29.12.2012 Views

Inkjet Printing of Well-Defined Polymer Dots and Arrays

Inkjet Printing of Well-Defined Polymer Dots and Arrays

Inkjet Printing of Well-Defined Polymer Dots and Arrays

SHOW MORE
SHOW LESS

Create successful ePaper yourself

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

<strong>Inkjet</strong> <strong>Printing</strong> <strong>of</strong> <strong>Well</strong>-<strong>Defined</strong> <strong>Polymer</strong> <strong>Dots</strong> <strong>and</strong> <strong>Arrays</strong><br />

Berend-Jan de Gans <strong>and</strong> Ulrich S. Schubert*<br />

Laboratory <strong>of</strong> Macromolecular Chemistry <strong>and</strong> Nanoscience, Eindhoven University <strong>of</strong><br />

Technology <strong>and</strong> Dutch <strong>Polymer</strong> Institute (DPI), P.O. Box 513,<br />

5600 MB Eindhoven, The Netherl<strong>and</strong>s<br />

Received March 1, 2004. In Final Form: June 9, 2004<br />

<strong>Inkjet</strong> printing represents a highly promising polymer deposition method, which is used for, for example,<br />

the fabrication <strong>of</strong> multicolor polyLED displays <strong>and</strong> polymer-based electronics parts. The challenge is to<br />

print well-defined polymer structures from dilute solution. We have eliminated the formation <strong>of</strong> ring stains<br />

by printing nonvolatile acetophenone-based inks on a perfluorinated substrate using different polymers.<br />

(De)pinning <strong>of</strong> the contact line <strong>of</strong> the printed droplet, as related to the choice <strong>of</strong> solvent, is identified as<br />

the key factor that determines the shape <strong>of</strong> the deposit, whereas the choice <strong>of</strong> polymer is <strong>of</strong> minor importance.<br />

Adding 10 wt % or more <strong>of</strong> acetophenone to a volatile solvent (ethyl acetate)-based polymer solution<br />

changes the shape <strong>of</strong> the deposit from ring-like to dot-like, which may be due to the establishment <strong>of</strong> a<br />

solvent composition gradient. <strong>Arrays</strong> <strong>of</strong> closely spaced dots have also been printed. The size <strong>of</strong> the dots<br />

is considerably smaller than the nozzle diameter. This may prove a potential strategy for the inkjet printing<br />

<strong>of</strong> submicrometer structures.<br />

Introduction<br />

<strong>Inkjet</strong> printing is considered one <strong>of</strong> the most promising<br />

methods for controlled deposition <strong>of</strong> polymers <strong>and</strong> functional<br />

materials, 1,2 especially in relation to the fabrication<br />

<strong>of</strong> multicolor polyLED displays <strong>and</strong> polymer-based electronics<br />

parts. Such devices cannot be prepared by conventional<br />

spin coating. Instead, a microscopic patterning<br />

technique is needed. <strong>Inkjet</strong> printing has the advantage <strong>of</strong><br />

simplicity, low cost, flexibility, <strong>and</strong> maturity. Outst<strong>and</strong>ing<br />

examples include inkjet printed all-polymer transistors3-5 <strong>and</strong> 130 ppi RGB polyLED displays. 6,7 However, the<br />

potential <strong>of</strong> inkjet printing is huge <strong>and</strong> goes far beyond<br />

what has been realized up to now. Future areas <strong>of</strong><br />

application may include the fabrication <strong>of</strong> sensors, polymerbased<br />

solar cells, or micr<strong>of</strong>luidics devices, all based on<br />

polymers as matrix or functional materials. In the field<br />

<strong>of</strong> combinatorial materials science, inkjet printed libraries<br />

