26.06.2014 Views

Film formation properties of inkjet printed poly ... - ResearchGate

Film formation properties of inkjet printed poly ... - ResearchGate

Film formation properties of inkjet printed poly ... - ResearchGate

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Thin Solid <strong>Film</strong>s 519 (2011) 3695–3702<br />

Contents lists available at ScienceDirect<br />

Thin Solid <strong>Film</strong>s<br />

journal homepage: www.elsevier.com/locate/tsf<br />

<strong>Film</strong> <strong>formation</strong> <strong>properties</strong> <strong>of</strong> <strong>inkjet</strong> <strong>printed</strong><br />

<strong>poly</strong>(phenylene-ethynylene)-<strong>poly</strong>(phenylene-vinylene)s<br />

Anke Teichler a,b , Rebecca Eckardt a,b , Christian Friebe a , Jolke Perelaer a,b , Ulrich S. Schubert a,b, ⁎<br />

a Laboratory <strong>of</strong> Organic and Macromolecular Chemistry, Friedrich-Schiller-University Jena, Humboldtstraβe 10, 07743 Jena, Germany<br />

b Dutch Polymer Institute (DPI), P.O. Box 513, 5600 MB Eindhoven, The Netherlands<br />

article<br />

info<br />

abstract<br />

Article history:<br />

Received 16 June 2010<br />

Received in revised form 18 January 2011<br />

Accepted 18 January 2011<br />

Available online 27 January 2011<br />

Keywords:<br />

Inkjet printing<br />

Poly(phenylene-ethynylene)<br />

Poly(phenylene-vinylene<br />

Thin films<br />

Combinatorial screening<br />

Inkjet printing was used here as a precise and fast dispensing technique to prepare thin-film libraries <strong>of</strong> a<br />

<strong>poly</strong>-(phenylene-ethynylene)-<strong>poly</strong>(phenylene-vinylene)s co<strong>poly</strong>mer. The films were prepared with a<br />

systematic variation <strong>of</strong> the ink composition, the dot spacing and the substrate temperature. Homogeneous<br />

films with a thickness <strong>of</strong> 100 nm were obtained when <strong>printed</strong> at room temperature and from a solvent<br />

mixture <strong>of</strong> toluene and ortho-dichlorobenzene in a volume ratio <strong>of</strong> 90/10. This approach can be used for<br />

optoelectronic applications, where the layer homogeneity is extremely important but where the ink<br />

compositions may vary per device, as well as the exact layer thickness. Our approach can be applied for the<br />

preparation <strong>of</strong> films by <strong>inkjet</strong> printing for any other (<strong>poly</strong>mer) ink solution and represents a fast and efficient<br />

screening <strong>of</strong> the parameters to obtain homogeneous films with a precise thickness.<br />

© 2011 Elsevier B.V. All rights reserved.<br />

1. Introduction<br />

Currently, there is a growing interest in conjugated <strong>poly</strong>mers for<br />

usage in optoelectronic devices because <strong>of</strong> their semiconducting<br />

<strong>properties</strong> [1]. Organic semiconductors show advanced processing<br />

<strong>properties</strong> due to solubility in organic solvents and their film<br />

<strong>formation</strong> characteristics. In recent research, conjugated <strong>poly</strong>mers,<br />

like <strong>poly</strong>(phenylene)s or <strong>poly</strong>thiophenes, have been used in optoelectronic<br />

devices, for example in organic light emitting diodes<br />

(OLED) [2,3], organic photovoltaics [4,5] and organic field-effect<br />

transistors (OFET) [6,7]. The possibility <strong>of</strong> altering the material<br />

<strong>properties</strong> and the usage <strong>of</strong> cheap processing methods such as <strong>inkjet</strong><br />

printing classify conjugated <strong>poly</strong>mers as interesting materials for<br />

optoelectronic applications [8,9]. The usage <strong>of</strong> semiconducting<br />

<strong>poly</strong>mers results in smaller, flexible and, most importantly, cheaper<br />

devices.<br />

Poly(phenylene-ethynylene)-<strong>poly</strong>(phenylene-vinylene)s (PPE-<br />

PPV) combine good photophysical characteristics <strong>of</strong> PPVs and charge<br />

transport <strong>properties</strong> <strong>of</strong> PPEs [10]. PPE-PPV derivatives are soluble in<br />

organic solvents due to the alkoxy side chains that are grafted to the<br />

<strong>poly</strong>mer backbone, hence improving the processability <strong>of</strong> the<br />

<strong>poly</strong>mers in solution [11]. The use <strong>of</strong> PPE-PPV co<strong>poly</strong>mers as<br />

electron-donating compounds in organic solar cells with an active<br />

⁎ Corresponding author at: Laboratory <strong>of</strong> Organic and Macromolecular Chemistry,<br />

Friedrich-Schiller-University Jena, Humboldtstraβe 10, 07743 Jena, Germany.<br />

E-mail address: ulrich.schubert@uni-jena.de (U.S. Schubert).<br />

layer consisting <strong>of</strong> PPE-PPV/PCBM (1-[3-(methoxycarbonyl)propyl]-<br />

1-phenyl)-[6,6]C 61 ) yielded device efficiencies up to 1.8% [12–14].<br />

The most common procedure for the preparation <strong>of</strong> the photoactive<br />

layer is by using spin-coating. However, this method is<br />

unfavorable in terms <strong>of</strong> materials efficiency, since most <strong>of</strong> the<br />

material is wasted during processing. Moreover, a combinatorial<br />

workflow is not possible [15]. In contrast, <strong>inkjet</strong> printing is<br />

characterized by reduced production costs due to its low material<br />

waste and controlled deposition <strong>of</strong> materials [16,17]. Additionally,<br />

