1K-2

σ S P

S / µV K-1

σ / S cm-1

10

10

6 4

60 2

σ S P

1

50

1

0

200

400

600

800

0

50

100

t/s

150

200

250

300

350

400

t/s

C

D

0.5 10

10

60

120

0.4

1

σ S P

0.1

σ S P

0.01

52

100 0.2 80

60

P / µW m-1 K-2

S / µV K-1

56

1

σ / S cm-1

1

P / µW m-1 K-2

S / µV K-1

0.3

σ / S cm-1

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

P / µW m-1 K-2

Page 19 of 28

0.1

0.1

0.1

1E-3 0

100

200

300

400

0

100

t/s

200

300

400

t/s

Figure 7. Doping level dependent thermoelectric properties of different P3ATs (A: P3BT, B: P3HT, C: P3OT, D: P3DT) with saturated AgClO4 (c=25 mM).

ACS Paragon Plus Environment

19

The Journal of Physical Chemistry

A

100

B

35

40

80

30

30

40

15

σ / S cm-1

60 20

S / µV K-1

σ / S cm-1

25

10

20

10

20 5 0

P3BT

P3HT

P3OT

P3DT

0

0

0.3

0.4

0.5

d

-1 doped

0.6

0.7

-1

/ nm

Figure 8. (A) Electrical conductivity and Seebeck coefficient values for highly doped P3ATs (B) Maximum electrical conductivity values in function of the reciprocal chain distance of the highly doped polymers (derived from XRD measurements).

B

P3BT NF + FeCl3

1.0

P3BT NF + FeCl3 P3BT NF

Normalized absorbance

0.8 0.6 0.4 0.2 0.0 1600

1400

1200

1000

800

600

3500

3000

Raman shift / cm-1

C

2500

1500

1000

500

Wavenumber / cm-1

P3BT/Ag P3BT-FeCl3

D 100

10 200

σ / S cm

-1

10

1

1

σ S P

0.1

150

100

0.01

0.1

0.01

P / µW m-1 K-2

Normalized intensity

P3BT NF

S / µV K-1

A

Normalized intensity

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 28

1E-3 50

1E-3

0

10

20

30

40

50

0

100

200

300

400

500

600

700

t/s

2 Θ / Deg

ACS Paragon Plus Environment

20

Page 21 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure 9 (A) Raman and (B) FTIR spectra (C) XRD pattern of the neutral and the FeCl3 doped P3BT nanofiber network (D). Doping level dependent thermoelectric properties of P3BT using 50 mM FeCl3 as oxidant. Supporting Information. Electronic Supporting Information (ESI) available: Scheme of the thermoelectric setup, Raman, FT-IR, TEM and thermoelectric results. This material is available free of charge via the Internet at http://pubs.acs.org. Corresponding author Csaba Janáky* Department of Physical Chemistry and Materials Science, University of Szeged, Rerrich Square 1., Szeged, H-6720, Hungary E-mail: [email protected] Fax: +36 62 546-482 Tel: +36 62 546-393 Author Contributions The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript. ACKNOWLEDGMENT Funding support of the Hungarian Academy of Sciences through its Momentum Excellence Grant is gratefully acknowledged.

ABBREVIATIONS

ACS Paragon Plus Environment

21

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 28

P3HT, poly(3-hexylthiophene); SEM, Scanning electron microscopy; CP, conducting polymer; P3AT, poly(3-alkylthiophene); XRD, X-ray diffraction; TEM, Transmission electron microscopy; FT-IR, Fourier-transform infrared spectroscopy; PT, polythiophene; P3MT, poly(3methylthiophene; P3BT, poly(3-butylthiophene; P3HT, poly(3-hexylthiophene; P3OT, poly(3octylthiophene; P3DT, poly(3-dodecylthiophene;