<strong>of</strong> polymer dots could bridge the gap between parallel<br />

synthesis <strong>and</strong> property characterization.<br />

As inkjet printing requires low viscosities (typically<br />

1-10 mPa s), polymeric materials can only be deposited<br />

from dilute solution. 8 However, a drying droplet usually<br />

deposits its solute as a ring stain that marks the perimeter<br />

<strong>of</strong> the droplet before drying. Ring formation, commonly<br />

known as the “c<strong>of</strong>fee drop effect”, is due to the combined<br />

* Corresponding author. Fax: 0031-40-247-4186. E-mail:<br />

u.s.schubert@tue.nl.<br />

(1) De Gans, B. J.; Duineveld, P. C.; Schubert, U. S. Adv. Mater.<br />

2004, 16, 203.<br />

(2) Calvert, P. Chem. Mater. 2001, 13, 3299.<br />

(3) Sirringhaus, H.; Kawase, T.; Friend, R. H.; Shimoda, T.; Inbasekaran,<br />

M.; Wu, W.; Woo, E. P. Science 2000, 290, 2123.<br />

(4) Stutzmann, N.; Friend, R. H.; Sirringhaus, H. Science 2003, 299,<br />

1881.<br />

(5) Paul, K. E.; Wong, W. S.; Ready, S. E.; Street, R. A. Appl. Phys.<br />

Lett. 2003, 83, 2070.<br />

(6) Funamoto, T.; Matsueda, Y.; Yokoyama, O.; Tsuda, A.; Takeshita,<br />

H.; Miyashita, S. Proc. 22 nd Int. Display Res. Conf., Boston 2002; Society<br />

for Information Display: San Jose, 2002; p 899.<br />

(7) Giraldo, A.; Duineveld, P. C.; Johnson, M. T.; Lifka, H.; Rubingh,<br />

J. E. J. M.; Childs, M. J.; Fish, D. A.; George, D. S.; Godfrey, S. D.;<br />

McCulloch, D. J.; Steer, W. A.; Trainor, M.; Young, N. D.; Hunter, I. M.<br />

Proc. 7 th Asian Symp. Information Display (ASID2002), Singapore 2002;<br />

Society for Information Display: San Jose, 2002; p 43.<br />

(8) De Gans, B. J.; Kazancioglu, E.; Schubert, U. S. Macromol. Rapid<br />

Commun. 2004, 25, 292.<br />

Langmuir 2004, 20, 7789-7793<br />

7789<br />

action <strong>of</strong> an increased evaporation rate at the droplet edge,<br />

<strong>and</strong> contact line pinning due to surface irregularities <strong>and</strong><br />

solute deposition (“self-pinning”). 9-11 A capillary-driven<br />

flow from the droplet center toward the edge compensates<br />

for evaporation losses <strong>and</strong> transports most <strong>of</strong> the solute<br />

toward the contact line. All previously mentioned applications<br />

<strong>of</strong> inkjet printing require well-defined, <strong>of</strong>ten<br />

dot-like deposits. Therefore, the elimination <strong>of</strong> ring<br />

formation is <strong>of</strong> great practical interest.<br />

To improve the homogeneity <strong>of</strong> the deposits, mixtures<br />

<strong>of</strong> low-boiling good solvents <strong>and</strong> high-boiling bad solvents<br />

for the polymer under consideration were used. During<br />

evaporation, the solvent quality decreases, <strong>and</strong> the<br />

polymer will precipitate. 12 A disadvantage is the poor<br />

(mechanical) properties <strong>of</strong> the deposit formed. It was<br />

shown that when a single solvent was used, the homogeneity<br />

increases with decreasing rate <strong>of</strong> evaporation. 13<br />

This could be due to the time scale for depinning <strong>of</strong> the<br />

contact line matching the time scale <strong>of</strong> evaporation.<br />

Polyvinylcarbazole dots having a Gaussian shape were<br />

inkjet printed, although the optical micrograph provided<br />

clearly shows irregularities, probably due to stick-slip<br />

(de)pinning <strong>of</strong> the contact line during evaporation. 14 ITO<br />

was used as the substrate. The velocity <strong>of</strong> the print head<br />

<strong>and</strong> the evaporation conditions are also known to affect<br />

the shape <strong>of</strong> the deposit. 15<br />

Casting experiments with macroscopic, aqueous drops<br />

have shown that ring formation can be (partially) eliminated<br />

by the use <strong>of</strong> a hydrophobic substrate, that is, a film<br />

<strong>of</strong> castor oil on glass. 16 In the field <strong>of</strong> inkjet printing, use<br />