<strong>inkjet</strong> printing is a non-contact processing technique and does not<br />

require expensive masks, which further reduce processing costs.<br />

The control over pattern <strong>formation</strong> <strong>of</strong> <strong>inkjet</strong> <strong>printed</strong> droplets, lines<br />

and films is required to form homogeneous features upon drying. The<br />

well-known c<strong>of</strong>fee drop effect forms a major problem for the usage <strong>of</strong><br />

the <strong>inkjet</strong> printing technique. An increased evaporation <strong>of</strong> solvent at<br />

the edges <strong>of</strong> the droplet as well as a pinned contact line lead to an<br />

outward flow <strong>of</strong> material and subsequently to accumulation <strong>of</strong><br />

material at the rim <strong>of</strong> the droplet [18,19]. A solution to this problem<br />

was found by using a solvent mixture that contains a minor part <strong>of</strong> a<br />

high-boiling solvent [20]. In addition to investigations on the behavior<br />

<strong>of</strong> single <strong>printed</strong> droplets on a surface, also the <strong>formation</strong> <strong>of</strong> lines was<br />

studied [21]. The control and optimization <strong>of</strong> line morphology is<br />

important for electronic applications, e.g. electrical contacts for<br />

transistors. Soltman et al. classified five different line morphologies<br />

depending on the dot spacing, which is the center-to-center distance<br />

between two adjacent droplets, and their drying time [22].<br />

Inkjet printing <strong>of</strong> thin-film libraries allows a combinatorial workflow<br />

to investigate structure-property relations [23,24]. With this<br />

0040-6090/$ – see front matter © 2011 Elsevier B.V. All rights reserved.<br />

doi:10.1016/j.tsf.2011.01.274


3696 A. Teichler et al. / Thin Solid <strong>Film</strong>s 519 (2011) 3695–3702<br />

approach, the influence <strong>of</strong> ink composition, substrate <strong>properties</strong>, and<br />

different printing parameters to the film <strong>properties</strong> can be studied in a<br />

fast and reproducible way. Tekin et al. investigated a selection <strong>of</strong><br />

parameters that influences the quality <strong>of</strong> <strong>inkjet</strong> <strong>printed</strong> <strong>poly</strong>styrene<br />

films, including solvent combinations, printing speed and dot spacing<br />

[25]. Furthermore, the optical <strong>properties</strong> <strong>of</strong> the <strong>inkjet</strong> <strong>printed</strong> films<br />

consisting <strong>of</strong> various PPE-PPV co<strong>poly</strong>mers were investigated as<br />

function <strong>of</strong> film thickness [26].<br />

In this contribution, we use <strong>inkjet</strong> printing as a tool to prepare<br />

two-dimensional thin films <strong>of</strong> a PPE-PPV co<strong>poly</strong>mer, in order to screen<br />

various parameters <strong>of</strong> the film <strong>formation</strong> <strong>properties</strong>. The influence <strong>of</strong><br />

dot spacing, solvent system, concentration as well as substrate<br />

temperature on the quality <strong>of</strong> <strong>inkjet</strong> <strong>printed</strong> films will be discussed.<br />

Topographical and optical measurements are performed to investigate<br />

the homogeneity, luminescence and reproducibility <strong>of</strong> the<br />

resulting film. This procedure enables a fast and efficient screening<br />

<strong>of</strong> thin-film PPE-PPV libraries.<br />

2. Experimental details<br />

2.1. Materials<br />

The synthetic route <strong>of</strong> the PPE-PPV co<strong>poly</strong>mer (Mn=8,600 g mol −1 ,<br />

Mw=34,000 g mol −1 , PDI=3.95) was described elsewhere [27]. The<br />

<strong>poly</strong>mer was dissolved in the solvent mixtures <strong>of</strong> toluene and orthodichlorobenzene<br />

(o-DCB), and the solutions were filtered before<br />

printing (pore size 1 μm) to prevent nozzle clogging. The solvents<br />

(toluene, o-DCB) were purchased from Biosolve (Valkenswaard, The<br />

Netherlands) and used as delivered. The <strong>poly</strong>mer has a viscosity <strong>of</strong><br />

0.72 and 0.79 mPa s for a concentration <strong>of</strong> 4 mg mL −1 and 8 mg mL −1<br />

in the solvent system toluene/o-DCB 90/10, respectively.<br />

2.2. Preparation <strong>of</strong> substrates<br />

Microscope slides (3×2 in.) from Marienfeld (Lauda-Königsh<strong>of</strong>en,<br />

Germany) were used as substrates. For cleaning, the microscope slides<br />

were ultrasonicated in a soap solution and washed afterwards with<br />

demineralized water to remove the soap [28]. This was followed by<br />

additional ultrasonication steps in acetone and iso-propanol. Finally,<br />

the substrates were rinsed with iso-propanol and dried with an air<br />

flow.<br />

2.3. Instrumentation<br />

The <strong>inkjet</strong> printing experiments were carried out with an<br />

Autodrop system from microdrop Technologies (Norderstedt,<br />

Germany). The printer was equipped with a piezo-based printhead<br />

(dispenser system) with an inner diameter <strong>of</strong> 70 μm. The microscope<br />

slides were placed onto a heatable table which can be moved in x- and<br />

y-direction. A voltage <strong>of</strong> 70 V and a pulse length <strong>of</strong> 35 μs revealed a<br />

stable droplet <strong>formation</strong> for the PPE-PPV in the solvent system<br />

toluene/o-DCB. These settings resulted in an in-flight droplet diameter<br />

<strong>of</strong> approximately 60 μm, which corresponds to a droplet volume <strong>of</strong><br />

