Sie sind auf Seite 1von 248

Process Development for Inkjet Printing of Organic

Photovoltaics

Prozessentwicklung für den Tintenstrahldruck


organischer Photovoltaik

Der Technischen Fakultät

der Friedrich-Alexander-Universität

Erlangen-Nürnberg

zur

Erlangung des Doktorgrades Dr.-Ing.

vorgelegt von

M.Sc. Philipp Maisch

aus Bad Neustadt a.d. Saale


Als Dissertation genehmigt von

der Technischen Fakultät

der Friedrich-Alexander-Universität Erlangen-Nürnberg

Tag der mündlichen Prüfung: 25.02.2019

Vorsitzender des Promotionsorgans: Prof. Dr.-Ing. Reinhard Lerch

1. Gutachter: Prof. Dr. Christoph J. Brabec

2. Gutachter: Prof. Dr.-Ing. Jörg Franke


Zusammenfassung
Aufgrund ihrer Flexibilität, Semitransparenz und potenziell günstiger
Herstellungsverfahren erfreut sich die organische Photovoltaik (OPV) nach wie vor eines
hohen Interesses in Wissenschaft und Industrie. Diese einzigartigen Eigenschaften
ermöglichen Anwendungen, wie z.B. die Integration in Kleidung, „Wearables“ oder
Gebäude, die der marktbeherrschenden Siliziumphotovoltaik weitgehend verschlossen
bleiben. Da bereits viele der gesetzten Meilensteine bei der Technologieentwicklung, wie
z.B. Wirkungsgrade von > 15 %,1 erreicht wurden, gilt es nun, einen industriell
relevanten Herstellungsprozess zu entwickeln, der perfekt auf die speziellen
Anwendungen zugeschnitten ist. Zu den Anforderungen an einen solchen Prozess
gehören unter anderem ein (i) hoher Produktionsdurchsatz, (ii) niedrige
Produktionskosten, (iii) effiziente Materialnutzung, die freie Wahl von (iv) Farbe und (v)
Form der Solarzellen, die (vi) Möglichkeit, diese schnell und einfach zu ändern sowie die
(vii) Beschichtung auf nahezu beliebigen Oberflächen. Da keine der etablierten OPV-
Produktionsmethoden im Stande ist dies zu leisten, gilt es neue Herstellungswege zu
erschließen. Speziell der Drop-on-Demand (DOD) Tintenstrahldruck erfüllt alle oben
aufgeführten Bedingungen. Moderne Druckmaschinen erreichen
Durchsatzgeschwindigkeiten von mehr als 1000 m²/h.2 Gleichzeitig erlaubt das gezielte
Aufbringen jedes einzelnen pikolitergroßen Tintentropfens eine sehr effiziente
Materialnutzung und direkte Strukturierung der aufgebrachten Schichten in jeglicher
Form.
Ziel dieser Arbeit ist es daher, einen vollständigen Prozess für den Tintenstrahldruck von
OPV auf industrierelevantem Maßstab zu entwickeln. Dies umfasst die Tintenherstellung,
den Druck funktioneller Schichten sowie die Produktion vollständig gedruckter
großflächiger Freiformsolarmodule.
Im ersten Teil der Arbeit (CHAPTER IV:) werden die notwendigen Voraussetzungen für
den Druck organischer Photovoltaik geschaffen. Dabei stehen die Tintenherstellung sowie
Druckbarkeitsuntersuchungen im Vordergrund. Durch Anwendung der Ohnesorgetheorie
werden stabile Elektroniktinten entwickelt, die ohne den Einsatz von
umweltgefährdenden halogenierten Lösemitteln einen satellitenfreien Druck erlauben.
Dies umfasst unter anderem eine alkoholbasierte Tinte für die Herstellung von
semitransparenten Silbernanodrahtelektroden (Englisch: Silver nanowire - AgNW), durch
deren Einsatz zum ersten Mal der Tintenstrahldruck von Nanodrähten mit einer Länge

i
von ~30 µm durch Düsen mit vergleichbarem Durchmesser demonstriert wird. Die
alkoholbasierten Tinten sind umweltfreundlich und eignen sich bestens für die
Massenproduktion gedruckter OPV. Durch den Verzicht auf Wasser ist auch ein Einsatz
in Perowskit-Solarzellen denkbar.
Ein weiterer wichtiger Aspekt bei der Herstellung von gedruckten Solarzellen ist die
Schichthomogenität. Um diese auch auf schwierig zu benetzenden Oberflächen, wie z.B.
der hydrophoben Aktivschicht, sicherzustellen, wird das Benetzungsverhalten von
verschiedenen Substraten und Tinten charakterisiert und durch den Einsatz von Tensiden
gezielt beeinflusst. Praktische Versuche zeigen jedoch, dass dieser Ansatz nicht immer zu
einer vollständigen Benetzung führt. Mit dem Ziel, auf nahezu beliebigen Oberflächen
homogen drucken zu können, wird daher eine alternative Strategie basierend auf
tintenstrahlgedruckten Ankerpunkten entwickelt. Diese sogenannten „Pinningzentren“
halten die Flüssigschicht auf dem Substrat in Position und verhindern so deren
Zusammenziehen oder Aufreißen. Durch lokale Oberflächenmodifikation mit
Pinningzentren lassen sich benetzende sowie entnetzende Areale auf der Oberfläche
definieren, was die Strukturierung von aufgebrachten Schichten mit einer Genauigkeit
von ~25 µm erlaubt.
Der zweite Teil der Arbeit (CHAPTER V:) beschreibt die Entwicklung von stabilen und
hochskalierbaren Druckprozessen für jede einzelne Lage der OPV-Multischichtstruktur.
Beginnend mit einer gut funktionierenden Referenzzelle wird der Übergang zu einer
vollständig tintenstrahldrucken Solarzelle durch sukzessives Ersetzen der einzelnen
Schichten demonstriert. Als semitransparente Elektroden werden gedruckte Silberstege
bzw. AgNW Schichten eingesetzt. Solarzellen mit einer geometrieoptimierten gedruckten
Silberstegbodenelektrode erreichen Wirkungsgrade von bis zu 5,66 %. Dieser Wert ist
deutlich höher als bisher in der Literatur publizierte Effizienzen ähnlicher Strukturen.
Noch bessere Ergebnisse werden durch den Einsatz digital gedruckter AgNW-Schichten
erzielt. Die visuell homogenen Elektroden zeigen ein hervorragendes Verhältnis von
Schichtwiderstand und Transparenz (𝑅𝑠𝑞 < 20 Ω/sq @ 𝑇 > 90 %). Solarzellen mit AgNW-
Kathode und -Anode erreichen Wirkungsgrade von bis zu 4,3 % und halten damit bisher
die Rekordeffizienz für vollständig tintenstrahlgedruckte OPV.
Der dritte Teil der Arbeit befasst sich mit einem Thema, welches von der Literatur kaum
behandelt wird. In diesem Kapitel (CHAPTER VI:) wird der Übergang von einzelnen
kleinen Solarzellen zu großflächig, vollständig tintenstrahlgedruckten Solarmodulen

ii
beschrieben. Hierbei werden erstmals Module mit vier gedruckten Schichten
demonstriert, welche über einen geometrischen Füllfaktor (𝐺𝐹𝐹 - Verhältnis von aktiver
Modulfläche zu gesamter Modulfläche) von 85 % und Wirkungsgraden bis zu 4,3 %
verfügen. Des Weiteren wird eine neue Methode zur monolithischen Serienverschaltung
einzelner Solarzellen zum Modul eingeführt, welche ohne visuell störende Lücken in der
Aktivschicht auskommt. Dies geschieht durch das Aufbringen von tintenstrahlgedruckten
„Silberbrücken“ auf die Bodenelektroden, welche die anschließend aufgebrachten
Schichten durchstoßen und so einen optisch unauffälligen Kontakt zwischen
benachbarten Zellen herstellen. Durch Einsatz dieser Technologie in Kombination mit
gedruckten AgNW-Elektroden sowie verschiedenfarbigen Aktivschichtmaterialien
werden erstmals tintenstrahlgedruckte semitransparente Solarmodule mit freier Farb- und
Formgestaltung auf Flächen von bis zu 150 cm² demonstriert.
Das letzte Unterkapitel der Arbeit beschreibt den Transfer der Drucktechnologie vom
Labormaßstab auf eine „Single-Pass“-Druckmaschine, welche im Rahmen einer
Industriekooperation entwickelt wurde. Die Maschine verfügt über vier Druckstationen
mit je vier Druckköpfen und Tintenzirkulation sowie Heißluft- und Infrarotöfen. Dies
ermöglicht den Druck aller Schichten der Solarzelle mit Durchsatzgeschwindigkeiten von
bis zu 5 m²/min. Vollständig gedruckte Solarmodule mit einer aktiven Fläche von 10 cm²
erzielen Wirkungsgrade von bis zu 4,7 % und übertreffen damit sogar noch den zuvor
aufgestellten Effizienzrekord der Einzelzellen mit AgNW-Kathode und -Anode.

iii
Abstract
Due to the exceptional properties, such as flexibility and semitransparency, organic
photovoltaics (OPV) is a topic of high interest in science and industry. These stand-alone
characteristics enable applications like the integration in cloths, wearables or even
buildings, which are out of reach for the market dominating silicon photovoltaics. As
many of the OPV technology milestones, such as efficiencies > 15 %,1 have already been
reached, industrially relevant production processes which perfectly suit the particular
applications have to be developed. Requirements for such process are (i) a high
throughput, (ii) low production costs, (iii) efficient use of material, (iv) free choice of
colour and (v) form, (vi) the option for fast and easy layout changes and (vii) the
possibility to coat on almost any substrate surface. As none of the established OPV
fabrication processes is able to fulfil these conditions, new processing techniques have to
be developed. In this context, Drop-on-Demand (DOD) inkjet printing seems to be a very
promising approach because it completely satisfies the above specified processing
requirements. Modern printing machines reach throughput speeds of more than
1000 m²/h.2 At the same time, the deposition of every single picoliter-sized droplet allows
extraordinarily efficient use of raw materials as well as direct patterning of the layers in
any arbitrary shape.
Therefore, the aim of this thesis is to develop a sophisticated inkjet process for
manufacturing of OPV on an industrially relevant scale. This comprises the ink
formulation, inkjet deposition of functional layers and the demonstration of fully printed
large area free-form solar modules.
In the first part of the work (CHAPTER IV:), the necessary foundation for digital printing
of OPV is provided. This includes the development of electronic inks and printing studies
on manufacturing of homogeneous layers. By applying the Ohnesorge theory, stable inks
are designed, which offer satellite-free drop formation. This comprises an alcohol based
silver nanowire (AgNW) ink, which is applied for manufacturing of semitransparent
electrodes. The successful printing is the first time proof that nanowires with a length of
~30 µm can be ejected through inkjet nozzles in the same size range. Furthermore, the
alcohol based inks are environmentally non-critical and perfectly suited for the mass
production of OPV. The water-free nature of the inks also enables the application in
perovskite solar cells.

iv
Another highly important aspect of OPV fabrication is the layer homogeneity. To be able
to print defect-free wet films also on low energy surfaces, such as the highly hydrophobic
polymer active layers, the wetting properties of inks and substrates are investigated and
modified with surface active agents. However, as shown by experiments, this strategy
does not always result in defect-free layers and well working solar cells. With the aim to
provide a process that allows printing of homogeneous wet films on almost every surface,
an alternative strategy, relying on inkjet printed anchoring points, is developed. These so-
called ‚pinning centres‘ fix the wet film at the desired position on the substrate, thus
preventing contraction or rupturing. Local modification of the substrate surface with
pinning centres results in a precise spatial definition of wetting as well as dewetting areas.
This allows convenient layer patterning with a resolution of ~6 µm.
The second part of the work (CHAPTER V:) describes the development of stable and
scalable printing processes for every single layer of the OPV structure. Starting with a
well working reference device, stepwise replacement of the single blade coated layers
with inkjet printed equivalents is demonstrated. Inkjet printed silver grids or AgNW
layers are applied as semitransparent electrodes. Solar cells with an optimised silver grid
base electrode reach efficiencies up to 5.66 %. This value is significantly higher than
literature reports of similar structures. Even better results are achieved by applying
digitally printed AgNW mesh electrodes. The visually homogeneous layers have an
excellent balance of sheet resistance and transmittance (𝑅𝑠𝑞 < 20 Ω/sq @ 𝑇 > 90 %).
Devices with AgNW-cathode and -anode reach efficiencies of 4.3 %, which is up to now
the record for fully inkjet printed OPV.
The third part of the work (CHAPTER VI:) describes the transition from small scale solar
cells to large area fully inkjet printed solar modules. For the first time, solar modules with
four inkjet printed layers are demonstrated, reaching a geometrical fill factor (𝐺𝐹𝐹 - ratio
of active module area to total module area) of 85 % and efficiency of 4.3 %. Furthermore,
a novel strategy to realise the monolithical cell-to-cell interconnection without visually
obstructive gaps in the active layer is introduced. This is achieved by inkjet printing of
‘silver bridges’, which penetrate subsequently applied layers, thereby forming a visually
inconspicuous contact between top and base electrodes of adjacent cells. Applying this
technology in combination with inkjet printed AgNWs and differently coloured active
materials, solar modules of unique design with sizes up to 150 cm² are demonstrated for
the first time.

v
The last subchapter of the work describes the transfer of the developed processes from
laboratory scale to a high-throughput single pass inkjet printer. The machine is equipped
with four print stations, which contain four printheads each, ink circulation, hot air- as
well as infrared (IR) drying stations and enables the deposition of all solar cell layers at
throughput speeds of up to 5 m²/min. Fully inkjet printed solar modules with an active
area of 10 cm² reach efficiencies of up to 4.7 %, which is even higher than the single cells
with AgNW electrodes (4.3 %) that held up to now the efficiency record for fully inkjet
printed devices.

vi
Acknowledgements
First of all, I would like to thank Prof. C.J. Brabec, who gave me the opportunity to do
my PhD under his supervision. His professional guidance and encouragements together
with the stimulating environment of his workgroup allowed me to grow personally and
scientifically.
I would also like to deeply thank Dr. H.-J. Egelhaaf, who guided and mentored me during
this work. The good words of advice and challenging questions he gave me broadened my
horizon. I am also grateful to Dr. M.M. Voigt, who brought me in first contact with this
interesting PhD topic.
Special thanks to H. Scheiber and E. Maier from Durst Phototechnik AG for their
personal support, valuable insights into the field of industrial inkjet printing and the
motivating scientific discussions during our project meetings in the beautiful city of
Lienz.
I would also like to thank all the colleagues at ZAE Bayern, EnCN and i-MEET. I want to
express my special gratitude to K.C. Tam for his friendship as well as his selfless and
responsible working attitude, which impressed and inspired me. Furthermore, I want to
thank L.M. Eisenhofer for her detailed scientific studies about inkjet printed pinning
centres, F.W. Fecher, who taught me how to perform finite element simulations, and S.
Keilwitz and M. Müller for taking the first steps in inkjet printing together with me.
Special thanks also to A. Distler, L. Lucera, P. Kubis, H.D. Schmidt, M. Heyder, I.
Döhrer, F. Hoga, M. Wagner, I. Levchuck, C.O. Ramírez Quiroz, A. Kidzun, F. Machui,
F. Högl, S. Strohm, T. Ahmad, I. Channa, S. Feroze, A. Karl, M. Steinberger, L. Wendt,
D. Jang, F. Winkler, A. Amin, N. Gawehns and H. Berr for supporting me during the
recent years.
I also want to express my gratitude to Alex, Dominik, Norbert, Hammers, Toni, Flo,
Maggi, Kili, Sonja and Jana. Your friendship and the great times spent together are
important pillars in my life. Sincere thanks also to my whole family, especially my
parents and sister. Your efforts and great confidence in me all through these years were
the main reason of my success. Finally, I owe my deepest thanks to my wife Helene and
our son Julius for their unconditional support, love and the permanent patience.

vii
Table of Contents
Zusammenfassung................................................................................................................. i

Acknowledgements ............................................................................................................vii

Table of Contents ............................................................................................................. viii

List of Tables ...................................................................................................................... xi

List of Figures ................................................................................................................... xiv

List of Abbreviations ....................................................................................................... xxv

List of Symbols and Constants ....................................................................................... xxix

CHAPTER I: Introduction .................................................................................................... 1

1.1. Green energy demand............................................................................................ 2

1.2. Solar technologies ................................................................................................. 3

1.3. New solar cell applications ................................................................................... 4

1.4. Printed electronics ................................................................................................. 6

1.5. Motivation and strategy......................................................................................... 7

CHAPTER II: Theoretical Background ............................................................................ 10

2.1. Organic solar cells ............................................................................................... 11

2.1.1. Working principle ........................................................................................ 11

2.1.2. Current density-voltage characteristics ........................................................ 14

2.1.3. Common materials ....................................................................................... 18

2.1.4. Solar modules............................................................................................... 25

2.1.5. Device fabrication ........................................................................................ 29

2.2. Inkjet technology ................................................................................................. 35

2.2.1. Historical background .................................................................................. 35

2.2.2. Scales of length and time ............................................................................. 36

2.2.3. Inkjet types ................................................................................................... 38

2.2.4. Aspects of ink formulation........................................................................... 41

2.2.5. Actuation principle....................................................................................... 46

viii
2.2.6. Advanced pulse design ................................................................................ 49

2.2.7. Drop formation............................................................................................. 50

2.2.8. Drop impact ................................................................................................. 53

2.3. State of the art - inkjet printed OPV .................................................................... 55

CHAPTER III: Experimental ............................................................................................ 63

3.1. Materials .............................................................................................................. 64

3.2. Fabrication methods ............................................................................................ 66

3.2.1. Reference processes ..................................................................................... 66

3.2.2. PiXDRO LP50 ............................................................................................. 68

3.2.3. Loop coater single-pass prototype inkjet ..................................................... 74

3.2.4. Encapsulation ............................................................................................... 78

3.3. Characterisation ................................................................................................... 79

3.3.1. Ink characterisation ...................................................................................... 79

3.3.2. Layer characterisation .................................................................................. 80

3.3.3. Solar cell and module characterisation ........................................................ 84

3.4. Software tools for simulation and optimisation .................................................. 84

3.4.1. Matlab .......................................................................................................... 84

3.4.2. COMSOL Multiphysics ............................................................................... 85

CHAPTER IV: From Inks to Layers ................................................................................. 91

4.1. Ink formulation .................................................................................................... 92

4.1.1. Ohnesorge theory ......................................................................................... 92

4.1.2. Ink stability and nozzle open time ............................................................... 96

4.2. Droplet control .................................................................................................. 102

4.3. Layer formation ................................................................................................. 104

4.3.1. Wetting....................................................................................................... 106

4.3.2. Pinning centres ........................................................................................... 111

4.4. Conclusion......................................................................................................... 121

ix
CHAPTER V:From Layers to Cells ................................................................................ 125

5.1. Inkjet printed active layers ................................................................................ 126

5.2. Inkjet printed buffer layers ................................................................................ 133

5.3. Inkjet printed electrodes .................................................................................... 140

5.3.1. Silver grid electrodes ................................................................................. 140

5.3.2. Silver nanowire electrodes ......................................................................... 145

5.4. Conclusion......................................................................................................... 151

CHAPTER VI: From Cells to Modules ........................................................................... 155

6.1. Different interconnection strategies .................................................................. 156

6.1.1. Conventional interconnection .................................................................... 156

6.1.2. Shy interconnection ................................................................................... 160

6.2. Arbitrary shape modules ................................................................................... 169

6.3. Large scale production ...................................................................................... 172

6.4. Conclusion......................................................................................................... 181

CHAPTER VII: Summary and Outlook .......................................................................... 184

7.1. Summary ........................................................................................................... 185

7.2. Outlook .............................................................................................................. 189

Bibliography .................................................................................................................... 192

Appendix .......................................................................................................................... 209

List of Publications .......................................................................................................... 211

x
List of Tables
Table II-1: Overview of high impact PSC donor polymers and performance. ............................... 21

Table II-2: Overview of the most relevant semitransparent flexible electrode materials for OPV
applications. Published in reference 27 and reproduced with permission from Wiley VCH. ......... 24

Table II-3: Literature overview and comparison of solution processed organic solar modules.
Ncells is the number of single solar cells in the module. Data partially extracted from 71 with
permission from P. Kubis. .............................................................................................................. 26

Table II-4: Comparison of printing techniques: Ink waste: 1 (none), 2 (little), 3 (some), 4
(considerable), 5 (significant). Pattern: 0 (0-dimensional), 1 (1-dimensional), 2 (2-dimensional), 3
(pseudo/quasi 2/3-dimensional), 4 (digital master). Speed: 1 (very slow), 2 (slow < m*min-1), 3
(medium 1–10 m*min-1), 4 (fast 10–100 m*min-1), 5 (very fast 100–1000 m*min-1). Ink
preparation: 1 (simple), 2 (moderate), 3 (demanding), 4 (difficult), 5 (critical). Ink viscosity: 1
(very low < 10 cP), 2 (low 10–100 cP), 3 (medium 100–1000 cP), 4 (high 1000–10000 cP), 5
(very high 10000–100000 cP). Reproduced from 88 with permission from Elsevier. .................... 30

Table II-5: Literature overview of important studies about inkjet printed OPV. ........................... 56

Table III-1: Materials used in this thesis. ....................................................................................... 64

Table III-2: Spectra S-class printhead specifications as given by the manufacturer.149 ................. 72

Table III-3: List of most important operation parameters at the PiXDRO LP 50. ......................... 73

Table III-4: SAMBA printhead specifications as given by the manufacturer.149 ........................... 76

Table III-5: List of most important operation parameters at the loop coater SAMBA prototype
inkjet. .............................................................................................................................................. 77

Table III-6: Simulation parameters applied in the 2D FEM model for silver grid loss analysis. ... 86

Table III-7: Parameters used in the FEM simulation of the 3-cell module. ................................... 89

Table IV-1: Parameters to qualify printability of inkjet inks. OPV inks which do not offer
satellite-free droplet formation (bold) are modified, and thus shifted into the stable printing
regime (nozzle diameter 35 µm)..................................................................................................... 93

Table IV-2: Rating of different P3HT:PC60BM active ink gelling tendencies (20 mg/ml, 30 °C, no
stirring). The inks are specified as liquid (L) with a grading system ranging from 1-4, to describe
opacity and precipitates (1: homogeneous ink, 4: significant amount of precipitates, gelling). If the
ink gels completely, it is specified as solid (S). Results were partly obtained by H. Scheiber and E.
Maier from Durst Phototechnik and are reported with permission. ............................................... 97

Table IV-3: Open time experiment for different active layer inks (15 mg/ml polymer content)
using a Spectra SE 128 AA printhead with 35 µm nozzle diameter. ........................................... 100

xi
Table IV-4: Open time experiment for different alcohol based ZnO N-10 inks (2.5 %wt
nanoparticles) using a Spectra SE 128 AA printhead with 35 µm nozzle diameter .................... 101

Table IV-5: List of printing parameters that allow stable Spectra SE 128 AA printhead operation.
..................................................................................................................................................... 104

Table IV-6: Surface energies (γS ) of commonly used OPV active layer materials and surface
tensions (γL ) of PEDOT:PSS inks with and without the fluorosurfactant Capstone FS 31 (CFS).
The polar (γP ) and dispersive parts (γD ) were determined following the Owens, Wendt, Rabel
and Kaelble method (OWRK). Each pendant drop and sessile drop measurement was repeated at
least three times. The presented data are mean values. The standard deviation is in all cases
< 0.2 mN/m. Published in reference 184 and reproduced with permission. .................................. 109

Table IV-7: Overview of OPV inks applied in this work and rating based on the critical
requirements for a robust inkjet printing process. Underlined inks were developed or modified in
this work to fulfil the processing requirements. ........................................................................... 122

Table V-1: P3HT:PC60BM solar cell key performance values of representative devices with active
layers blade coated (DB) or inkjet printed (IJ) from different ink formulations. The charge carrier
mobility µ and the bimolecular recombination coefficient k rec,Bimol are extracted from simulated
fit curves using a single diode model as described in chapter 2.1.2. ........................................... 127

Table V-2: Device performance parameters of the PV2000:PC70BM solar cells from o-X:THN 1:1
(vol.) ink shown in Figure V-5..................................................................................................... 132

Table V-3: Device performance parameters of the solar cells shown in Figure V-8. .................. 135

Table V-4: Device performance parameters of the solar cells shown in Figure V-11. ................ 138

Table V-5: P3HT:PC60BM solar cell key performance values of representative devices shown in
Figure V-13. Published in reference 184 and reproduced with permission. .................................. 139

Table V-6: Device performance parameters of the solar cells shown in Figure V-15. ................ 143

Table V-7: Photovoltaic parameters of the fully inkjet printed organic solar cell with AgNW
electrodes. Published in reference 164 and reproduced with permission from Elsevier B.V.. ...... 151

Table V-8: Overview of inkjet printed OPV layers and rating based on the critical requirements
for a robust inkjet printing process. ............................................................................................. 152

Table VI-1: Device performance parameters of the solar cells and module shown in Figure VI-1,
NOTE: PCE and JSC are calculated from active device area ........................................................ 158

Table VI-2: Device performance parameters of the solar cells and module shown in Figure VI-2.
NOTE: PCE and JSC are calculated from active device area. ....................................................... 159

Table VI-3: Device performance parameters of the modules shown in Figure VI-5. NOTE: PCE
and JSC are calculated from active device area. Published in reference 159 and reproduced with
permission of John Wiley and Sons. ............................................................................................ 165

xii
Table VI-4: Simulated device performance parameters of OPV modules with different ρIC . The
resulting values for ρIC = 0.0020 Ω*cm2 (obtained from CKBR measurements) are highlighted.
Published in reference 159 and reproduced with permission of John Wiley and Sons. ................. 167

Table VI-5: Key device performance parameters of free-form OPV modules............................. 171

Table VI-6: Device performance parameters of the solar cells shown in Figure VI-11. .............. 174

Table VI-7: Device performance parameters of the solar cells shown in Figure VI-13 c)........... 176

Table VI-8: Device performance parameters of the solar cells shown in Figure VI-14. .............. 179

Table VI-9: Device performance parameters of the solar cells shown in Figure VI-15. .............. 180

Table VI-10: Overview of inkjet printed OPV layers and rating based on the critical requirements
for a robust inkjet printing process. .............................................................................................. 181

xiii
List of Figures
Figure I-1: Share of energy sources (preliminary information) in gross German power production
in 2017 according to the German Federal Statistical Office – ‘Arbeitsgemeinschaft
Energiebilanzen’ (AGEB). Data extracted from 3............................................................................ 2

Figure I-2: a) German pavilion at the international EXPO 2015 in Milan, Italy showing integrated
organic solar modules from Belectric OPV (now OPVIUS). Reproduced from 21 with permission
from OPVIUS GmbH. b) The rollable OPV HeLi-on compact solar charger from InfinityPV can
be used to charge smartphones. Reproduced from 22 with permission from InfinityPV ApS. c)
Solar art in form of a garden lamp from Belectric OPV. Reproduced with permission from the
Energy Campus Nuremberg and K. Fuchs. d) Plant growth test with semitransparent OPV
modules. Reproduced from 23 with permission from the Energy Campus Nuremberg and F.
Machui. ............................................................................................................................................ 5

Figure I-3: Critical requirements on inks (red), layers (green) and devices (blue) for inkjet printing
of OPV. ............................................................................................................................................ 8

Figure II-1: a) Typical structure of a flexible organic solar cell consisting of a substrate, base
electrode, electron extraction layer (EEL), active layer, hole extraction layer (HEL) and top
electrode (Al: Aluminum, PEDOT:PSS: Poly(3,4-ethylenedioxythiophene) polystyrene sulfonate,
ZnO: Zinc oxide, IMI: ITO-Metal(Ag)-ITO, PET: Polyethylene terephthalate). Published in
reference 27 and reproduced with permission from Wiley VCH. b) Energy band diagram of a
normal structure and c) inverted structure OPV cell. Data extracted from 28. ............................... 11

Figure II-2: From light absorption to photocurrent in a bulk heterojunction solar cell. a) From a
kinetic point of view. b) Simplified energy diagram (binding energies for excitons and polaron
pairs are not shown). Numbers i-vi correspond to the six characteristic steps of the energy
harvesting process as described above. Reproduced from 31 with permission from IOP Publishing.
....................................................................................................................................................... 13

Figure II-3: Typical J-V characteristics of an organic solar cell (P3HT:PC60BM measured under 1
sun illumination). Published in reference 27 and reproduced with permission from Wiley VCH. . 14

Figure II-4: Equivalent circuit model of a solar cell. a) Simplified structure. b) Standard model
with R S and R P . Redrawn from 42. ................................................................................................. 16

Figure II-5: Logarithmic plot of J-V curve measured in the dark (P3HT:PC60BM). ..................... 18

Figure II-6: Top: Chemical structures of frequently used donor and acceptor molecules. a)
PCPDTBT: poly[2,6-(4,4-bis(2-ethylhexyl)-4H-cyclopenta[2,1-b;3,4-b´]-dithiophene)-alt-4,7-
(2,1,3-benzothiadiazole)]. b) P3HT. c) PC60BM. Reproduced from 48 with permission from A.
Distler. d) Comparison between solar spectrum and the photoresponse of a P3HT:PC60BM solar
cell. Reproduced from 49 with permission from Springer Nature. ................................................. 19

Figure II-7: Contour plot showing the calculated energy-conversion efficiency (contour lines and
colors) versus the band gap and the LUMO level of the donor polymer according to a model
described in.50 Straight lines starting at 2.7 eV and 1.8 eV indicate HOMO levels of –5.7 eV and –
4.8 eV, respectively. A schematic energy diagram of a donor PC60BM system with the band gap

xiv
energy (Eg ) and the energy difference (ΔE) is also shown. Reproduced from 50 with permission
from John Wiley and Sons. ............................................................................................................ 20

Figure II-8: a) Typical graph of transmittance (generally measured at 550 nm) plotted versus sheet
resistance for thin films of nanostructured materials. Note that each curve is divided into two
regimes, the bulk regime (solid line is a fit to equation 12) and the percolation regime (dashed line
is a fit to equation 14). Reproduced from 63 with permission from Cambridge University Press. b)
Plot of sheet resistance versus number of bends for CuNW films (85 % transparent) and ITO on
PET foil. Inset shows the radius of curvature before (10 mm) and after bending (2.5 mm).
Reproduced from 64 with permission from John Wiley and Sons. ................................................. 23

Figure II-9: Sketch of a common solar module layout with three cells connected in series. ......... 25

Figure II-10: a) Flexible OPV solar module fabricated by slot die coating of stripes with a GFF of
80 %. b) Module fabricated by coating of extensive layers and LASER patterning with a GFF
> 95 %. Reproduced from 87 with permission from Elsevier. c) Microscope image of the
interconnection region showing the P1, P2 and P3 LASER structuring line. Reproduced from 85
with permission from the Royal Society of Chemistry. ................................................................. 29

Figure II-11: Schematic illustration of a a) spin coating and b) doctor blading process. b) is
reproduced from reference 19 with permission from L. Lucera. ..................................................... 31