ACS Paragon Plus Environment

22

Page 23 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

REFERENCES (1) DiSalvo, F. J. Thermoelectric Cooling and Power Generation. Science 1999, 285, 703-706. (2) Zhang, Q.; Sun, Y.; Xu, W.; Zhu, D. Organic Thermoelectric Materials: Emerging Green Energy Materials Converting Heat to Electricity Directly and Efficiently. Adv. Mater. 2014, 26, 6829-6851. (3) He, M.; Qiu, F.; Lin, Z. Towards High-Performance Polymer-Based Thermoelectric Materials. Energy Environ. Sci. 2013, 6, 1352-1361. (4) Chen, Y.; Zhao, Y.; Liang, Z. Solution Processed Organic Thermoelectrics: Towards Flexible Thermoelectric Modules. Energy Environ. Sci. 2015, DOI: 10.1039/c4ee03297g (5) McGrail, B. T.; Sehirlioglu, A.; Pentzer, E. Polymer Composites for Thermoelectric Applications. Angew. Chem. Int. Edit. 2014, DOI: 10.1002/anie.201408431 (6) Dubey, N.; Leclerc, M. Conducting Polymers: Efficient Thermoelectric Materials. J. Polym. Sci. Pol. Phys. 2011, 49, 467-475. (7) Poehler, T. O.; Katz, H. E. Prospects for Polymer-Based Thermoelectrics: State of the Art and Theoretical Analysis. Energy Environ. Sci. 2012, 5, 8110-8115. (8) Bubnova, O.; Khan, Z. U.; Malti, A.; Braun, S.; Fahlman, M.; Berggren, M.; Crispin, X. Optimization of the Thermoelectric Figure of Merit in the Conducting Polymer Poly(3,4Ethylenedioxythiophene). Nat. Mater. 2011, 10, 429-433. (9) Kim, G.; Shao, L.; Zhang, K.; Pipe, K. P. Engineered Doping of Organic Semiconductors for Enhanced Thermoelectric Efficiency. Nat. Mater. 2013, 12, 719-723. (10) Heeger, A. J. Semiconducting Polymers: The Third Generation. Chem. Soc. Rev. 2010, 39, 2354-2371. (11) Horowitz, G. Organic Field-Effect Transistors. Adv. Mater. 1998, 10, 365-377. (12) Sirringhaus, H. 25th Anniversary Article: Organic Field-Effect Transistors: The Path Beyond Amorphous Silicon. Adv. Mater. 2014, 26, 1319-1335. (13) Tseng, H.; Phan, H.; Luo, C.; Wang, M.; Perez, L. A.; Patel, S. N.; Ying, L.; Kramer, E. J.; Nguyen, T.; Bazan, G. C. et al. High-Mobility Field-Effect Transistors Fabricated with Macroscopic Aligned Semiconducting Polymers. Adv. Mater. 2014, 26, 2993-2998. (14) Yun, H.; Lee, G. B.; Chung, D. S.; Kim, Y.; Kwon, S. Novel Diketopyrroloppyrrole Random Copolymers: High Charge-Carrier Mobility from Environmentally Benign Processing. Adv. Mater. 2014, 26, 6612-6616.