113 pL. The printing speed was set to 10 mm s −1 for all experiments.<br />

Surface topography as well as film thicknesses were measured<br />

using an optical interferometric pr<strong>of</strong>iler Wyko NT9100 (Veeco,<br />

Mannheim, Germany). In each film a scratch was made with a scalpel<br />

in a controlled manner. At five different positions <strong>of</strong> the film the depth<br />

<strong>of</strong> the scratch was measured with the optical pr<strong>of</strong>iler. Two separate<br />

films per dot spacing were measured in order to obtain an average<br />

film thickness.<br />

The contact angles from alle solution were measured, using a<br />

Dataphysics OCA 30 (Filderstadt, Germany), and found to be below<br />

10°.<br />

A UV-vis/fluorescence plate reader from Analytik Jena (FLASHScan<br />

530, Jena, Germany) was used to measure the UV-vis absorption and<br />

emission spectra <strong>of</strong> the <strong>printed</strong> films. The films (5×5 mm 2 ) were<br />

<strong>printed</strong> into a microtiter plate pattern, i.e. the films were <strong>printed</strong> with<br />

an interdistance <strong>of</strong> 4 mm to cover the positions <strong>of</strong> the wells. This<br />

allowed the characterization <strong>of</strong> a microtiter plate format with 96 UVvis<br />

spectra within 50 seconds. For measuring the microscope slides an<br />

adapter was used. All measurements were referenced to air. UV-vis<br />

absorption measurements <strong>of</strong> the solutions were carried out with a<br />

Specord 250 (Analytik Jena, Jena, Germany) and UV-vis emission<br />

measurements with a FP 6500 from JASCO Inc. (Easton, USA). Quartz<br />

cuvettes with a diameter <strong>of</strong> 1 cm and filled with pristine solvent as<br />

reference were used. All emission spectra were obtained by excitation<br />

at the wavelength <strong>of</strong> the maximum absorption.<br />

Viscosity measurements were performed using an AMVn viscosimeter<br />

(Anton Paar, Graz, Austria). The dynamic viscosity was<br />

measured at three different angles 30°, 50° and 70° and an average<br />

value was calculated.<br />

3. Results and discussion<br />

A schematic representation <strong>of</strong> the chemical structure <strong>of</strong> the PPE-<br />

PPV co<strong>poly</strong>mer used in the experiments is shown in Fig. 1. The<br />

octyloxy side chains are grafted to the PPE part and octadecyloxy<br />

chains to the PPV part <strong>of</strong> the <strong>poly</strong>mer and improve the solubility<br />

significantly [11].<br />

Many parameters influence the quality and homogeneity <strong>of</strong> <strong>inkjet</strong><br />

<strong>printed</strong> thin films. Jabbour et al. reported <strong>inkjet</strong> printing <strong>of</strong> <strong>poly</strong>(3,4-<br />

ethylenedioxythiophene) <strong>poly</strong>(styrenesulfonate) (PEDOT:PSS) layers<br />

using a standard desktop printer in gray-scale mode. An increased<br />

intensity <strong>of</strong> the “black” ink, which in fact was replaced with a PEDOT:<br />

PSS ink, resulted in an increased PEDOT:PSS layer thickness. The<br />

<strong>printed</strong> conductive <strong>poly</strong>mer layer is used as an anode in OLEDs [29].<br />

This simple but straight-forward approach revealed the influence <strong>of</strong><br />

the film thickness on the device <strong>properties</strong>.<br />

Investigation <strong>of</strong> each single parameter would cost too much time<br />

and, moreover, synergies between parameters remained unnoticed.<br />

Therefore, we use here a two-dimensional and combinatorial approach<br />

to study multiple parameters at the same time: for example, on one axis<br />

the dot spacings are systematically varied, while on the other axis the<br />

concentration <strong>of</strong> the <strong>poly</strong>mer is varied. In the next sections, the effect <strong>of</strong><br />

solute concentration, substrate temperature, and varied ratios <strong>of</strong> the<br />

solvent mixture toluene/ortho-dichlorobenzene (toluene/o-DCB) are<br />

described as function <strong>of</strong> the dot spacing.<br />

3.1. Influence <strong>of</strong> the dot spacing<br />

The <strong>poly</strong>mer was dissolved in the solvent system toluene/o-DCB in<br />

a ratio <strong>of</strong> 90/10 and a concentration <strong>of</strong> 4 mg mL −1 . <strong>Film</strong>s were <strong>inkjet</strong><br />

<strong>printed</strong> with dot spacings between 80 μm and 160 μm. Fig. 2 shows<br />

the influence <strong>of</strong> the dot spacing on the film <strong>formation</strong>. The optical<br />

pr<strong>of</strong>iler images showed that with a dot spacing <strong>of</strong> 80 μm (Fig. 2a)<br />

homogeneously dried film were obtained, but when using a dot<br />

spacing <strong>of</strong> 120 μm a more homogeneous film <strong>formation</strong> was obtained<br />