Figure II-12: Schematic illustration of the R2R techniques a) slot die coating, b) rotary screen
printing, c) gravure printing and d) flexographic printing for solar cell fabrication as described in
references 88,92 (redrawn). ............................................................................................................... 33

Figure II-13: The Siphon recorder is the first practical continuous inkjet device. It was used for
automatic recordings of telegraph messages and invented by William Thomson. Reproduced from
98
with permission from Elsevier. ................................................................................................... 36

Figure II-14: Length scales involved in inkjet printing. Reproduced from 104 with permission from
Wiley VCH. .................................................................................................................................... 37

Figure II-15: Time scales involved in inkjet printing. Reproduced from 104 with permission from
Wiley VCH. .................................................................................................................................... 37

Figure II-16: Different types of inkjet printers as described in 98,107 (redrawn). ............................ 38

Figure II-17: Schematic illustration of continuous and drop-on-demand inkjet printing as


described in 92,104 (redrawn). ........................................................................................................... 39

Figure II-18: Schematic illustration of the most common DOD inkjet types a) thermal and b)
piezo as described in 98,104 (redrawn). ............................................................................................. 40

Figure II-19: A schematic diagram showing the operating regime for stable operation of DOD
inkjet printing. The diagram uses the Ohnesorge number as the ordinate axis in place of the
Weber number We = (Re ∗ Oh)2 . The criterion for a drop to possess sufficient kinetic energy to
be ejected from the nozzle is given by Derby as Wecrit ≥ 4.111 The criterion for onset of splashing

xv
following impact is given by Derby as Oh ∗ Re5/4 ≥ 50. Reproduced from 112 with permission
from AIP Publishing. ..................................................................................................................... 42

Figure II-20: Schematic of a liquid drop showing the quantities in the Young equation. ............. 43

Figure II-21: Schematic illustration of polar and dispersive part for two phase contact model
according to 117 (redrawn). ............................................................................................................. 44

Figure II-22: Wetting envelope of a 50 J/m² surface for a perfect wetting condition (θ = 0°).
Liquids with polar and dispersive surface tension coordinates inside the wetting envelope are
predicted to wet perfectly. For coordinates outside the envelope lower wettability is expected. .. 45

Figure II-23: Piezoelectric drop generation pressure waves as described by references 98,118
(redrawn). Only the part of the wave which drives out a droplet is shown. .................................. 46

Figure II-24: a) Calculated pressure at the entrance of the nozzle and shape of the driving
waveform. b) Measured drop speed as a function of the drop repetition rate (DOD frequency).
The largest peak corresponds to the channel resonance frequency. Reproduced from 98 with
permission from Elsevier. .............................................................................................................. 47

Figure II-25: Parameterisation of a) unipolar and b) bipolar pulse. Reproduced from 104 with
permission from John Wiley and Sons. ......................................................................................... 49

Figure II-26: Photograph sequence taken from a DOD drop formation cycle. The droplets are
recorded by single flash stroboscopy. In the left sequence, the actuation signal has low amplitude,
resulting in a slow droplet. The tail remains stable until the tail droplet is merged with the head
droplet due to capillary contraction of the ligament. In the sequence on the right, the actuation
signal amplitude is 25 % higher. The droplet is so fast that the tail droplet cannot merge with the
head droplet before the tail has broken up. Eventually, the tail droplet merges with the head
droplet. This figure demonstrates the challenge of producing satellite-free droplets at a high
velocity. Image courtesy of Marc van den Berg, Océ Technologies B.V.. Reproduced from 104
with permission from John Wiley and Sons. ................................................................................. 50

Figure II-27: The droplet velocity is plotted versus the maximum ejection velocity. The velocities
are scaled with the capillary velocity, so that the square root of the Weber number with which the
liquid is ejected is on the horizontal axis. The droplet velocity increases with the velocity in the
nozzle. The dashed line is the analytical prediction. The solid line is the droplet velocity obtained
from a numerical simulation using the slender-jet approximation. For Weber numbers below the
jetting threshold, no droplet is formed. Reproduced from 104 with permission from John Wiley and
Sons. ............................................................................................................................................... 53

Figure II-28: A variety of morphologies of liquid drop impact onto a dry surface. Reproduced
from 123 with permission from Begell House, Inc.. ........................................................................ 54

Figure II-29: a) High-speed imaging of a water drop impact on a superhydrophobic surface (Time
interval between the images: 2.7 ms). b) Maximum diameter of the spreading drop normalised by
the drop radius as a function of the Weber number. The open squares correspond to those
obtained with water drops and the filled ones correspond to those obtained with mercury drops.

xvi
The solid line indicates the slope 1/4. Reproduced from 124 with permission from Cambridge
University Press. ............................................................................................................................. 55

Figure II-30: a) J-V curves of 2x2 cm2 devices with ITO, PEDOT:PSS and Ag-grid/PEDOT:PSS
electrodes. b)-d) Inkjet printed Ag grids covered by an inkjet printed layer of PEDOT:PSS for
collection of current. b) Maximum height of the grid lines is 400 nm, PEDOT:PSS thickness is
100 nm. c) Maximum height of the grid lines is 600 nm, PEDOT:PSS thickness is 100 nm. d)
Maximum height of the grid lines is 600 nm, PEDOT:PSS thickness is 200 nm. Reproduced from
137
with permission from Elsevier. .................................................................................................. 60

Figure III-1: Reference solar cell layout. ....................................................................................... 66

Figure III-2: a) Doctor blade and b) LASER at the EnCN laboratories. Reproduced with
permission from the Energy Campus Nuremberg and K. Fuchs. ................................................... 67

Figure III-3: Schematic drawing of PiXDRO LP50 modules. Reproduced from 148 with permission
from Meyer Burger. ........................................................................................................................ 69

Figure III-4: PiXDRO LP50 Printer. a) Drawing of the printhead assembly. b) Drawing of the
motion system. Reproduced from 148 with permission from Meyer Burger. c) Photograph of the
printing setup at the EnCN laboratories. Reproduced with permission from the Energy Campus
Nuremberg and K. Fuchs................................................................................................................ 70

Figure III-5: a) Dimatix FUJIFILM Spectra SE 128 AA printhead used in this work. b) Schematic
illustration of trapezoidal voltage pulse applied to the inkjet printhead piezoelectric elements. ... 71

Figure III-6: Schematic illustration of relation between printhead angle and DPI number. .......... 73

Figure III-7: Top: a) Schematic illustration of loop coater system. b) Photograph of SAMBA
prototype print stations. c) Photograph of recirculation system. .................................................... 75

Figure III-8: a) Photograph of SAMBA printhead. b) Example of VersaDrop multipulses. ......... 77

Figure III-9: a) Lamination process of a flexible OPV module. b) OPV module entering the UV
station. Reproduced with permission from the Energy Campus Nuremberg and K. Fuchs. .......... 78

Figure III-10: Transmission line method, top left: a) Schematic drawing of transmission line
structure. b) Photograph of structure used for measuring the contact resistance between
PEDOT:PSS and printed silver. The PEDOT:PSS layer is located between the dotted lines. c)
Schematic R vs. l plot used to extrapolate the contact resistance. .................................................. 82

Figure III-11: a) Schematic drawing of 2D FEM model used for the silver grid electrode loss
analysis. b) FEM meshing of silver line grid. FEM model Adapted with permission from F.W.
Fecher.156 ........................................................................................................................................ 86

Figure III-12: Schematic illustration of the electrical 2D FEM model used to simulate the 3 cell
OPV module similar to references 156,158. Published in reference 159 and reproduced with
permission. ..................................................................................................................................... 88

xvii
Figure III-13: Geometry of the 3-cell OPV module with dot type interconnection used in the FEM
simulations. Inset: Enlarged simulation mesh at the interconnection region. Published in reference
159
and reproduced with permission. .............................................................................................. 89

Figure IV-1: Regime for stable operation of DOD inkjet printing according to Derby111 and
qualification of the most important inkjet inks used in this thesis. ................................................ 93

Figure IV-2: Drop watcher images of a) IPA based pristine ZnO nanoparticle N-10 ink showing
satellite formation, b) 1-pentanol based N-10 ink and c) 1-pentanol based AgNW ink forming a
stable droplet (printing parameters for all inks: 80 V, 6-6-6 µs, 500 Hz, 16 mbar, pictures were
taken 50 µs, 100 µs, 150 µs after the voltage pulse). ..................................................................... 94

Figure IV-3: P3HT:PC60BM (20 mg/ml) dissolved in a) o-X, b) Indane, c) THN and d) TMB at
60 °C and stored at 30 °C for 5 h without stirring. The o-X and TMB inks formed gels, while the
Indane and THN inks are still liquid. ............................................................................................. 97

Figure IV-4: Viscosity of different P3HT:PC60BM (20 mg/ml, 30 °C) ink formulations over time
at a constant shear rate of 50 s-1, Inset: Photograph of used double gap measurement geometry
with the arrow highlighting gelled active layer ink of the o-X:THN 3:1 (vol.) mixture. Results
were partly obtained by H. Scheiber and E. Maier from Durst Phototechnik and are reported with
permission. ..................................................................................................................................... 98

Figure IV-5:Relative drying times of o-X, o-X:THN 1:1 (vol.) mixture and o-DCB simulated with
the HSPiP software. ....................................................................................................................... 99

Figure IV-6: Relative drying times of IPA:1-pentanol mixtures simulated with the HSPIP
software. Compared to pristine IPA a sevenfold increase of drying time is achieved with a 1:2
IPA:1-pentanol mixture which was applied for inkjet printing of semitransparent electrodes.
Results are partly published in 164 and reproduced with permission. ........................................... 101

Figure IV-7: Drop velocity and volume in dependence of the a) piezoelectric element voltage, b)
hold time, c) ramp-up time and d) ramp-down time for a 1-pentanol based ZnO N-10 nanoparticle
ink. Variation of each parameter was done keeping the other parameters at their initial condition
(Spectra SE 128 AA printhead 80 V, 6-6-6 µs, 500 Hz, 25 °C). ................................................. 102

Figure IV-8: Drop velocity and volume in dependence on the firing frequency for the Spectra SE
128 AA printhead using a model fluid XL-30 (1.08 g/cm³, 31.2 +/-0.5 mN/m, ~4.6 mPas @
70 °C, 55 V, 2-4-2 µs). Results were partly obtained by H. Scheiber and E. Maier from Durst
Phototechnik and are reported with permission. .......................................................................... 103

Figure IV-9: a) Wetting envelopes (θ = 0 °) of commonly used OPV active layer materials
P3HT:PC60BM and PV2000:PC70BM and surface tension data of commercially available
PEDOT:PSS inks P VP Al 4083, F HC Solar, P JET N V2. An amount of 0.1 %wt wetting agent
CFS was added to the PEDOT:PSS inks to lower the surface tension and improve wetting
behaviour. Inks within the wetting envelope are expected to show good wetting on the
corresponding surface. Inks located outside are expected to show only partial wetting or
dewetting. b) Photograph (size: 1x1cm) of a dry inkjet printed PEDOT:PSS layer from F HC
Solar ink, c) F HC Solar + 0.1 %wt CFS, d) P JET N V2 and e) P JET N V2 + 0.1 %wt CFS on a
P3HT:PC60BM surface. Published in reference 184 and reproduced with permission. ................. 108

xviii
Figure IV-10: Surface tension and conductivity of PEDOT:PSS F HC Solar with different
amounts of fluorosurfactant CFS. Published in reference 184 and reproduced with permission. .. 110

Figure IV-11: Schematic illustration of the creation and use of inkjet printed pinning centres.
Published in reference 184 and reproduced with permission. ........................................................ 112

Figure IV-12: a) Optical microscope image of a dry PEDOT:PSS P Jet N V2 layer on a


P3HT:PC60BM surface. The dashed and dotted lines mark the edges of the pinning and printing
areas, respectively. Printing in an area covered by pinning centres results in stable film formation.
Printing on a blank surface results in dewetting and shrinkage of the layer. b) PEDOT:PSS
pinning centres dried at 60 °C do not provide sufficient adhesion to pin the subsequently applied
continuous wet film. c) The same sample after annealing at 140 °C shows stable film formation.
d) Photograph of PVDF surface with water droplets. The square shaped droplet is held in position
by P Jet N V2 pinning centres. e) Microscope image the water droplet’s pinned three phase
contact line. Published in reference 184 and reproduced with permission. .................................... 113

Figure IV-13: Demonstration of locally defined pinning centres on a P3HT:PC60BM surface for
patterning applications. a) Schematic illustration of pinning centre deposition in form of the ZAE
research institute lettering as negative image and b) blade coating of continuous PEDOT:PSS
liquid film. c-h) Photographs of samples after liquid film deposition recorded in 5 s intervals. The
PEDOT:PSS film ruptures and uncovers the surface area without pinning centres, thus creating
the desired layer patterning. Samples and images were made by L. M. Eisenhofer and are
reproduced in this work with permission. Published in reference 184 and reproduced with
permission. ................................................................................................................................... 115

Figure IV-14: Optical microscope image of dry PEDOT:PSS P Jet N V2 layer pinned on a
P3HT:PC60BM surface. b) The zoom shows a 3D confocal microscope image of the pinned
PEDOT:PSS layer edge. The pinning centres printed from 30 pl droplets have an average
diameter of 10.2 ± 0.5 µm after drying. Published in reference 184 and reproduced with
permission. ................................................................................................................................... 116

Figure IV-15: Pinning centres printed with P Jet NV2 ink on P3HT:PC60BM layers at different
resolutions. a) 250 DPI – regular pattern of pinning centres. b) 400 DPI – pinning centres start to
merge. c) 800 DPI – droplets have merged to a wet film which dewets from the surface by
contracting. Published in reference 184 and reproduced with permission. .................................... 118

Figure IV-16: Schematic illustration of the model used to estimate the minimum pinning centre
density: a) Substrate surface covered with a regular pattern of pinning centres and a continuous
wet film. b) If the pinning centre density is too low, dewetting results in formation of spherical
caps of ink surrounding the pinning centres. c) Zoom-in on spherical cap. d) PEDOT:PSS P Jet N
V2 at a P3HT:PC60BM surface with different pinning centre densities. A pinning centre resolution
of 100 DPI (< Resmin) leads to dewetting of the wet film printed on top. e) A pinning centre
resolution of 200 DPI (> Resmin) stabilizes the wet film and prevents layer rupturing and
subsequent shrinkage during drying. Published in reference 184 and reproduced with permission.
...................................................................................................................................................... 119

Figure V-1: Representative J-V characteristics of solar cells with doctor bladed (DB)
P3HT:PC60BM active layers from different ink formulations. The inset shows the J-V

xix
characteristics measured in the dark. b) Normalised UV-Vis absorbance spectra of P3HT:PC60BM
layers blade coated on a glass substrate. ...................................................................................... 127

Figure V-2: a)-c) Microscope images of inkjet printed P3HT:PC60BM layers from o-DCB ink. a)
Substrate temperature 60 °C, inset: Confocal microscope height profile measured along the red
line. b) Substrate temperature 50 °C. c) Substrate temperature RT. All layers were printed with a
speed of 100 mm/s and QF 6. d) AFM image of inkjet printed P3HT:PC60BM layer from o-DCB
(substrate temperature 60 °C – dot pattern. e) AFM image of P3HT:PC60BM reference layer from
CB + 5 %vol BrAni ink. .............................................................................................................. 128

Figure V-3: a) Representative J-V characteristics of solar cells with inkjet printed (IJ)
P3HT:PC60BM active layers from different ink formulations and simulated fit curves used to
extract the charge carrier mobility. The inset shows the J-V characteristics measured in the dark.
b) Normalised UV-Vis absorbance spectra of P3HT:PC60BM layers on a glass substrate. ......... 129

Figure V-4: Box plot PCE, FF, JSC and VOC comparison of P3HT:PC60BM solar cells with doctor
bladed reference (black) and inkjet printed (red) active layer from a) CB + 5 %vol BrAni ink and
b) o-X:THN 1:1 (vol.) ink. Each box contains data of at least 5 solar cells. ............................... 130

Figure V-5: a) Representative J-V characteristics of PV2000:PC70BM solar cells blade coated and
ink printed from different ink formulations. The inset shows the J-V characteristics measured in
the dark. b) Normalised UV-Vis absorbance spectra of PV2000:PC70BM layers on a glass
substrate. ...................................................................................................................................... 131

Figure V-6: Box plot PCE, FF, JSC and VOC comparison of solar cells with doctor bladed (black)
and inkjet printed (red) PV2000:PC70BM active layers (o-X:THN 1:1 vol.). Each box contains
data of six solar cells. ................................................................................................................... 132

Figure V-7: Confocal microscope images of Inkjet printed ZnO droplets on an ITO substrate
surface printed from a) IPA + 10 %vol PG and b) 1-pentanol ink. c) 3D confocal microscope
image of a ZnO layer printed at 40 °C from IPA + 10 %vol PG ink (QF 6, printing speed
100 mm/s). d) 3D confocal microscope image ZnO droplets printed from 1-pentanol ink. e)
Optical microscope image of ZnO layer printed at room temperature from IPA + 10 %vol PG ink
and f) 1-pentanol ink. ................................................................................................................... 134

Figure V-8: Representative J-V characteristics of P3HT:PC60BM solar cells with ZnO
nanoparticle EEL inkjet printed from 1-pentanol based ink (red) and blade coated reference
(black). Also shown is a solar cell with a ZnO layer (blue) inkjet printed at elevated substrate
temperature from IPA + 10 %vol PG, resulting in a ‘dot’-shaped pattern as displayed in Figure
V-7 c). The inset shows the J-V characteristics measured in the dark. ........................................ 135

Figure V-9: Marangoni flow induced spreading and fingering instability of a droplet of IPA +
10 %vol PG placed with a pipette tip on a on an ITO surface (40 °C). Droplet after a) 1 s, b) 5 s,
c) 10 s, d) 20 s. ............................................................................................................................. 136

Figure V-10: Box plot PCE, FF, JSC and VOC comparison of P3HT:PC60BM solar cells with ZnO
nanoparticle EEL inkjet printed from 1-pentanol based ink (red) and blade coated reference
(black). Each box contains data of six solar cells. ....................................................................... 136

xx
Figure V-11: Representative J-V characteristics of solar cells with inkjet printed HEL
PEDOT:PSS F HC Solar + 0.1 %wt CFS (red) and blade coated reference (black). ................... 137

Figure V-12: Box plot PCE, FF, JSC and VOC comparison of solar cells with inkjet printed HEL
PEDOT:PSS F HC Solar + 0.1 %wt CFS (red) and blade coated reference (black). Each box
contains data of at least six solar cells. ......................................................................................... 138

Figure V-13: Representative J-V characteristics of solar cells with inkjet printed HEL
PEDOT:PSS from P Jet N V2 ink. The layer does not allow wetting with small amounts of
surfactant. The red curve is from a device with a PEDOT:PSS layer fabricated according to the
pinning centre strategy (see chapter 4.3.2 and inset). The black curve shows a reference solar cell
where high amounts of fluorosurfactant are applied to enforce wetting. Published in reference 184
and reproduced with permission. .................................................................................................. 139

Figure V-14: Box plot PCE, FF, JSC and VOC comparison of solar cells with inkjet printed HEL
PEDOT:PSS P Jet N V2 applied by using the pinning centre strategy (red) and reference with
high amount of fluorosurfactant (black) to enforce wetting. Each box contains data of at least six
solar cells.. .................................................................................................................................... 140

Figure V-15: a) Photograph of PV2000:PC70BM solar cells with inkjet printed silver grid base
electrode and evaporated silver grid top electrode. b) Confocal microscope 3D image of an inkjet
printed silver finger on a glass substrate ending in a conductive trace Published in reference 200
and reproduced with permission from IEEE. c) Representative J-V characteristics of solar cells
with inkjet printed silver grid base electrode (red) and ITO reference (black). Additionally, a
simulated J-V curve of a solar cell with silver grid base electrode (blue) is shown. The solar cells
are illuminated through the inkjet printed electrode. The inset shows the J-V characteristics
measured in the dark..................................................................................................................... 142

Figure V-16: Box plot PCE, FF, JSC and VOC comparison of PV2000:PC70BM solar cells with
inkjet printed silver grid base electrode and ITO reference. Same device structure and
measurement parameters as described in Figure V-15 apply. ...................................................... 143

Figure V-17: a) Normalised PCE of solar cells with ITO-, inkjet printed silver grid- and simulated
silver grid electrode. b) Analysis of PCE loss mechanisms derived by FEM simulation. c)
Transmission line measurement of HC PEDOT:PSS (F HC Solar) and inkjet printed silver
nanoparticles (ink I50DM106) revealing a contact resistance of 0.11 Ω*cm². d) Comparison of
transmission spectra of glass, ITO coated glass and glass covered with 100 nm PEDOT:PSS +
silver grid according to the above specified grid layout (10 % area loss by shading).................. 144

Figure V-18: a) Optical micrograph and b) SEM image of inkjet printed AgNW layer (500 DPI,
four passes). Published in reference 164 and reproduced with permission from Elsevier B.V.. The
SEM image was recorded by K.C. Tam and is reproduced in this work with permission. c) Optical
micrograph of a liquid 10 µl AgNW ink droplet on a microscope slide directly after deposition. d)
Droplet during drying 1 min after deposition. e) Inhomogeneous dry AgNW layer resulting from
droplet drying. .............................................................................................................................. 146

Figure V-19: Percolation analysis of inkjet printed AgNW meshes by image processing of
stitched SEM pictures. a) Single layer (500 DPI - largest connected region: 99.4 %). b) Two
layers printed on top of each other (500 DPI - largest connected region: 99.4 %). c) Four layers

xxi
printed on top of each other (500 DPI - largest connected region: 98.7 %). Connected AgNWs
forming a network are marked in red colour. SEM images and analysis were provided by Manuela
Göbelt from the Max Planck Institute for the Science of Light and is gratefully acknowledged. 147

Figure V-20: a) Total transmittance of inkjet printed AgNW layers (1-4 printing passes) corrected
for the absorption of the glass substrate and haze of printed AgNW layer (4 printing passes). b)
Specular transmittance at 550 nm vs. sheet resistance of inkjet printed AgNW layers (1-4 printing
passes) corrected for the absorption of the glass substrate. The solid line is a fit curve according to
equation 12. Published in reference 164 and reproduced with permission from Elsevier B.V.. .... 148

Figure V-21: a) Specular transmittance of the encapsulated device. b) Photograph of


semitransparent fully inkjet printed cell. c) Cross-sectional SEM image of the whole device
structure. The thickness of each individual layer is included in the image. The nanoparticle flakes
on the active and glass layer are due to the glass cutting process. The SEM image was recorded by
K.C. Tam and is reproduced in this work with permission. The images are published in reference
164
and reproduced with permission from Elsevier B.V.. ............................................................. 150

Figure V-22: Electrical characteristics of the fully inkjet printed OPV cell. a) Current density vs.
voltage characteristics of the cell during illumination by a sun simulator through a mask of 1 cm²
area. Inset: J-V characteristics measured in the dark. b) EQE of the device. The integrated
response using a standard reference spectrum results in a JSC of 10.7 mA/cm². Published in
reference 164 and reproduced with permission from Elsevier B.V.. ............................................. 150

Figure VI-1: Layout and J-V characteristics of inkjet printed P3HT:PC60BM solar module with
silver grid top electrode and single cells. a) Layout of patterned ITO base electrode. b) Printing
pattern of ZnO and P3HT:PC60BM. c) Printing pattern of HC PEDOT:PSS F HC Solar. d)
Printing pattern of silver grid top electrode and contact pads. e) J-V characteristics of module and
single cells. f) Photograph of the device. ..................................................................................... 157

Figure VI-2: Layout and J-V characteristics of inkjet printed PV2000:PC70BM solar module with
AgNW top electrode. a) Layout of the device consisting of patterned ITO base electrode (white),
ZnO/P3HT:PC60BM (red) and PEDOT:PSS/AgNWs layers (black). b) Photograph of the device.
c) J-V characteristics of module and single cells. ........................................................................ 159

Figure VI-3: a) Schematic illustration of a conventional interconnection (Ic) region using a ‘gap’
in the EEL and AL to realise the cell-to-cell contact. b) ‘Bridge’ arrangement proposed in this
work. c) Photograph of a ‘gap’ type solar module. d) Photograph of a ‘bridge’ dot type solar
module. e) Schematic drawing showing the 3-cell OPV module layout. Each solar cell can be
contacted in a way that allows measurement from bottom to top electrode (cell 1: contact 1-2) or
through the interconnection region (cell 1: contact 1-3). c)-e) Are published in reference 159 and
reproduced with permission of John Wiley and Sons. ................................................................. 161

Figure VI-4: Schematic illustration of the a) ‘lines’ interconnection and b) ‘dots’ interconnection.
c) Dark field microscope image of the ‘lines’ interconnection and d) ‘dots’ interconnection. e)
SEM image of the cross-section at the edge of an interconnection dot. f) Higher magnification
SEM image of the dot edge shows all solar cell layers as printed on top of each other. g) Higher
magnification SEM image closer to the middle of the dot shows direct contact formation of ITO,
silver and the AgNW:PEDOT:PSS hybrid top electrode. The SEM pictures were recorded by K.C.

xxii
Tam, are published in reference 159 and are reproduced with permission of John Wiley and Sons
and the authors of the paper.......................................................................................................... 163

Figure VI-5: Typical J-V characteristics of inkjet printed OPV modules with different types of
interconnections. Published in reference 159 and reproduced with permission of John Wiley and
Sons. ............................................................................................................................................. 164

Figure VI-6: Experimental (triangles) and simulated (lines) J-V curves of a solar module with
dotted cell-to-cell interconnection. Various specific interconnection resistances are simulated
ranging from 0.0002 to 0.1 Ω*cm2. The simulated curves with ρIC = 0.0020 Ω*cm2 (obtained
from CKBR measurements), show excellent agreement with the measured module J-V
characteristics. Published in reference 159 and reproduced with permission of John Wiley and
Sons. ............................................................................................................................................. 166

Figure VI-7: Simulated potential distribution (colour bar) and lateral current flow (arrows) at the
‘dot’ type interconnection area when the module is operated at the maximum power point
(ρIC = 0.0020 Ω*cm2). Different colour scales for the potential drop in bottom and top electrode
are necessary for visualisation, due to their different resistance (bottom 15 Ω/sq, top: 10 Ω/sq).
Published in reference 159 and reproduced with permission of John Wiley and Sons. ................. 168

Figure VI-8: a) Photo of Alan Heeger used for the design. Obtained from 211 with permission. b)
Printing pattern for ZnO and P3HT:PC60BM layer in red colour, PEDOT:PSS and AgNW layer in
blue colour. The white background is the ITO substrate. Black lines indicate the interconnection
region. c) Additional AgNP layer printed on the encapsulation glass to change visual appearance.
d) The completed OPV module portrait of Prof. Alan J. Heeger. Published in reference 159 and
reproduced with permission of John Wiley and Sons. ................................................................. 170

Figure VI-9: J-V characteristics of inkjet printed 84 cm² OPV portrait module. The inset shows
the integration of the module into a radio clock demonstrator where it provides the necessary
power for operation under room light (stray light – no direct illumination). Published in reference
159
and reproduced with permission of John Wiley and Sons. ...................................................... 171

Figure VI-10: Inkjet printed arbitrary shape OPV devices. a) P3HT:O-IDTBR* (same printing
parameters as for P3HT:PC60BM) OPV module in shape of the Durst Phototechnik ‘Crystal’
building. b) Photograph of the building obtained from ck-projekt.it. c) P3HT:O-IDTBR OPV
module in shape of the Durst Phototechnik company logo integrated into a digital clock. d)
P3HT:PC60BM solar cell portrait of Prof. C.-J. Brabec. .............................................................. 172

Figure VI-11: a) Single pass prototype inkjet integrated in a quasi-R2R machine, which runs a 7
m long PET substrate in an endless loop. b) Homogeneous PEDOT:PSS F HC Solar layer printed
on top of a P3HT:PC60BM surface with the single pass inkjet system. The white arrows show the
web movement direction. c) Representative J-V characteristics of P3HT:PC60BM solar cells with
ZnO (red) and PEDOT:PSS (F HC Solar) layers inkjet printed with the single pass machine. The
inset shows the J-V characteristics measured in the dark. ............................................................ 174

Figure VI-12: Box plot PCE, FF, JSC and VOC comparison of P3HT:PC60BM solar cells with ZnO
(N10 - red) and PEDOT:PSS (F HC Solar - blue) layers inkjet printed at the single pass machine
as well as doctor bladed reference (black).................................................................................... 175

xxiii
Figure VI-13: a) Printing process of homogeneous PV2000:PC70BM layers from THN ink with
the single pass inkjet. b) Dry non-functional PV2000:PC70BM layer showing printing defects
from nozzle malfunction. c) Representative J-V characteristics of solar cells with
PV2000:PC70BM active layers doctor bladed (black) and printed at the single pass loop coater
(red) from THN based ink. The inset shows the J-V characteristics measured in the dark. d) Box
plot PCE, FF, JSC and VOC comparison of PV2000:PC70BM solar cells with doctor bladed (black)
and inkjet printed (red) active layer from THN ink. Each box contains data of at least five solar
cells. ............................................................................................................................................. 176

Figure VI-14: a) Layout of fully inkjet printed 10 cm² three-layer modules, black: AgNW:ZnO
base electrode, green: PV2000:PC70BM active layer, red: AgNW:PEDOT:PSS top electrode. b)
Photograph of modules in production. c) J-V characteristics of PV2000:PC70BM three-layer solar
module and single module cells. d) SEM cross section image of an inkjet printed
AgNW:PEDOT:PSS blend electrode and e) inkjet printed AgNW:ZnO blend electrode. The red
arrows mark AgNWs, which appear brighter compared to the matrix material. The SEM images
were recorded by K.C. Tam and are reproduced in this work with permission. .......................... 178

Figure VI-15: Fully single pass inkjet printed flexible three-layer OPV demonstrator consisting of
four 3.5 cm² solar cells showing pictograms of the renewable energy sources solar power, wind
power, hydro power and power from regrowing biomass. a) Layout of the device, black:
AgNW:ZnO base electrode, green: PV2000:PC70BM, red: AgNW:PEDOT:PSS top electrode. b)
Photograph of the demonstrator before encapsulation. c) During encapsulation the UV-curable
epoxy glue (Katiobond LP 655) dissolves the PC70BM in areas which are not protected by the top
electrode, thereby changing the active layer colour in the semitransparent devices from brown to
green. d) J-V characteristics of the 3.5 cm² PV2000:PC70BM solar cells.................................... 180