ACS Paragon Plus Environment

23

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 28

(15) Long, Y.; Li, M.; Gu, C.; Wan, M.; Duvail, J.; Liu, Z.; Fan, Z. Recent Advances in Synthesis, Physical Properties and Applications of Conducting Polymer Nanotubes and Nanofibers. Prog. Polym. Sci. 2011, 36, 1415-1442. (16) Lang, U.; Müller, E.; Naujoks, N.; Dual, J. Microscopical Investigations of PEDOT:PSS Thin Films. Adv. Funct. Mater. 2009, 19, 1215-1220. (17) Nardes, A. M.; Kemerink, M.; Janssen, R. A. J.; Bastiaansen, J. A. M.; Kiggen, N. M. M.; Langeveld, B. M. W.; van Breemen,A. J. J. M.; de Kok, M. M. Microscopic Understanding of the Anisotropic Conductivity of PEDOT:PSS Thin Films. Adv. Mater. 2007, 19, 11961200. (18) Martin, C. R. Template Synthesis of Electronically Conductive Polymer Nanostructures. Acc. Chem. Res. 1995, 28, 61-68. (19) Alemu, D.; Wei, H.; Ho, K.; Chu, C. Highly Conductive PEDOT:PSS Electrode by Simple Film Treatment with Methanol for ITO-Free Polymer Solar Cells. Energy Environ. Sci. 2012, 5, 9662-9671. (20) Kim, J. Y.; Jung, J. H.; Lee, D. E.; Joo, J. Enhancement of Electrical Conductivity of Poly(3,4-Ethylenedioxythiophene)/Poly(4-Styrenesulfonate) by a Change of Solvents. Synth. Met. 2002, 126, 311-316. (21) Kim, Y. H.; Sachse, C.; Machala, M. L.; May, C.; Müller-Meskamp, L.; Leo, K. Highly Conductive PEDOT:PSS Electrode with Optimized Solvent and Thermal Post-Treatment for ITO-Free Organic Solar Cells. Adv. Funct. Mater. 2011, 21, 1076-1081. (22) Mengistie, D. A.; Ibrahem, M. A.; Wang, P.; Chu, C. Highly Conductive PEDOT:PSS Treated with Formic Acid for ITO-Free Polymer Solar Cells. ACS Appl. Mater. Interfaces 2014, 6, 2292-2299. (23) Ouyang, J. Solution-Processed PEDOT:PSS Films with Conductivities as Indium Tin Oxide through a Treatment with Mild and Weak Organic Acids. ACS Appl. Mater. Interfaces 2013, 5, 13082-13088. (24) Ihn, K. J.; Moulton, J.; Smith, P. Whiskers of Poly(3-Alkylthiophene)S. J. Polym. Sci. Pol. Phys. 1993, 31, 735-742. (25) Samitsu, S.; Shimomura, T.; Heike, S.; Hashizume, T.; Ito, K. Effective Production of Poly(3-Alkylthiophene) Nanofibers by Means of Whisker Method using Anisole Solvent: Structural, Optical, and Electrical Properties. Macromolecules 2008, 41, 8000-8010. (26) Oosterbaan, W. D.; Vrindts, V.; Berson, S.; Guillerez, S.; Douheret, O.; Ruttens, B.; D'Haen, J.; Adriaensens, P.; Manca, J.; Lutsen, L. et al. Efficient Formation, Isolation and Characterization of Poly(3-Alkylthiophene) Nanofibres: Probing Order as a Function of Side-Chain Length. J. Mater. Chem. 2009, 19, 5424-5435.