(Fig. 2b). When further increasing the dot spacing adjacent droplets<br />

were not able to merge into a continuous film. This resulted first in<br />

R 1 O<br />

OR 1<br />

R 2 O<br />

OR 2<br />

Fig. 1. Schematic representation <strong>of</strong> the chemical structure <strong>of</strong> the investigated PPE-PPV<br />

co<strong>poly</strong>mer (R 1 =octyl, R 2 =octadecyl).<br />

n


A. Teichler et al. / Thin Solid <strong>Film</strong>s 519 (2011) 3695–3702<br />

3697<br />

Fig. 3 shows a comparison <strong>of</strong> the solvent mixtures with a dot<br />

spacing <strong>of</strong> 120 μm. For a solvent ratio <strong>of</strong> 90/10 and 95/5 (Fig. 3a, b)<br />

homogeneous films were obtained. In contrast, a solvent ratio <strong>of</strong> 99/1<br />

(Fig. 3c) led to the <strong>formation</strong> <strong>of</strong> inhomogeneous films due to a fast and<br />

irregular drying, which can be ascribed to the low amount <strong>of</strong> o-DCB in<br />

the solvent mixture. Moreover, when using a solvent ratio <strong>of</strong> 99/1 no<br />

homogeneous films could be formed at any dot spacing between 80<br />

and 160 μm. For all three solvent ratios, a dot spacing <strong>of</strong> 80 μm<br />

revealed irregular films since too much material was dispensed, while<br />

a dot spacing <strong>of</strong> 160 μm did not result in continuous films, since the<br />

droplets were not able to merge. Therefore, the content <strong>of</strong> o-DCB in<br />

Fig. 2. Optical pr<strong>of</strong>iler images <strong>of</strong> PPE-PPV films <strong>inkjet</strong> <strong>printed</strong> with a dot spacing <strong>of</strong><br />

a) 80 μm, b) 120 μm, and c) 160 μm. Solvent system was toluene/o-DCB in a ratio <strong>of</strong><br />

90/10 and a concentration <strong>of</strong> 4 mg mL −1 .<br />

line instead <strong>of</strong> film <strong>formation</strong> (Fig. 2c), while even larger dot spacings<br />

led to single droplets, which have a diameter <strong>of</strong> approximately<br />

140 μm.<br />

3.2. Influence <strong>of</strong> the solvent system<br />

The influence <strong>of</strong> the solvent system was investigated by <strong>inkjet</strong><br />

printing the PPE-PPV co<strong>poly</strong>mer with a concentration <strong>of</strong> 4 mg mL −1 in<br />

the solvent system toluene/o-DCB while the ratio between the two<br />

solvents systematically varied. The ratios 90/10, 95/5 and 99/1 were<br />

used with dot spacings between 80 μm and 160 μm. Ratios with a<br />

larger amount <strong>of</strong> the lower boiling solvent toluene are known to show<br />

irregular films after being <strong>printed</strong> [25].<br />

Fig. 3. Optical pr<strong>of</strong>iler images <strong>of</strong> PPE-PPV films <strong>inkjet</strong> <strong>printed</strong> from the solvent system toluene/<br />

o-DCB in the ratios a) 90/10, b) 95/5, and c) 99/1 (dot spacing 120 μm, c=4 mg mL −1 ).


3698 A. Teichler et al. / Thin Solid <strong>Film</strong>s 519 (2011) 3695–3702<br />

the experiments performed here is recommended to be at least 5 vol%,<br />

but preferably 10 vol%, in order to obtain homogeneous films.<br />

3.3. Influence <strong>of</strong> the concentration<br />

The effect <strong>of</strong> the solute concentration on their film forming<br />

<strong>properties</strong> was investigated for three different concentrations,<br />

including 4 mg mL −1 , 6 mg mL −1 and 8 mg mL −1 . The solvent<br />

system toluene/o-DCB 90/10 was chosen following the good results<br />

from the previous section. Fig. 4 shows optical pr<strong>of</strong>iler images <strong>of</strong><br />

typical <strong>inkjet</strong> <strong>printed</strong> films from the solutions with a systematically<br />

varied concentration and dot spacing. It can be seen that a variation<br />

<strong>of</strong> the concentration <strong>of</strong> the solutions strongly affects the film<br />

<strong>formation</strong>.<br />

A concentration <strong>of</strong> 4 mg mL −1 (Fig. 4a) and a dot spacing <strong>of</strong> 80 μm<br />

led to inhomogeneous films because too much material was<br />

deposited. Homogeneous films were obtained for a dot spacing <strong>of</strong><br />

120 μm, while a dot spacing <strong>of</strong> 160 μm resulted in inhomogeneous<br />

films and line <strong>formation</strong>. When having a closer look at the dot spacing<br />

<strong>of</strong> 120 μm still a faint line <strong>formation</strong> can be observed. This may be<br />

ascribed to the velocity <strong>of</strong> the printhead: at low printing velocities a<br />

<strong>printed</strong> line may dry just before the following line will be <strong>printed</strong>,<br />

hampering a smooth merging <strong>of</strong> the two lines (in our experiments we<br />

have not varied the printhead velocity.).<br />

A concentration <strong>of</strong> 6 mg mL −1 (Fig. 4b) revealed inhomogenous<br />

films when <strong>printed</strong> with a dot spacing <strong>of</strong> 80 μm, while 120 μm<br />

resulted in homogeneous films. Again at a dot spacing <strong>of</strong> 160 μm line<br />