Figure VII-1: Potential application scenarios for the inkjet processes developed in this work,
including integration into wearables, BIPV, OLED lightning and display applications,
photodetectors (image reproduced with permission from 215), solar parks (image reproduced with
permission from 216) and perovskite PV (image reproduced with permission from 204) .............. 190

xxiv
List of Abbreviations

AFM Atomic force microscope


Ag Silver
AgNW Silver nanowire
Al Aluminum
AM Air mass
BE Backscattered electron
BHJ Bulk heterojunction
BIPV Building integrated photovoltaics
bis-PC60BM Bis(1-[3-(methoxycarbonyl)-propyl]-1-phenyl)-[6,6]C62
BP Boiling point
BrAni 4-Bromoanisole
C60 Buckminsterfullerene
Ca Calcium
CB Chlorobenzene
CCD Charge-coupled device
CdS Cadmium selenide
CdTe Cadmium telluride
CFS Fluorosurfactan Capstone FS 31
CIGS Copper indium gallium (di)selenide
CIJ Continuous inkjet
CN 1-Chloronaphthalene
CNT Carbon nanotube
CO2 Carbon dioxide
Cr Chromium
CTC Charge-transfer complex
CuNW Copper nanowire
CVD Chemical vapour deposition
1D 1-Dimensional
2D 2-Dimensional
3D 3-Dimensional
DB Doctor blade

xxv
DI Deionised
DOD Drop-on-demand
DPI Dots per inch
DPP-TT-T Thieno[3,2-b]thiophene-diketopyrrolopyrrole
DSSC Dye sensitised solar cells
EEG Erneuerbare Energien Gesetz
EEL Electron extraction layer
EPBT Energy payback times
EQE External quantum efficiency
FEM Finite element method
FoM Figure of merit
FTO Fluorine-doped tin oxide
H2O Water
HC Highly conductive
HEL Hole extraction layer
HMI Human machine interface
HOMO Highest occupied molecular orbital
HSPIP Hansen solubility parameters in practice
ICBA Indene-C60 bisadduct
IEO International Energy Outlook
IJ Inkjet
IMI ITO-Metal(Ag)-ITO
IPA Isopropyl alcohol
ITO Indium doped tin oxide
J-V Current density-voltage
LASER Light amplification by stimulated emission of radiation
LED Light emitting diode
LiF Lithium fluoride
LUMO Lowest unoccupied molecular orbital
MEH-PPV Poly[2-methoxy-5-(2-ethylhexyloxy)-1,4-
phenylenevinylene]
MEMS Microelectromechanical systems
NIR Near infrared
o-DCB 1,2-Dichlorobenzene

xxvi
ODT 1,8-Octanedithiol
OFET Organic field effect transistor
O-IDTBR (5Z,5'Z)-5,5'-((7,7'-(4,4,9,9-tetraoctyl-4,9-dihydro-s-
indaceno[1,2-b:5,6-b']dithiophen-2,7-
diyl)bis(benzo[c][1,2,5]thiadiazole-7,4-
diyl))bis(methanylylidene))bis(3-ethyl-2-thioxothiazolidin-
4-one)
OLED Organic light emitting diode
OPV Organic photovoltaics
OPV12 Trade name active layer polymer from Polyera
OWRK Owens, Wendt, Rabel and Kaelble
o-X o-Xylene
P1, P2, P3 Structuring lines in a solar module interconnection region

P3HT Poly(3-hexylthiophene)
P3MHOCT Poly-(3-(2-methylhexan-2-yl)-oxy-carbonyldithiophene)
PBTZT-stat-BDTT-8 Poly[(5,6-bis-R1-(2,1,3-benzothiadiazole-4,7-diyl)-
thiophene-2,5-diyl)-stat-(4,8-bis-R2-(benzo[1,2-b;4,5-
b´]dithiophene-2,6-diyl)-thiophene-2,5-diyl)], side chains
R1 and R2 are not published
PC60BM Phenyl-C61-butyric acid methyl ester
PC70BM Phenyl-C71-butyric acid methyl ester
PCDTBT Poly[ N -9´´-hepta-decanyl-2,7-carbazole-alt-5,5-(4´,7´-di-
2-thienyl-2´,1´,3´-benzothiadiazole)
PCPDTBT Poly[2,6-(4,4-bis(2-ethylhexyl)-4H-cyclopenta[2,1-b;3,4-
b´]-dithiophene)-alt-4,7-(2,1,3-benzothiadiazole)]
PDMS Polydimethylsiloxane
pDPP5T-2 Diketopyrrolopyrrole-quinquethiophene alternating
copolymer
PEDOT:PSS Poly(3,4-ethylenedioxythiophene) polystyrene sulfonate
PET Polyethylene terephthalate
PFDTBTP Poly[9,9-dioctylfluorenyl-2,7-dyil-co-(10,12-bis(thiophen-
2-y)-3, 6-dioxooctyl-11-thia-9,13-diaza-
cyclopenta[b]triphenylene]

xxvii
PffBT4T-C9C13 Poly[(5,6-difluoro-2,1,3-benzothiadiazol-4,7-diyl)-alt-
(3,3´´´-di(2-nonyltridecyl)-2,2´;5´,2´´;5´´,2´´´-
quaterthiophen-5,5´´´-diyl)]).
PG 1,2-Propanediol
pH Potential of hydrogen
PIJ Piezo inkjet
PN 1-Phenylnaphthalene
PSBTBT Poly[(4,4′-bis-(2-ethylhexyl)-dithieno(3,2-b;2′,3′-d)silole]-
2,6-diyl-alt-(2,1,3-benzothiadiazole)-4,7-diyl]
PTB7 Poly[[4,8-bis[(2-ethylhexyl)oxy]benzo[1,2-b:4,5-
b']dithiophene-2,6-diyl][3-fluoro-2-[(2-
ethylhexyl)carbonyl]thieno[3,4-b]thiophenediyl]];
PTFE Polytetrafluoroethylene
PV Photovoltaics
PV2000 Tradename active layer polymer from Raynergy TEK
PVD Physical vapour deposition
PVDF Polyvinylidene fluoride
R2R Roll-to-Roll
RCA Radio Company of America
RT Room temperature
SEM Scanning electron microscopy
SPICE Simulation program with integrated circuit emphasis
THN 1,2,3,4-Tetrahydronaphtalene
TIJ Thermal inkjet
TMB 1,2,4-Trimethylbenzene
USP Unique selling points
UV Ultraviolet light
Vol. Volume
WP Watt peak
ZnO Zinc oxide

xxviii
List of Symbols and Constants
𝐴 Area (m²)
𝐴𝑎𝑐𝑡𝑖𝑣𝑒 Active cell or module area (mm²-cm²)
𝐴𝑑𝑒𝑎𝑑 Non-photoactive cell or module area (mm²)
𝐴𝑛 Cross section of the nozzle (µm²)
𝐴𝑡𝑜𝑡𝑎𝑙 Total cell or module area (mm²-cm²)
𝑐 Speed of sound (m/s)
𝑐𝑜 Concentration (mg/ml)
𝐶1 , 𝐶2 Empirical constants (spin coating) ()
𝑑 Layer thickness (nm-µm)
𝑑𝑎 Damping constant (N/m)
𝐷0 Initial diameter of droplet (µm)
𝐷𝑑𝑜𝑡 Distance between silver ‘bridge’ dots (µm)
𝐷𝑚𝑎𝑥 Maximum spreading diameter of droplet (µm)
𝐷𝑃1,𝑃2,𝑃3 Distance between P1, P2, P3 lines (µm)
eV Electronvolt 1.602*10-19 (kg*m²/s²)
𝐸 Energy (J)
𝐸𝐶𝑇 Charge transfer state energy (eV)
𝐸𝑔 Band gap energy (eV)
𝐸𝑥𝑡 Charge extraction rate (1/(m3*s))
𝛥𝐸 Energy difference (LUMO donor-acceptor) (eV)
𝐹𝐹 Fill factor (%)
𝐹𝑟𝑒𝑠 Resonance frequency (1/s)
𝑔 Gravitational acceleration 9.81 (m/s²)
𝐺 Gibbs free energy (J)
𝐺𝑑 Gibbs free energy in dewetted state (J)
𝐺𝑤 Gibbs free energy in wetted state (J)
𝐺𝑒𝑛 Charge generation rate (1/(m3*s))
𝐺𝐹𝐹 Geometrical fill factor (%)
∆G Change of Gibbs free energy (J)

xxix
ℎ Gap height (doctor blade) (µm)
ℎ𝑆𝐶 Height of spherical cap (m)
𝐽 Current density (mA/cm²)
𝐽0 Diode saturation current density (mA/cm²)
𝐽𝐷 Diode current density (mA/cm²)
𝐽𝑚𝑝𝑝 Maximum power point current density (mA/cm²)

𝐽𝑃𝐻 Photocurrent density (mA/cm²)


𝐽𝑅𝑃 Parallel resistance current density (mA/cm²)
𝐽𝑆𝐶 Short circuit current density (mA/cm²)
𝑘𝐵 Boltzmann constant 1.38*10-23 (J/K)
𝑘𝑟𝑒𝑐,𝐵𝑖𝑚𝑜𝑙 Bimolecular recombination coefficient (1/(m3*s))
𝑘𝑅 Reflection coefficient ()
𝑘𝑇 Transmission coefficient ()
𝑙 Length (m)
𝑙𝑐𝑒𝑙𝑙 Length of single solar cell (mm)
𝑙𝑖𝑐 Length of interconnection region (µm)
𝑙𝑇 Transfer length (mm)
𝑚 Mass (kg)
𝑚𝑠 Slope of transmission line curve (Ω /m)
𝑀𝑤 Molecular weight (kg/mol)
𝑛 Charge carrier density (1/m³)
𝑛𝐷 Diode ideality factor ()
𝑛𝑝 Percolation exponent ()
𝑁𝑐𝑒𝑙𝑙𝑠 Number of solar cells ()
𝑁𝑚 number of molecules ()
𝑂ℎ Ohnesorge number ()
𝑝 Momentum (kg*m/s)
𝑝𝑎 Momentum transferred due to advection (kg*m/s)
𝑝𝑐 Momentum transferred due to surface tension (kg*m/s)
𝑝𝜇 Momentum transferred due to viscous tension (kg*m/s)

xxx
𝑃 Pressure (Pa)
𝑃𝑚𝑒𝑛𝑖𝑠𝑐𝑢𝑠 Printhead meniscus pressure (Pa)
𝑃𝑠𝑒𝑡 Negative pressure applied to ink reservoir (Pa)
𝑃𝐶𝐸 Power conversion efficiency (%)
𝑃𝑜𝐼𝑁 Input power density form solar irradiation (W/m²)
𝑃𝑜𝑚𝑝𝑝 Maximum power point power density (W/m²)
𝛥𝑃 Laplace pressure difference (Pa)
𝑞 Elementary charge 1.602*10-19 (C)
QF Quality factor; Number of nozzles used to print ()
one line in in-scan direction
𝑟1 , 𝑟2 Droplet curvature radii (pendant drop) (m)
𝑟𝑑𝑜𝑡 Radius of silver ‘bridge’ dots µm
𝑟𝑛 Nozzle radius (µm)
𝑟𝑆𝐶 Radius spherical cap base circle (µm)
𝑅𝐶 Contact resistance (Ω)
𝑅𝑃 Parallel resistance (Ω)
𝑅𝑆 Series resistance (Ω)
𝑅𝑠𝑞 Sheet resistance (Ω/sq)
𝑅𝑒 Reynolds number ()
𝑅𝑒𝑐 Charge recombination rate (1/(m3*s))
𝑅𝐸𝑅 Relative evaporation rate ()
𝑅𝑒𝑠𝑚𝑎𝑥 Maximum pinning centre resolution (DPI)
𝑅𝑒𝑠𝑚𝑖𝑛 Minimum pinning centre resolution (DPI)
𝑅𝑅 Regioregularity (%)
𝑠𝑝 Spring constant (N/m)
𝑡 Time (s)
𝑇 Optical transmittance (%)
𝑇𝑒 Temperature (K)
𝑢 Velocity (m/s)
𝑢𝑑𝑟𝑜𝑝 Velocity of droplet (m/s)
𝑢𝑛 Velocity of ink at the nozzle (m/s)

xxxi
𝑉 Voltage (V)

𝑉𝐵𝐼 Built-in potential (V)

𝑉𝐷 Diode voltage (V)


𝑉𝑚𝑝𝑝 Maximum power point voltage (V)

𝑉𝑂𝐶 Open circuit voltage (V)


𝑉𝑡𝑜𝑡 Bias voltage (simulation) (V)
𝑉𝑜𝑙𝐷 Volume droplet (pl)
∆𝑉𝑛𝑜𝑛𝑟𝑎𝑑 Non-radiative 𝑉𝑂𝐶 losses (V)
∆𝑉𝑟𝑎𝑑 Radiative 𝑉𝑂𝐶 losses (V)
𝛥𝑉𝑜𝑙 Volume change (µl)
𝑤 Width (m)
𝑤𝑐𝑒𝑙𝑙 Width of solar cell (mm)
𝑤𝑔𝑟𝑖𝑑𝑓𝑖𝑛𝑔𝑒𝑟 Width of silver grid fingers (µm)
𝑤𝑃1,𝑃2,𝑃3 Width of P1, P2, P3 lines (µm)
𝑊𝑒 Weber number ()
𝑥𝑈𝐶 Unit cell length (µm)
𝑍 Acoustic impedance (kg/(s*m4))
𝑍0 Electrical impedance of the free space 377 (Ω)
α Absorption coefficient (1/m)
𝛽 Printhead angle (°)
𝛾 Free energy of an interface (J/m²)
𝛾𝐷 Dispersive part of surface energy/tension (J/m²)
𝛾𝑃 Polar part of surface energy/tension (J/m²)
𝛾𝐿 Surface tension of a liquid surface (J/m²)
𝛾𝑆 Surface energy of a solid surface (J/m²)
ε Relative permittivity ()
θ Contact angle (°)
Θ Vector containing voltage pulse parameter (µs,V)
λ Wavelength (m)
μ Dynamic viscosity (Pa*s)

xxxii
υ Kinematic viscosity (m²/s)
П Percolative FoM ()
𝜌 Density (g/cm3)
𝜌𝐶 Specific contact resistance (Ω*cm²)
𝜌𝐼𝐶 Specific interconnection resistance (Ω*cm²)
𝜌𝑃 Specific parallel resistance (Ω*cm²)
𝜌𝑆 Specific series resistance (Ω*cm²)
σ or σDC,B Conductivity (direct current, bulk material) (S/m)
σOP Optical conductivity (S/m)
ψ FoM determining the FF ()
𝜔 Angular velocity (rad/s)

Å Ångström 10-10 (m)

xxxiii
CHAPTER I: Introduction

Abstract:

In this introductory chapter, the future development of worldwide energy demand and
generation is considered, leading to the insight that there is clearly a need for new,
environmentally friendly and cheap electricity sources. Photovoltaics (PV) has already
shown that it can contribute a significant share. However, due to their stiffness and
opacity, the application spectrum of conventional silicon PV is limited. New applications
in building facades, wearables and clothes require the solar cells to be semitransparent
and flexible. Emerging printed PV and especially organic photovoltaics (OPV) is able to
satisfy these demands almost perfectly. At the same time, the change in PV applications
makes a free choice of solar cell colour and form more and more important. As these
specifications cannot be met satisfactorily by established manufacturing methods, such as
slot die coating, new ways for solar module production have to be opened up. In this
context, especially drop-on-demand (DOD) inkjet printing is appealing because it offers
the possibility for literally waste-free printing of arbitrary shapes at throughput speeds
exceeding 1000 m²/h.2 Digital printing of whole solar cells and modules would open the
way for printed PV to enter highly customised consumer driven applications, thus
providing the motivation for this work.

Part of this chapter has been published in:


P. Maisch, L. Lucera, C.J. Brabec, H.-J. Egelhaaf, Book Section Flexible Solar Cells, in: Flexible Carbon
Based Electronics, 2018, ISBN: 978-3-527-34191-7, reproduced with permission from Wiley VCH.
Only parts of the book chapter authored by P. Maisch are reproduced in subchapter 1.3. Additionally,
permission to utilise the whole content of the book chapter as part of this thesis was given from all co-
authors involved through author ownership declarations.

1
Introduction

1.1. Green energy demand


Due to economic progress and population growth, the worldwide energy demand is
continuously increasing. According to the International Energy Outlook (IEO) from the
US Energy Administration, the total world consumption of marketed energy will expand
from 579 EJ in 2012 to 664 EJ in 2020 and further to 860 EJ in 2040, corresponding to a
48 % increase from 2012 to 2040.3 The fastest growing end-use energy type within this
statistics is electricity, which is expected to grow by 69 % within the same time frame to a
total of 36.5*1012 kWh.
Today’s power generation is still mainly based on combustion of fossil fuels, such as coal
and gas. This results in annual CO2 emissions in the range of 30-35*1012 tons. Since these
emissions are accused of being the main driver behind an ongoing climatic change with
unpredictable consequences, there is a rising demand for alternative, environmentally
friendly energy sources. Germany plays a pioneering role in supporting the change
towards this new and clean energy, which comprises hydroelectric-, wind- and solar
power as well as power from biomass and burning of waste.

Figure I-1: Share of energy sources (preliminary information) in gross German power
production in 2017 according to the German Federal Statistical Office – ‘Arbeitsgemein-
schaft Energiebilanzen’ (AGEB). Data extracted from 4.

The German ‘Renewable Energies Act’ (Erneuerbare Energien Gesetz, (EEG)) passed in
the year 2000 and guarantees fixed feed-in tariffs for the power generated from green
sources for 20 years. This act stimulated not only the rise of large on- and off-shore wind

2
Introduction

parks, but also decentralised power generation with solar cells on roofs of private
households. Resulting from this versatile mix, the renewable energies share in German
net power production grew from about 6 % in 2000 to more than 33 % in 2017 (Figure
I-1). Further increase without losing social acceptance seems possible by introducing new
inexpensive electricity generation technologies, such as printed PV. At the same time,
novel concepts for energy harvesting areas which do not occupy and deteriorate large
countrysides have to be identified. One possible strategy is the integration of printed PV
into already existing and vastly available structures, such as building facades (584 km²
available in Germany),5 green houses or car roofs. Further elucidation of such new solar
applications is provided in chapter 1.3.

1.2. Solar technologies


Existing solar technologies are typically categorised into three generations. The most
common mono- or polycrystalline silicon solar modules, which can already be found on
the rooftops of many households, are so-called 1st generation PV. While laboratory scale
monocrystalline devices reach power conversion efficiencies (𝑃𝐶𝐸) of around 25 %, most
commercial modules are in the range of 15-20 %.6 The low power production costs of
only ~5 €Ct/kWh,7 arising from the high efficiency, low investment costs and long
lifetime, explain the market domination of more than 90 %.8 The manufacturing of silicon
modules, however, includes an energy intensive melting process and sawing of the
> 100 µm thick wafers, leading to high material losses and energy payback times (EPBT)
of more than 2 years in northern Europe.8
The 2nd generation PV includes amorphous silicon, CdTe, CdS and CIGS (Copper indium
gallium (di)selenide), which are in the range of 10 - 22 % 𝑃𝐶𝐸.9 By avoiding silicon
wafer technology, lower material consumption and reduced production cost than 1st
generation PV can be realised. Furthermore, the thin film technology allows deposition on
flexible substrates. The limiting factor for the price of these technologies is often related
to the processing, which mostly relies on chemical or physical vapour deposition (CVD,
PVD), requiring vacuum equipment and the use of scarce elements.
3rd generation PV comprises high performance multi-junction solar cells as well as
emerging solar technologies with the potential to provide electricity at a lower cost than
1st and 2nd generation PV, such as dye sensitised solar cells (DSSC),10 small-molecule
solar cells,11 polymer solar cells (PSC)12 and perovskites.13 Organic solar cells are of

3
Introduction

particular interest because they offer the production from very cheap and environment-
friendly raw materials. Record efficiencies are already in the range of 17 %1 and lifetimes
of several years under outdoor conditions have already been reported.14 The low printing
or coating temperatures of the solution based fabrication of usually less than 140 °C
enable the use of flexible substrates and high throughput Roll-to-Roll (R2R) production.
Market analysis based on advanced industrial R2R coating scenarios describe that costs of
less than 0.05–0.6 € per watt peak (WP) could be possible for OPV production.15 At the
same time, the EPBT for this technology can by extraordinarily low. Some authors claim
that values of only one day can be reached.16 However, all these positive results have not
been achieved simultaneously, i.e. with the same structures and devices. Lifetimes of
several years have only been achieved for materials which already start at comparatively
low efficiencies. At the same time, very high efficiencies from solution processed OPV
above 10 % 𝑃𝐶𝐸 range have so far only been reported for very small areas of a few mm².
Solar modules > 10 cm² with 𝑃𝐶𝐸 beyond 4 % can scarcely be found. Lucera et al.
identify reasons for the upscaling losses like (i) accelerated processing times, (ii) limited
temperature windows, (iii) smaller or larger current densities leading to a larger or smaller
cell width, (iv) interface and charge extraction layers with distinct absorption features or
(v) inhomogeneities as well as varying film quality and provided strategies for future
research to close this efficiency gap.17 Considering the new fields of use for printed OPV
(see chapter 1.3), he expanded a model by C.J. Brabec18 and defined 4 critical aspects that
OPV has to fulfil simultaneously in order to become a major success story. These are
‘efficiency’, ‘cost’, ‘lifetime’ and ‘design’.19 Especially the parameter ‘design’ is
considered the most powerful stand-alone feature that cannot be met by any other solar
technology. Any short- or midterm commercially relevant OPV products heavily rely on
this cornerstone.

1.3. New solar cell applications


Although the stability and performance of OPV as well as other third generation
technologies is still lower compared to first and second generation devices, they already
have great potential to be used in many emerging applications. The unique selling points
(USP) for OPV comprise the superior aesthetics (i), the flexibility (ii), the tunability of
colour by band gap adjustment (iii), the light weight (iv), the freedom of form (v) and the
scalability (vi), which is realised by the versatility of different processing techniques.

4
Introduction

Especially the light weight and semitransparency of OPV are key factors for integration
of solar modules into buildings (building integrated photovoltaics, BIPV). In comparison
to the most common roof top installations of silicon PV, BIPV applications summarise
the PV integration into the building envelope to create multifunctional structures. OPV
may be incorporated into building facades, windows or awnings. Besides the structural
purpose of the building element itself and electricity generation, other architectural
benefits like passive shading and added value to the appearance of the building can be
achieved. Compared to other technologies, OPV offers superior low light operation,
which is important for vertical installations with reduced access to direct sunlight.20

Figure I-2: a) German pavilion at the international EXPO 2015 in Milan, Italy showing
integrated organic solar modules from Belectric OPV (now OPVIUS). Reproduced from
21
with permission from OPVIUS GmbH. b) The rollable OPV HeLi-on compact solar
22
charger from InfinityPV can be used to charge smartphones. Reproduced from with
permission from InfinityPV ApS. c) Solar art in form of a garden lamp from Belectric
OPV. Reproduced with permission from the Energy Campus Nuremberg and K. Fuchs. d)
23
Plant growth test with semitransparent OPV modules. Reproduced from with
permission from the Energy Campus Nuremberg and F. Machui.

5
Introduction

A new concept, which is currently under investigation by the ZAE Bayern, is to


implement semitransparent OPV modules into green houses.23 The goal of this approach
is to absorb photon wavelengths that are less relevant for photosynthesis and to use the
generated electric power for climatisation or watering of the plants. Economic analysis by
Emmott et al. suggests that there will be a huge potential for OPV in green houses if
aggressive cost targets are met. The modelling shows that even well established active
layer materials, such as Poly(3-hexylthiophene) (P3HT) blended with Phenyl-C61-butyric
acid methyl ester (PC60BM), could be used for economically viable systems if electrical
power is produced at less than 1.71 €/Wp.24
Other possible applications for OPV are consumer products like flexible and rollable
smart phone chargers (see Figure I-2), integration into backpacks, umbrellas, wristbands
for watches or even cloths. While the above-mentioned USPs balance out the lower
efficiency of OPV compared to other competing solar technologies, the limited operation
time for such consumer products allows lower requirements regarding the lifetime of
solar modules. Furthermore, the raw materials are environmentally uncritical, which is an
absolutely essential requirement for such applications that cannot be met by every solar
technology, like in the case of the emerging perovskites, which contain lead, or CdTe PV.

1.4. Printed electronics


In order to produce electronic devices like thin film solar cells, flexible displays or smart
labels at very low cost, liquid film deposition methods are a promising choice. Depending
on the application, there is a vast variation of possible processing technologies which can
be used for fabrication, such as slot die coating, flexographic printing, screen printing and
inkjet printing. A detailed analysis on this topic is provided in chapter 2.1.5 of this work.
Compared to other electronic industry standards, printed electronics is considered a cheap
and large area technology. Switching times and integration density are generally lower. In
fact, conventional and printed electronics can be considered complementary and the
decision which fabrication route to take will depend on the application. In the case of
solar cells, where no high integration density and switching times are needed, the
attraction of printing arises from the possibility to prepare multilayer stacks in a very
simple and cost-effective way. Furthermore, the ability to implement improved
functionalities, like mechanical flexibility, and the use of R2R technology are considered
as major advantages.

6
Introduction

Inkjet printing, which is the main topic of this thesis, is one of the most versatile
deposition methods in the field of printed electronics. The main advantages of this
technology are (i) the possibility to print any arbitrary shape, (ii) the easy change of
layout by just uploading another image file for printing, (iii) the direct patterning with
DOD technology, which makes any pre or post structuring unnecessary and leads to (iv)
an almost material waste-free process as well as the fact that the (v) non-contact method
can be used to print on basically any kind of substrate. The achievable resolution of
advanced industrial printheads, like the Samba G3L, which is also used in this work, of
1200 dots per inch (DPI) and a nominal drop size of only 2.4 pl enable the printing of
highly precise patterns in the µm range.25 All these arguments make small inkjet
machines a most powerful tool for lab scale device fabrication. At the same time, modern
industrial inkjet printers reach a throughput of more than 1200 m²/h,2 which makes this
technology also highly relevant for industrial production and clearly proves that inkjet
printing is not the low throughput method that it used to be when introduced to the
consumer market in the late 1980s. In view of the fast technological progress, many
consultants believe that digital print is still at a very early stage of development,
especially in industrial markets. According to BCC Research, the global market for inkjet
technologies is expected to grow from 4.2 billion $ in 2016 to 12.6 billion $ by 2021,
which corresponds to a five-year annual growth rate of 24.6 %.26

1.5. Motivation and strategy


Chapter 1.3 elucidates the huge potential of printed OPV in emerging applications like
BIPV and portable electronics. At the same time, inkjet technology offers device
fabrication with free choice of shape, colour and substrate which perfectly satisfies the
demands of such utilisation scenarios. The motivation of this work is to bring these two
technologies together, and thus to demonstrate a R2R compatible and industrially
applicable inkjet process for manufacturing of fully printed free-form and free-colour
organic solar cells and modules.
First approaches to realise the combination of inkjet and OPV technology have already
been reported in literature. A detailed summary of the state of the art, which also reveals
still missing aspects, is provided after the necessary backgrounds of both technologies are
introduced (see chapter 2.3). Based on the literature overview, several critical
requirements for the successful realisation of inkjet (IJ) printed OPV which have to be

7
Introduction

fulfilled simultaneously are identified. A summary of these requirements is provided in


Figure I-3. They can be categorised into ink-, layer- and device-aspects, predetermining
to the structure of this work.

Figure I-3: Critical requirements on inks (red), layers (green) and devices (blue) for
inkjet printing of OPV.

Firstly, the foundation for a robust inkjet process has to be set by developing OPV inks
which offer a satellite-free printing process and long nozzle open times (time until
printing nozzles clog when idle). In order to enable industrial processing, the inks should
be free of environmentally highly critical halogenated solvents, such as chlorobenzene
(CB) and orthodichlorobenzene (o-DCB). In literature, very little information about the
choice of OPV inkjet ink solvents and their impact on drop formation and nozzle open
time is provided. As a consequence, most works are limited to small area printing and
report significant performance losses when switching from well established fabrication
methods, such as spin coating or doctor blading, to inkjet fabrication. This situation
should be improved by following the rules of fluid dynamics, like the Ohnesorge theory
(see chapter 2.2.4.1), during the ink formulation process.

8
Introduction

Secondly, a printing process for every single layer of the OPV structure has to be
developed. Printed layers should be of high homogeneity and free of pinholes even on
low energy surfaces in order to guarantee good device performance and a large choice of
usable substrates. By applying the previously developed inks and fine tuning of the
processing conditions, such as printing resolution, drop size, drying times and annealing
temperature, the inkjet printed layers should prove their functionality by yielding the
same OPV device efficiency as blade coated references (see chapter 3.2.1 for details of
the processing). Fully inkjet printed solar cells are created by stepwise replacement of
blade coated layers with inkjet printed equivalents. To replace the sputtered indium tin
oxide (ITO) and the evaporated silver (Ag) electrode of the reference device, inkjet
printed silver grids or silver nanowire (AgNW) meshes are chosen. This variety enables
printing of OPV with different structures and layouts. Even semitransparent devices,
which showcase the USPs that OPV technology has to offer, can be realised.
The third milestone of this work is to apply the developed inks and robust printing
processes on an industrially relevant scale. This requires the first time demonstration of
highly efficient, larger area, fully inkjet printed solar modules. In order to print optically
attractive OPV devices with real life applicability, new strategies to mask the visually
conspicuous cell-to-cell interconnections, to optimise the area usage and to provide
current matching between arbitrarily shaped single cells have to be developed. Finally,
fabrication of solar modules with a high throughput single pass inkjet printer has to be
demonstrated. Therefore, such a sophisticated machine has to be set up (see chapter 3.2.3)
and applied to print highly efficient OPV modules.

9
CHAPTER II: Theoretical Background

Abstract:

This chapter introduces the fundamentals of OPV, comprising sections on device physics,
commonly used materials, cell architectures and fabrication methods. It furthermore
provides a background in inkjet technology, including descriptions of different machine
types, drop generation mechanisms and the underlying physical principles. At the end of
the chapter, the state of the art in digitally printed OPV is summarised and discussed.

Part of this chapter has been published in:


P. Maisch, L. Lucera, C.J. Brabec, H.-J. Egelhaaf, Book Section Flexible Solar Cells, in: Flexible Carbon
Based Electronics, 2018, ISBN: 978-3-527-34191-7, reproduced with permission from Wiley VCH.
Only parts of the book chapter authored by P. Maisch are reproduced in subchapter 2.1 of this work.
Additionally, permission to utilise the whole content of the book chapter as part of this thesis was given
from all co-authors involved through author ownership declarations.

10
Theoretical Background

2.1. Organic solar cells

2.1.1. Working principle


A typical device structure of a flexible organic solar cell is given in Figure II-1. It consists
of two electrodes, at least one of them being semitransparent, to allow the light to couple
into the device. An active layer sandwiched between them uses the incoming photons
energy to generate free charge carriers. Buffer layers between active layer and electrodes
are usually applied to achieve selectivity of the contacts and to support charge extraction.

Figure II-1: a) Typical structure of a flexible organic solar cell consisting of a substrate,
base electrode, electron extraction layer (EEL), active layer, hole extraction layer (HEL)
and top electrode (Al: Aluminum, PEDOT:PSS: Poly(3,4-ethylenedioxy-
thiophene) polystyrene sulfonate, ZnO: Zinc oxide, IMI: ITO-Metal(Ag)-ITO, PET:
27
Polyethylene terephthalate). Published in reference and reproduced with permission
from Wiley VCH. b) Energy band diagram of a normal structure and c) inverted structure
OPV cell. Data extracted from 28.