(9) Deegan, R. D.; Bakajin, O.; Dupont, T. F.; Huber, G.; Nagel, S.<br />

R.; Witten, T. A. Nature 1997, 389, 827.<br />

(10) Deegan, R. D.; Bakajin, O.; Dupont, T. F.; Huber, G.; Nagel, S.<br />

R.; Witten, T. A. Phys. Rev. E 2000, 62, 756.<br />

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

(12) Lyon, P. J.; Carter, J. C.; Bright, J. C.; Cacheiro, M. WO Patent<br />

02/069119 A1, 2002.<br />

(13) Madigan, C. F.; Hebner, T. R.; Sturm, J. C.; Register, R. A.;<br />

Troian, S. MRS Symp. Proc. Vol. 625; Materials Research Society:<br />

Warrendale, 2000; p 123.<br />

(14) Hebner, T. R.; Wu, C. C.; Marcy, D.; Lu, M. H.; Sturm, J. C. Appl.<br />

Phys. Lett. 1998, 72, 519.<br />

(15) Shimoda, T.; Morii, K.; Seki, S.; Kiguchi, H. MRS Bull. 2003,<br />

28, 821.<br />

(16) Takhistov, P.; Chang, H.-C. Ind. Eng. Chem. Res. 2002, 41, 6256.<br />

10.1021/la049469o CCC: $27.50 © 2004 American Chemical Society<br />

Published on Web 07/30/2004


7790 Langmuir, Vol. 20, No. 18, 2004 de Gans <strong>and</strong> Schubert<br />

Table 1. Solvent Properties at Room Temperature<br />

liquid Tb (°C) Pv (mmHg) θc (deg) γLV (mN m -1 )<br />

ethyl acetate 77.1 76 37.4 ( 2.2 23.39<br />

anisole 153.7 3.51 57.2 ( 1.3 35.10<br />

acetophenone 202 0.35 69.1 ( 1.4 39.04<br />

was made <strong>of</strong> hydrophobic/hydrophilic patterns to control<br />

the position <strong>of</strong> inkjet printed droplets, 1,15,17 but not yet for<br />

influencing the shape <strong>of</strong> the deposit. In this contribution,<br />

we will demonstrate how homogeneous deposits <strong>of</strong> polymeric<br />

material can be obtained via use <strong>of</strong> a hydrophobic,<br />

perfluorinated surface <strong>and</strong> a single solvent rather than<br />

a mixture <strong>of</strong> a good <strong>and</strong> a bad solvent. In addition, we will<br />

show that the use <strong>of</strong> a mixture <strong>of</strong> two good solvents for the<br />

polymer allows even further control over the shape <strong>of</strong> the<br />

deposit <strong>and</strong> that arrays <strong>of</strong> dots can be inkjet printed.<br />

Experimental Section<br />

Materials. Polydisperse polystyrene (N5000, Shell Nederl<strong>and</strong>,<br />

Den Haag, The Netherl<strong>and</strong>s; Mn 80 kD, Mw 282 kD) was used.<br />

Poly(methyl methacrylate) was purchased from Sigma-Aldrich<br />

(Steinheim, Germany; Mn 130 kD, Mw 240 kD). Monodisperse<br />

polystyrenes were purchased from <strong>Polymer</strong> St<strong>and</strong>ards Service<br />

(Mainz, Germany; Mn/Mw 17/17.5, 30/34, 60/64, 120/125, 214/<br />

226 kD). Ethyl acetate (Biosolve, Valkenswaard, The Netherl<strong>and</strong>s),<br />

anisole (Acros Organics, Geel, Belgium), <strong>and</strong> acetophenone<br />

(Sigma-Aldrich, Steinheim, Germany) were used as solvents, or<br />

mixtures there<strong>of</strong>. Solvent properties are listed in Table 1.<br />

Solutions used for printing contained 1.0% polymer by weight.<br />

Viscosities were measured with an Ubbelohde viscometer at<br />

20.0 °C. Despite the high molecular weight, these solutions could<br />

easily be printed without persistent filament formation. 8<br />

Substrate Preparation. Glass slides (Marienfeld, Lauda-<br />

Königsh<strong>of</strong>en, Germany) were ultrasonicated in acetone for 5 min,<br />

rubbed with sodium dodecyl sulfate solution, ultrasonicated in<br />

sodium dodecyl sulfate solution for 5 min, flushed with demineralized<br />

water to remove the soap, treated with 2-propanol<br />

vapor to remove the water in a reflux setup, dried with a flow<br />

<strong>of</strong> nitrogen, <strong>and</strong> subsequently treated in a UV-ozone photoreactor<br />

(PR-100, UVP, Upl<strong>and</strong>, CA) for 30 min to remove any remaining<br />

organic contamination. Substrates were used immediately after<br />

cleaning. Alternatively, glass slides were coated with (tridecafluoro-1,1,2,2-tetrahydro-octyl)trichlorosilane<br />