<strong>formation</strong> appeared. A solute concentration <strong>of</strong> 8 mg mL −1 (Fig. 4c)<br />

was not suitable to create homogeneous films since separate lines<br />

were formed at any dot spacing above 120 μm.<br />

As a conclusion, the best solute concentration was found to be<br />

4mg mL −1 , since a broader range <strong>of</strong> dot spacings could be used to<br />

obtain homogeneous films.<br />

3.4. Influence <strong>of</strong> the substrate temperature<br />

To study the effect <strong>of</strong> the substrate temperature, a <strong>poly</strong>mer<br />

concentration <strong>of</strong> 4 mg mL −1 in toluene/o-DCB 90/10 was chosen,<br />

Fig. 4. Optical pr<strong>of</strong>iler images <strong>of</strong> PPE-PPV films <strong>inkjet</strong> <strong>printed</strong> with concentrations <strong>of</strong> a)4 mg mL −1 , b) 6 mg mL −1 , and c) 8 mg mL −1 (solvent system toluene/o-DCB 90/10).


A. Teichler et al. / Thin Solid <strong>Film</strong>s 519 (2011) 3695–3702<br />

3699<br />

Fig. 5. Optical pr<strong>of</strong>iler images <strong>of</strong> PPE-PPV films <strong>inkjet</strong> <strong>printed</strong> with a dot spacing <strong>of</strong> 80 μm (a), 120 μm (b) and 160 μm (c). The columns show, from left to right, a substrate<br />

temperature <strong>of</strong> room temperature, 35 °C and 50 °C, respectively (solvent system toluene/o-DCB 90/10, c=4 mg mL −1 ).<br />

based on the results from previous experiments. The temperature<br />

varied between room temperature and 50 °C. Fig. 5a–c shows the<br />

comparison between the different substrate temperatures at a dot<br />

spacing <strong>of</strong> 80 μm, 120 μm, and 160 μm, respectively.<br />

At a dot spacing <strong>of</strong> 80 μm (Fig. 5a) homogeneous films could only<br />

be obtained at room temperature, while increasing the temperature<br />

showed irregular films or films with line <strong>formation</strong>. For a dot spacing<br />

<strong>of</strong> 120 μm (Fig. 5b) the <strong>formation</strong> <strong>of</strong> lines took place already at a<br />

slightly increased temperature <strong>of</strong> 35 °C. At room temperature smooth<br />

films were <strong>printed</strong>. With a dot spacing <strong>of</strong> 160 μm (Fig. 5c) no<br />

homogeneous films could be <strong>printed</strong>; already at room temperature<br />

line <strong>formation</strong> was observed. The increase <strong>of</strong> the temperature to 50 °C<br />

did not improve the film <strong>formation</strong>.<br />

Increased temperatures did not have a positive effect on the film<br />

<strong>formation</strong>; only at room temperature well-defined films were formed.<br />

Furthermore, when using a substrate temperature greater than 50 °C,<br />

the films consisted only <strong>of</strong> individual lines due to a faster drying<br />

process, which prevents the lines from merging into a homogeneous<br />

film. It has been shown in literature that the nozzle-to-substrate<br />

distance as well as the substrate temperature strongly influences the<br />

in-flight droplet diameter [30]. An increased substrate temperature<br />

stimulates solvent evaporation during the dispensing <strong>of</strong> the ink,<br />

resulting in a reduced droplet diameter on the substrate. We believe<br />

that the effect <strong>of</strong> <strong>inkjet</strong> printing a wet droplet (line) next to a semidried<br />

droplet (line), i.e. an as-<strong>printed</strong> droplet (line) that is evaporating,<br />

is the cause <strong>of</strong> the less homogeneous film <strong>formation</strong>, as seen here, at<br />

elevated temperatures.<br />

3.5. Influence on the film thickness<br />

For all homogeneous films the thickness was measured using<br />

optical pr<strong>of</strong>ilometry; the results are summarized in Table 1. For a<br />

concentration <strong>of</strong> 4 mg mL −1 and a solvent ratio 90/10 film thicknesses<br />

between 104 nm and 79 nm were obtained for a dot spacing between<br />

100 and 130 μm, respectively. Smaller film thicknesses between<br />

94 nm and 68 nm were obtained for films <strong>printed</strong> from toluene/o-DCB<br />

in a ratio <strong>of</strong> 95/5 at the same concentration and dot spacing.<br />

<strong>Film</strong>s <strong>inkjet</strong> <strong>printed</strong> either from the solvent ratio 90/10 or 95/5


3700 A. Teichler et al. / Thin Solid <strong>Film</strong>s 519 (2011) 3695–3702<br />

Table 1<br />

<strong>Film</strong> thickness, maximum absorption and standard deviation <strong>of</strong> the absorption <strong>of</strong> <strong>inkjet</strong><br />

<strong>printed</strong> PPE-PPV films for different dot spacings, concentrations, and solvent ratios.<br />

Concentration<br />

(mg mL −1 )<br />

Solvent ratio<br />

(toluene/o-DCB)<br />

showed the same homogeneity as can be seen by the standard<br />

deviation between 13 nm and 24 nm. This confirms the observation in<br />

a previous section that the amount <strong>of</strong> high boiling solvent o-DCB<br />

should be at least 5 vol%.<br />

The films <strong>printed</strong> from a concentration <strong>of</strong> 6 mg mL −1 and a dot<br />

spacing between 115 and 130 μm revealed a film thickness <strong>of</strong> approximately<br />