11
Theoretical Background

Depending on the layer sequence, there are two possible main architectures. In the so-
called ‘normal structure’ positive charges (holes) are extracted through the substrate side
(bottom) electrode, while negative charges leave the device through the top electrode. For
the so-called ‘inverted structure’, the device polarity is reversed.
Compared to inorganic solar cells, such as silicon PV, the OPV active layer consists of
semiconducting organic molecules. The conductivity of such materials comes from the
conjugation of alternating single and double bonds between the carbon atoms.29 Due to
the sp2 hybridisation, the pz electron wave functions can overlap and delocalise, thus
enabling charge transport. Bonding π and antibonding π* orbital are also referred to as
highest occupied molecular orbital (HOMO) and lowest unoccupied molecular orbital
(LUMO). Caused by Peierls instability, a band gap 𝐸𝑔 is opening between bonding and
antibonding states.30 The electronic transition between HOMO and LUMO can be
stimulated by visible light. The whole energy harvesting process from the incoming
photons to electric current generation can be summarised in six steps (see Figure II-2):31

 Light absorption: An incident photon travels through the semitransparent electrode


and stimulates the transition of an electron from HOMO to LUMO.
 Exciton creation: The transition results in an excited state, which may be described as
an electron-hole pair on the same molecule (‘exciton’). Due to their small distance and
the weak relative permittivity of organic compounds of 𝜀 ≈ 2-4, electron and hole are
strongly Coulomb bound.32
 Exciton diffusion: The neutral exciton diffuses through the material until it decays or
finds an interface where it can dissociate into separate charges.
 Charge carrier separation: In order to overcome the Coulomb binding of electron
and hole, a second material with lower LUMO is commonly introduced into the
system. This acceptor material forms a second phase and provides a favourable energy
level for the electron. At the interface of the photoactive donor and the acceptor
material, the electron can be transferred between the phases, thus separating it from the
hole. If an absorption event happens at the donor-acceptor interface, free charge
carriers can directly be created on a timescale of femtoseconds without the
aforementioned exciton diffusion processes.
 Movement of carriers to the electrodes: The free electrons and holes can travel by a
hopping mechanism from molecule to molecule until they reach the electrodes.

12
Theoretical Background

 Charge collection: By overcoming the potential barrier at the interface, charges are
transferred into the electrodes, resulting in a photocurrent.

Figure II-2: From light absorption to photocurrent in a bulk heterojunction solar cell. a)
From a kinetic point of view. b) Simplified energy diagram (binding energies for excitons
and polaron pairs are not shown). Numbers i-vi correspond to the six characteristic steps
of the energy harvesting process as described above. Reproduced from 31 with permission
from IOP Publishing.

First generation OPV devices consisted of only a donor layer, sandwiched between two
electrodes.31 The resulting 𝑃𝐶𝐸𝑠 were relatively low, which can be attributed to
ineffective exciton separation. A breakthrough was achieved by introducing a second,
electron accepting material.33 In the 1980s, the first cells were built with a donor-acceptor
bilayer structure and a 𝑃𝐶𝐸 of ~1 %. One weakness of this setting comes with the
diffusion of the neutral exciton to the donor-acceptor interface. The diffusion length of
excitons is often only in the 10-20 nm range.34 However, for a quite complete absorption
a layer thickness of about 100 nm is necessary for semiconductors with a molar extinction
coefficient of 105 cm-1. In the 1990s, another major step on the way to high efficiencies
was taken by introduction of the bulk heterojunction (BHJ) concept.35 By casting donor-
acceptor solvent blends or by co-evaporation, an interpenetrating network of the two
phases is formed (see Figure II-2 a). This results in significantly larger interface area and
shorter diffusion paths for the excitons. Consequently, the ratio of the two competing
mechanisms of exciton separation and decay is increased drastically. Since the
introduction of the BHJ, significant effort has been spent in optimising its morphology.
This is done with the aim to achieve the shortest possible exciton diffusion path to the

13
Theoretical Background

donor-acceptor interface. At the same time, a percolated network of each phase has to be
formed in order to enable efficient charge transport to the electrodes. The approaches for
BHJ morphology optimisation are highly dependent on the material used. Commonly
followed strategies involve annealing of the active layer and the use of processing
additives for solution processed OPV in order to achieve higher phase purities as well as
crystallinity.36

2.1.2. Current density-voltage characteristics

In order to characterise the performance of solar cells, their current density-voltage (J-V)
curves under illumination are measured.

Figure II-3: Typical J-V characteristics of an organic solar cell (P3HT:PC60BM


27
measured under one sun illumination). Published in reference and reproduced with
permission from Wiley VCH.

From a typical curve as depicted in Figure II-3, the figures of merit (FoM) can be
extracted. These are:

14
Theoretical Background

 The power conversion efficiency (𝑷𝑪𝑬) is the most important parameter and
defines the actual maximum electric power obtained divided by the radiation
power which the device is exposed to.

 The short circuit current density (𝑱𝑺𝑪 ) is the maximum current density of the
cell at zero voltage, which occurs when the two contacts are connected without
any consumer. It corresponds to the number of charge carriers collected at the
electrodes, and thus depends on the number of incident photons as well as on the
yields of each of the six elementary steps described in the previous chapter
‘Working principle’.

 The open circuit voltage (𝑽𝐎𝐂 ) is the maximum potential difference receivable
from the cell when the current output is zero. In a blended donor-acceptor organic
solar cell, 𝑉OC can be explained by formation of a charge-transfer complex (CTC)
between the two materials. Upon excitation of either phase, the CT excitonic state,
which is lower in energy than the pristine excitons, is occupied at the interface
between the donor and acceptor. The energy of the CT state ECT is at best equal to
the difference of the donor’s HOMO and the acceptor’s LUMO level,37 but
frequently found to be smaller than this. Further 𝑉OC losses can mainly be
attributed to radiative ∆𝑉𝑟𝑎𝑑 as well as non-radiative recombination ∆𝑉𝑛𝑜𝑛𝑟𝑎𝑑 and
are expressed in equation 1 with 𝑞 being the elementary charge. A detailed
discussion of the loss mechanisms is provided by Vandewal et al..37,38

𝐸𝐶𝑇
𝑉𝑂𝐶 = − ∆𝑉𝑟𝑎𝑑 − ∆𝑉𝑛𝑜𝑛𝑟𝑎𝑑 equation 1
𝑞

 The fill factor (𝑭𝑭) is the ratio of the maximum obtainable power (gray area in
Figure II-3) divided by the product of 𝑉OC and 𝐽𝑆𝐶 . Compared to 𝑉𝑂𝐶 and 𝐽𝑆𝐶 , the
𝐹𝐹 is a more sensitive indicator for voltage dependent recombination processes
within the device. Bartesaghi et al. give a quantitative description of the
competition between charge carrier extraction rate 𝐸𝑥𝑡 and recombination rate
𝑅𝑒𝑐 in BHJ solar cells by introducing a single parameter 𝜓. Using simulations as
well as experimental proof, they show that 𝜓 determines the 𝐹𝐹.39

𝑅𝑒𝑐
𝜓∝ equation 2
𝐸𝑥𝑡

15
Theoretical Background

The relation between the aforementioned FoMs is expressed by equation 3 with 𝑃𝑜𝐼𝑁
being the power density from incident solar irradiation, 𝐽𝑚𝑝𝑝 the current density and
𝑉𝑚𝑝𝑝 the voltage at the maximum power point.

𝐽𝑚𝑝𝑝 ∗ 𝑉𝑚𝑝𝑝 𝐽𝑆𝐶 ∗ 𝑉𝑂𝐶 ∗ 𝐹𝐹


𝑃𝐶𝐸 = = equation 3
𝑃𝑜𝐼𝑁 𝑃𝑜𝐼𝑁

It is also possible to approximate the J-V behaviour of organic solar cells by the Shockley
equation, which was originally derived for inorganic pn-junctions.40,41 In this model, an
exponential term describes the injection of charge carriers with increasing voltage,
leading to a rectifying behaviour. To provide better matching to real solar cells, a parallel
resistance (𝑅𝑃 ) and a series resistance (𝑅𝑆 ) are introduced, resulting in equation 4.42

(
𝑞∗(𝑉−𝐽∗𝐴∗𝑅𝑆 )
) 𝑉 − 𝐽 ∗ 𝐴 ∗ 𝑅𝑆
𝐽(𝑉) = 𝐽0 ∗ [𝑒 𝑛𝐷 ∗𝑘𝐵 ∗𝑇𝑒 − 1] − − 𝐽𝑃𝐻 equation 4
𝐴 ∗ 𝑅𝑃

In this standard model, 𝐽0 is the diode saturation current density in reverse bias direction,
𝑞 is the elementary charge, 𝑛𝐷 is the diode ideality factor, 𝐴 the area of the solar cell and
𝑘𝐵 ∗ 𝑇𝑒 the thermal energy (𝑘𝐵 : Boltzmann constant; 𝑇𝑒: temperature). As the generated
photocurrent density 𝐽𝑃𝐻 under illumination is in opposing direction, it is subtracted from
the diode therm.
For graphical illustration, an equivalent circuit model, as in Figure II-4, is often used.

Figure II-4: Equivalent circuit model of a solar cell. a) Simplified structure. b) Standard
model with 𝑅𝑆 and 𝑅𝑃 . Redrawn from 42.

As described by Waldauf et al.,43 𝐽𝑃𝐻 can be derived from continuity description of the
charge carrier density 𝑛 under steady state conditions (equation 5), with 𝐺en being the
charge carrier generation rate, 𝐸xt the charge carrier extraction rate and 𝑅e𝑐 the charge
carrier recombination rate.

16
Theoretical Background

𝑑n
= 𝐺en − 𝐸xt − 𝑅ec = 0 equation 5
𝑑𝑡

The charge carrier extraction rate is given by equation 6.43 𝑉𝐵𝐼 refers to the built-in potential
of the device, 𝑑 to the thickness and 𝜇 to the charge carrier mobility. To describe 𝑅𝑒𝑐,
bimolecular recombination is assumed (equation 7), with 𝑘𝑟𝑒𝑐,𝐵𝑖𝑚𝑜𝑙 being the second
order recombination coefficient. A detailed analysis of the recombination mechanisms in
polymer solar cells is provided in references 44,45.

(𝑉𝐵𝐼 − 𝑉 ) ∗ 𝜇
𝐸𝑥𝑡 = ∗𝑛
𝑑2 equation 6

𝑅𝑒𝑐 = 𝑘𝑟𝑒𝑐,𝐵𝑖𝑚𝑜𝑙 ∗ 𝑛2
equation 7

(𝑉𝐵𝐼 − 𝑉) ∗ 𝜇
0 = −𝐺𝑒𝑛 + ∗ 𝑛 + 𝑘𝑟𝑒𝑐,𝐵𝑖𝑚𝑜𝑙 ∗ 𝑛2 equation 8
𝑑2

−(𝑉𝐵𝐼 −𝑉)∗𝜇
𝐽𝑃𝐻 = 𝑞 ∗ 𝑑
*
0,5
(𝑉𝐵𝐼 − 𝑉) ∗ 𝜇 (𝑉𝐵𝐼 − 𝑉) ∗ 𝜇 2 equation 9
− + (( ) + 4 ∗ 𝑘𝑟𝑒𝑐,𝐵𝑖𝑚𝑜𝑙 ∗ 𝐺𝑒𝑛)
𝑑2 𝑑2

2 ∗ 𝑘𝑟𝑒𝑐,𝐵𝑖𝑚𝑜𝑙

By inserting equation 6 and equation 7 into equation 5, equation 8 is generated. Solving


the quadratic equation leads to an expression for the steady state charge carrier density
and finally 𝐽𝑃𝐻 (equation 9).
In good solar cells, 𝐽0 is low, as it is a direct measure for recombination losses. 𝑅𝑆 , which
summarises all effects responsible for a higher total resistance of the device, such as
contact- and electrode resistances, should also be as low as possible. Too high values will
lead to decreased 𝐹𝐹 and 𝐽𝑆𝐶 . 𝑅𝑃 should be as high as possible. Low values indicate
shunting defects, resulting in decreased 𝐹𝐹 and eventually 𝑉𝑂𝐶 loss. The influence of 𝑅𝑆
and 𝑅𝑃 on the J-V characteristics can easily be seen in the logarithmic scale of the dark
curve as shown in Figure II-5.

17
Theoretical Background

Figure II-5: Logarithmic plot of J-V curve measured in the dark (P3HT:PC60BM).

It is also possible to further expand the described standard model to a two diode system.
This provides even better matching with the measured J-V characteristics. However, for
most purposes the standard model is found to give sufficiently good approximation.

2.1.3. Common materials

2.1.3.1. Active layer

Innovations in materials science have always been the key driver behind the steadily
increasing record efficiencies for OPV devices. Since first observations of
photoconductivity in solid anthracene in the beginning of the 20th century, 46 many works
have pushed the development of new active layer polymers and small molecules. Among
the most important milestones on this way are the achievement of a 𝑃𝐶𝐸 of more than
3 % with MEH-PPV poly[2-methoxy-5-(2-ethylhexyloxy)-1,4-phenylenevinylene] and
the development of highly regioregular (RR) P3HT with the McCulloch, Rieke or GRIM
methods,47 which is up to today still a frequently used donor material. A comparison of a
P3HT:PC60BM solar cell’s external quantum efficiency (EQE) with 𝐸𝑔 = 1.9 eV and the
solar spectrum is given in Figure II-6. It is easily visible that the cell barely harvests
photons at wavelengths above 650 nm.

18
Theoretical Background

With the growing understanding of device physics, specific requirements for materials
were identified. These are:

 appropriate HOMO and LUMO levels of donor and acceptor materials,


respectively, for maximum 𝑉𝑂𝐶
 absorption spectra with a good match to the solar irradiation for maximum light
absorption
 good processability leading to optimised BHJ morphology for optimum charge
carrier extraction

Figure II-6: Top: Chemical structures of frequently used donor and acceptor molecules.
a) PCPDTBT: poly[2,6-(4,4-bis(2-ethylhexyl)-4H-cyclopenta[2,1-b;3,4-b´]-dithiophene)-
48
alt-4,7-(2,1,3-benzothiadiazole)]. b) P3HT. c) PC60BM. Reproduced from with
permission from A. Distler. d) Comparison between solar spectrum and the
49
photoresponse of a P3HT:PC60BM solar cell. Reproduced from with permission from
Springer Nature.

19
Theoretical Background

The first two requirements were theoretically investigated by Scharber et al.,50 who
formulated a relation between power conversion efficiency of a BHJ solar cell, band gap
and the LUMO level of the donor (Figure II-7). Their model predicted device efficiencies
of more than 10 % in the year 2006 and is still frequently used as a guideline for BHJ
material selection and material development.

Figure II-7: Contour plot showing the calculated energy-conversion efficiency (contour
lines and colors) versus the band gap and the LUMO level of the donor polymer
according to a model described in 50. Straight lines starting at 2.7 eV and 1.8 eV indicate
HOMO levels of –5.7 eV and –4.8 eV, respectively. A schematic energy diagram of a
donor PC60BM system with the band gap energy (𝐸𝑔 ) and the energy difference (𝛥𝐸) is
also shown. Reproduced from 50 with permission from John Wiley and Sons.

Regarding the second requirement, a lot of effort has been spent on extending the donor’s
absorption range in order to effectively harvest the full solar spectrum including the near
infrared (NIR) region. This can be achieved by a so-called push-pull strategy, in which
polymer chains consist of monomeric repeat units with covalently linked electron
donating and accepting moieties, leading to a charge transfer nature of the
photoexcitations.51,52 In chemical terms, electron donating, i.e. electron rich, monomers
are characterised by shallow HOMOs, whereas electron accepting, i.e. electron deficient,
moieties possess deep LUMOs. Prominent examples are PCPDTBT (see Table II-1) and
PCDTBT: poly[ N -9´´-hepta-decanyl-2,7-carbazole-alt-5,5-(4´,7´-di-2-thienyl-2´,1´,3´-

20
Theoretical Background

benzothiadiazole), which combine electron rich thiophene and electron poor


benzothiadiazole units in the polymer backbone.53 Other methods for increasing the
absorption focus on the light–matter interaction (absorption oscillator strength). Current
research observations identify that optical absorption is strongly related to the chain
stiffness of conjugated polymers.54 This effect, which was for example shown for
thieno[3,2-b]thiophene-diketopyrrolopyrrole (DPP-TT-T),52 may lead to further
improvement of OPV efficiencies in the future.

Table II-1: Overview of high impact PSC donor polymers and performance.
PTB7: poly[[4,8-bis[(2-ethylhexyl)oxy]benzo[1,2-b:4,5-b']dithiophene-2,6-diyl][3-fluoro
-2-[(2-ethylhexyl)carbonyl]thieno[3,4-b]thiophenediyl]]. PV5001: trade name from
Raynergy Tek Incorporation – unpublished structure. PBTZT-stat-BDTT-8: poly[(5,6-bis-
R1-(2,1,3-benzothiadiazole-4,7-diyl)-thiophene-2,5-diyl)-stat-(4,8-bis-R2-(benzo[1,2-b;4,
5-b´]dithiophene-2,6-diyl)-thiophene-2,5-diyl)], side chains R1 and R2 are not published
(trade name from Merck: lisicon). PffBT4T-C9C13: (poly[(5,6-difluoro-2,1,3-
benzothiadiazol-4,7-diyl)-alt-(3,3´´´-di(2-nonyltridecyl)-2,2´;5´,2´´;5´´,2´´´-quaterthio
phen-5,5´´´-diyl)]). Published in reference 27 and reproduced with permission from Wiley
VCH.

Donor Material 𝐽𝑆𝐶 𝑉𝑂𝐶 𝐹𝐹 𝑃𝐶𝐸 Ref.


(mA/ cm2) (V) (%) (%)
55
P3HT 10.6 0.61 67.4 4.37
56
14.5 0.74 69.0 7.40
PTB7
57
PCPDTBT 15.7 0.61 53.0 5.12
58
PCDTBT 12.7 0.90 59.0 6.79
59
PV5001 18.8 0.78 79.5 11.5
PBTZT-stat- 60
16.3 0.78 74.0 9.30
BDTT-8
61
PffBT4T-C9C13 19.8 0.78 73.0 11.7

Besides scientific progress in polymer chemistry to create new materials, the morphology
of the BHJ has always been a major focus in PSC technology in order to push the
maximum achievable efficiency. With the aim to avoid halogenated solvents, such as
chloroform, CB and o-DCB, many less environmentally critical inks have been developed

21
Theoretical Background

in order to fulfil the requirements of large scale R2R production. To identify suitable
formulations, solubility investigations based on the Hansen solubility parameters are a
frequently applied strategy.36 Many inks used for high efficiency OPV contain high
boiling point additives with low solubility for the donor and high solubility for the
acceptor material. Following this route, Zhao et al. demonstrated organic solar cells using
a PffBT4T-C9C13 active layer processed from a hydrocarbon-based solvent system. With
their ink, containing 1,2,4-trimethylbenzene (TMB) and 1-phenylnaphthalene (PN) as
additive, solar cells with optimised active layer morphology reached efficiencies as high
as 11.7 %.61

2.1.3.2. Semitransparent electrodes

For most highly efficient OPV reported in literature a semitransparent indium tin oxide
(ITO) bottom electrode is used in combination with an evaporated metal top electrode.
This combination is not practical for high throughput R2R fabrication on flexible
substrates because the high processing temperature and the brittleness of ITO limit the
film quality. Furthermore, the scarcity of indium makes the ITO electrode to one of the
cost drivers for an upscaled industrial process. Research focusing on ITO alternatives led
to a variety of different electrode types. Among them are conductive polymers, such as
PEDOT:PSS, which can also be combined with metal grids to enhance their performance,
graphene, carbon nanotubes (CNTs) and metal nanowires made for example from copper
(CuNWs) or silver (AgNWs). Furthermore, an extremely thin three layer electrode
consisting of a metal film (typically silver < 10 nm) sandwiched between two ITO layers
(typically ~30 nm) has proven to overcome the brittleness limitations of pristine
conductive metal oxides and is a promising candidate for R2R applications.62 In order to
compare these materials for their capability to be used as semitransparent electrodes, a
ranking system based on sheet resistance 𝑅𝑠𝑞 and transmittance 𝑇 is an often applied
strategy. A detailed description of such a ranking is reported by De and Coleman.63
Combining physical descriptions of 𝑅𝑠𝑞 (equation 10) and 𝑇 (equation 11) based on their
thickness (𝑑) dependence, they derived equation 12 for bulk materials. 𝑍0 is the
impedance of the free space (377 Ω), 𝜎𝑂𝑃 the optical conductivity of the material (related
to the absorption coefficient α as given in equation 13) and 𝜎𝐷𝐶,𝐵 the direct current bulk
conductivity.

22
Theoretical Background

−1
𝑅𝑠𝑞 = [𝜎𝐷𝐶,𝐵 ∗ 𝑑] equation 10
𝑍0 ∗ 𝜎𝑂𝑃 ∗ 𝑑 −2
𝑇 = [1 + ] equation 11
2
−2
𝑍0 ∗ 𝜎𝑂𝑃
𝑇 = [1 + ] equation 12
2 ∗ 𝑅𝑠𝑞 ∗ 𝜎𝐷𝐶,𝐵
𝛼
𝜎𝑂𝑃 = equation 13
𝑍0

The conductivity ratio 𝜎𝐷𝐶,𝐵 /𝜎𝑂𝑃 is used as FoM for the transparent electrode. However,
thin films of nanowires or exfoliated graphene show experimental deviations from this
model calculations (see Figure II-8).

Figure II-8: a) Typical graph of transmittance (generally measured at 550 nm) plotted
versus sheet resistance for thin films of nanostructured materials. Note that each curve is
divided into two regimes, the bulk regime (solid line is a fit to equation 12) and the
63
percolation regime (dashed line is a fit to equation 14). Reproduced from with
permission from Cambridge University Press. b) Plot of sheet resistance versus number
of bends for CuNW films (85 % transparent) and ITO on PET foil. The inset shows the
radius of curvature before (10 mm) and after bending (2.5 mm). Reproduced from 64 with
permission from John Wiley and Sons.

Based on the percolation theory, De and Coleman show how the relationship of 𝑅𝑠𝑞 and 𝑇
can be rewritten to describe thin layers, leading to equation 14 where П is the percolative
FoM and 𝑛𝑝 the percolation exponent.

23
Theoretical Background

1 −2
1 𝑍0 𝑛𝑝 +1
𝑇 = [1 + ( ) ] equation 14
П 𝑅𝑠𝑞

1
𝜎𝐷𝐶,𝐵 /𝜎𝑂𝑃 𝑛𝑝 +1
П= 2∗[ 𝑛 ] equation 15
(𝑍0 ∗ 𝑑𝑚𝑖𝑛 ∗ 𝜎𝑂𝑃 ) 𝑝

In this new model, 𝑑𝑚𝑖𝑛 is the thickness above which the direct current conductivity
becomes thickness independent (equation 14 applies below 𝑑𝑚𝑖𝑛 , but equation 12 applies
above 𝑑𝑚𝑖𝑛 ).

Table II-2: Overview of the most relevant semitransparent flexible electrode materials for
OPV applications. Published in reference 27 and reproduced with permission from Wiley
VCH.

Flexible semitransparent Sheet resistance Transmittance at Ref.


electrode (Ω/sq) ∼ 550 nm (%)

65
PEDOT:PSS ~50 ~75
66
Silver grid (without PEDOT:PSS) ~1 ~92
67
ITO (flexible) ~47 ~83
62
ITO-Metal(Ag)-ITO (IMI) ~6 ~90
68
SWNT ~150 ~78
69
CVD graphene ~30 ~90
64
CuNWs ~30 ~85
70
AgNWs ~13 ~85

Besides the information of the electrode quality based on the transmittance vs. sheet
resistance rating, further application related aspects have to be considered. Among them
are the long term stability, cost, visual appearance and mechanical durability. The latter
was addressed by Rathmell et al., who investigated the electrodes ability to withstand
severe mechanical bending stress (Figure II-8, right).64 They showed the superior
durability of nanowires electrodes compared to ITO.

24
Theoretical Background

2.1.4. Solar modules


For most applications of solar cells, it is necessary to leave the small laboratory scale
devices (typically only a few square millimetres) and go to larger areas. In order to
achieve this, it is not possible to just increase the area of a single cell to the desired scale.
The main reasons behind this are resistive losses, which scale with the electrode size, and
thus with the cell length, as well as with the increase of generated current. Furthermore,
most organic solar cells deliver a 𝑉𝑂𝐶 of < 1 V, which is barely enough to power any
electronics.
In order to keep the generated current and resistive losses low while increasing the output
voltage, a monolithic series connection of several equally sized single cells is introduced.
By doing this, the current remains at the single cell level while the voltages of the
individual cells sum up. A schematic drawing of a solar module is shown in Figure II-9.

Figure II-9: Sketch of a common solar module layout with three cells connected in series.

A typical interconnection between two neighbouring solar cells is established by three


characteristic structuring lines.

 The P1 line separates the bottom electrodes of two neighbouring solar cells.
 The P2 line opens a gap in the absorber layers to enable a contact between the top
electrode of one cell with the bottom electrode of the neighbouring cell.
 The P3 line separates the top electrodes of two neighbouring cells.

As the interconnection region does not generate any photocurrent, it is in the


manufacturers interest to keep it as small as possible. The ratio of the module active area
𝐴𝑎𝑐𝑡𝑖𝑣𝑒 and the total area 𝐴𝑡𝑜𝑡𝑎𝑙 , which includes the non-photoactive region 𝐴𝑑𝑒𝑎𝑑 , is
defined as the geometrical fill factor 𝐺𝐹𝐹. For regular modules with rectangular shaped

25
Theoretical Background

cells, this relation is expressed by equation 16. In this equation 𝑙𝑖𝑐 corresponds to the
length of the interconnection region and 𝑙𝑐𝑒𝑙𝑙 to the length of a single solar cell.

𝐴𝑎𝑐𝑡𝑖𝑣𝑒 𝐴𝑑𝑒𝑎𝑑 𝑙𝑖𝑐


𝐺𝐹𝐹 = = (1 − ) = (1 − ) equation 16
𝐴𝑡𝑜𝑡𝑎𝑙 𝐴𝑡𝑜𝑡𝑎𝑙 𝑙𝑐𝑒𝑙𝑙 + 𝑙𝑖𝑐

A non-exhaustive literature overview of solar modules and the achieved 𝐺𝐹𝐹 is provided
in Table II-3. The summary is a continuation of a work by P. Kubis expanded to state of
the art publications of the recent years.71

Table II-3: Literature overview and comparison of solution processed organic solar
modules. 𝑁𝑐𝑒𝑙𝑙𝑠 is the number of single solar cells in the module. Data partially extracted
from 71 with permission from P. Kubis.

Ref 𝐺𝐹𝐹 𝑁𝑐𝑒𝑙𝑙𝑠 𝑙𝑐𝑒𝑙𝑙 𝑙𝑖𝑐 Materials Deposition 𝑃𝐶𝐸


(%) () (mm) (mm) Methods (%)
72
80 13 24 6 ITO - < 0.1
MEH-PPV Screen printing
Al Evaporation
73
66 8 6 3 ITO - 0.37
P3MHOCTa):ZnO Slot die coating
Ag Screen printing/
Slot die coating

74
- 2-8 - - ITO - 1.09
P3HT:PC60BM Slot die coating
Ag Screen printing
75
- 2-8 - - ITO - 0.31
P3HT:PC60BM Slot die coating
Ag Screen printing
76
50 16 9 9 ITO - 1.5
P3HT:PC60BM Slot die coating
Ag Screen printing
76
67 16 13 6.5 ITO - 1.6
P3HT:PC60BM Slot die coating
Ag Screen printing

26
Theoretical Background

Ref 𝐺𝐹𝐹 𝑁𝑐𝑒𝑙𝑙𝑠 𝑙𝑐𝑒𝑙𝑙 𝑙𝑖𝑐 Materials Deposition 𝑃𝐶𝐸


(%) () (mm) (mm) Methods (%)
76
75 16 18 6 ITO - 0.6
P3HT:PC60BM Slot die coating
Ag Screen printing
77
70- 16 5 1 ITO - 1.6-
80 P3HT:PC60BM Slot die coating 2.0
Ag Screen printing
78
- 3 10 - ITO - 0.7
P3HT:PC60BM Slot die coating
LiFb)/Al Evaporation
79
66 16 10 5 Al/Crc) Evaporation 0.41
P3HT:PC60BM Slot die coating
Ag Screen printing
80
- 5-8 - - ITO - 1.68-
P3HT:PC60BM Gravure printing 1.92
Cad)/Al Evaporation
81
70- 16 5 1 ITO, Ag - 0.23-
80 P3HT:PC60BM Slot die coating 0.44
Ag Screen printing
82
70 4 10 3 Ag/PEDOT:PSS Screen printing 1.36
P3HT:PC60BM Slot die coating
Ag/PEDOT:PSS Screen printing
82
70 7 10.5 2.5 Ag/PEDOT:PSS Screen printing 1.60
P3HT:PC60BM Slot die coating
Ag/PEDOT:PSS Screen printing
82
70 9 10.5 2.5 Ag/PEDOT:PSS Screen printing 1.62
P3HT:PC60BM Slot die coating
Ag/PEDOT:PSS Screen printing
83
70- 16 5 1 Ag Inkjet/Screen 1.7
80 printing
P3HT:PC60BM Slot die coating
Ag Inkjet

27
Theoretical Background

Ref 𝐺𝐹𝐹 𝑁𝑐𝑒𝑙𝑙𝑠 𝑙𝑐𝑒𝑙𝑙 𝑙𝑖𝑐 Materials Deposition 𝑃𝐶𝐸


(%) () (mm) (mm) Methods (%)
84
87 4 - - ITO - 4.9
PCDTBT:PC70BM Spin casting
Al Evaporation
85
98.5 14 5 0.08 FTOe), IMI - 4.18-
PBTZT-stat-BDTT- Slot die coating 5.28
8:PC60BM
Ag Evaporation
86
> 90 3 - < 0.5 IMI - 5.4-
OPV12f):PC60BM Doctor blading 5.7
pDPP5T-2g): PC70BM Doctor blading
Ag Evaporation

a) P3MHOCT: Poly-(3-(2-methylhexan-2-yl)-oxy-carbonyldithiophene)
b) LiF: Lithium fluoride
c) Cr: Chromium
d) Ca: Calcium
e) FTO: Fluorine-doped tin oxide
f) OPV12: Trade name from Polyera company
g) pDPP5T-2: Diketopyrrolopyrrole-quinquethiophene alternating copolymer

The first reports given in the literature overview date back to the year 2007 and show
relatively low 𝐺𝐹𝐹s of less than 80 %.72 This big loss of active area can be explained by
the device fabrication (detailed description of different processing technologies, their
advantages and limitations, is provided in paragraph 2.1.5). First modules were almost
exclusively produced by slot die coating of overlapping stripes. The limited accuracy of
this process forces the manufacturer to design wide cell-to-cell interconnections, resulting
in the comparatively low 𝐺𝐹𝐹s. With the introduction of LASER patterning (Light
amplification by stimulated emission of radiation) to OPV module fabrication, the
achievable 𝐺𝐹𝐹s were pushed far beyond 90 %, allowing dramatically improved area
yield.87 The improved 𝐺𝐹𝐹, however, comes at the price of higher production complexity,
requiring more labour and investment cost, which is a critical aspect for mass production
and commercialisation.