(ABCR, Karlsruhe,<br />

Germany), 18 resulting in strongly hydrophobic surfaces. Contact<br />

angles were measured with an OCA30 device from Dataphysics<br />

(Filderstadt, Germany). The results are listed in Table 1.<br />

<strong>Inkjet</strong> Printer. A microdrop Autodrop device (Norderstedt,<br />

Germany) was used, consisting <strong>of</strong> an automated XYZ-stage in<br />

combination with a AD-K-501 micropipet print head. 19 The<br />

printer <strong>of</strong>fers a workspace <strong>of</strong> 200 × 200 × 80 mm. The positioning<br />

accuracy <strong>of</strong> the print head is 3 µm. The diameter <strong>of</strong> the micropipet<br />

nozzle is 70 µm.<br />

The Autodrop is a piezoelectric inkjet printer. Droplet ejection<br />

occurs via contraction <strong>of</strong> a piezoelectric actuator that generates<br />

an acoustic wave. The signal that drives the actuator is a<br />

rectangular pulse, with an amplitude <strong>of</strong> 80 V, <strong>and</strong> a width <strong>of</strong> 29<br />

µs. The ejection frequency is 200 Hz, <strong>and</strong> the print head velocity<br />

is5mms-1 . Variable amounts <strong>of</strong> material can be deposited by<br />

changing the number <strong>of</strong> droplets that are printed at a spot. The<br />

inkjet printer produces highly uniform droplets with a typical<br />

radius error <strong>of</strong> less than 1%. Unless stated otherwise, all samples<br />

were printed as a 2 × 10 array, with 1.00 mm spacing between<br />

the dots. Each dot is the result <strong>of</strong> five drops that coalesce on the<br />

surface to form one large drop.<br />

Surface Topography. A white light confocal microscope<br />

(Nan<strong>of</strong>ocus µSurf, Duisburg, Germany) was used to determine<br />

(17) Wang, J. Z.; Zheng, Z. H.; Li, H. W.; Huck, W. T. S.; Sirringhaus,<br />

H. Nat. Mater. 2004, 3, 171.<br />

(18) Trimbach, D.; Feldman, K.; Spencer, N. D.; Broer, D. J.;<br />

Bastiaansen, C. W. M. Langmuir 2003, 19, 10957.<br />

(19) microdrop GmbH, Norderstedt, Germany; http://<br />

www.microdrop.de.<br />

Figure 1. Drying pattern <strong>of</strong> an inkjet printed droplet <strong>of</strong> a<br />

1 wt % solution <strong>of</strong> polystyrene in acetophenone on clean glass.<br />

the surface topography. The setup uses an external 100 W xenon<br />

cold light source <strong>and</strong> a 20× objective.<br />

Results <strong>and</strong> Discussion<br />

Figure 1 shows a typical drying pattern <strong>of</strong> a drop (i.e.,<br />

consisting <strong>of</strong> 5 droplets printed on top <strong>of</strong> each other) <strong>of</strong> a<br />

1.0 wt % solution <strong>of</strong> polystyrene (Mn 80 kD; Mw 282 kD)<br />

in acetophenone on an uncoated, cleaned glass substrate.<br />

The viscosity <strong>of</strong> the solution is 3.5 mPa s. Despite its high<br />

boiling point, the droplet dries within a few seconds, due<br />

to the high surface-to-volume ratio. The pattern consists<br />

<strong>of</strong> a series <strong>of</strong> rings that originate from the increase in<br />

surface energy when drying with the contact line fixed.<br />

Eventually, the droplet will contract in a stick-slip manner,<br />

producing the pattern. 11 Changing the solvent does not<br />

improve homogeneity.<br />

Figure 2a shows the deposit that is left by the same<br />

solution <strong>of</strong> polystyrene in acetophenone on a perfluorinated<br />

glass slide. After being dried, a regular polystyrene dot<br />

is formed. Figure 2b shows cross-sections <strong>of</strong> the dot in the<br />

x- <strong>and</strong> y-directions. Its diameter is 86 µm, <strong>and</strong> its height<br />

3.0 µm. The dot is shaped as a volcano; that is, it has a<br />

small crater on top. The experiment was repeated using<br />

a 1.0 wt % solution <strong>of</strong> poly(methyl methacrylate) (Mn 130<br />

kD; Mw 240 kD) in acetophenone, which had a viscosity<br />

<strong>of</strong> 3.9 mPa s. In this case, the diameter <strong>of</strong> the dot is 75 µm,<br />