95 nm. As expected, these films showed a greater thickness<br />

than those <strong>printed</strong> at the same dot spacings from a 4 mg mL −1<br />

solution. Standard deviations were for both concentrations in the same<br />

range; hence their film homogeneity was comparable.<br />

As mentioned before, it was found that the range <strong>of</strong> dot spacings,<br />

wherein homogeneous films could be <strong>printed</strong>, decreased with a<br />

higher solute concentration in the ink. For a 4 mg mL −1 solution<br />

homogeneous films were obtained with dot spacings between 100<br />

and 130 μm, whereas a 6 mg mL −1 solution led to homogeneous films<br />

between 115 and 130 μm. An 8 mg mL −1 solution could not be <strong>printed</strong><br />

into homogeneous films at any dot spacing.<br />

3.6. Optical characterization<br />

Dot<br />

spacing<br />

(μm)<br />

<strong>Film</strong><br />

thickness<br />

(nm)<br />

A max<br />

(a.u.)<br />

Standard<br />

deviation,<br />

A max (%)<br />

4 90/10 100 104±14 0.400 9.1<br />

105 102±20 0.372 3.5<br />

110 91 ±20 0.367 11.3<br />

115 88 ±14 0.306 11.0<br />

120 83 ±13 0.306 8.7<br />

125 80 ±16 0.403 14.0<br />

130 79 ±14 0.369 4.2<br />

4 95/5 100 93 ±24 0.032 4.2<br />

105 94 ±18 0.063 9.3<br />

110 83 ±15 0.025 4.0<br />

115 75 ±16 0.155 22.2<br />

120 83 ±19 0.084 16.2<br />

125 68 ±18 0.057 10.5<br />

6 90/10 115 95 ±20 0.856 15.1<br />

120 95 ±18 0.775 14.3<br />

125 94 ±17 0.698 10.8<br />

130 93 ±15 0.751 17.2<br />

UV-vis absorption spectra were recorded for the PPE-PPV<br />

co<strong>poly</strong>mer in solution with a concentration <strong>of</strong> 0.1 μg mL −1 in<br />

toluene and in solid state from the <strong>printed</strong> films (toluene/o-DCB 90/<br />

10, 4 mg mL −1 ) and shown in Fig. 6a. The maximum absorbance<br />

peak in solution was found at a wavelength <strong>of</strong> 453 nm, while the<br />

<strong>printed</strong> films showed a maximum at 470 nm and a shoulder around<br />

515 nm. The observed shoulder can be ascribed to band splitting<br />

that is attributed to different arrangements <strong>of</strong> the <strong>poly</strong>mer, including<br />

H- (475 nm) and J-aggregates (515 nm) [26]. These aggregates are<br />

formed by interactions between the <strong>poly</strong>mers or the <strong>poly</strong>mer<br />

segments [31].<br />

The photoluminescence spectrum <strong>of</strong> the solution (Fig. 6b) shows a<br />

maximum at 514 nm and a small shoulder at a higher wavelength<br />

(550 nm). The <strong>inkjet</strong> <strong>printed</strong> films revealed emission at a wavelength<br />

<strong>of</strong> 588 nm. A bathochromic shift was observed for absorption as well<br />

as photoluminescence <strong>of</strong> the films when compared to the solutions.<br />

This can be attributed to the <strong>formation</strong> <strong>of</strong> aggregates in the films, and,<br />

thus, an enhanced planarization <strong>of</strong> the <strong>poly</strong>mer backbone takes place<br />

[27].<br />

The reproducibility <strong>of</strong> the <strong>inkjet</strong> <strong>printed</strong> films was estimated by<br />

comparing the optical <strong>properties</strong> <strong>of</strong> multiple films. Hereto, five films<br />

were <strong>inkjet</strong> <strong>printed</strong> next to each other and with equal print settings.<br />

Table 1 summarizes the solution <strong>properties</strong>, the printing parameters,<br />

the average maximum absorbance values as well as the<br />

Absorption (normalized)<br />

Photoluminescence (normalized)<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

a)<br />

300 350 400 450 500 550<br />

b)<br />

Wavelength (nm)<br />

solution<br />

film<br />

500 550 600 650 700 750 800<br />

Wavelength (nm)<br />

solution<br />

film<br />

Fig. 6. a) Normalized absorption and b) photoluminescence spectra <strong>of</strong> the PPE-PPV<br />

co<strong>poly</strong>mer in solution (solid line, c=0.1 μg mL −1 in toluene) and in thin film (dashed<br />

line, <strong>printed</strong> from 4 mg mL −1 in toluene/o-DCB with a ratio <strong>of</strong> 90/10).<br />

standard deviations <strong>of</strong> the respective maximum absorption. It was<br />

found that for a concentration <strong>of</strong> 4 mg mL −1 the most reproducible<br />

films were obtained for a solvent ratio <strong>of</strong> 90/10 and a dot spacing<br />

between 100 μm and 130 μm. The standard deviations obtained for<br />

the maximum absorption were between 4% and 14%, which<br />

corroborates in the optical pr<strong>of</strong>iler results in the proceeding section.<br />

In contrast, a solvent ratio <strong>of</strong> 95/5 was less reproducible, indicated by<br />

the slightly larger standard deviations between 4% and 22%. Finally,<br />

standard deviations between 11% and 17% were obtained for films<br />

that were <strong>printed</strong> from a higher concentration (6 mg mL −1 ) in<br />

toluene/o-DCB 90/10.<br />

The absorption spectra <strong>of</strong> five samples <strong>inkjet</strong> <strong>printed</strong> with equal<br />

conditions from the PPE-PPV co<strong>poly</strong>mer with a concentration <strong>of</strong><br />