28
Theoretical Background

Figure II-10: a) Flexible OPV solar module fabricated by slot die coating of stripes with
a 𝐺𝐹𝐹 of 80 %. b) Module fabricated by coating of extensive layers and LASER
87
patterning with a 𝐺𝐹𝐹 > 95 %. Reproduced from with permission from Elsevier. c)
Microscope image of the interconnection region showing the P1, P2 and P3 LASER
85
structuring line. Reproduced from with permission from the Royal Society of
Chemistry.

2.1.5. Device fabrication


The versatility of materials and structures in OPV lead to a broad range of processing
methods which can be applied for layer deposition and structuring. This includes for
example thermal evaporation of electrodes or small molecule active layers, sputtering of
ITO and printing or coating of functional inks. In a review paper by F. Krebs
requirements for an ideal OPV fabrication process are summarised with the following
words:88

‘The ideal process should involve solution processing of all layers on flexible substrates
by the combination of as few coating and printing steps as possible. The process should
be free from costly indium, toxic solvents and chemicals and the final polymer solar cell
product should have a low environmental impact and a high degree of recyclability’

29
Theoretical Background

As a comprehensive description of all suitable OPV fabrication methods would go far


beyond the scope of a doctoral thesis, a focus is set on the most important state of the art
solution based processes. In this respect, the different techniques are compared regarding
their speed, material yield, ability to create 2-dimensional structures and R2R
compatibility. An overview of the most important printing and coating processes is
provided in Table II-4.

Table II-4: Comparison of printing techniques: Ink waste: 1 (none), 2 (little), 3 (some), 4
(considerable), 5 (significant). Pattern: 0 (0-dimensional), 1 (1-dimensional), 2
(2-dimensional), 3 (pseudo/quasi 2/3-dimensional), 4 (digital master). Speed: 1 (very
slow), 2 (slow < m*min-1), 3 (medium 1–10 m*min-1), 4 (fast 10–100 m*min-1), 5 (very
fast 100–1000 m*min-1). Ink preparation: 1 (simple), 2 (moderate), 3 (demanding), 4
(difficult), 5 (critical). Ink viscosity: 1 (very low < 10 cP), 2 (low 10–100 cP), 3 (medium
100–1000 cP), 4 (high 1000–10000 cP), 5 (very high 10000–100000 cP). Reproduced
from 88 with permission from Elsevier.
Thickness
Technique Waste Pattern Speed Ink prep. Viscosity R2R
(μm)

Spin coating 5 0 - 1 1 0-100 No


Doctor blade 2 0 - 1 1 0-100 Yes

Casting 1 0 - 2 1 5-500 No

Spraying 3 0 1-4 2 2-3 1-500 Yes


Knife-over-
1 0 2-4 2 3-5 20-700 Yes
edge
Meniscus 1 0 3-4 1 1-3 5-500 Yes

Curtain 1 3 4-5 5 1-4 5-500 Yes

Slide 1 3 3-5 5 1-3 25-250 Yes

Slot-die 1 1 3-5 2 2-5 10-250 Yes

Screen 1 2 1-4 3 3-5 10-500 Yes

Inkjet 1 4 1-3 2 1 1-500 Yes

Gravure 1 2 3-5 4 1-3 5-80 Yes

Flexo 1 2 3-5 3 1-3 5-200 Yes

Pad 1 2 1-2 5 1 5-250 Yes

30
Theoretical Background

Considering the current situation of OPV in between research and market readiness, it is
useful to distinguish between laboratory and industrial scale processing before turning to
in-depth description of the most relevant methods.

2.1.5.1. Laboratory scale

Spin coating is the most commonly used deposition method in OPV research. A liquid is
applied to a substrate located on a rotatable disc. The solution can either be given onto the
substrate before it is accelerated or while it is already spinning at several thousand rpm.
Due to the centrifugal force, the liquid spreads across the surface, leaving a thin film.
Spin coating allows a very homogeneous distribution of ink and optimal morphology of
the layers, which is also highly reproducible. Due to these reasons, the highest 𝑃𝐶𝐸s for
solution processed OPV devices are usually obtained with spin coating systems.89
However, due to the unique approach of using rotational speed for the ink spreading and
film formation, spin coating is not R2R compatible. Furthermore, to ensure a complete
layer, the ink has to be applied at large excess, which is then lost by flowing over the
substrate edge. This wastefulness prevents spin coating from OPV processing at any
industrially relevant scale. An often used empirical approach for thickness estimation,
with 𝜔 being the angular velocity and 𝐶1 , 𝐶2 empirical constants, is provided in equation
17. In this equation 𝐶2 is typically in the range of -0.5.

𝑑 = 𝐶1 ∗ 𝜔 𝐶2 equation 17

Figure II-11: Schematic illustration of a a) spin coating and b) doctor blading process. b)
is reproduced from reference 19 with permission from L. Lucera.

31
Theoretical Background

Doctor blading is another important deposition method frequently used in OPV research.
In Figure II-11, a sketch of this technology is shown. The ink is applied close the edge of
a blade where a meniscus is formed. By moving the blade or the substrate, the ink is
distributed and forms a homogeneous liquid film. In this process, the blade has a fixed
distance from the substrate (usually several 100 μm). An empirical model for thickness
estimation is expressed in equation 18, with ℎ being the gap height (distance between
substrate and blade), 𝑐𝑜 the concentration of the solid material in the ink and 𝜌 the density
of the material in the final film.88

ℎ ∗ 𝑐𝑜
𝑑= equation 18
2∗𝜌

As the equation does not account for other influencing factors such as viscosity, surface
energy, temperature, blading speed and possible shear thinning effects, is can only be
used for a rough estimation and starting point for experiments. One advantage of doctor
blading is the large coating window regarding viscosity and choice of solvents, leading to
comparably little link formulation work. Compared to spin coating, the transfer of doctor
blading from small scale to a R2R process is relatively straightforward. Several works
have proven that organic solar cells manufactured by doctor blading can achieve
performances comparable to those prepared by spin coating.90,91

2.1.5.2. Industrial scale

Since the establishment of OPV technology, there has been the idea to introduce
continuous processes for cost efficient mass production. In contrast to the laboratory scale
techniques described in the previous paragraph, which all operate on relatively small
substrates with a batch to batch process, the methods presented in this section involve a
long flexible sheet wound on a roll. This substrate is referred to as the ‘web’. In the first
processing step the web is unwound from a roll. After passing through all stages of the
machine, which usually involve several printing and drying stations, the web is rewound
on a second roll. This leads to the commonly used term ‘Roll-to-Roll’ (R2R) production.
In the ideal scenario, the substrate should enter the processing machine at one end and a
fully functional flexible polymer solar module should emerge on the other end. The high
level of integration and automatisation of such machines brings the advantages of high

32
Theoretical Background

throughput and low work force requirements, resulting in very low-cost production. In
practice, implementation of such production is often challenging. With the increase of
active area also the probability for layer defects rises. Therefore, the mass production
requires an extremely clean and dust-free processing environment and good processing
control. Furthermore, the different inks applied in the OPV multilayer stack have to be
designed in such a way that printing and drying of the single layers is achieved at the
same web speed. At the same time, the printing methods that can be applied for the
different layers are versatile. Slot die coating, rotary screen-, gravure- and flexographic
printing are among the most commonly applied methods and should briefly be described.
A schematic illustration of these processes is found in Figure II-12.

Figure II-12: Schematic illustration of the R2R techniques a) slot die coating, b) rotary
screen printing, c) gravure printing and d) flexographic printing for solar cell fabrication
as described in references 88,92 (redrawn).

33
Theoretical Background

Slot die coating is a method applied to produce extensive wet films or, with the help of
shims, 1-dimensionally structured layers in the form of stripes. The slot die head usually
consists of two secular stainless steel blocks ending in a triangle vertex.92 A pump or
other pressure system transfers the ink from an external reservoir into a chamber inside
the coating head. From there it flows through the slot die lips onto the substrate. The layer
thickness is controlled by the speed of the web, the feeding rate and the solid content of
the ink. Slot die coating is generally considered a simple and robust technique88 and has
been extensively studied in the field of OPV.87,93
Rotary screen printing involves a printing cylinder which is equipped with a squeegee
and a mask. The ink is stored at the inner part of the cylinder and transferred to the web
by direct contact. The squeegee presses on the substrate in the contact point, transferring
the ink through the meshes of the screen, thereby replicating the mask pattern. The fast
processing speed and high achievable wet film thicknesses make this method attractive
for mass production. Therefore, it is widely used in the PV field to deposit PEDOT:PSS
and silver grids.94
Gravure printing involves two cylinders. While the first one is patterned with gravures of
a defined shape and depth, carrying the ink, the second one supplies the pressure for
imprinting on the web. A doctor blade is located at the first cylinder in order to remove
the excess of ink taken from the reservoir. A main advantage of this technology is the
possibility to print even low viscosity inks at high speeds. For this reason it was
successfully applied in OPV active layer printing.95
Flexographic printing involves four cylinders. The first one takes the ink from a reservoir
and transfers it to a so-called anilox roller, which is characterised by gravures to collect
and meter the liquid. In the next step, the ink is transferred onto a raised pattern on the
printing cylinder. A direct contact to the substrate, ensured by a fourth impression
cylinder, transfers the ink pattern onto the web. Different OPV layers have been
demonstrated to be compatible with this technology, including silver grids for front
electrodes.94
Another technology which is emerging in the field of OPV and printed electronics in
general is inkjet printing. Since the main focus of this thesis is dedicated to further
develop inkjet printing of thin film PV, this method is described in more detail in the
following chapter.

34
Theoretical Background

2.2. Inkjet technology


Due to the unique characteristics of inkjet printing, using single droplet deposition, it is
perfectly suited for batch- as well as R2R processing and cannot be categorised easily into
lab or industrial scale methods. In the following paragraph and subchapters, the function
principles, advantages and limitations of this technology are introduced.

2.2.1. Historical background


Many important physical observations had to be made and formulated into laws to pave
the way for the emergence of first inkjet machines. In the 1750s Leonhard Euler
published his equations describing ideal elastic liquids.96 His work can be understood as
adaption of Isaac Newtons second law to describe friction-free flow. Seventy years later,
in 1822, the French physicist and engineer Claude-Louis Navier formulated his equations
on the motion of fluids, which also consider the inner friction, and thus viscosity. Further
work of George Gabriel Stokes in 1845 led to the formulation of the Navier-Stokes
equations, which are still used extensively to describe viscous flow behaviour.97,98
Another important work on the way to first inkjet machines is from Plateau in 1843.99 He
described the decay of a liquid column, noticing that perturbations become unstable,
resulting in the evolution of single drops. Later, in the 1870s, Lord Rayleigh published
several papers describing the instability and flow dynamics of jets mathematically.100,101
The result of both works is the well known Plateau–Rayleigh instability. The important
role of surface tension for the jet breakup and formation of single droplets was described
later in 1804 by Young102 and 1805 by Laplace.103
Due to the complex physics behind it, the concept of inkjet machines is relatively new
when compared to other printing technologies. The earliest reported inkjet-like systems
are Siphon recorders, which used electrostatic forces for automatic recording of telegraph
messages. These machines were developed in 1858 by William Thomson as well as Lord
Kelvin and patented in 1867 (UK Patent 2147). They worked with a siphon, which was
shifted back and forth according to a signal, thus producing a continuous stream of ink
flowing onto a moving web of paper. An illustration of such a machine is provided in
Figure II-13.

35
Theoretical Background

Figure II-13: The Siphon recorder is the first practical continuous inkjet device. It was
used for automatic recordings of telegraph messages and invented by William Thomson.
Reproduced from 98 with permission from Elsevier.

Extensive commercially driven development of inkjet technology began much later in the
early 1950s. In 1951 the Siemens-Elema company patented the first device, using a
stream of droplets created by Rayleigh breakup.98 The first printers which were able to
reproduce digital images generated by computers started to emerge in the 1970s and were
mainly engineered by the companies Epson, Hewlett-Packard and Canon. Since then,
many different inkjet types have been developed.

2.2.2. Scales of length and time


Before going into detail of different inkjet machine types and their working mechanism, it
is useful to analyse characteristic length and time scales in the field of modern inkjet
printing machines. A good overview of this field is given by Hutchings et al. in the Book
‘The Fundamentals of Inkjet Printing’ and can be seen in Figure II-14 and Figure II-15.
As shown in this comparison, inkjet technology involves length scales ranging from the
nanometer (nanoparticles, pigments) up to the meter scale (print width).104

36
Theoretical Background

104
Figure II-14: Length scales involved in inkjet printing. Reproduced from with
permission from Wiley VCH.

104
Figure II-15: Time scales involved in inkjet printing. Reproduced from with
permission from Wiley VCH.

With the nozzle and drop size typically being in the range of 10-100 µm, it seems clear
that any particles dispersed in the printing ink should be much smaller in order to avoid

37
Theoretical Background

clogging. An interesting insight on this topic is provided in the results part of this work,
where nanowires of ~30 µm length are successfully printed through nozzles of only
10 µm diameter (chapter 6.3).
Typical inkjet time scales exhibit a similarly wide range. Some of the timescales which
have to be considered are related to the machine, while others are intrinsic to the ink. In
this context, the impact of the drop onto a substrate is of very high importance because
capillary spreading and the curing/drying timescale dictate the deposition quality.

2.2.3. Inkjet types


Modern inkjet applications are versatile ranging from common desktop printers to high
throughput industrial machines for textile, label, poster or ceramic printing and many
more. In order to suit the different requirements of this multitude of applications, several
types of inkjet machines were created (Figure II-16), which all consist of three
fundamental inkjet components:104–106
(i) Ink
(ii) Print engine (printhead)
(iii) Imaging medium (e.g. paper)

Figure II-16: Different types of inkjet printers as described in 98,107 (redrawn).

In the following, the most important modern inkjet types should be described briefly.
Generally, inkjets can be categorised in continuous mode inkjet (CIJ) and drop-on
demand inkjet (DOD). A schematic illustration of both working principles is provided in
Figure II-17.
In Continuous inkjet technology, a constant stream of ink is generated by applying
pressure to an ink reservoir. The resulting jet is inherently unstable (Rayleigh instability)
and breaks up into droplets. This step can be modulated by a periodic stimulation, which

38
Theoretical Background

is commonly realised by a piezoelectric or thermal transducer. If modulated properly, a


continuous stream of hundreds of thousands single droplets per nozzle and second with
defined volume and velocity of typically 10-20 m/s is achieved.104 The droplets pass
through an electric field and are charged according to the image information. Further
downstream, they pass high voltage deflection plates, which guide them according to their
charge either onto the correct spot on the substrate or to an ink collection unit from where
they are recycled back to the feeding system. The CIJ technology is commonly used for
high volume and relatively low quality fabrication, such as label printing, packaging or
personalised advertising applications.

Figure II-17: Schematic illustration of continuous and drop-on-demand inkjet printing as


described in 92,104 (redrawn).

In Drop-on-Demand inkjet (DOD) technology, droplets are only produced if required.


The printers can further be classified according to their droplet ejection mechanism. The
two most important types are the thermal inkjet (also known as ‘bubble jet’) and piezo
inkjet (Figure II-18).
Thermal inkjet (TIJ) is, with approximately three quarters of the market share, the widest
spread DOD machine type.104 It is used in most desktop inkjet printers in private

39
Theoretical Background

households. As the name suggests, the drop ejection is triggered by heating, which results
in localised vaporisation, and thus volume expansion of the ink inside a jet chamber.
Nowadays, this technology is not only used in desktop printers, but also in the publishing
market with high-speed continuous feed printers, such as the Hewlett-Packard InkJet Web
Press, reaching incredible throughput speeds of up to 244 m/min.104 Besides that, TIJ
technology was never able to significantly enter the industrial manufacturing market,
which is reasoned in the necessity to heat the often temperature sensitive inks. The high
firing frequencies require very high heating rates of more than 107 K/s.104 This
corresponds to a temperature increase of 100 °C in less than 10 µs, highlighting the
technical challenges that come with this technology.

Figure II-18: Schematic illustration of the most common DOD inkjet types a) thermal and
b) piezo as described in 98,104 (redrawn).

Piezo inkjets (PIJ) are used in most industrial and some consumer printers (for example
Epson and Brother). The function principle of this technology is to apply a voltage to a
piezoelectric element located at the ink channel. The deformation of the piezoelectric
crystal stimulates a pressure wave which forces a droplet to eject from the nozzle. First
works on PIJ printers were already performed in the late 1940s by the Radio Company of
America (RCA).104 In this initial concepts, pressure waves were introduced to an ink
filled conical nozzle via a coaxially arranged piezoelectric disc leading to a spray of ink
droplets. The first pioneering patents to protect this idea are from 1950 and describe the
use of a piezoelectric unit to convert an electrical driving voltage into a mechanical
deformation of an ink chamber. Compared to TIJ, PIJ inkjets allow a wider variety of inks

40
Theoretical Background

as there are no strict requirements regarding the need for a volatile component,
temperature stability and kogation (build-up of dried ink residue at the heating
element).104 However, PIJ printheads are generally more expensive to manufacture than
TIJ.

2.2.4. Aspects of ink formulation

2.2.4.1. Printability

In order to achieve a stable PIJ DOD printing process, i.e. formation of one single round
shaped droplet per actuation pulse, the physical properties of the inkjet ink, like viscosity
and surface tension, have to be within distinct processing windows, which highly depend
on the used printhead. A brief introduction on how the correlation of ink properties and
printing behaviour can be theoretically accessed is provided the following.
Based on the Navier-Stokes equations, which are a powerful mathematical model to
describe the motion of fluids arising from the adoption of Newtons second law
(conservation of momentum) to a continuum system, several dimensionless figures of
fluid mechanics can be derived. One of them is the Reynolds number (𝑅𝑒), which is a
measure for the ratio of inertial forces to viscous forces.

𝑖𝑛𝑒𝑟𝑡𝑖𝑎𝑙 𝑓𝑜𝑟𝑐𝑒𝑠 𝜌 ∗ 𝑢 ∗ 𝑙 𝑢 ∗ 𝑙
𝑅𝑒 = = = equation 19
𝑣𝑖𝑠𝑐𝑜𝑢𝑠 𝑓𝑜𝑟𝑐𝑒𝑠 𝜇 υ

𝑖𝑛𝑒𝑟𝑡𝑖𝑎𝑙 𝑓𝑜𝑟𝑐𝑒𝑠 𝜌 ∗ 𝑢2 ∗ 𝑙
𝑊𝑒 = = equation 20
𝑠𝑢𝑟𝑓𝑎𝑐𝑒 𝑓𝑜𝑟𝑐𝑒𝑠 𝛾𝐿

𝑣𝑖𝑠𝑐𝑜𝑢𝑠 𝑓𝑜𝑟𝑐𝑒𝑠 √𝑊𝑒 𝜇


𝑂ℎ = = = equation 21
√𝑖𝑛𝑡𝑒𝑟𝑖𝑎𝑙 𝑓𝑜𝑟𝑐𝑒𝑠 ∗ 𝑠𝑢𝑟𝑓𝑎𝑐𝑒 𝑓𝑜𝑟𝑐𝑒𝑠 𝑅𝑒 √𝜌 ∗ 𝛾𝐿 ∗ 𝑙

𝑅𝑒 is expressed by equation 19, with 𝜌 referring to the liquid density, 𝑢 to the speed, 𝑙 to
a characteristic length scale (for inkjet typically the nozzle or drop diameter), 𝜇 to the
dynamic viscosity and υ to the kinematic viscosity. Initially, the concept of the 𝑅𝑒
number was introduced by George Gabriel Stokes in 1851.108 However, the name goes
back to Osborne Reynolds, who used it to describe the transition of laminar to turbulent
flow.109 In general, low 𝑅𝑒 numbers predict laminar flow, as viscous forces are dominant.

41
Theoretical Background

Turbulent flow behaviour with vortices and eddies are characterised by high 𝑅𝑒 numbers,
as inertial forces are dominating the system. Another dimensionless number which is
frequently used to describe the behaviour of liquids is the Weber number (𝑊𝑒). It was
introduced by Moritz Weber (1871–1951) and compares the fluids inertial forces with the
surface forces.110 In equation 20, 𝛾𝐿 refers to the surface tension of the ink. 𝑊𝑒 is often
applied to analyse the formation of bubbles or droplets. A third dimensionless number of
high importance in fluid mechanics was introduced by Wolfgang Ohnesorge in 1936. The
Ohnesorge number (𝑂ℎ) relates viscous, inertial and surface forces as given in equation
21.

Figure II-19: A schematic diagram showing the operating regime for stable operation of
DOD inkjet printing. The diagram uses the Ohnesorge number as the ordinate axis in
place of the Weber number 𝑊𝑒 = (𝑅𝑒 ∗ 𝑂ℎ)2. The criterion for a drop to possess
sufficient kinetic energy to be ejected from the nozzle is given by Derby as 𝑊𝑒𝑐𝑟𝑖𝑡 ≥ 4.111
The criterion for onset of splashing following impact is given by Derby as 𝑂ℎ ∗ 𝑅𝑒 5/4
≥ 50. Reproduced from 112 with permission from AIP Publishing.

In inkjet technology, the 𝑂ℎ number reflects only the size scale of the jet or drop and the
physical properties of the liquid. In Contrast to 𝑅𝑒 and 𝑊𝑒 numbers, the velocity, and
thus the dependence on the driving conditions is removed from the mathematical

42
Theoretical Background

description.104 Good printability is expected in an 𝑂ℎ range from 0.1-1. For higher


numbers, too high viscosity will prevent separation of droplets. Too low 𝑂ℎ values can
result in a large numbers of satellites.104,111 Further considerations of the necessary kinetic
energy for droplets to eject from the nozzle and splashing of the droplet when impacting
on the substrate were reported by Derby.111 All these conditions can be put together to
create a map illustrating the regime where stable DOD inkjet printing is possible (Figure
II-19). It is worth mentioning that some authors prefer to use 1/𝑂ℎ to describe the
processing window of printability.111,113 A convenient online software tool to estimate the
printing behaviour of inks similarly to Figure II-19 is provided by Steven Abbott and is
found in reference 114.

2.2.4.2. Wetting

Wetting of the ink on the substrate surface is an important aspect for printing of
homogeneous layers which has to be considered during ink formulation. This topic is
typically addressed by investigation of the interfacial free energies 𝛾 between the three
phases Liquid (𝐿), Solid (𝑆) and Gas (𝐺).115 The parameter 𝛾 is thereby defined as the
work 𝐸 which has to be expended to increase the size of the interface 𝐴 between two
adjacent phases. Based on 𝛾, the contact angle 𝜃 of a liquid droplet on a solid surface at
thermodynamic equilibrium is given by the Young equation (equation 22) (see Figure
II-20).

Figure II-20: Schematic of a liquid drop showing the quantities in the Young equation.

𝐸
𝛾= equation 22
𝐴

𝛾𝑆𝐺 − 𝛾𝑆𝐿
𝑐𝑜𝑠𝜃 = equation 23
𝛾𝐿𝐺

Depending on the contact angle the wettability is commonly quantified as:


 𝜃 = 0° Perfect wetting
43
Theoretical Background

 0° < 𝜃 < 90° High wettability


 90° < 𝜃 < 180° Low wettability
 𝜃 = 180° Perfect non-wetting

In the following, the focus is set on experiments performed in ambient atmosphere.


Therefore, the gas-interface is considered as a surface and the term ‘surface energy’ 𝛾𝑆
instead of 𝛾𝑆𝐺 is used for description of the solid surface. The commonly used unit for the
surface energy is J/m². For the liquid-gas interface the term ‘surface tension’ and the
symbol 𝛾𝐿 instead of 𝛾𝐿𝐺 is applied. The reason behind the different formalism is the
derivation of the surface tension based on a mechanical description of forces acting on the
liquids molecules parallel to the surface. These forces keep the surface area small and
under continuous tension. The unit of 𝛾𝐿 is N/m, which is equivalent to J/m².
A more detailed description of the wetting behaviour was introduced by Fowkes.116 Based
on the well understood different intermolecular forces as cause of the interfacial energies,
he proposed to split 𝛾 into a dispersive 𝛾 𝐷 and a non-dispersive part (see Figure II-21).
Later, Owens, Wendt, Rabel and Kaelble (OWRK) further specified the non-dispersive
part as polar 𝛾 𝑃 . According to their model, the interfacial tension 𝛾𝑆𝐿 is calculated based
on the two surface tensions 𝛾𝑆 and 𝛾𝐿 , with the similar interactions between the phases
interpreted as geometrical mean of 𝛾 𝐷 and 𝛾 𝑃 . This leads to equation 24 and equation
25.117

Figure II-21: Schematic illustration of polar and dispersive part for two phase contact
model according to 117 (redrawn).

𝛾 = 𝛾𝑃 + 𝛾𝐷 equation 24

𝛾𝑆𝐿 = 𝛾𝑆 + 𝛾𝐿 − 2 (√𝛾𝑆𝐷 𝛾𝐿𝐷 + √𝛾𝑆𝑃 𝛾𝐿𝑃 ) equation 25

44
Theoretical Background

(1+cos 𝜃)𝛾𝐿 𝛾𝑃
= √𝛾𝑆𝑃 √𝛾𝐿𝐷 + √𝛾𝑆𝐷
2√𝛾𝐿𝐷 𝐿 equation 26

(cos 𝜃 + 1) (2√𝛾𝐷 𝛾𝐷 − 𝛾𝐷𝐿 (cos 𝜃 + 1)) + 2𝛾𝑃𝑆 ± ⋯


𝐿 𝑠
𝛾𝑃𝐿 =
(cos 𝜃 + 1)2 equation 27

… 2√(cos 𝜃 + 1) (2√𝛾𝐷 𝛾𝐷 − 𝛾𝐷𝐿 (cos 𝜃 + 1)) 𝛾𝑃𝑆 + (𝛾𝑃𝑆 )2


𝐿 𝑠

By combining the Young equation 22 and OWRK equation 25, equation 26 is obtained.117
It should be mentioned that this results in the equation of a line, which can be used to
determine the polar and dispersive parts of an unknown surface. In order to do so, the
contact angles of at least two test liquids with known polar and dispersive parts have to be
determined. By plotting the line equation, 𝛾𝑆𝑃 and 𝛾𝑠𝐷 can be extracted from the slope and
y-axis intersection, respectively.

Figure II-22: Wetting envelope of a 50 J/m² surface for a perfect wetting condition
(𝜃 = 0°). Liquids with polar and dispersive surface tension coordinates inside the wetting
envelope are predicted to wet perfectly. For coordinates outside the envelope lower
wettability is expected.

Furthermore, it is possible to combine equation 24 and equation 26 in order to describe


𝛾LP as function of 𝛾LD . This results in equation 27, describing the so-called wetting

45
Theoretical Background

envelope (Figure II-22), which is a powerful tool to characterise and predict the wetting
behaviour of inks on a surface. In this work, it is used to describe the challenging wetting
behaviour of hydrophilic PEDOT:PSS inks on hydrophobic polymeric active layers.
Based on the findings, ink formulations with wetting agents were developed (see chapter
4.3.1).

2.2.5. Actuation principle


In the next step, the well printable inks which offer good wetting on the surface, are
deposited by inkjet. A simplified illustration of the PIJ printhead operation with a long
channel and large reservoir, as used for example by Dimatix/Spectra, is given in Figure
II-23.98 By applying a voltage at the piezoelectric element, the ink channel enlarges,
creating a negative pressure wave (t 1). This wave propagates (t 2) and will be reflected
when the acoustic impedance 𝑍 of the channel changes.

Figure II-23: Piezoelectric drop generation by pressure waves as described by references


98,118
(redrawn). Only the part of the wave which drives out a droplet is shown.

𝑍 depends on the channel cross section 𝐴, the ink density 𝜌 and the speed of sound 𝑐
(equation 28). The effective speed of sound is influenced by the compliance of the
channel cross section. The reflection 𝑘𝑅 and transmission 𝑘𝑇 coefficients at the interfaces
between domain 1 (nozzle) and 2 (reservoir) are calculated according to equation 29.

46
Theoretical Background

When the compliance does not change, the expressions can be simplified as given in
equation 30.

𝜌∗𝑐
𝑍= equation 28
𝐴

𝑍2 − 𝑍1 2 ∗ 𝑍2
𝑘𝑅 = 𝑘𝑇 = equation 29
𝑍1 + 𝑍2 𝑍1 + 𝑍2

𝐴1 − 𝐴2 2 ∗ 𝐴1
𝑘𝑅 = 𝑘𝑇 = equation 30
𝐴1 + 𝐴2 𝐴1 + 𝐴2

In the given example of Figure II-23 a very large reservoir 𝐴2 >> 𝐴1 is supposed.
Consequently, the reflection coefficient at this open end is -1 and the transmission
coefficient is 0. This means that the pressure wave will be completely reflected at the
reservoir side with changing pressure amplitude (t 3). In the next step, the piezoelectric
element is discharged and returns to its initial shape. This creates a positive pressure
which amplifies the reflected wave (t 4). If the pulse is tuned properly, according to the
channel structure, the positive pressure pulse peaks at the nozzle (t 5) and viscous forces,
inertia and ink surface tension are overcome. This results in droplet formation.

Figure II-24: a) Calculated pressure at the entrance of the nozzle and shape of the
driving waveform. b) Measured drop speed as a function of the drop repetition rate
(DOD frequency). The largest peak corresponds to the channel resonance frequency.
Reproduced from 98 with permission from Elsevier.

47
Theoretical Background

Subsequently, the pressure at the nozzle oscillates, as the acoustic wave damps out. A
graph showing a typical nozzle pressure during such a process is given in Figure II-24 a).
In modern inkjet technology, microelectromechanical systems (MEMS)-based printheads
with very small geometries are often used.104 When the inlet to the pressure channel also
becomes a narrow section, a more closed reflection and completely changed acoustic
properties are the result. The nozzle and inlet act as partially closed boundaries, and thus
the standing wave inside the channel is truncated.
When geometries are small, the delay time of travelling waves plays no role and the
acoustic properties of the channel can be described as a Helmholtz resonator. In this
model, the large volume of the actuation channel can act as a spring, and the ink in the
inlet and nozzle as a vibrating mass 𝑚.104 The resonance frequency 𝐹𝑟𝑒𝑠 of such systems
without external load can in general be described by equation 31 with 𝑠𝑝 being the spring-
and 𝑑𝑎 being the damping constant.

1 𝑠𝑝 𝑑𝑎
𝐹𝑟𝑒𝑠 = √ − equation 31
2𝜋 𝑚 2𝑚

If the channel geometry is known (𝐴 cross section, 𝑙 length), 𝑚 and 𝑠𝑝 are expressed by
equation 32 and equation 33.