<strong>and</strong> its height 4.1 µm; that is, the dimensions <strong>of</strong> the<br />

polystyrene <strong>and</strong> the poly(methyl methacrylate) deposits<br />

<strong>and</strong> their shape are similar. This implies that the dominant<br />

factor that determines the deposition process is the solvent<br />

<strong>and</strong> its physical properties, rather than the solute.<br />

To check this assumption, polystyrene solutions were<br />

printed using anisole as a solvent. Anisole (methoxybenzene)<br />

has a boiling point <strong>and</strong> a vapor pressure that are<br />

considerably lower <strong>and</strong> higher, respectively, than those<br />

<strong>of</strong> acetophenone. The anisole dot is similar to the acetophenone<br />

dot, although the crater is more prominent.<br />

The shape <strong>of</strong> the deposit changes completely when using<br />

a low-boiling solvent like ethyl acetate. Figure 3a shows<br />

the deposit that is left by a drop (i.e., 5 printed droplets)<br />

<strong>of</strong>a1wt%solution <strong>of</strong> polystyrene in ethyl acetate. This<br />

solution has a viscosity <strong>of</strong> 0.68 mPa s. A ring stain that<br />

marks the perimeter <strong>of</strong> the original drop is formed. The<br />

cross-section displayed in Figure 3b shows that indeed<br />

most <strong>of</strong> the polymeric material is present at the ring rather<br />

than at the center. In principle, this effect can be used to


<strong>Inkjet</strong> <strong>Printing</strong> <strong>of</strong> <strong>Well</strong>-<strong>Defined</strong> <strong>Polymer</strong> <strong>Dots</strong> Langmuir, Vol. 20, No. 18, 2004 7791<br />

Figure 2. (a) Volcano-shaped polymer dot, formed by an inkjet<br />

printed droplet <strong>of</strong>a1wt%solution <strong>of</strong> polystyrene in acetophenone<br />

on perfluorinated glass; (b) cross-sections in the x- <strong>and</strong><br />

y-directions.<br />

Figure 3. (a) <strong>Polymer</strong> dot, formed by a droplet <strong>of</strong> a1wt%<br />

solution <strong>of</strong> polystyrene in ethyl acetate on perfluorinated glass;<br />

(b) cross-sections in the x- <strong>and</strong> y-directions.<br />

print thin rings or lines with a resolution better than the<br />

limit that is set by the nozzle diameter. 20<br />

Drop-casting experiments showed that on the perfluorinated<br />

surface the contact line <strong>of</strong> a droplet <strong>of</strong> this solution<br />

<strong>of</strong> polystyrene in ethyl acetate is pinned, whereas the<br />

contact lines <strong>of</strong> acetophenone <strong>and</strong> anisole solution droplets<br />

are at least initially unpinned. This may be due to self-<br />

(20) Cuk, T.; Troian, S. M.; Hong, C. M.; Wagner, S. Appl. Phys. Lett.<br />

2000, 77, 2063.<br />

Figure 4. (a) <strong>Polymer</strong> dot, formed by a droplet <strong>of</strong> a1wt%<br />

solution <strong>of</strong> polystyrene in an 80/20 wt % ethyl acetate/<br />

acetophenone mixture on perfluorinated glass; (b) cross-sections<br />

in the x- <strong>and</strong> y-directions.<br />

pinning <strong>of</strong> the ethyl acetate droplet: The vapor pressure<br />

Pv <strong>of</strong> ethyl acetate, to which the rate <strong>of</strong> evaporation is<br />

proportional, is 1 order <strong>of</strong> magnitude larger than the vapor<br />

pressure <strong>of</strong> anisole, which is again 1 order <strong>of</strong> magnitude<br />

larger than the vapor pressure <strong>of</strong> acetophenone (see Table<br />

1). A ring-like deposit that pins the contact line is therefore<br />

more easily formed. In addition, the contact angle θc <strong>of</strong><br />

ethyl acetate is smaller than the contact angle <strong>of</strong> anisole,<br />

which is again smaller than the contact angle <strong>of</strong> acetophenone.<br />