4mgmL −1 , a solvent ratio <strong>of</strong> 90/10 and a dot spacing <strong>of</strong> 130 μm are<br />

shown in Fig. 7. At the maximum peak, the difference between the<br />

spectra is approximately 4%.<br />

Concluding the combinatorial screening, optimal settings for a<br />

homogeneous film <strong>formation</strong> for the PPE-PPV co<strong>poly</strong>mer with a good<br />

reproducibility are: a solvent concentration <strong>of</strong> 4 mg mL −1 , a solvent<br />

system <strong>of</strong> toluene/o-DCB in a ratio 90 to 10, a dot spacing <strong>of</strong> 120 μm,<br />

and processing at room temperature. Fig. 8 shows a three-dimensional<br />

optical pr<strong>of</strong>iler image <strong>of</strong> the optimal settings (Fig. 8a) and, as a<br />

comparison, films <strong>of</strong> a 4 mg mL −1 solution, toluene/o-DCB ratio 95/5<br />

with a dot spacing <strong>of</strong> 110 μm (Fig. 8b) and <strong>of</strong> a 6 mg mL −1 solution<br />

with a solvent ratio <strong>of</strong> 90/10 and a dot spacing <strong>of</strong> 120 μm (Fig. 8c).


A. Teichler et al. / Thin Solid <strong>Film</strong>s 519 (2011) 3695–3702<br />

3701<br />

Fig. 7. Absorption spectra <strong>of</strong> five films <strong>inkjet</strong> <strong>printed</strong> with equal conditions from<br />

PPE-PPV co<strong>poly</strong>mer in toluene/o-DCB 90/10 with a concentration <strong>of</strong> 4 mg mL −1 and<br />

a dot spacing <strong>of</strong> 130 μm (samples1–5).<br />

4. Conclusions<br />

In this study, <strong>inkjet</strong> printing was used as a precise and reproducible<br />

patterning technique to create thin-film libraries <strong>of</strong> the in-house<br />

synthesized co<strong>poly</strong>mer <strong>poly</strong>(phenylene-ethynylene)-<strong>poly</strong>(phenylenevinylene)<br />

(PPE-PPV). Inkjet printing <strong>of</strong> thin-film libraries allow a<br />

combinatorial workflow to investigate structure-property relationships.<br />

Here, we systematically varied a set <strong>of</strong> parameters that strongly<br />

influence the film forming <strong>properties</strong>, including the ink composition<br />

(solute concentration, solvent mixture ratios and solute concentration),<br />

as well as the substrate temperature and dot spacing.<br />

All parameters were studied as function <strong>of</strong> the dot spacing, which<br />

directly relates to the film thickness. When using the solvent system<br />

toluene and ortho-dichlorobenzene (o-DCB), a ratio between the two<br />

solvents <strong>of</strong> 90 to 10 and a solute concentration <strong>of</strong> 4 mg mL −1 leads to<br />

the <strong>formation</strong> <strong>of</strong> homogeneous films when <strong>printed</strong> at room temperature<br />

with a dot spacing between 100 and 130 μm. The resulting film<br />

thicknesses are between 80 and 100 nm, which is promising for<br />

<strong>printed</strong> electronics applications, such as organic light emitting diodes<br />

(OLEDs) and organic photovoltaics.<br />

With a decreased amount <strong>of</strong> the high boiling solvent o-DCB or<br />

increased substrate temperature, the <strong>formation</strong> <strong>of</strong> lines became<br />

visible, which indicates a too fast drying <strong>of</strong> the solvent, resulting in<br />

inhomogeneous films.<br />

Optical characterization was performed in a fast and reproducible<br />

manner using a UV-vis plate reader. The films were <strong>printed</strong> in a<br />

microtiter plate format where each film is located at the position <strong>of</strong> a<br />

well. It was found that the <strong>poly</strong>mer had its maximum absorption in<br />

solution at a wavelength <strong>of</strong> 453 nm and emitted at 514 nm. In the<br />

solid state, the absorption and emission were red-shifted due to the<br />

increased planarization <strong>of</strong> the <strong>poly</strong>mer backbone.<br />

By using <strong>inkjet</strong> printing, film forming <strong>properties</strong> could be<br />

optimized in a combinatorial approach, which did not only save<br />

time and (materials) costs, but also revealed synergies between<br />

different parameters that would not have been discovered when<br />

optimizing each parameter individually.<br />

In optoelectronic devices the homogeneity is <strong>of</strong> utmost importance,<br />

but the film thickness and used inks may vary. Our approach<br />

can now be applied for film preparation by <strong>inkjet</strong> printing for any<br />

other (<strong>poly</strong>mer) ink solution and represents a fast and efficient<br />

screening <strong>of</strong> the parameters to obtain homogeneous films with a<br />

precise thickness.<br />

Acknowledgements<br />

The authors would like to thank the DPI for financial support<br />

(projects #620 and #589).<br />

References<br />

Fig. 8. 3D optical pr<strong>of</strong>iler images <strong>of</strong> films prepared from a) 4 mg mL −1 , toluene/o-DCB<br />