𝑚 =𝜌∗𝐴∗𝑙 equation 32

𝜌 ∗ 𝑐 2 ∗ 𝐴2 equation 33
𝑠𝑝 =
𝛥𝑉𝑜𝑙

It is important to consider the resonance behaviour of the printhead, as it can heavily


influence parameters such as droplet velocity and volume. The so-called DOD curve
(Figure II-24 b)) shows the velocity of an ejected droplet in dependence of the time delay
to the previously fired droplet. For low frequencies, the drop velocity is constant because
pressure waves had enough time to damp out. At higher frequencies, the remaining
acoustic pressure waves inside the channel as well as the position and movement (in or
out of phase) of the meniscus lead to variations of drop velocity or volume.118

48
Theoretical Background

2.2.6. Advanced pulse design


The performance of an inkjet printhead is clearly limited by the previously described
residual oscillations. One possible way of improving the printing consistency is to add an
additional pulse which actively damps the residual oscillations. In advanced systems, this
is done using either a unipolar or a bipolar approach (Figure II-25).

104
Figure II-25: Parameterisation of a) unipolar and b) bipolar pulse. Reproduced from
with permission from John Wiley and Sons.

The unipolar pulse consists of the jetting pulse plus an additional trapezoidal pulse with
the same polarity (quenching pulse).119 The bipolar pulse consists of the jetting pulse and
a quenching pulse of opposite sign.120 The advantage of the bipolar strategy is the faster
damping of the residual oscillations allowing higher printing frequencies. However,
driving electronics are more complex to realise.104 In both cases, the pulse parameters can
be expressed by a vector 𝛩 = [𝑡𝑟𝑅 𝑡𝑤𝑅 𝑡𝑓𝑅 𝑉𝑅 𝑡𝑑𝑄 𝑡𝑟𝑄 𝑡𝑤𝑄 𝑡𝑓𝑄 𝑉𝑄 ] (see Figure II-25 for
explanation of single parameters). To determine well working vector coordinates, first
hand information about the printhead acoustics is required or exhaustive experiments
have to be performed.104

49
Theoretical Background

2.2.7. Drop formation


After the driving pulse is provided, the droplet forms outside the nozzle and pinches off.
In this process, it is important to achieve one single spherical drop per actuation pulse in
order to print high quality images. This is especially critical when printing electronics
because any imperfection like satellite droplets or tails could easily lead to complete
failure of the whole device. Good droplet formation depends on both, ink properties such
as density, viscosity and surface tension and the printhead itself. Only if the printhead
specifications (for example the viscosity range given by the head manufacturer) and a
carefully chosen pulse shape exactly match the ink properties, good drop formation is
achieved. An example on how the driving pulse amplitude effects the droplet formation is
given in Figure II-26.

Figure II-26: Photograph sequence taken from a DOD drop formation cycle. The
droplets are recorded by single flash stroboscopy. In the left sequence, the actuation
signal has low amplitude, resulting in a slow droplet. The tail remains stable until the tail
droplet is merged with the head droplet due to capillary contraction of the ligament. In
the sequence on the right, the actuation signal amplitude is 25 % higher. The droplet is so
fast that the tail droplet cannot merge with the head droplet before the tail has broken up.
Eventually, the tail droplet merges with the head droplet. This figure demonstrates the
challenge of producing satellite-free droplets at a high velocity. Image courtesy of Marc
104
van den Berg, Océ Technologies B.V.. Reproduced from with permission from John
Wiley and Sons.
50
Theoretical Background

With low pulse amplitude, a droplet tail is forming which can catch up with the head
droplet already shortly after leaving the nozzle. For higher amplitude, the so-called head
droplet is too fast for the tail to catch up and a satellite is forming due to Rayleigh-Plateau
instability. Since head and substrate are moving with respect to each other during the
printing process, head droplet and satellite might land on different positions of the
substrate, resulting in a printing pattern with unwanted features. In general, printing
accuracy and productivity are increasing with droplet velocity, but this usually comes
with a rising tendency to form satellites.
It is also possible to theoretically analyse the droplet formation process (a more detailed
description is found in chapter 4 of reference 104). For the model, the formation of a single
droplet is regarded, which emerges as ink is flowing out of the nozzle. This ink is joined
by fast moving ink leaving the nozzle a bit later. As long as the ink stream is fast enough,
it joins the droplet, which thereby grows and accelerates. Eventually, the ink inside the
nozzle decreases in velocity and the droplet stops growing. Ink which is still ejected but
cannot overtake the droplet due to too low speed forms a ligament (tail). This ligament
can form a tail drop which may either merge with the head drop or break off due to
capillary instability.
In such case, the volume of the head drop 𝑉𝑜𝑙𝐷 can be calculated from the flow rate
through the nozzle during the time frame within which the droplet grows (equation 34). In
the given equation, 𝐴𝑛 is the nozzle exit area, 𝑢𝑛 the ink velocity at the nozzle, 𝑡0 the
time when the meniscus starts emerging and 𝑡𝑒 the time when the velocity of the ink gets
too low to reach the head droplet anymore.104

𝑡𝑒
𝑉𝑜𝑙𝐷 = ∫ 𝐴𝑛 ∗ 𝑢𝑛 ∗ 𝑑𝑡 equation 34
𝑡0

To estimate the droplet velocity, the momentum transfer due to viscous tension, surface
tension and advection into the droplet has to be considered.
The total amount of transferred momentum 𝑝𝜇 due to viscous tension is expressed by
equation 35, with 𝜇 being the dynamic viscosity.104

𝑝𝜇 = −3 ∗ 𝜇 ∗ 𝐴𝑛 equation 35

51
Theoretical Background

It is worth mentioning that the minus sign describes a pulling back of the droplet towards
the nozzle.
The momentum transfer 𝑝𝑐 due to surface tension 𝛾 can be calculated by simple
geometric considerations of the jet. The capillary tension is proportional to the jet radius,
which is equal to the nozzle size 𝑟𝑛 during the ejection process (equation 36).104

𝑝𝑐 = (𝑡𝑒 − 𝑡0 ) ∗ 𝜋 ∗ 𝑟𝑛 ∗ 𝛾 equation 36

The third momentum transfer mechanism to consider is advection 𝑝𝑎 (equation 37).

𝑡𝑒
𝑝𝑎 = ∫ 𝜌 ∗ 𝐴𝑛 ∗ 𝑢𝑛2 ∗ 𝑑𝑡 equation 37
𝑡0

Combining all the momentum contributions, the droplet velocity 𝑢𝑑𝑟𝑜𝑝 can be estimated
according to equation 38.104

𝑡𝑒
1
𝑢𝑑𝑟𝑜𝑝 = (∫ 𝜌 ∗ 𝐴𝑛 ∗ 𝑢𝑛2 ∗ 𝑑𝑡 − 3 ∗ 𝜇 ∗ 𝐴𝑛
𝜌 ∗ 𝑉𝑑𝑟𝑜𝑝 𝑡0
equation 38
− (𝑡𝑒 − 𝑡0 ) ∗ 𝜋 ∗ 𝑟𝑛 ∗ 𝛾)

Momentum contributions which enter the speed equation with a minus sign (e.g. viscous
tension) slow down or even pull back the droplet towards the nozzle plate. As a
consequence, it is even possible to create a free-standing droplet with a velocity towards
the nozzle at the time it pinches off. This happens if the bracket in equation 38 becomes
negative at all times, which leads to equation 39.104

𝛾 > 𝜌 ∗ 𝑟𝑛 ∗ 𝑢𝑛2 ⇔ 𝑊𝑒 < 1 equation 39

An impressive video showing this counterintuitive process of forming a droplet with


121
negative velocity is found in reference . The point where the jet velocity passes 0 is
usually referred to as the ‘jetting threshold’. An illustration of this process is provided in
Figure II-27.
Similarly to the above referenced calculation, it is also possible to theoretically describe
processes like tail formation and tail break-up. For further information the motivated
reader is referred to references.98,104

52
Theoretical Background

Figure II-27: The droplet velocity is plotted versus the maximum ejection velocity. The
velocities are scaled with the capillary velocity, so that the square root of the Weber
number with which the liquid is ejected is on the horizontal axis. The droplet velocity
increases with the velocity in the nozzle. The dashed line is the analytical prediction. The
solid line is the droplet velocity obtained from a numerical simulation using the slender-
jet approximation. For Weber numbers below the jetting threshold, no droplet is formed.
Reproduced from 104 with permission from John Wiley and Sons.

2.2.8. Drop impact


The drop impact on the substrate is critical for the quality of the resulting printing pattern.
Rioboo et al. describes six possible consequences for a droplet falling on a dry surface
(Figure II-28). These are:122

 deposition
 prompt splash
 corona splash
 receding breakup
 partial rebound
 complete rebound

53
Theoretical Background

For the description of these effects, interfacial hydrodynamics have to be considered. The
main influencing factors, as identified by Jung et al. (chapter 8 of reference 104), are liquid
properties like viscosity, elasticity, surface tension, droplet size, droplet velocity and
surface properties such as surface energy, roughness, planarity and chemical
homogeneity.

Figure II-28: A variety of morphologies of liquid drop impact onto a dry surface.
Reproduced from 123 with permission from Begell House, Inc..

A systematic investigation of drop impact on super-hydrophobic surfaces was carried out


by Clanet et al..124 For low viscosity liquids, such as water (dominant 𝑊𝑒), they found out
that the maximum spreading diameter 𝐷𝑚𝑎𝑥 scales with the initial drop diameter 𝐷0 and
𝑊𝑒 as given in equation 40.

1
𝐷𝑚𝑎𝑥 ~ 𝐷0 ∗ 𝑊𝑒 4 equation 40

They further experimentally verified that this law holds also on partially wettable
surfaces.

54
Theoretical Background

Figure II-29: a) High-speed imaging of a water drop impact on a superhydrophobic


surface (Time interval between the images: 2.7 ms). b) Maximum diameter of the
spreading drop normalised by the drop radius as a function of the Weber number. The
open squares correspond to those obtained with water drops and the filled ones
correspond to those obtained with mercury drops. The solid line indicates the slope 1/4.
Reproduced from 124 with permission from Cambridge University Press.

On the other hand, in a regime where the viscous dissipation dominates the surface
tension (dominant 𝑅𝑒), the kinetic energy of the impacting drop is dissipated by viscosity.
In their work, Clanet et al. describe this correlation mathematically and derive an
equation which relates 𝐷𝑚𝑎𝑥 to 𝑅𝑒 (equation 41).104,124

1
𝐷𝑚𝑎𝑥 ~ 𝐷0 ∗ 𝑅𝑒 5 equation 41

2.3. State of the art - inkjet printed OPV


In order to bring the results of this thesis into the right context and to identify the
contribution to the field, the state of the art in inkjet printed OPV should be described in
the following. This is done in close relation to a statement by F. Krebs in one of the most
cited reviews on OPV printing and coating techniques, which was published shortly after
first reports of inkjet entering the field in the year 2008.88

55
Theoretical Background

It is difficult at the current stage to determine whether inkjet printing will play an
important role in the high volume fabrication of polymer solar cells. The advantages
offered by the technique and the good literature examples of its use makes it likely that it
will become of use for creating complex patterns and perhaps devices with a small outline
of the active area.

Ideally, an answer to these speculations will be given at the end of the thesis, considering
the vast technological progress since then.
The first literature report using inkjet as production method for OPV dates back to the
year 2004 and was published by Shah and Wallace.125 With printed BHJ P3HT:C60 layers
they reached a 𝑃𝐶𝐸 of 0.00033 % and short circuit current densities of 0.0048 mA/cm2.
After publication of their conference proceedings, there was no significant progress until
the year 2007. The true starting point for inkjet to enter the OPV field can be attributed to
Hoth et al..126 They demonstrated first efficient inkjet printed P3HT:PC60BM layers
resulting in high 𝑃𝐶𝐸 values of 2.9 %. This was achieved by suitable solvent formulation
with mixtures of o-DCB and mesitylene as well as characterisation of morphological and
interfacial properties. Since then there has been continuously increasing interest in inkjet
production of OPV. This resulted in a significant number of publications up to today. A
non-exhaustive summary is given in Table II-5.

Table II-5: Literature overview of important studies about inkjet printed OPV.
𝑃𝐶𝐸
Ref Year Inkjet printed material Comment
(%)
125
2004 P3HT:C60a) First inkjet printed BHJ < 0.1
126
2007 P3HT:PC60BM First efficient inkjet printed BHJ 2.9
127
2008 P3HT:PC60BM New record efficiency 3.5
128
2008 P3HT:PC60BM Investigation of coffee ring 1.4
effect and pinholes
129
2009 P3HT:PC60BM Study of different solvent 3.5
mixtures and layer morphologies
130
2009 PEDOT:PSS Use of additives and surfactants 3.2
131
2009 PEDOT:PSS Comparison to spin and spray 3.3
coating

56
Theoretical Background

𝑃𝐶𝐸
Ref Year Inkjet printed material Comment
(%)
132
2010 P3HT:PC60BM Study of different solvent 2.4
mixtures and layer morphologies
133
2010 PEDOT:PSS Use of high boiling point 3.7
additives ODT b) and CNc)
P3HT:PC60BM
134
2010 ITO nanoparticles Thermal annealing at 450 °C 2.13
resulting in 84 % transmittance
and 207 Ω/sq sheet resistance
135
2011 PCPDTBT:bis-PC60BMd), Combinatorial screening with -
PCPDTBT:PC60BM morphology investigation 1.48
PSBTBTe):PC60BM 0.64
136
2012 Ag grid front electrode R2R fabrication and comparison 0.75
to flexography and thermal
imprinting
137
2012 Ag grid front electrode Effect of grid and busbar height 1.54
PEDOT:PSS measured and simulated
83
2013 Ag grid front electrode R2R fabrication and 1.7
demonstration of module with
16 cells
138
2013 Ag grid back electrode Alternative to evaporated metal 1.61
Ag full back electrode
1.96
139
2013 PFDTBTP f):PC60BM Chlorine-free solvent system 2.7
based on anisole and tetralin
140
2014 PEDOT:PSS First report of fully inkjet 2
PCDTBT:PC70BM printed OPV
ZnO nanoparticles
Ag full back electrode
141
2015 AgNW back electrode Poor mesh uniformity resulting 2.7
in 44.9 Ω/sq at 86 % average
transparency

57
Theoretical Background

𝑃𝐶𝐸
Ref Year Inkjet printed material Comment
(%)
142
2015 PEDOT:PSS Modules on 92 cm² active area 0.98
P3HT:PC60BM with non-halogenated solvents
ZnO nanoparticles (evaporated electrodes)
143
2015 Ag grid front electrode Fully inkjet printed devices, 1.7
ZnO nanoparticles Freedom of form cell
PEDOT:PSS demonstrated with Christmas 4.1
P3HT:PC60BM or tree shape
PV2000g):PC60BM
PEDOT:PSS
Ag grid back electrode
144
2016 ZnO precursor Use of precursor ZnO and 3.2
P3HT:PC60BM comparison to spin coating
145
2016 P3HT:ICBAh) Printing of eco-friendly active 2.9
nanoparticles layer nanoparticle dispersion
146
2018 PEDOT:PSS Combination of solvents and 1.9
surfactants to reduce surface
tension
147
2018 Cu nanoparticles Copper grids as ITO 3.4
replacement
a)
C60: Buckminsterfullerene
b)
ODT: 1,8-Octanedithiol
c)
CN: 1-Chloronaphthalene
d)
bis-PC60BM: Bis(1-[3-(methoxycarbonyl)-propyl]-1-phenyl)-[6,6]C62
e)
PSBTBT: Poly[(4,4′-bis-(2-ethylhexyl)-dithieno(3,2-b;2′,3′-d)silole]-2,6-diyl-alt-(2,1,3-
benzothiadiazole)-4,7-diyl]
f)
PFDTBTP: Poly[9,9-dioctylfluorenyl-2,7-dyil-co-(10,12-bis(thiophen-2-y)-3, 6-dioxooctyl-11-thia-
9,13-diaza-cyclopenta[b]triphenylene]
g)
PV2000: Tradename active layer polymer from Raynergy TEK
h)
ICBA: indene-C60 bisadduct

As shown in the table, tremendous progress has been made since the introduction of
inkjet technology to the field of OPV and since this thesis work was started in 2013.

Active layer:

Most reports focus on printing the active layer. This includes the analysis of pinhole
formation,139 coffee ring effects,128 the transfer from chlorinated to environmentally non-

58
Theoretical Background

critical solvents139,145 and the use of processing additives.133 Results are usually ink
formulations which differ quite significantly from those used for doctor blading, spin
coating or slot die coating. In general, solvents or solvent mixtures with high boiling point
and low evaporation rate are preferred because they lead to extended open time (time
until a nozzle clogs when idle). Reported efficiency losses, when switching from blade-
or spin coated active layers to inkjet printed equivalents, are widely distributed.132,142
These variations highlight the importance of good processing control for achieving well
working OPV devices.
Most publications focus on very small printing areas of only a few mm² and provide little
information on the robustness and scalability of the applied inkjet process. Many critical
parameters like the printability (Ohnesorge theory) and stability of the ink against gel
formation are barely addressed. This missing information should be provided by reporting
the development of OPV active layer inkjet inks with the focus on both, a stable and
scalable printing process and high device efficiencies.

EEL and HEL:

Reports of inkjet printed EELs and HELs, such as ZnO and PEDOT:PSS, show similarly
wide distribution of OPV 𝑃𝐶𝐸𝑠. While some authors achieve lower efficiencies when
switching from their reference process to inkjet technology,131 others demonstrate only
negligible performance losses.142,144 An example is the work by Singh et al. They were
able to inkjet print ZnO EELs from a precursor with similar efficiencies to spin coated
references and highlighted the critical parameters that affect film formation. These are ink
concentration, viscosity, substrate surface treatment, drop spacing, substrate temperature
annealing temperature, the surface tension of the ink and the rate of solvent
evaporation.144 Inkjet printing of PEDOT:PSS HELs was demonstrated by Eom et al.. 130
They modified their inks with additives of glycerol and surfactants, resulting in improved
surface morphology, high conductivity and OPV 𝑃𝐶𝐸 of 3.16 %. In this work, the first
aim is to achieve this state of the art and use it as building block for manufacturing of
highly efficient fully inkjet printed OPV devices. In a second step, the state of the art
should be surpassed by introduction of a novel strategy to print homogeneous and directly
structured HELs on surfaces with challenging wetting behaviour, such as the active layer
polymers P3HT:PC60BM, without the use of environmentally critical surfactants.

59
Theoretical Background

Electrodes:

Regarding the literature reports, it is clear that the electrodes are the most critical and
challenging layers for inkjet application. Performance losses when switching from
sputtered ITO front or evaporated metal back electrodes to inkjet are usually quite drastic.
Works about inkjet printing of transparent conductive oxide layers from nanoparticles
rely on very high sintering temperatures of more than 400 °C.134 The resulting electrodes
show still comparatively high sheet resistances of more than 200 Ω/sq at 84 %
transmittance, indicating that this strategy is not the best to follow if cost efficient
printing of electronics is targeted. Most works about inkjet printed electrodes focus on a
combination of PEDOT:PSS and silver grid structures. This combination allows to make
maximum use of the core advantage that comes with inkjet technology, namely to directly
produce the grid pattern without loss of the expensive silver ink.

Figure II-30: a) J-V curves of 2x2 cm2 devices with ITO, PEDOT:PSS and Ag-
grid/PEDOT:PSS electrodes. b)-d) Inkjet printed Ag grids covered by an inkjet printed
layer of PEDOT:PSS for collection of current. b) Maximum height of the grid lines is
400 nm, PEDOT:PSS thickness is 100 nm. c) Maximum height of the grid lines is 600 nm,
PEDOT:PSS thickness is 100 nm. d) Maximum height of the grid lines is 600 nm,
PEDOT:PSS thickness is 200 nm. Reproduced from 137 with permission from Elsevier.

In a paper by Galagan et al., silver grids were inkjet printed and the effect of the height
and width of lines and busbars on the performance of P3HT:PC60BM solar cells was

60
Theoretical Background

investigated.137 In their work, an increase in the height resulted in a simultaneous increase


in width of the grid-lines and following enhanced shadowing effects. Furthermore, they
found out that too high grids (> 600 nm) limit the possibility for over-coating by
PEDOT:PSS and active layer und consequently also the applicability as OPV front
electrode. Compared to ITO references with 1.99 % 𝑃𝐶𝐸, their Ag-grid/PEDOT:PSS
combination achieved a performance of 1.54 %. Another very informative study of the
influence of inkjet metal grid/PEDOT:PSS combinations as front and rear electrodes was
published by Eggenhuisen et al..143 In their work, they started with a spin coated cell,
using ITO front and evaporated silver back contacts. Subsequently they substituted layer
by layer by an inkjet printed equivalent. Their PV2000 active ink cells dropped in
efficiency from 4.9 % to 4.7 % when replacing the ITO by inkjet printed metal grids and
further to 4.1 % when also replacing the back contact by an inkjet printed grid structure.
Finally, a work by Lu et al. about inkjet printing of AgNWs as transparent electrode for
OPV applications should be mentioned.141 One main advantage of this approach
compared to silver grids is the optical homogeneity of the devices. Considering the
potential applications (see chapter 1.3), the unwanted optical features introduced into an
OPV product by a grid structure could be of disadvantage. By applying their inkjet
printed nanowires as back contact, they were able to achieve 2.7 % 𝑃𝐶𝐸 P3HT:PC60BM
devices. Compared to their evaporated reference cells with 3.15 % 𝑃𝐶𝐸, the nanowires
devices suffer from limited 𝐹𝐹 of 33-54 %. Microscopic images reveal relatively short
average AgNW length of less than 10 µm and poor AgNW mesh uniformity which results
in a sheet resistance of 44.9 Ω/sq at an average transmission in the visible range of 86 %.
In summary, there is still significant room for improvement in the field of inkjet printed
electrodes. One goal of this work is to push the record efficiencies for inkjet printed silver
grid electrodes. Furthermore, optically inconspicuous inkjet printed AgNW mesh
electrodes which offer a balance of sheet resistance and transmittance comparable to
those manufactured by conventional spray- or blade coating should be introduced.
Therefore, it is necessary to develop new inkjet inks which offer stable drop formation
and ejection of AgNWs with an average length in the ~30 µm range.

Solar modules:

To achieve the main goal of this work, namely to demonstrate inkjet printing of OPV on
an industrially relevant scale, the transition from small to large areas, and thus from single

61
Theoretical Background

cells to solar modules, is necessary. In many of the above-mentioned publications authors


speculate about large scale application of inkjet technology in OPV.126,128 However, very
little works address demonstration of such process.83,136 Some reports are even limited to
single nozzle printing.140 Intuitively, with increasing printing area the processing
complexities increase. Larger areas require printheads with hundreds or thousands of
nozzles. This leads to a rising probability for nozzle clogging and printing defects.
Furthermore, ink sedimentation and evaporation at the nozzle plate make the use of ink
circulation systems necessary. One of the most significant works in the field was
published by Angmo et al., who demonstrated the use of inkjet printed silver grids as
front electrode of solar modules.83 Their 15.4 cm² big P3HT:PC60BM solar modules with
inkjet printed silver grid achieved up to 1.7 % efficiency. Except the base electrode, all
other layers were realised by conventional deposition methods, such as slot-die coating or
rotary screen printing. Further progress was reported by Eggenhuisen et al., who
demonstrated 92 cm² size P3HT:PC60BM modules with inkjet printed buffer- and active
layers.142 As base electrode they applied sputtered Mo-Al-Mo grids which were patterned
by photolithography. Up to now, fully inkjet printed solar modules, have not been
demonstrated and are thus a main topic of this work. At best, the fully inkjet printed
devices should be of arbitrary shape and colour. It is furthermore necessary, to mask the
visually obstructive cell-to-cell interconnection area in conventional modules in order to
meet the high demands of modern OPV applications (see chapter 1.3). Therefore, new
concepts of contact formation between adjacent cells have to be developed. Finally, to
demonstrate industrial feasibility, the printing processes have to be transferred to high-
throughput single-pass inkjet machine.
Based on the above given literature overview, a list of critical requirements for the
successful implementation of an industrial OPV inkjet printing process has been created
(chapter 1.5, Figure I-3). The target of this thesis it the simultaneous fulfilment of these
requirements by outlining the complete process chain from ink formulation to large area
device application, which makes this work a ‘tool box’ for everybody who is working on
printed electronics.

62
CHAPTER III: Experimental

Abstract:

This chapter presents the materials, machines and methods applied in this thesis for the
purpose of OPV manufacturing and characterisation. It furthermore contains descriptions
of finite element simulation models for dimensioning and loss analysis of silver grid
electrode structures as well as solar module interconnection regimes.

Part of this chapter has been published in:


P. Maisch, K.C. Tam, P. Schilinsky, H.-J. Egelhaaf, C.J. Brabec, Shy Organic Photovoltaics: Digitally
Printed Organic Solar Modules with Hidden Interconnects, Sol. RRL 2, 1-9 (2018). doi
10.1002/solr.201800005, reproduced with permission from John Wiley and Sons.
Only parts of the publication authored by P. Maisch are reproduced in subchapter 3.4.2 of this work.
Additionally, permission to utilise the whole content of the publication as part of this thesis was given from
all co-authors involved through author ownership declarations.

63
Experimental

3.1. Materials
All materials used for the experiments are commercially available. A summary is given in
Table III-1.

Table III-1: Materials used in this thesis.

Type Material Trade name Supplier

Substrates Glass - Weidner Glas


GmbH

PET foil Melinex ST504 DuPont Teijin


Films UK Ltd

Electrodes ITO sputtered glass - Weidner Glas


15-20 Ω/sq GmbH

Ag evaporation pellets EVMAG40EXE-B Kurt J. Lesker Co.

AgNWs in water or ClearOhm Cambrios Advanced


isopropyl alcohol (IPA) Materials Co.

AgNW:PEDOT:PSS Clevios HY E Heraeus Clevios


mixture GmbH

Silver ink EMD 5800 SunChemical Co.

Silver ink Sicrys I50T13 P.V. Nano Cell Ltd.

Silver ink Sicrys I40DM106 P.V. Nano Cell Ltd.

Buffer layer ZnO nanoparitcles N-10 Avantama AG

PEDOT:PSS Clevios P VP Al Heraeus Clevios


4083 GmbH

PEDOT:PSS Clevios F HC Solar Heraeus Clevios


GmbH

PEDOT:PSS P Jet N V2 Heraeus Clevios


GmbH

64
Experimental

Type Material Trade name Supplier

Active layer P3HT RR = 94.2 % Lisicon SP001 Merck Chemicals


Mwa) = 54.2 kg/mol GmbH

PBTZT-stat-BDTT-8 Lisicon PV-D4610 Merck Chemicals


Mw = 49.5 kg/mol GmbH

Not disclosed ActivInk PV 2000 Raynergy Tek Inc.

PC60BM > 99 % - Solenne BV.

PC70BM > 99 % - Solenne BV.

Solvents Water DIb) - -

IPA > 99 % - Th. Geyer GmbH &


Co. KG.

Aceton > 99 % - Th. Geyer GmbH &


Co. KG.

1-Pentanol - Sigma-Aldrich Co.

1,2-Propanediol (PG) - Sigma-Aldrich Co.


> 99.5 %

Toluene > 98 % - VWR


International GmbH

o-DCB > 99 % - Sigma-Aldrich Co.

CB > 99.5 % - VWR


International GmbH

o-Xylene (o-X) > 98 % - Merck KGaA

1,2,3,4- - Sigma-Aldrich Co.


Tetrahydronaphtalene
(THN) > 99%

TMB > 98 % - Merck KGaA

4-Bromoanisole (BrAni) - Alfa Aesar GmbH &


≥ 99 % Co KG

Indane > 95 % - Sigma-Aldrich Co.

65
Experimental

Type Material Trade name Supplier

Wetting Fluorosurfactant Capstone FS 31 DuPont Co.


agents (CFS)

Surfactant (ethoxylated Dynol 604 Air Products


acetylenic dioles) Chemicals Europe
B.V.

Encapsulation Epoxy resin (UV) c) Katiobond LP655 DELO Industrie


curable) Klebstoffe GmbH &
Co. KGaA

Encapsulation foil, Rolic Barrier Films Rolic Tehnologies


water vapour Ltd.
transmission rate down
to 10-6 g/(m2*d)
a)
Mw = molecular weight
b)
DI = deionised
c)
UV = ultraviolet light

3.2. Fabrication methods

3.2.1. Reference processes


Well working small scale organic solar cells were manufactured on a frequent routine and
are used as references in this work. The commonly applied layer sequence is:

ITO/ZnO/P3HT:PC60BM or PV2000:PC70BM/PEDOT:PSS/Ag

Figure III-1: Reference solar cell layout.

66
Experimental

The associated layout with six solar cells of 10.4 mm² (2x5.2 mm²) active area each on a
25x25 mm² substrate is shown in Figure III-1.

The processing steps for fabrication of a reference device are the following:

 LASER patterning of ITO substrate


 Ultrasonic cleaning with water, acetone and IPA
 Blade coating of ZnO nanoparticles (N-10 Avantama AG, 30 °C, 50 nm)
 Blade coating of active layer (P3HT:PC60BM from CB + 5 %vol BrAni, 65 °C,
120 nm), (PV2000:PC70BM from o-X:THN 1:1 (vol.), 70 °C, 200 nm)
 Blade coating of PEDOT:PSS (Clevios F HC Solar:H2O 1:1 (vol.) + 0.2 %wt
CFS, 60 °C, 100 nm)
 Annealing in nitrogen atmosphere (140 °C)
 Thermal evaporation of silver (100 nm)
 LASER patterning of PEDOT:PSS
 Encapsulation with UV curable epoxy resin (Delo LP 655)

A picture of the used equipment can be seen in Figure III-2.

Figure III-2: a) Doctor blade and b) LASER at the EnCN laboratories. Reproduced with
permission from the Energy Campus Nuremberg and K. Fuchs.

67
Experimental

A doctor blade from Zehntner GmbH was used for layer coating. It is equipped with a hot
plate (max. 140 °C), a speed control (1 - 99 mm/s) and applicators with a width of 6 cm
or 20 cm.
For LASER structuring a LS-7xxP setup from LS Laser Systems GmbH was used. It is
equipped with a High Q LASER (femtoREGEN UC – 1040-8000 fs Yb SHG) emitting at
520 nm (first harmonic) and 1040 nm with a pulse duration < 350 fs. The advantage of
such ultra short pulses is the avoidance of heating damage to sensitive substrates. The
maximum repetition rate is 960 kHz. By galvanometric control, the LASER beam can
pattern the substrate surface at speeds up to 4000 mm/s. A camera system is included for
positioning. For the last year of this thesis LASER setup LS-6KP4P520 from LS Laser
Systems GmbH was used. It is equipped with Spectra-Physics Spirit 1040-8-SHG laser
source emitting at 520 nm (first harmonic) with a pulse width < 400 fs. The maximum
repetition rate is 960 kHz. It is using the same galvanometric control as LS-7xxP
achieving maximum speeds up to 4000 mm/s.
Silver back electrodes were thermally evaporated using a Univex 250 machine from
Oerlikon Leybold Vacuum GmbH. Shadowing masks made from steel foil were used to
define the electrode area. Evaporation was commonly done at a pressure < 5*10-6 mbar
with a rate of 0.5 Å/s - 3.0 Å/s until a final thickness of 100 nm was reached.