Near the contact line, the evaporation flux<br />

J is proportional to (R-r) -λ , with λ ) (π - 2θc)/(2π - 2θc),<br />

where r is the distance from the edge toward the center<br />

<strong>of</strong> the droplet <strong>and</strong> R is its radius. λ increases with<br />

decreasing θc, <strong>and</strong> therefore the flux. 9 Viscosity effects<br />

cannot explain the observed behavior. The viscosity <strong>of</strong> the<br />

acetophenone solution is considerably higher than the<br />

viscosity <strong>of</strong> the ethyl acetate-based solution. A large<br />

acetophenone dot as compared to ethyl acetate is therefore<br />

expected, but the opposite is found.<br />

The effect <strong>of</strong> mixing the two solvents is particularly<br />

interesting. A typical result, obtained with a mixture <strong>of</strong><br />

80% ethyl acetate <strong>and</strong> 20% acetophenone by weight, is<br />

shown in Figure 4a. Figure 4b displays the corresponding<br />

cross-sections. Again, a regular polystyrene dot is formed,<br />

with a height <strong>of</strong> 6.2 µm, that is, twice as high as the pure<br />

acetophenone dot, <strong>and</strong> with a diameter <strong>of</strong> 68 µm. Assuming<br />

a cylindrical shape, it follows that the total amount <strong>of</strong><br />

deposited material must be equal in both cases, as it<br />

should, as both concentration <strong>and</strong> number <strong>of</strong> deposited<br />

droplets are equal in both cases. A striking difference<br />

between the pure acetophenone deposit <strong>and</strong> the deposit<br />

left by the mixture is the absence <strong>of</strong> a crater on top in the<br />

case <strong>of</strong> the latter. Perhaps most surprising is the fact that<br />

a small amount <strong>of</strong> acetophenone already suffices to change<br />

the shape <strong>of</strong> the deposit from ring-like to dot-like, that is,<br />

to depin the contact line. <strong>Dots</strong>, identical to that shown in<br />

Figure 4a, were already obtained with mixtures containing<br />

10% acetophenone by weight.


7792 Langmuir, Vol. 20, No. 18, 2004 de Gans <strong>and</strong> Schubert<br />

Figure 5. Dot height as a function <strong>of</strong> polymer molecular weight,<br />

using monodisperse polystyrene st<strong>and</strong>ards. The inset shows<br />

the droplet mass (0) <strong>and</strong> the specific viscosity (2) as a function<br />

<strong>of</strong> molecular weight. A power law was fitted to the viscosity<br />

data, the exponent <strong>of</strong> which is 0.77.<br />

To underst<strong>and</strong> the origin <strong>of</strong> this phenomenon, the<br />

contact angle <strong>of</strong> the 80/20% w/w ethyl acetate/acetophenone<br />

solution mixture on the perfluorinated surface was<br />

measured. It was found that the (initial) contact angle θc<br />

(t ) 0) is 41.5 ( 1.0°, which is very close to the value <strong>of</strong><br />

pure ethyl acetate, that is, 37.4°. This proves that the<br />

results cannot be explained by a change in contact angle.<br />

We then realized that the process <strong>of</strong> droplet evaporation<br />

<strong>and</strong> solute deposition depends on the composition <strong>of</strong> the<br />

droplet at the contact line rather than the overall<br />

composition <strong>of</strong> the droplet. When using a mixture <strong>of</strong> a<br />

low- <strong>and</strong> a high-boiling solvent, the composition at the<br />

contact line will shift toward a higher fraction <strong>of</strong> highboiling<br />

solvent than in the bulk, due to the increased rate<br />

<strong>of</strong> evaporation at the edge. Therefore, the rate <strong>of</strong> evaporation<br />

at the contact line decreases, <strong>and</strong> a surface tension<br />

gradient is established. A flow will be induced from regions<br />

with low to regions with high surface tension when the<br />

Marangoni number M ) ∆γL/ηD is sufficiently large. 21<br />

Here, ∆γ denotes the surface tension difference, L is the<br />

length scale involved, η is the viscosity, <strong>and</strong> D is the<br />

diffusion coefficient. Taking for ∆γ the difference between<br />

the bulk values <strong>of</strong> ethyl acetate <strong>and</strong> acetophenone, <strong>and</strong><br />

using typical values for L, η, <strong>and</strong> D, it follows that M is<br />

<strong>of</strong> the order 10 6 . This result shows that even a very small<br />

concentration gradient will be sufficient to cause a<br />

Marangoni flow, which will homogenize the liquid droplet<br />

<strong>and</strong> reduce the concentration gradient between the contact<br />

line <strong>and</strong> the bulk. 16<br />

To gain insight into the process <strong>of</strong> dot formation, we<br />

studied the height <strong>of</strong> the dots as a function <strong>of</strong> molecular<br />

weight <strong>of</strong> the dissolved polymer, using solutions containing<br />