90/10, dot spacing 120 μm, b) 4 mg mL −1 , toluene/o-DCB 95/5, dot spacing 110 μm and<br />

c) 6 mg mL −1 , toluene/o-DCB 90/10, dot spacing 120 μm.<br />

[1] A.C. Grimsdale, K.L. Chan, R.E. Martin, P.G. Jokisz, A.B. Holmes, Chem. Rev. 109<br />

(2009) 897.<br />

[2] J.H. Burroughes, D.D.C. Bradley, A.R. Brown, R.N. Marks, K. Mackay, R.H. Friend, P.L.<br />

Burns, A.B. Holmes, Nature 347 (1990) 539.<br />

[3] A. Menon, H. Dong, Z.I. Niazimbetova, L.J. Rothberg, M.E. Galvin, Chem. Mater. 14<br />

(2002) 3668.<br />

[4] S. Günes, H. Neugebauer, N.S. Sariciftci, Chem. Rev. 107 (2007) 1324.<br />

[5] G. Dennler, M.C. Scharber, C.J. Brabec, Adv. Funct. Mater. 21 (2009) 1323.<br />

[6] T. Kawase, T. Shimoda, C. Newsome, H. Sirringhaus, R.H. Friend, Thin Solid <strong>Film</strong>s<br />

437 (2003) 279.<br />

[7] H.N. Tsao, D. Cho, J.W. Andreasen, A. Rouhanipour, D.W. Breiby, W. Pisula, K.<br />

Müllen, Adv. Mater. 21 (2009) 209.<br />

[8] E. Tekin, D.A.M. Egbe, J.M. Kranenburg, C. Ulbricht, S. Rathgeber, E. Birckner, N.<br />

Rehmann, K. Meerholz, U.S. Schubert, Chem. Mater. 20 (2008) 2727.<br />

[9] B.-J. de Gans, P.C. Duineveld, U.S. Schubert, Adv. Mater. 16 (2004) 203.<br />

[10] D.A.M. Egbe, C.P. Roll, E. Birckner, U.-W. Grummt, R. Stockmann, E. Klemm,<br />

Macromolecules 35 (2002) 3825.<br />

[11] D.A.M. Egbe, H. Tillmann, E. Birckner, E. Klemm, Macromol. Chem. Phys. 202<br />

(2001) 2712.<br />

[12] M. Al-Ibrahim, A. Konkin, H.K. Roth, D.A.M. Egbe, E. Klemm, U. Zhokhavets, G.<br />

Gobsch, S. Sensfuss, Thin Solid <strong>Film</strong>s 474 (2005) 201.<br />

[13] D.A.M. Egbe, L.H. Nguyen, H. Hoppe, D. Mühlbacher, N.S. Sariciftci, Macromol.<br />

Rapid Commun. 26 (2005) 1389.<br />

[14] H. Hoppe, N.S. Sariciftci, D.A.M. Egbe, D. Mühlbacher, M. Koppe, Mol. Cryst. Liq.<br />

Cryst. 426 (2005) 255.<br />

[15] F.C. Krebs, Sol. Energy Mat. Sol. C 93 (2009) 394.<br />

[16] E. Tekin, P.J. Smith, U.S. Schubert, S<strong>of</strong>t Matter 4 (2008) 703.<br />

[17] M. Singh, H.M. Haverinen, P. Dhagat, G.E. Jabbour, Adv. Mater. 22 (2010) 673.<br />

[18] R.D. Deegan, Phys. Rev. E 61 (2000) 475.<br />

[19] R.D. Deegan, O. Bakajin, T.F. Dupont, G. Huber, S.R. Nagel, T.A. Witten, Phys. Rev. E<br />

62 (2000) 756.<br />

[20] B.-J. de Gans, U.S. Schubert, Langmuir 20 (2004) 7789.<br />

[21] E. Tekin, E. Holder, D. Kozodaev, U.S. Schubert, Adv. Funct. Mater. 17 (2007) 277.<br />

[22] D. Soltman, V. Subramanian, Langmuir 24 (2008) 2224.


3702 A. Teichler et al. / Thin Solid <strong>Film</strong>s 519 (2011) 3695–3702<br />

[23] B.-J. de Gans, E. Kazancioglu, W. Meyer, U.S. Schubert, Macromol. Rapid Commun.<br />

25 (2004) 292.<br />

[24] B.-J. de Gans, U.S. Schubert, Macromol. Rapid Commun. 24 (2003) 659.<br />

[25] E. Tekin, B.-J. de Gans, U.S. Schubert, J. Mater. Chem. 14 (2004) 2627.<br />

[26] E. Tekin, H. Wijlaars, E. Holder, D.A.M. Egbe, U.S. Schubert, J. Mater. Chem. 16<br />

(2006) 4294.<br />

[27] D.A.M. Egbe, C. Bader, E. Klemm, L. Ding, F.E. Karasz, U.-W. Grummt, E. Birckner,<br />

Macromolecules 36 (2003) 9303.<br />

[28] A.M.J. van den Berg, A.W.M. de Laat, P.J. Smith, J. Perelaer, U.S. Schubert, J. Mater.<br />

Chem. 17 (2007) 677.<br />

[29] Y. Yoshioka, G.E. Jabbour, Synth. Met. 156 (2006) 779.<br />

[30] J. Perelaer, P.J. Smith, M.M.P. Wijnen, E. van den Bosch, R. Eckardt, P.H.J.M.<br />

Ketelaars, U.S. Schubert, Macromol. Chem. Phys. 210 (2009) 387.<br />

[31] S. Siddiqui, F.C. Spano, Chem. Phys. Lett. 308 (1999) 99.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!