3.2.2. PiXDRO LP50


Sheet-to-sheet inkjet printing was done with a PiXDRO LP50 inkjet system. PiXDRO is a
brand of Roth & Rau B.V., a member of the Meyer Burger Group. A schematic picture of
the machine is provided in Figure III-3.
The PiXDRO LP 50 setup consists of eight modules, which are briefly described in the
following.
The motion system consists of the five axes listed below. A close look at the motion
system the is provided in Figure III-4:148
 Y-axis: Main in-scan print axis - substrate table movement
 X-axis: Cross scan axis - printhead assembly movement
 Z-axis: Printhead height - printhead assembly movement
 Phi-P-axis: Printhead rotation
 Phi-S-axis: Substrate rotation

68
Experimental

148
Figure III-3: Schematic drawing of PiXDRO LP50 modules. Reproduced from with
permission from Meyer Burger.

The substrate table provides enough space for samples up to A4 size. It is connected to
the Y-axis of the motion system and equipped with a vacuum system. Due to the pressure
difference between the vacuum and the outside air, the substrate is held in place during
movement of the table.
The printhead assembly (see Figure III-4) includes an ink reservoir, which can be filled
with up to ~15 ml of ink. A hose connects the reservoir cap to a vacuum pump. This is
necessary to provide a meniscus pressure 𝑃𝑚𝑒𝑛𝑖𝑠𝑐𝑢𝑠 that prevents ink from flowing
through the head due to gravity. The necessary printhead meniscus pressure depends on
the ink as well as the printhead and is typically in the range of 2.5 to 10 mbar. To
compensate for the gravitational force of the ink inside the reservoir, the pressure 𝑃𝑠𝑒𝑡 to
set as a machine is adjusted according to equation 42, with 𝜌 being the density, 𝑔 the
gravitational acceleration and ℎ the height of the reservoir above the nozzle plate. For the
setup used, h is approximately 20 cm, leading to 𝑃𝑠𝑒𝑡 values in the range of -15 to
-25 mbar.

𝑃𝑠𝑒𝑡 = −(𝑃𝑚𝑒𝑛𝑖𝑠𝑐𝑢𝑠 + 𝜌 ∗ 𝑔 ∗ ℎ) equation 42

69
Experimental

The ink flows from the reservoir through a hose into a heating block, which is located
directly on top of the printhead. Ink temperatures of up to 100 °C can be realised, but low
temperatures as are usually preferred because increased evaporation rates at the nozzle
plate can cause nozzle clogging.

Figure III-4: PiXDRO LP50 Printer. a) Drawing of the printhead assembly. b) Drawing
148
of the motion system. Reproduced from with permission from Meyer Burger. c)
Photograph of the printing setup at the EnCN laboratories. Reproduced with permission
from the Energy Campus Nuremberg and K. Fuchs.

The PC with LP50 HMI (human machine interface) provides the necessary control unit
for the user. It is used for optimising and checking all the LP 50 functions ranging from
printing pattern management to exact pressure and printhead voltage control.
The dropview system consists of a camera and a light emitting diode (LED), which
provides short light pulses. When matching the droplet firing frequency with the light
pulses, the camera records a ‘standing droplet’, which can be analysed at the computer
screen (see for example Figure II-26). By changing parameters like the pressure setting,

70
Experimental

piezo voltage or frequency, the droplet can be optimised in real time to achieve a well-
defined spherical shape.
The waste station removes excess ink from the capping station. It consists of a waste
bottle and vacuum membrane pump.
The wiping and capping station includes a tissue roll where the printhead can
automatically wipe across to remove dirt or excess ink from the nozzle plate.
Furthermore, a system of rubber lips connected to the waste station can be used to remove
ink from purging or to apply vacuum directly to the nozzle plate. In some cases, the
sucking, purging and wiping mechanisms can help to recover clogged nozzles.

The printview and alignment system consists of a light source and a camera, which moves
together with the printhead. It is used to define the print position by automatic or manual
alignment at cross or dot shaped markers. The positioning system can be used to correct x
and y axes offsets as well as tilting of the substrate using the Phi-S-axis.
The PiXDRO LP50 printer can be equipped with different printheads. In this work,
Spectra SE 128 AA and Spectra SL 128 AA heads were used. The head specifications as
given by the manufacturer FUJIFILM Dimatix are summarised in Table III-2.

Figure III-5: a) Dimatix FUJIFILM Spectra SE 128 AA printhead used in this work. b)
Schematic illustration of trapezoidal voltage pulse applied to the inkjet printhead
piezoelectric elements.

Stimulation of the printhead piezoelectric element is done using a trapezoidal voltage


pulse (see Figure III-5 b)). The pulse shape is adjusted by specifying the voltage
amplitude, the ramp-up time (µs), hold time (µs) and ramp-down time (µs). In the
following, the three characteristic times are summarised as pulse shape (µs-µs-µs).

71
Experimental

Table III-2: Spectra S-class printhead specifications as given by the manufacturer.149

Printhead SE 128 AA SL 128 AA

Number of nozzles 128 128

Nozzle spacing 508 µm (50 DPI) 508 µm (50 DPI)

Nozzle diameter 35 µm 50 µm

Calibrated drop size 30 pl 80 pl

Adjustment range for drop size 25 - 30 pl 65 - 90 pl

Drop size variation 4% 4%

Jet straightness 3 mrad [0.17°] 3 mrad [0.17°]

Nominal drop velocity 10 m/sec 8 m/sec

Drop velocity variation 5% 5%

Crosstalk, maximum 5% 5%

Operating temperature range up to 90°C up to 90°C

Fluid viscosity range 8 - 20 cP 8 - 20 cP

Compatible jetting fluids Organic solvents, UV Organic solvents, UV


curables curables

Maximum operating frequency 40 kHz 30 kHz

Both printhead types have a native resolution of 50 DPI. However, for most inks used in
this work, 400 - 1000 DPI are necessary to create homogeneous wet films with sufficient
thickness. In order to achieve higher DPI numbers, the printer can pass over the substrate
multiple times (multipass printing). With the first pass, each active nozzle creates a row
of droplets on the substrate while moving in in-scan (y) direction. Subsequently, the
printhead is shifted in cross-scan (x) direction by 1/DPI and a second parallel row of
droplets is created. This process is repeated until the pattern is finished.
Another approach is to adjust the printhead angle (Phi-P-axis). A schematic illustration of
this strategy is given in Figure III-6.

72
Experimental

Figure III-6: Schematic illustration of relation between printhead angle and DPI number.

The relation between native head resolution, the head angle (𝛽) and printing pattern
resolution is given by a simple trigonometric equation 43.

𝐷𝑃𝐼𝑛𝑎𝑡𝑖𝑣𝑒 ℎ𝑒𝑎𝑑
𝑐𝑜𝑠(𝛽) = equation 43
𝐷𝑃𝐼𝑝𝑟𝑖𝑛𝑡 𝑝𝑎𝑡𝑡𝑒𝑟𝑛

Using this strategy, it is possible to print a small image in the correct resolution with only
one pass. Due to the limited printhead width, larger images are built up in blocks.
Compared to the multi-pass procedure described above, the single-pass approach is
significantly faster. Higher industrial relevance and the speed advantage are the main
reasons for focusing on single pass printing in this work.
Finally, a short list of the most critical printing parameters is provided in Table III-3.

Table III-3: List of most important operation parameters at the PiXDRO LP 50.

Parameter Unit Explanation

Printhead control

Pulse shape µs-µs-µs Shape of voltage pulse applied to printhead


piezoelectric element specified by ramp-up –
hold – ramp-down

Piezoelement voltage V Maximum voltage applied to printhead


piezoelectric elements

73
Experimental

Parameter Unit Explanation

Meniscus pressure mbar Negative pressure to prevent ink from flowing


through the nozzles due to gravity (see equation
42)

Head temperature °C Temperature at printhead ink channel (RT if not


specified otherwise)

Firing frequency Hz Number of droplets fired from one nozzle per


second (typically 3 to8 kHz)

Print pattern control

Resolution X and Y DPI Resolution in X and Y direction can be chosen


independently

Quality factor (QF) - Number of nozzles used to print one line in in-
scan direction (1 if not specified otherwise)

Step size - Number of pixels to skip in cross-scan direction


for a nozzle before using that nozzle again (1 if
not specified otherwise)

Printhead angle ° Can be used to print at higher DPI numbers than


native printhead resolution (see equation 43)

Number of nozzles - Every nozzle can be addressed individually. If a


nozzle is blocked it can be excluded from
printing

Substrate temperature °C Heating of substrate table up to 60 °C possible

3.2.3. Loop coater single-pass prototype inkjet


To demonstrate inkjet printing on an industrially relevant level, a single pass prototype
inkjet module was used. It was constructed by Durst Phototechnik Digital Technology
GmbH based on the OPV printing expertise which was gained from the laboratory scale
experiments of this work. The setup of this sophisticated inkjet machine at the EnCN
laboratories and the transfer of the printing processes developed at the PiXDRO LP50 to
industrial scale was one of the major goals of this thesis. The inkjet module was

74
Experimental

integrated into a quasi-R2R system with the web running in an endless loop without any
un- or rewinder. In the following, it is referred to as ‘loop coater’ (see Figure III-7).

Figure III-7: Top: a) Schematic illustration of loop coater system. b) Photograph of


SAMBA prototype print stations. c) Photograph of recirculation system.

The loop coater is equipped with a splicing section to cut and replace the 7 m long and up
15 cm wide substrate web. A pressurised air driven movable roller controls the tension of
the web and an edge guide system with 150 µm precision keeps the web running straight.
For drying and curing purposes, the machine is equipped with two ovens combining hot
air and infrared drying (up to 140 °C) and a 6 kW infrared heating system from Adphos
(NIR126-125-400V). Two slot die coating sections together with the inkjet make the loop
coater a highly versatile machine.
The inkjet module consists of four print stations for different inks. Each station can be
equipped with up to four printheads, which are connected to an ink reservoir with
circulation (Megnajet CIMSII). The reservoir, as depicted in Figure III-7, contains a
floating switch and refilling pump, a temperature control, two recirculation pumps, a
damping unit to decay any pressure peaks from the pumps, an air pump and two pressure

75
Experimental

sensors at the in- and outlet. The meniscus pressure (see previous chapter) is set indirectly
as difference between infeed and recirculation pressure. The printing command can be
triggered manually or by a light barrier. A camera system in combination with printed
markers can detect any misalignment and correct the printing pattern by pixel shift. This
enables the exact positioning of every layer to create complex multilayer structures like
solar modules.
SAMBA heads from FUJIFILM Dimatix were used for printing. Specifications from the
manufacturer are listed in Table III-4.

Table III-4: SAMBA printhead specifications as given by the manufacturer.149

Printhead SAMBA

Number of nozzles (matrix array) 2048

Native resolution 1200 DPI

Drop volume 2.4 pl

Frequency Up to 100 kHz

Fluid viscosity range 2 - 10 cP

Versa drop technology Capable of multipulses and gray scale

Ink recirculation Continuous

Shape Designed for large arrays

Construction MEMS

The SAMBA technology is based on silicon microelectromechanical systems (MEMS)


manufacturing. It is an extensible ‘printhead on a chip’ technology (terms used by
manufacturer).149 The development can be compared to the evolution of integrated
circuits from single chips and limited function to extremely high integration densities.
The material used is silicon (Si) because of its resistance to mechanical abrasion, high
temperature tolerance and chemical robustness. However, the brittle nature of this
material demands extremely careful handling.

76
Experimental

Figure III-8: a) Photograph of SAMBA printhead. b) Example of VersaDrop multipulses.

Table III-5: List of most important operation parameters at the loop coater SAMBA
prototype inkjet.

Parameter Unit Explanation

Printhead control

VersaDrop pulse µs, V Shape of voltage pulse (single or multiple) applied to


design printhead piezoelectric elements

Infeed pressure mbar Pressure at the ink reservoir outlet

Recirculation pressure mbar Pressure at the ink reservoir inlet

Firing frequency Hz Number of droplets fired from one nozzle per second

Print pattern control

Resolution DPI 600 or 1200 (no free choice because of nozzle


alignment in matrix and static printhead position)

Delay time s Time between light barrier trigger and printing event

Pixel shift px Shift of printing pattern to correct for web


misalignment calculated from print marker recording

77
Experimental

The piezoelectric crystals are incorporated into a thin barrier layer and placed at the top of
the ink channel, which acts as pumping chamber. This enables the design of an entire
nozzle micro array matrix with high resonance frequency from a single silicon monolithic
structure. The printhead has the shape of a parallelogram, allowing the placement of
multiple heads in close proximity to each other to create a printbar for large areas.
The VersaDrop technology from FUJIFILM Dimatix is used for controlling the printing.
It allows the activation of a piezoelectric crystal with several pulses of varying amplitude
to produce metered amounts of ink, which are pumped into a single drop that detaches
from the nozzle. The modulation of drop size by multipulses is possible due to the high
frequency response of the printheads and provides a robust strategy to achieve high
resolution grayscale printing.
All components of the single pass prototype machine described above can directly be
integrated into modern large format R2R machines. The gained results and coating
parameters are therefore considered highly transferrable to industrial formats and of high
relevance to the field.

3.2.4. Encapsulation
All devices were encapsulated prior to solar simulator measurement. For this purpose, an
UV-curable epoxy glue (Katiobond LP 655) from DELO GmbH was applied on the
encapsulation glass or foil.

Figure III-9: a) Lamination process of a flexible OPV module. b) OPV module entering
the UV station. Reproduced with permission from the Energy Campus Nuremberg and K.
Fuchs.

78
Experimental

For large devices (> 5 cm²) a precision laminator (model UVL 60.4 pro) from FETZEL
Maschinenbau GmbH was used to ensure bubble-free and homogeneous distribution of
the glue. The resin was cured in a home-made UV station with strong emission at 410 nm
(absorption wavelength of the glue) for 10 minutes. For smaller devices < 5 cm² the
encapsulation glass or foil were manually positioned on the substrate. Curing was done
with an UV setup UVACUBE 100 from Hönle AG located in a nitrogen filled glovebox.

3.3. Characterisation
In order to compare and optimise inks, layers and devices, a variety of different
characterisation methods was used. A brief description is provided in following sub-
chapters.

3.3.1. Ink characterisation

3.3.1.1. Viscosity

Ink viscosities were measured using a rotational rheometer HAAKE RheoStress 6000
from ThermoScientific. In dependence of the expected viscosity range, suitable
measurement geometries (cone and plate or coaxial cylinder) were selected. The
temperature was kept at 23 °C for all measurements using a thermostat. The shear rate
was steadily increased for 300 s to a value of 10000 s-1, held constant for 300 s and
afterwards steadily decreased to 0 s-1.

3.3.1.2. Surface energy/tension

To obtain surface energies of solid surfaces and surface tensions of inks, a video based
optical contact angle measuring instrument OCA 20 system from DataPhysics was used.
Contact angles were measured according to the sessile drop method.150 For this purpose, a
droplet was placed on the surface, illuminated by diffuse light and recorded with a CCD
camera. The contact angle was obtained by fitting the droplet shape with the device
software. To gain information about polar and dispersive parts of the surface energy,
contact angles of known polar and dispersive liquid droplets on the substrate were
measured. Diiodomethane was used as highly dispersive, ethylene glycol as intermediate
and water as polar reference liquid. Values were used as pre-set in the software
79
Experimental

(Diiodomethane: Busscher et al. 𝛾𝐿𝑃 = 2.60 mN/m and 𝛾𝐿𝐷 = 47.40 mN/m; Water: Sell et
al. 𝛾𝐿𝑃 = 50.65 mN/m and 𝛾𝐿𝐷 = 22.10 mN/m; Ethylene glycol: Janczuk et al. 𝛾𝐿𝑃 =
18.91 mN/m and 𝛾𝐿𝐷 = 29.29 mN/m). From the measurements, polar and dispersive parts
of the surface energy were calculated according to the OWRK method (see chapter
2.2.4.2).
To obtain information about the surface tension of the liquids, the pendant drop method
was used.151 To this end, a picture of a droplet hanging from a cannula was recorded. The
droplet shape, which is dependent on gravity as well as surface tension, was analysed by
an imaging software.

1 1
∆𝑃 = 𝛾 ( + ) equation 44
𝑟1 𝑟2

The surface tension was obtained from the Yong-Laplace equation (equation 44), which
relates the main droplet curvature radii 𝑟1 and 𝑟2 to the pressure difference ∆𝑃 at the
curved surface. For the determination of the dispersive and polar parts of the surface
tension, contact angle measurements on a highly dispersive polydimethylsiloxane
(PDMS) (𝛾𝑆𝐷 = 21.53 mN/m) or polytetrafluoroethylene (PTFE) (𝛾𝑆𝐷 = 16.76 mN/m)
reference substrate were performed.
To guarantee reproducibility, all surface energy, -tension and contact angle measurements
were performed at least three times at the same conditions. Reported numbers are
averaged values of these measurements.

3.3.2. Layer characterisation

3.3.2.1. Optical microscopy

For imaging two different optical microscope systems were used. An MX51 from
Olympus Co. provides imaging capabilities in bright- as well as dark field and polarised
modes.
A confocal microscope µsurf custom from NanoFocus AG was used for the
characterisation of layer topographies and thickness measurement. The microscope
consists of an LED light source, a rotating multi-pinhole disc acting as an aperture for the
incoming and reflected beams, an objective lens with a piezo drive and a CCD camera
(charge-coupled device). During measurement, the objective lens position is changed to

80
Experimental

record a series of images with different focal plane. A 3-dimensional (3D) topography can
be obtained by combining the single images. For transparent layers, such as PEDOT:PSS,
a thin layer of silver was evaporated prior to confocal microscope measurement. This led
to better image quality and prevented inconsistent data due to multireflections at the
transparent layer interfaces.

3.3.2.2. Scanning electron microscopy

Scanning electron microscopy (SEM) images were recorded using a JSM-7610F from
JEOL with secondary electron image detector. For operation 2 kV accelerating voltage, a
working distance 6 mm in low probe current mode (~65 nA) were used. An r-filter
expands the capabilities by allowing selective electron detection from the specimen
according to energy range. Secondary electrons, backscattered electrons or a mixture of
both were detected to create images. SEM cross section samples were prepared by
braking at low temperature (liquid nitrogen cooling) or using an IB-19500CP cross
section polisher.

3.3.2.3. Atomic force microscopy

Atomic Force Microscope (AFM) measurements were performed using a Dimension


5000 from Veeco Instruments Inc. or Solver Nano from NT-MDT. Pictures with
dimensions of 5x5 μm² were recorded by scanning the surface with a piezoelectrically
controlled cantilever. The measurements were performed in the intermittent contact mode
or short ‘tapping mode’ with the tip oscillating close to the surface. Height- as well as
phase contrast images were obtained.

3.3.2.4. Sheet and contact resistance

The sheet resistance 𝑅𝑠𝑞 of conductive layers, such as highly conductive (HC)
PEDOT:PSS, AgNWs or silver, was characterised using a 4 point probe system R-Chek
from EDTM Inc.. From sheet resistance and layer thickness 𝑑 data, the bulk conductivity
𝜎 can be calculated according to equation 45.

81
Experimental

1
𝜎= equation 45
𝑑 ∗ 𝑅𝑠𝑞

In order to gain information about contact resistances between two materials (specifically
152
PEDOT:PSS and silver), transmission line measurements as described in were
performed. The measurement strategy is illustrated in Figure III-10.

Figure III-10: Transmission line method, top left: a) Schematic drawing of transmission
line structure. b) Photograph of structure used for measuring the contact resistance
between PEDOT:PSS and printed silver. The PEDOT:PSS layer is located between the
dotted lines. c) Schematic 𝑅 vs. 𝑙 plot used to extrapolate the contact resistance.

Silver fingers of different distances 𝑙 to each other were applied on a PEDOT:PSS coated
substrate. The resistance is measured between two neighbouring fingers and plotted over
the related spacing. By fitting a straight line to the data points, the contact resistance 𝑅𝐶 is
obtained as half the interpolated resistance value at 𝑙 = 0. The so-called transfer length
𝑙 𝑇 , which describes the length underneath a silver finger within which most current is
transferred from the PEDOT:PSS to the silver, can also be found in a similar way. It is
half of the length at 𝑅 = 0. As a basic design rule for the experiment, the silver fingers
have to be longer than the 𝑙 𝑇 in current flow direction in order to guarantee an
undisturbed current transfer. From 𝑅𝐶 , 𝑙 𝑇 and 𝑤 the specific contact resistance 𝜌𝐶 is
calculated using equation 46.

82
Experimental

𝜌𝐶 = 𝑅𝐶 ∗ 𝑙 𝑇 ∗ 𝑤 equation 46

Additionally, the sheet resistance 𝑅sq can be found with the transmission line method. It
is calculated from the slope 𝑚𝑠 of the fitted line and the contact width 𝑤.

𝑅𝑠𝑞 = 𝑚𝑠 ∗ 𝑤 equation 47

With additional profilometer or confocal microscope measurements, the PEDOT:PSS


thickness can be determined and used to calculate the conductivity 𝜎 according to
equation 45.
Shadowing masks were designed and cut from stainless steel foil by LASER to create the
transfer length structure. Evaporated 100 nm thick silver fingers have a width of 1 mm
and distances of 0.125 mm, 0.25 mm, 0.5 mm, 1 mm, 2 mm, 3 mm and 5 mm (see Figure
III-10). For the measurement a source measure unit 236 from Keithley Instruments Co.
was used. J-V curves were recorded between -0.6 and 0.6 V in steps of 0.5 V for every
finger to gain information about the quality of the contact. Resulting straight lines
confirmed good ohmic contact and were used to extract the resistance values from the
slope.

3.3.2.5. UV-Vis spectroscopy

Transmittance and absorbance data were acquired using a Lambda 950 UV-Vis
spectrophotometer from Perkin Elmer, which is equipped with a tungsten-halogen and a
deuterium light source (operating range 175 - 3300 nm). Spectra were recorded between
300 nm and 800 nm with a step size of 1 nm. To gain data about diffuse transmittance and
haze, an integrating sphere was used.

3.3.2.6. Stylus profilometry

Layer thickness measurements were performed with a stylus profilometer Tencor D100
from KLA-Tencor Co.. Velocity and stylus force operation ranges are 0.01 - 40 mm/s and
1 - 10 mg.

83
Experimental

3.3.3. Solar cell and module characterisation

3.3.3.1. Current density-voltage characteristics

The J-V curves were recorded using a solar simulator with xenon arc lamp and source
measurement system from LOT-QuantumDesign GmbH. The AAA class machine flux
density of 1000 W/m² with terrestrial AM (air mass) 1.5G spectrum was calibrated prior
to each measurement using a certified silicon reference cell. In this work, the mostly
applied sweeping ranges for solar cells are -0.5 V to 1.5 V and for solar modules -2 V to
3 V, in 20 mV steps increment.

3.3.3.2. External quantum efficiency

External quantum efficiency (EQE) was measured with a QE-R machine from Enli
Technology Co., Ltd.. Prior to each measurement the machine was calibrated with a
certified reference silicon cell. EQE spectra were recorded between 300 nm and 900 nm
using a step size of 10 nm. The 𝐽𝑆𝐶 values of the solar cells were extracted from the EQE
spectra and compared to the solar simulator measurements.

3.4. Software tools for simulation and optimisation

3.4.1. Matlab
The software Matlab from The MathWorks Inc. provides a numerical computing
environment and language allowing matrix manipulations, implementation of algorithms,
plotting and interfacing with other programs. The software was used for multiple
purposes in this work. This includes data analysis and 3D plotting.
Furthermore, Matlab was used to program a simple software tool for slicing arbitrary
shapes into equally sized areas. The tool was applied in the designing process of inkjet
printable solar modules (chapter 6.2). It provided the necessary layouts for each layer to
print under consideration of module voltage and single cell current matching by
subsequently executing the following steps:

 Import arbitrary shape black and white image

84
Experimental

 Ask the user for input of the required module voltage and open circuit voltage of
single solar cell (depends on material)
 Calculate the number of solar cells
 Count pixels of the black or white area to determine total available area
 Determine size of a single cell in pixels
 Specify and subtract an inactive area from each cell which is used for the
monolithic series connection to the subsequent cell
 Define cells and interconnection region by counting pixels from left to right until
size of single cell + interconnection region is reached
 Display, export and save results

3.4.2. COMSOL Multiphysics


Finite element method (FEM) simulations using the software COMSOL Multiphysics
from COMSOL Inc. were performed for the following purposes:
 Loss analysis in solar cells with silver grid electrodes
 Investigation of the interconnection resistance influence on solar module
performance
These tasks were fulfilled by simulations of 2-dimensional (2D) FEM models. Compared
to SPICE- (simulation program with integrated circuit emphasis) or 3D FEM simulations
as reported by Acciani et al.,153 Pieters154 and Koishiyev,155 the 2D FEM simulations offer
reduced calculation time and memory resources.

Silver grid

The FEM model for silver grid electrode loss analysis was constructed based on a work
by F.W. Fecher, which was originally designed for grid geometry optimisation.156 The
permission for adoption and modification of F.W. Fechers work is gratefully
acknowledged. A schematic drawing of the FEM model applied in this work is given in
Figure III-11. The model consists of the three individual layers bottom electrode, HC
PEDOT and top silver grid. Lateral current densities within each layer are calculated by
Ohm’s law. The perpendicular current density through the solar cell is introduced using
experimental light and dark J-V curves of a well working reference cell. In the model,
photocurrent is only generated in areas which are not shaded by the grid structures
(Figure III-11 b) red area). The influence of the PEDOT:PSS thickness on device

85
Experimental

performance is implemented using experimental data obtained by the author of this thesis.
This comprises conductivity and transmittance data for PEDOT:PSS layers of different
thicknesses. By accounting for the PEDOT:PSS absorbance, the solar cell EQE and the
AMG 1.5 sun spectrum, the current loss due to parasitic absorption is estimated and
implemented into the simulation model.

Figure III-11: a) Schematic drawing of 2D FEM model used for the silver grid electrode
loss analysis. b) FEM meshing of silver line grid. FEM model adapted with permission
from F.W. Fecher.156

Furthermore, contact resistances between the HC PEDOT:PSS layer and the silver finger
as well as the resistance at the cell interconnection were obtained and considered in this
simulation.

Table III-6: Simulation parameters applied in the 2D FEM model for silver grid loss
analysis.

Parameter Value

𝑙𝑐𝑒𝑙𝑙 5 mm
(cell length)
𝑤𝑐𝑒𝑙𝑙 1.0 mm; 1.5 mm
(unit cell width = distance between linear grid fingers)

86
Experimental

Parameter Value

𝑤𝑃1,𝑃2,𝑃3 50 µm
(width of interconnection lines)
𝐷𝑃1,𝑃2,𝑃3 100 µm
(distance between interconnection lines)
𝑤𝑔𝑟𝑖𝑑𝑓𝑖𝑛𝑔𝑒𝑟 100 µm
(width of grid fingers limited by inkjet drop size)
𝑑𝐻𝐶 𝑃𝐸𝐷𝑂𝑇:𝑃𝑆𝑆 100 nm
(thickness HC PEDOT:SS)
𝑅𝑠𝑞 𝐴𝑔 𝑒𝑣𝑎𝑝 0.20 Ω/sq
(Sheet resistance evaporated silver electrode 100 nm)
𝑅𝑠𝑞 𝐴𝑔 𝐼𝐽 0.22 Ω/sq, 2.2 Ω/sq
(Sheet resistance inkjet printed silver grid 500 nm)
𝑅𝑠𝑞 𝐼𝑇𝑂 15 Ω/sq
(Sheet resistance ITO electrode)
𝜎𝑃𝐸𝐷𝑂𝑇:𝑃𝑆𝑆 500 S/cm
(Conductivity PEDOT:PSS F HC Solar)
𝜌𝐶 𝑃𝐸𝐷𝑂𝑇:𝑃𝑆𝑆−𝐴𝑔 0.11 Ω*cm²
(Specific contact resistance PEDOT:PSS-silver)
𝜌𝐼𝐶 156,157 5*10-3 Ω*cm²
(Specific interconnection resistance)
𝑉𝑡𝑜𝑡 2V
(bias voltage)

Experimental cell characteristics; Basic J-V curve PV2000:PC70BM; P3HT:PC60BM

𝑉𝑂𝐶 0.81 V; 0.55 V


𝐹𝐹 58.4 %; 63.5 %
𝐽𝑆𝐶 14.3 mA/cm²; 8.0 mA/cm²

𝑃𝐶𝐸 6.72 %; 2.80 %

Interconnection resistance:

To investigate the influence of the interconnection resistance on the solar module


performance, a 2D FEM model was constructed in analogy to previously described work
about silver grids. The model consists of three solar cells monolithically connected in
series similar to a structure reported in reference 158. Two layers are used to describe base

87
Experimental

and top electrode of the module. The lateral current density within these layers follows
Ohm’s law. The perpendicular current density through the solar cell stack is modelled
using an equivalent circuit approach as described in chapter 2.1.2 and expressed in
equation 48.

𝐽(𝑥, 𝑦, 𝑉𝑖 ) = −𝐽𝑃𝐻 + 𝐽𝐷 + 𝐽𝑅𝑃


𝑉𝑖 +𝜌𝑆 ∗𝐽 𝑉𝑖 + 𝜌𝑆 ∗ 𝐽 equation 48
= −𝐽𝑃𝐻 + 𝐽0 ∗ (𝑒 𝑛𝐷 ∗𝑘𝐵 ∗𝑇𝑒 − 1) +
𝜌𝑃

The position and voltage 𝑉𝑖 dependent current density 𝐽(𝑥, 𝑦, 𝑉𝑖 ) consists of three
components. These are the photocurrent density 𝐽𝑃𝐻 , the diode current density 𝐽𝐷 and the
parallel resistance current density 𝐽𝑅𝑃 . The term 𝐽𝐷 is described using the diode saturation
current density 𝐽0 , diode ideality factor 𝑛𝐷 , the Boltzmann constant 𝑘𝐵 , the specific solar
cell series resistance 𝜌𝑆 and the temperature 𝑇𝑒, while 𝐽𝑅𝑃 is expressed by the quotient of
𝑉𝑖 , 𝜌𝑆 , 𝐽 and the specific parallel resistance 𝜌𝑃 .

Figure III-12: Schematic illustration of the electrical 2D FEM model used to simulate the
3 cell OPV module similar to references 156,158. Published in reference 159 and reproduced
with permission.

Good agreement between experimental curves and simulation is found when modelling
the perpendicular current through the solar cell with 𝜌𝑆 = 0. The series resistance is
introduced to the system via the lateral current, which is limited by the sheet resistance of
the electrodes 𝑅𝑠𝑞 PEDOT:AgNW, 𝑅𝑠𝑞 𝐼𝑇𝑂 and the specific contact resistance of the
interconnection region 𝜌𝐼𝐶 . The interconnection area is formed by connecting single cells

88
Experimental

from the top electrode of one cell to the bottom electrode of the adjacent cell following
the P1, P2, P3 concept as described in chapter 2.1.4. The contacts between ITO, Ag and
AgNWs are considered ohmic. Figure III-13 shows the geometry used for the simulation.
It exactly replicates the solar modules experimentally realised and described in chapter
6.1.2. A List of the used simulation parameters is provided in Table III-7.