1.0% monodisperse polystyrene by weight in acetophenone.<br />

One single substrate was used on which all solutions<br />

were printed as 1 × 10 arrays. In addition, the reduced<br />

viscosities <strong>of</strong> the polymer solutions were measured, <strong>and</strong><br />

their contact angles on the fluorinated substrate were<br />

used. The mass <strong>of</strong> the droplets as generated by the inkjet<br />

printer was measured as follows: The total amount <strong>of</strong><br />

printed material was collected during a certain period <strong>of</strong><br />

time, weighed, <strong>and</strong> divided by the number <strong>of</strong> droplets.<br />

The results are shown in Figure 5. It should be emphasized<br />

that, although the errors in the dot height are generally<br />

small (1-3.5%), the errors between different surfaces are<br />

much larger (up to 30%). As for the three highest molecular<br />

weight samples, the dot height shows very little variation<br />

(21) Pesach, D.; Marmur, A. Langmuir 1987, 3, 519.<br />

Figure 6. (a) Array <strong>of</strong> polymer dots printed at a mutual distance<br />

<strong>of</strong> 150 µm, formed by droplets <strong>of</strong>a1wt%solution <strong>of</strong> polystyrene<br />

in acetophenone. The typical size <strong>of</strong> the dots shown here is<br />

29 ( 2 µm, which is smaller than the nozzle diameter; (b) crosssection<br />

in the x-direction, corresponding to the line in<br />

Figure 6a.<br />

with molecular weight. However, a jump in dot height<br />

occurs between 34 <strong>and</strong> 64 kD. The inset <strong>of</strong> Figure 5 shows<br />

that the mass <strong>of</strong> the droplets decreases with molecular<br />

weight. The specific viscosity, also shown in the inset,<br />

increases with molecular weight. A power law, fitted to<br />

these data, yields [η]c ∝ Mw 0.77 , in perfect agreement with<br />

the prediction <strong>of</strong> Zimm for polymer chains in a good<br />

solvent. 22 Finally, it was found that within the experimental<br />

error the presence <strong>and</strong> the molecular weight <strong>of</strong><br />

the polymer do not affect the contact angle. It is expected<br />

that dot height decreases with viscosity as the dot diameter<br />

increases. Furthermore, it is expected that the dot height<br />

scales linearly with droplet size, that is, with the cube<br />

root <strong>of</strong> the droplet mass, which also decreases with<br />

molecular weight. However, the reason for the discontinuous<br />

decrease <strong>of</strong> droplet height with molecular weight<br />

remains unclear.<br />

For most applications, the formation <strong>of</strong> dot arrays will<br />

be <strong>of</strong> interest. Figure 6a shows a dot array that was<br />

obtained by printing a rectangular array <strong>of</strong> single,<br />

acetophenone-based droplets at a mutual distance <strong>of</strong> 150<br />

µm on a perfluorinated surface. Figure 6b displays a<br />

corresponding cross-section. Here, the average dot diameter<br />

is 29 µm, <strong>and</strong> no craters can be observed. This could<br />

be due to different evaporation conditions when printing<br />

an array <strong>of</strong> closely spaced droplets. Similar results were<br />

obtained by Shimoda et al., who studied shape differences<br />

<strong>of</strong> polyLED dots: Crater formation decreased with increasing<br />

velocities. 15 The effect was attributed to partial<br />

drying before the droplet hits the surface. However,<br />

Duineveld et al. have shown that droplet mass reduction<br />

during flight is negligible, assuming typical values for<br />

traveling distance, droplet size, <strong>and</strong> solvent vapor pressure.<br />

23 Both our <strong>and</strong> Shimoda’s results can be explained<br />

(22) Doi, M.; Edwards, S. F. The Theory <strong>of</strong> <strong>Polymer</strong> Dynamics; Oxford<br />

University Press: Oxford, 1986; p 144.