ACS Paragon Plus Environment

24

Page 25 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(27) Kiriy, N.; Jähne, E.; Adler, H.; Schneider, M.; Kiriy, A.; Gorodyska, G.; Minko, S.; Jehnichen, D.; Simon, P.; Fokin, A. A. et al. One-Dimensional Aggregation of Regioregular Polyalkylthiophenes. Nano Lett. 2003, 3, 707-712. (28) Samitsu, S.; Shimomura, T.; Ito, K. Nanofiber Preparation by Whisker Method using Solvent-Soluble Conducting Polymers. Thin Solid Films 2008, 516, 2478-2486. (29) Endrodi, B.; Mellar, J.; Gingl, Z.; Visy, C.; Janaky, C. Reasons Behind the Improved Thermoelectric Properties of Poly(3-Hexylthiophene) Nanofiber Networks. RSC Adv. 2014, 4, 55328-55333. (30) Järvinen, H.; Lahtinen, L.; Näsman, J.; Hormi, O.; Tammi, A.-L. A New Method to Prepare 3-Octylthiophene and Poly-(3-Octylthiophene). Synth. Met. 1995, 69, 299-300. (31) Coakley, K. M.; Srinivasan, B. S.; Ziebarth, J. M.; Goh, C.; Liu, Y.; McGehee, M. D. Enhanced Hole Mobility in Regioregular Polythiophene Infiltrated in Straight Nanopores. Adv. Funct. Mater. 2005, 15, 1927-1932. (32) Lu, G.; Bu, L.; Li, S.; Yang, X. Bulk Interpenetration Network of Thermoelectric Polymer in Insulating Supporting Matrix. Adv. Mater. 2014, 26, 2359-2364. (33) Baibarac, M.; Lapkowski, M.; Pron, A.; Lefrant, S.; Baltog, I. SERS Spectra of Poly(3Hexylthiophene) in Oxidized and Unoxidized States. J. Raman Spectrosc. 1998, 29, 825832. (34) Pintér, E.; Fekete, Z. A.; Berkesi, O.; Makra, P.; Patzkó, Á; Visy, C. Characterization of Poly(3-Octylthiophene)/Silver Nanocomposites Prepared by Solution Doping. J. Phys. Chem. C 2007, 111, 11872-11878. (35) Roussel, F.; Chen Yu King, R.; Kuriakose, M.; Depriester, M.; Hadj-Sahraoui, A.; Gors, C.; Addad, A.; Brun, J. Electrical and Thermal Transport Properties of Polyaniline/Silver Composites and their use as Thermoelectric Materials. Synth. Met. 2015, 199, 196-204. (36) Xuan, Y.; Liu, X.; Desbief, S.; Leclère, P.; Fahlman, M.; Lazzaroni, R.; Berggren, M.; Cornil, J.; Emin, D.; Crispin, X. Thermoelectric Properties of Conducting Polymers: The Case of Poly(3-Hexylthiophene). Phys. Rev. B 2010, 82, 115454. (37) Russ, B.; Robb, M. J.; Brunetti, F. G.; Miller, P. L.; Perry, E. E.; Patel, S. N.; Ho, V.; Chang, W. B.; Urban, J. J.; Chabinyc, M. L. et al. Power Factor Enhancement in SolutionProcessed Organic N-Type Thermoelectrics through Molecular Design. Adv. Mater. 2014, 26, 3473-3477. (38) Mai, C. K.; Schlitz, R. A.; Su, G. M.; Spitzer, D., Wang, X.; Fronk, S. L.; Cahill, D. G.; Chabinyc, M. L.; Bazan, G. C. Side-Chain Effects on the Conductivity, Morphology, and Thermoelectric Properties of Self-Doped Narrow-Band-Gap Conjugated Polyelectrolytes J. Am. Chem. Soc. 2014, 136, 13478-13481.

ACS Paragon Plus Environment

25

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 26 of 28

(39) Mai, C.; Zhou, H.; Zhang, Y.; Henson, Z. B.; Nguyen, T.; Heeger, A. J.; Bazan, G. C. Facile Doping of Anionic Narrow-Band-Gap Conjugated Polyelectrolytes during Dialysis. Angew. Chem. Int. Edit. 2013, 52, 12874-12878. (40) Nielsen, C. B.; McCulloch, I. Recent Advances in Transistor Performance of Polythiophenes. Prog. Polym. Sci. 2013, 38, 2053-2069. (41) Glaudell, A. M.; Cochran, J. E.; Patel, S. N.; Chabinyc, M. L. Impact of the Doping Method on Conductivity and Thermopower in Semiconducting Polythiophenes. Adv. Energy Mater. 2014, DOI: 10.1002/aenm.201401072 (42) Zhang, Q.; Sun, Y.; Xu, W.; Zhu, D. Thermoelectric energy from flexible P3HT films doped with a ferric salt of triflimide anions. Energy Environ. Sci. 2012, 5, 9639-9644.

ACS Paragon Plus Environment

26

Page 27 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

TABLE OF CONTENTS

ACS Paragon Plus Environment

27

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Proposed cover art graph 1058x1058mm (72 x 72 DPI)

ACS Paragon Plus Environment

Page 28 of 28