<strong>Inkjet</strong> <strong>Printing</strong> <strong>of</strong> <strong>Well</strong>-<strong>Defined</strong> <strong>Polymer</strong> <strong>Dots</strong> Langmuir, Vol. 20, No. 18, 2004 7793<br />

assuming a correlation between rate <strong>of</strong> evaporation <strong>and</strong><br />

dot shape: Slow evaporation occurs when the time for<br />

printing is small as compared to drying times (i.e., at high<br />

printing velocities), as the droplets evaporate in an<br />

atmosphere with a high partial solvent pressure, due to<br />

evaporation <strong>of</strong> neighboring droplets. This prevents crater<br />

formation.<br />

Minimum spacing at which dots can be printed without<br />

coalescence <strong>of</strong> the droplets on the surface is at present<br />

about 125 µm. It is seen from Figure 6a that the deposition<br />

accuracy is far worse than its theoretical limit, which is<br />

set by the positioning accuracy <strong>of</strong> the print head, that is,<br />

3 µm. This is probably due to r<strong>and</strong>om motion <strong>of</strong> the droplets<br />

during evaporation. Confinement <strong>of</strong> the droplets, either<br />

mechanically or by surface energy patterning, will be<br />

necessary to produce regular arrays. Finally, it is interesting<br />

to note that, by using dilute solutions, structures<br />

can be created <strong>of</strong> size smaller than the diameter <strong>of</strong> the<br />

nozzle (29 vs 70 µm). This may prove to be a potential<br />

strategy for the inkjet printing <strong>of</strong> defined submicrometer<br />

structures.<br />

Conclusions<br />

We have inkjet printed well-defined (arrays <strong>of</strong>) polymer<br />

dots on a hydrophobic, perfluorinated substrate using a<br />

nonvolatile solvent such as acetophenone. Different<br />

polymers yield qualitatively identical deposits, suggesting<br />

that the choice <strong>of</strong> the solvent may be the key factor<br />

determining the deposition process. When a volatile<br />

solvent such as ethyl acetate is used, ring-like deposits<br />

are formed. Drop-casting experiments show that the<br />

(23) Duineveld, P. C.; De Kok, M. M.; Buechel, M.; Sempel, A. H.;<br />

Mutsaers, K. A. H.; Van de Weijer, P.; Camps, I. G. J.; Van den Biggelaar,<br />

T. J. M.; Rubingh, J.-E. J. M.; Haskal, E. I. Proc. SPIE Vol. 4464; SPIE:<br />

Bellingham, 2002; p 59.<br />

contact line <strong>of</strong> acetophenone droplets on the surface used<br />

is unpinned, whereas the contact line <strong>of</strong> ethyl acetate<br />

droplets is not, which may explain the results. Mixing a<br />

small amount <strong>of</strong> acetophenone, that is, 10 wt % or more,<br />

with ethyl acetate changes the shape <strong>of</strong> the deposit from<br />

ring-like to dot-like, probably due to a local increase <strong>of</strong> the<br />

acetophenone concentration at the contact line. Investigation<br />

<strong>of</strong> the dot height as a function <strong>of</strong> molecular weight<br />

using monodisperse polystyrene samples showed a discontinuous<br />

decrease with molecular weight. Simultaneously,<br />

the mass <strong>of</strong> the ejected droplets decreases with<br />

molecular weight, whereas the solution viscosity increases<br />

<strong>and</strong> the contact angle <strong>of</strong> the droplets remains constant.<br />

Thus, none <strong>of</strong> these three factors provides an explanation<br />

for the observed behavior. Finally, arrays <strong>of</strong> dots were<br />

printed. The size <strong>of</strong> the dots that are printed is considerably<br />

smaller than the nozzle diameter. This may represent a<br />

potential strategy for the inkjet printing <strong>of</strong> submicrometer<br />

structures.<br />

Acknowledgment. We thank Carlos Sanchez for his<br />

help in preparing perfluorinated substrates, Caroline<br />

Abeln for the GPC measurements, Pr<strong>of</strong>. Bert de With<br />

(Laboratory <strong>of</strong> Coatings Technology, TU/e) for the use <strong>of</strong><br />

the confocal scanning microscope, Niek Lousberg <strong>and</strong><br />

Alper Tiftikci for sharing their expertise, <strong>and</strong> Wilhelm<br />

Meyer from Microdrop for fruitful collaboration. This work<br />

is part <strong>of</strong> the Dutch <strong>Polymer</strong> Institute (DPI) research<br />

program (Project 400).<br />

Supporting Information Available: Scanning confocal<br />

micrograph <strong>of</strong> poly(methyl methacrylate) dot <strong>and</strong> corresponding<br />

cross-sections. Scanning confocal micrograph <strong>and</strong> corresponding<br />

cross-sections <strong>of</strong> polystyrene dot, printed from anisole solution.<br />

This material is available free <strong>of</strong> charge via the Internet at<br />

http://pubs.acs.org.<br />

LA049469O

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

Saved successfully!

Ooh no, something went wrong!