Figure III-13: Geometry of the 3-cell OPV module with dot type interconnection used in
the FEM simulations. Inset: Enlarged simulation mesh at the interconnection region.
Published in reference 159 and reproduced with permission.

Table III-7: Parameters used in the FEM simulation of the 3-cell module.
Parameter Value

𝑙𝑐𝑒𝑙𝑙 8.5 mm
(cell length)
𝑤𝑐𝑒𝑙𝑙 10 mm
(cell width)
𝑤𝑃1,𝑃2,𝑃3 40 µm, 100 µm, 500 µm
(width of interconnection lines)
𝐷𝑃1,𝑃2,𝑃3 200 µm
(distance between interconnection lines)
𝑟𝑑𝑜𝑡 50 µm
(radius of silver ‘bridge’ dots)
𝐷𝑑𝑜𝑡 254 µm (100 DPI)
(distance between silver ‘bridge’ dots)
𝑅𝑠𝑞 PEDOT:AgNW 10 Ω/sq
(sheet resistance PEDOT:PSS:AgNW)

89
Experimental

Parameter Value

𝑅𝑠𝑞 𝐼𝑇𝑂 15 Ω/sq


(sheet resistance ITO)
𝜌𝐼𝐶 159 0.002 Ω*cm²
(specific interconnection resistance)

Solar cell characteristics (Shockley one-diode model ~P3HT:PC60BM)


𝐽𝑃𝐻 8.9 mA/cm²
(photocurrent density)
𝐽0 10-7 A/cm²
(diode saturation current density)
𝜌𝑃 1550 Ω*cm²
(specific parallel resistance)
𝑛𝐷 1.9
(diode ideality factor)
𝑇𝑒 300 K
(temperature)
𝑘𝐵 1.38*10−23 J/K
(Boltzmann constant)
𝑉𝑡𝑜𝑡 3.0 V
(bias voltage)

It should be mentioned that the FEM model described above is quite generic and can
easily be extrapolated to other types of thin film solar cells, such as perovskites, by
adjusting the parameters of the above given equation.

90
CHAPTER IV: From Inks to Layers

Abstract:

In this chapter, OPV inkjet inks which fulfil the critical requirements for an industrially
applicable inkjet printing process (see Figure I-3) are developed. The inks are free of (i)
environmentally highly critical halogenated solvents and can be ejected from the nozzles
as (ii) well-defined spherical droplets, with (iii) sufficient open time. This is achieved by
applying the Ohnesorge theory as guideline for solvent selection. Based on the theoretical
predictions, 1-pentanol based ZnO and AgNW inkjet inks are created. Their excellent
printability proves that nanowires with a length of 10s of µm can be ejected from inkjet
nozzles of similar dimension. Furthermore, the layer formation of printed wet films on the
substrate is investigated. Wetting problems of water based PEDOT:PSS inks on
hydrophobic polymeric active layers are solved by introduction of a fluorosurfactant,
resulting in homogenous layers (iv). However, his approach does not work for every
combination of ink and surface. Therefore, a novel strategy of pinning the wet film onto
the substrate with inkjet printed anchoring points is developed. The new concept allows
definition of wetting and non-wetting areas on different surfaces and can thus be used for
patterning of wet films.

Parts of this chapter have been published in:


P. Maisch, K.C. Tam, L. Lucera, H.-J. Egelhaaf, H. Scheiber, E. Maier, et al., Inkjet printed silver
nanowire percolation networks as electrodes for highly efficient semitransparent organic solar cells, Org.
Electron. (2016). doi:10.1016/j.orgel.2016.08.006, reproduced with permission from Elsevier.
P. Maisch, L.M. Eisenhofer, K.C. Tam, A. Distler, M.M. Voigt, C.J. Brabec, H.-J. Egelhaaf, A generic
surfactant-free approach to overcome wetting limitations and its application to improve inkjet-printed
P3HT:non-fullerene acceptor PV, J. Mater. Chem. A. (submitted)
Only parts of the publication text authored by P. Maisch are reproduced in subchapters 4.1.2 , 4.3 and 4.4 of
this work. Contributors and specific contributions in form of images and measurement data to this chapter
are clearly stated in the figure captions of subchapters 4.1.2 and 4.3. Additionally, permission to utilise the
whole content of both publications as part of this thesis was given from all co-authors involved through
author ownership declarations.

91
From Inks to Layers

4.1. Ink formulation

4.1.1. Ohnesorge theory


The first step on the way to fully inkjet printed organic solar cells and modules is the
creation of well printable OPV inks from environmentally non-critical solvents. As
described in chapter 2.3, there are several publications which demonstrate inkjet printing
of different OPV layers. Typically, these works provide only brief information about the
materials they used for printing. Details about ink viscosity, surface tension or drop
formation are rarely found, even though these parameters are highly critical when a stable
printing process that allows precise structuring (no satellites) and up-scaling is targeted.
Interestingly, according to 𝑂ℎ-theory (see chapter 2.2.4.1), none of the frequently
reported pristine solvents used for OPV active material deposition is located in the range
of stable drop formation (𝑂ℎ𝑜-𝐷𝐶𝐵 = 0.03, 𝑂ℎ𝑜-𝑋 = 0.03, 𝑂ℎTHN = 0.06). This can mainly
be attributed to their low viscosities of typically ≤ 2 mPas, which results in 𝑂ℎ numbers
< 0.1. Nevertheless, formation of well-defined spherical ink droplets is possible by
careful control of the ink loading. It is intuitively clear that a sufficient amount of
dissolved active layer polymer can increase the ink viscosity and consequently also the
𝑂ℎ number. Among the very few works about inkjet printed OPV bothering with
printability aspects, a publication by Lamont et al. should be highlighted.160 They
identified a too low viscosity of their active layer inkjet ink as a limiting factor for the
processing. By adding polystyrene, they increased the ink viscosity by 70 %, thus
improving printability. This strategy resulted in a reduction of device efficiency by 20 %,
which showcases the difficulty of simultaneously achieving good droplet formation and
electrical functionality of the printed layer.
In this work, the focus is set on both, a reliable inkjet printing process and electrical
functionality of every layer. Electronic inks are designed accordingly. Therefore, a
printability rating system according to the Ohnesorge theory is applied. The evaluation
(Figure IV-1 and Table IV-1) contains commercial inkjet inks, such as PEDOT:PSS P Jet
N V2 and the silver ink Sicrys I50DM106. It is not surprising that both are in an 𝑂ℎ range
which promises good printability.

92
From Inks to Layers

Figure IV-1: Regime for stable operation of DOD inkjet printing according to Derby111
and qualification of the most important inkjet inks used in this thesis.

Table IV-1: Parameters to qualify printability of inkjet inks. OPV inks which do not offer
satellite-free droplet formation (bold) are modified, and thus shifted into the stable
printing regime (nozzle diameter 35 µm).
Ink 𝛾𝐿 𝜌 𝜇 𝑅𝑒 𝑊𝑒 𝑂ℎ
(mN/m) (g/cm³) (mPas) () () ()

Silver ink Sicrys 31.0 1.87 20.0 22.4 115.8 0.48


I50DM106

PEDOT:PSS P Jet N V2 19.8 0.98 9.0 26.1 95.0 0.37

PEDOT:PSS F HC Solar 27.1 0.98 15.0 15.7 69.4 0.53

AgNW:PEDOT:PSS H YE 28.0 1.0 19.0 12.6 68.6 0.66

P3HT:PC60BM 28.5 1.1 1.2 233.7 75.5 0.04


(CB + 5 %vol BrAni)
P3HT:PC60BM 28.0 0.95 3.9 58.5 65.1 0.14
(o-X:THN 1:1)

93
From Inks to Layers

Ink 𝛾𝐿 𝜌 𝜇 𝑅𝑒 𝑊𝑒 𝑂ℎ
(mN/m) (g/cm³) (mPas) () () ()

ZnO N-10 (IPA) 21.7 0.81 2.4 81.0 71.7 0.10

ZnO N-10 (1-pentanol) 24.0 0.82 4.0 49.2 65.6 0.16

AgNW ClearOhm (IPA) 25.0 0.90 2.4 90.0 69.1 0.09

AgNW ClearOhm 25.0 0.90 4.2 51.4 69.1 0.16


(1-pentanol)

Interestingly, formulations which are not intended for inkjet printing purposes, such as
PEDOT:PSS F HC Solar and H YE, exhibit 𝑂ℎ values between 0.1 and 1 as well. Printing
experiments at the PiXDRO LP50 confirm good drop formation without any further ink
modification.
Other OPV inks have been designed originally for spin coating, doctor blading or slot die
coating purposes and cannot be printed well by inkjet. Examples are the IPA based ZnO
N-10 ink from Avantama, ClearOhm AgNW ink from Cambrios and the active layer
P3HT:PC60BM ink with CB + 5 %vol BrAni solvent. As shown in Figure IV-1, these inks
are located at the border or outside of the stable printing area. According to the 𝑂ℎ theory
satellite formation is expected. Printing experiments confirm this prediction (Figure
IV-2).

Figure IV-2: Drop watcher images of a) IPA based pristine ZnO nanoparticle N-10 ink
showing satellite formation, b) 1-pentanol based N-10 ink and c) 1-pentanol based AgNW
ink forming a stable droplet (printing parameters for all inks: 80 V, 6-6-6 µs, 500 Hz, 16
mbar, pictures were taken 50 µs, 100 µs, 150 µs after the voltage pulse).

94
From Inks to Layers

In order to improve the printing behaviour, the IPA based ZnO and AgNW inks were
modified with solvents that are, according to 𝑂ℎ theory, located inside the stable printing
regime. The liquids 1-pentanol (𝑂ℎ = 0.15), propylene glycol (PG, 𝑂ℎ = 1.0) and
ethylene glycol (EG, 𝑂ℎ = 0.39) were chosen for testing. Addition of > 10 %vol of PG
and EG to the IPA based ZnO nanoparticle ink resulted in formation of well-defined
spherical droplets. Addition of the high boiling point alcohol 1-pentanol to the IPA based
ZnO ink also improves the drop formation. Testing of different mixing ratios revealed
that well-defined spherical droplets can easily be obtained for inks with 1-pentanol
contents of > 50 %vol. Best results were achieved with an almost pure 1-pentanol
formulation, which was created by removing IPA from the ink by rotary evaporation.
This ink formulation is used as standard in the following chapters.
Following the same strategy, 1-pentanol based inkjet inks containing AgNWs with an
average length of ~30 µm were created. After filtration of the AgNW ink with a 30 µm
polycarbonate syringe filter to remove agglomerates, stable drop formation, as predicted
by the 𝑂ℎ theory, was achieved (see Figure IV-2 c)). This is the first time proof that
AgNWs with a length similar to the nozzle diameter can be printed. Application of the
AgNW ink to fabricate semitransparent electrodes is presented in chapter 5.3.2.
For inkjet printing of the active materials PV2000:PC70BM and P3HT:PC60BM, CB
based inks, which are applied in the blade coating reference process (see chapter 3.2.1),
are not preferable. As predicted by the 𝑂ℎ theory, satellite formation is observed for
printing of such inks. Furthermore, the environmental hazards associated with CB prevent
any large scale industrial use. Based on ink stability considerations, which are elucidated
in the following chapter, o-X:THN solvent blends were identified as best suited for inkjet
printing of PV2000:PC70BM and P3HT:PC60BM active layers. Despite the low viscosities
of < 2 mPas and 𝑂ℎ numbers of < 0.1 of the selected pristine solvents, inkjet inks with
𝑂ℎ > 0.1 were created. This is possible because the dissolution of the active layer
polymers leads to an increase in viscosity, and thus also 𝑂ℎ number. Inks consisting of
o-X:THN mixed in a ratio of 1:1 with polymer concentration > 15 mg/ml (P3HT (𝑀𝑤 :
94100 g/mol) or PV2000 (𝑀𝑤 : N/A)) reach viscosities of > 3 mPas, resulting in 𝑂ℎ
numbers > 0.14 and offer stable drop formation. The functionality of the inks is
demonstrated by manufacturing of highly efficient solar cells in chapter 5.1.

95
From Inks to Layers

4.1.2. Ink stability and nozzle open time


Besides the ability to form well-defined spherical droplets, inkjet inks have to fulfil
further criteria in order to qualify for defect-free printing of functional layers on large
areas. One property that has been identified as highly critical for a robust printing process
is a sufficient resistance of the OPV inks towards chemical or physical changes over time,
such as sedimentation or gel formation. It is intuitively clear that such stability problems
or fast evaporation of volatile ink components can result in a disturbed drop formation or
even clogging of the printing nozzles. In this context, an important rating factor for
assessing the robustness of an inkjet process is the time nozzles can stay uncovered and
idle before they can no longer print (open time). Even though ink stability and nozzle
open time are of very high importance for a robust and up-scalable OPV inkjet printing
process, very little information is provided in the state-of-the-art literature. The most
129
significant report to this topic is by C. Hoth et al. . They found that inkjet printing of
P3HT:PC60BM active layers from o-DCB/mesitylene or pristine THN inks at room
temperature (RT) is limited by the shelf life of the ink. At ambient conditions they
reported rapid gelling for inks containing 98 % RR (regioregular) P3HT. Furthermore,
they found higher aggregation tendency and decreased device performance for printing
the active layers from pristine THN inks compared to o-DCB/mesitylene, highlighting the
challenge of finding suitable halogen-free solvents for inkjet printing of OPV.
In the following, ink stability problems that were observed in this work, suitable
countermeasures as well as strategies to increase the nozzle open times are presented.
Similar to Hoth et al., limited ink stability was found when developing PV2000:PC70BM
and P3HT:PC60BM active layer inkjet inks. o-X was chosen as starting point for
formulation experiments because both active layer blends are known to give excellent
device efficiency when deposited from this solvent. However, the relatively high vapour
pressure of pristine o-X (6.6 mmHg @25 °C, 1 bar) and the high tendency of
P3HT:PC60BM inks to form gels (see Figure IV-3) at temperatures ≤ 30 °C (higher
processing temperatures are not preferred, because they lead to decreased open times due
to solvent evaporation from the printing nozzles) makes pristine o-X based active layer
inks not suitable for deposition via inkjet. To improve the printability, high boiling point
(BP) co-solvents are added. Indane (BP = 176.5 °C, 𝑂ℎ = 0.01), THN (BP = 207 °C,
𝑂ℎ = 0.06) and TMB (BP = 170 °C, 𝑂ℎ = 0.04) were selected for experimental testing
because they have repeatedly been applied for processing of highly efficient OPV

96
From Inks to Layers

devices.129,143,161,162 In the following, the focus is set on P3HT:PC60BM because


aggregation problems are much more prominent compared to PV2000:PC70BM.

Figure IV-3: P3HT:PC60BM (20 mg/ml) dissolved in a) o-X, b) Indane, c) THN and d)
TMB at 60 °C and stored at 30 °C for 5 h without stirring. The o-X and TMB inks formed
gels, while the Indane and THN inks are still liquid.

Table IV-2: Rating of different P3HT:PC60BM active ink gelling tendencies (20 mg/ml,
30 °C, no stirring). The inks are specified as liquid (L) with a grading system ranging
from 1-4, to describe opacity and precipitates (1: homogeneous ink, 4: significant amount
of precipitates, gelling). If the ink gels completely, it is specified as solid (S). Results were
partly obtained by H. Scheiber and E. Maier from Durst Phototechnik and are reported
with permission.

t/h o-X o-X:THN o-X:TMB o-X:Indane


1:1 (vol.) 1:1 (vol.) 1:1 (vol.)

0 L1 L1 L1 L1

1 L3 L1 L2 L1

2.5 S L1 L3 L1

3.5 S L1 S L1

4.5 S L1 S L2

20 S S S S

97
From Inks to Layers

The impact of the co-solvents on the gelling tendency of the o-X based P3HT:PC60BM
inks was first assessed visually. For this reason, inks with different mixing ratios of the
co-solvents were prepared and stirred at 60 °C for more than 2 h to ensure complete
dissolution of the P3HT. Subsequently, the inks were stored at 30 °C without any stirring
and inspected visually. Results are listed in Figure IV-3 and Table IV-2. Inks with THN
co-solvent yield the lowest gelling tendency and were chosen for further experimental
testing. The qualitative results obtained by visual inspection of the inks are corroborated
by quantitative measurements of the ink viscosities over extended periods of time (Figure
IV-4). Mixtures of o-X:THN 3:1 (vol.) still show gelling within a few hours which leads
to a drastic increase in dynamic viscosity of more than 2000 %.

Figure IV-4: Viscosity of different P3HT:PC60BM (20 mg/ml, 30 °C) ink formulations
over time at a constant shear rate of 50 s-1, Inset: Photograph of used double gap
measurement geometry with the arrow highlighting gelled active layer ink of the
o-X:THN 3:1 (vol.) mixture. Results were partly obtained by H. Scheiber and E. Maier
from Durst Phototechnik and are reported with permission.

The 1:1 (vol.) ink shows significantly improved long term stability during the viscosity
measurement. The slower gelling and viscosity increase allows the application of this ink
in an inkjet printing process. It seems possible to slow down the gelation process even

98
From Inks to Layers

further by increasing the THN content. Nevertheless, the 1:1 (vol.) mixture was chosen
for printing of solar cells because C. Hoth demonstrated that P3HT:PC60BM solar cells
inkjet printed in a mono-solvent approach from pristine THN suffer from poor
performance due to microstructure limitations.129 Further tests reveal that the o-X:THN
1:1 solvent blend can be used to create PV2000:PC70BM inkjet inks. Gelling of this high
efficiency material is much less pronounced compared to P3HT:PC60BM.
PV2000:PC70BM inks with 15 mg/ml polymer content can be stored for several days at
RT before significant gelling is observed. As reference for the viscosity measurement, a
P3HT:PC60BM ink using o-DCB as solvent was investigated. Even though this ink shows
no significant viscosity increase during the whole 24 h viscosity measurement,
environmental risks make this ink unsuitable for any industrial process.
Besides the gelling tendency, also the evaporation rates of the solvents were considered
during ink development because they are a determining factor for the nozzle open time
and process stability. The relative evaporation rates (RER; n-butyl acetate = 100) of the
single liquid compounds were obtained from the software ‘Hansen solubility parameters
in practice’ (HSPiP).163 Simulated drying characteristics of pristine o-DCB (RER 7.2),
pristine o-X (RER 80) and the o-X blend with THN (RER 2.2) are displayed in Figure
IV-5.

Figure IV-5:Relative drying times of o-X, o-X:THN 1:1 (vol.) mixture and o-DCB
simulated with the HSPiP software.

99
From Inks to Layers

According to the simulation, the o-X:THN 1:1 mixture offers a longer drying time than
pristine o-X and even pristine o-DCB. Experimentally determined open times for the
o-X:THN and o-DCB inks are > 30 min (Table IV-3). This prevents fast clogging of
nozzles during idle time and allows easy handling of the inkjet.

Table IV-3: Open time experiment for different active layer inks (15 mg/ml polymer
content) using a Spectra SE 128 AA printhead with 35 µm nozzle diameter.

Idle time Number of printing nozzles


(min)
P3HT:PC60BM P3HT:PC60BM P3HT:PC60BM PV2000:PC70BM
o-X o-DCB o-X:THN 1:1 o-X:THN 1:1

0 128 128 128 128


0.25 121 128 128 128
0.5 56 128 128 128
1 5 128 128 128
2 0 128 128 128
5 0 128 128 128
10 0 128 127 128
30 0 128 127 128
60 0 127 12 98
900 0 23 0 0

Simulated drying characteristics and open time experiments were also performed for the
alcohol based ZnO nanoparticle and AgNW inkjet inks developed according to the
𝑂ℎ theory (see previous chapter). Experiments confirm a very short nozzle open time of
only a few seconds (see Table IV-4) for pristine IPA (RER 150) inks. It is possible to
recover clogged nozzles by purging cycles. However, the high effort for recovery and
large amount of waste material makes the inkjet process very inefficient. Effects of
adding the high boiling point alcohol 1-pentanol to IPA based inks, as done in the case of
ZnO N-10 and AgNW ClearOhm, are presented in Figure IV-6. The modification does
not only prevent satellite formation and shifts the inks into the regime of stable DOD
operation (see Ohnesorge theory, Figure IV-1), but also increases the drying time
significantly.

100
From Inks to Layers

Figure IV-6: Relative drying times of IPA:1-pentanol mixtures simulated with the HSPIP
software. Compared to pristine IPA a sevenfold increase of drying time is achieved with a
1:2 IPA:1-pentanol mixture which was applied for inkjet printing of semitransparent
electrodes. Results are partly published in 164 and reproduced with permission.

Table IV-4: Open time experiment for different alcohol based ZnO N-10 inks (2.5 %wt
nanoparticles) using a Spectra SE 128 AA printhead with 35 µm nozzle diameter.

Idle time Printing nozzles


(min) IPA ZnO ink 1-pentanol ZnO ink

0 128 128
0.25 123 128
0.5 3 128
1 0 128
2 0 128
4 0 11
8 0 0

With modified inks containing 1-pentanol (RER 15), open times in the range of several
minutes are reached for both, ZnO and AgNW inks. This fits well with the simulated
drying behaviour and allows easy handling and a high processing yield.

101
From Inks to Layers

4.2. Droplet control


The next step on the way from inks to layers is the identification of suitable printing
parameters by extensive experimental testing and systematic variation of machine
parameters. Hereby, the Ohnesorge theory provides a rough estimation of the expected
width of the processing window. Printing experiments confirm that satellites can hardly
be avoided for inks with 𝑂ℎ = 0.1, such as IPA based ZnO. Compared to this, printing
parameters which allow the formation of a stable droplet for the modified 1-pentanol
N-10 ink (𝑂ℎ = 0.16) can easily be identified (see Figure IV-2). A list of printing
parameters that were found for the different inks is provided in Table IV-5. Explanation
of these parameters is provided in the materials and methods section (Table III-3). Within
the range of stable printing, the speed and size of the droplet can be adjusted. Information
on how the trapezoidal voltage pulse applied to the Spectra SE 128 AA printhead
piezoelements influences these parameters is provided in Figure IV-7, using the
1-pentanol based ZnO N-10 ink as illustrative example.

Figure IV-7: Drop velocity and volume in dependence of the a) piezoelectric element
voltage, b) hold time, c) ramp-up time and d) ramp-down time for a 1-pentanol based
ZnO N-10 nanoparticle ink. Variation of each parameter was done keeping the other
parameters at their initial condition (Spectra SE 128 AA printhead 80 V, 6-6-6 µs,
500 Hz, 25 °C).

102
From Inks to Layers

It should be highlighted that this as well as other electronic inks used in this work have
significantly lower viscosity (see Table IV-1) than the ideal range of 8 - 20 mPas for
Spectra SE 128 AA operation. Nevertheless, printing conditions that result in formation
of well-defined spherical droplets were identified for every ink in the 𝑂ℎ number range
between 0.1 and 1. Besides the pulse shape and voltage amplitude, the firing frequency
also influences the droplet speed and volume. This is illustrated in Figure IV-8 for the
Spectra SE 128 AA printhead and a model fluid XL-30. The measurement data was
provided by Durst Phototechnik using a drop watcher from Jet Expert (Own in-house
PiXDRO LP50 drop watcher system only allows measurements up to 10 kHz). For
frequencies below 12 kHz only little influence on the droplet is observed because residual
pressure oscillations have sufficient time to damp out (see theoretical background chapter
2.2.5). To keep the influence of the frequency on drop formation low, layer processing
was typically done at values of 8 kHz or less. This corresponds to printing speeds of
~200 mm/s at a resolution of 1000 DPI.

Figure IV-8: Drop velocity and volume in dependence on the firing frequency for the
Spectra SE 128 AA printhead using a model fluid XL-30 (1.08 g/cm³, 31.2 +/-0.5 mN/m,
~4.6 mPas @ 70 °C, 55 V, 2-4-2 µs). Results were partly obtained by H. Scheiber and E.
Maier from Durst Phototechnik and are reported with permission.

103
From Inks to Layers

Table IV-5: List of printing parameters that allow stable Spectra SE 128 AA printhead
operation.

Ink Piezo Pulse Meniscus Firing Ink


voltage shape pressure frequency temperature
(V) (µs-µs-µs) (mbar) (kHz) (°C)

Silver Sicrys 80 10-8-4 -33 6 RT


I50DM106

ZnO N-10 (1-pentanol) 80 6-6-6 -21 5 RT

PEDOT:PSS Clevios P 70 3-6-3 -18 5 RT


Jet N V2

PEDOT:PSS Clevios F 80 3-6-3 -20 5 RT


HC Solar

P3HT:PC60BM in 100 1-8-1 -18 3 RT


o-X:THN 1:1

PV2000:PC70BM in 100 1-8-1 -18 3 RT


o-X:THN 1:1

AgNW ClearOhm 130 6-8-6 -16 4 RT


(1-pentanol)

AgNW:PEDOT:PSS 130 4-6-2 -19 4 RT


Clevios H YE

4.3. Layer formation


The final step on the way from inks to functional layers is the formation of thin
homogeneous films of well-defined thickness. Their quality is essential for the
functionality of printed electronic devices such as solar modules, LEDs and thin film
transistors.17,165,166 As these devices consist of several different layers which are coated on
top of each other, the concept of ‘orthogonal solvents’ is an often applied strategy to
ensure that coating a layer does not damage the layer deposited before.167 This implies
that the surface energy of the substrate and the surface tension of the ink are often grossly
different. In systems with high surface tension and low surface energy values, however,
energy minimisation often causes the applied wet film to shrink or rupture, a phenomenon

104
From Inks to Layers

which is referred to as ‘dewetting’. Several strategies have therefore been developed to


help this issue. One possible strategy is increasing the surface energy by, e.g., plasma or
corona treatment,168 but these techniques are not applicable to many organic surfaces as
they cause serious damage. Another strategy is reducing the surface tension by the
addition of wetting agents. This strategy is pursued in the first subchapter of this
paragraph.
A further requirement for printed electronic devices is the fabrication of distinct layer
patterns. Well working and highly reproducible structuring processes are necessary to
achieve for example source and drain contacts in organic field effect transistors
(OFETs),166 organic light emitting diodes (OLED) display patterning169 or monolithic
cell-to-cell interconnection in OPV modules.85 It is possible to achieve such patterning
subsequent to layer deposition by lithography process or LASER scribing.87 However, the
separation of coating and structuring steps increases production complexity and is often a
major cost driver. Coating and patterning in the 10s of µm scale should ideally be
combined in a single process. Approaches offering such direct patterning of the surface,
i.e. by locally defined deposition of the coating, have been described in various
publications. Multiple works make use of the coffee-ring effect to generate line
structures.170,171 Zhang et al. followed this strategy to deposit silver nanoparticles in
various patterns on substrates with different wettabilities.170 By means of inkjet printing
they were able to fabricate micro circuits formed by two parallel conductive lines of silver
nanoparticles each of the lines having a width of 5 - 10 μm. Due to the nature of the
coffee-ring effect, such fabrication methods are limited to line structures. Manufacturing
of larger continuous films is not possible. Another strategy to create patterned layers is to
actively control wetting and dewetting by introducing heterogeneities onto the surface.
The starting point for many works relying on this principle dates back to the year 1944
when Cassie and Baxter introduced their famous model to predict equilibrium contact
angles of liquids on composite surfaces based on the area fractions and individual contact
angles.172 Significant progress to this field was made by Raj et al., who investigated
circular hydrophilic wetting defects and introduced a model to predict receding as well as
advancing contact angles of liquids on heterogeneous and superhydrophobic surfaces.173
They demonstrated later how this technology can be used to influence the contact area
and 3D shape of single droplets on such surfaces.174 The application of wetting/dewetting
phenomena to structuring purposes in device manufacturing was demonstrated by Wang
et al..175 They investigated the direct patterning of single dots of PEDOT:PSS by

105
From Inks to Layers

controlled rupturing of droplets of PEDOT:PSS ink on top of an ultrathin hydrophobic


line of evaporated SiO2 modified with a self assembled monolayer of 1H, 1H, 2H, 2H-
perfluorodecyltrichlorosilane. Using the resulting two PEDOT:PSS semi-circles as source
and drain contact, they were able to manufacture OFETs with a channel width of only
500 nm. Another work demonstrating the strength of wetting/dewetting phenomena for
device manufacturing was reported by Chen et al..176 They achieved controlled rupturing
of silver ink wet films by sandwiching the liquid between a surface covered with wetting
defects and a non-patterned substrate to form a space confined assembly. With
evaporation starting from the sample edge they were able to form 1-dimensional (1D)
nanoscale bridges on superhydrophobic surfaces. A technique which combines wetting of
larger low energy surfaces and the possibility of direct patterning is the multiple step
spray coating process introduced by Colsmann et al..177 By spraying at low flow rates
they were able to deposit small seeds of surfactant-free PEDOT:PSS on hydrophobic
P3HT:PC60BM surfaces. In a second step, spray coating PEDOT:PSS at increased flow
rates produced continuous layers which were pinned to the previously deposited seeds. In
order to provide direct patterning, shadowing masks are used. Limitations of this strategy
are the time consuming seeding process (> 1 min) and the material wastage due to
spraying on the mask. All the methods described above provide direct patterning
techniques by overcoming or even benefiting from low surface energy and dewetting
effects. However, none of these techniques provides R2R compatible and material loss-
free processing of structured layers applicable for printed electronics mass production.
In the following chapters, ink-substrate interactions and resulting wetting phenomena are
described. Besides the conventional approach of using surfactants to overcome wetting
difficulties, a new additive-free strategy relying on inkjet printed anchoring points is
introduced. The anchoring points can pin a wet film and enable the template-free
deposition of arbitrarily shaped functional films with resolutions in the 10 µm range.

4.3.1. Wetting
In OPV production, especially the application of water based PEDOT:PSS dispersions on
top of the often hydrophobic active layer can result in dewetting. One way of overcoming
this is a modification of the ink with surfactants. Well known wetting agents for such
applications are found under the tradenames Dynol, Surfynol, Zonyl and Capstone.178–180
The latter was used in this work to improve the wetting of different commercially

106
From Inks to Layers

available PEDOT:PSS inks by lowering their surface tensions. Three different


PEDOT:PSS inks from Heraeus Clevios GmbH, (i) P VP 4083, (ii) F HC Solar and (iii)
P JET N V2, were selected for investigation because they cover a wide range of different
physical properties, applications and processing techniques. P VP Al 4083 is low
conductive (𝜎 = 0.002 - 0.0002 S/cm) and often applied in organic or perovskite PV and
LED technology.38,181 F HC solar is a highly conductive material (𝜎 ≈ 500 S/cm), which
can be used as transparent electrode.182 P Jet N V2 is a special inkjet formulation. It has
high conductivity (σ ≈ 250 S/cm) and is tailored in such a way that the dispersion shows
low viscosity and slow drying to avoid clogging of the inkjet nozzle.183 Furthermore, P Jet
N V2 ink has a neutral pH value to protect corrosion sensitive parts of the printheads that
might be attacked by acidic media. One thing that all these different aqueous inks have in
common is the challenging wetting behaviour on hydrophobic surfaces, such as
P3HT:PC60BM and PV2000:PC70B active layers. Surface energies/tensions of both, the
surfaces to be wetted as well as the liquids to be applied, were measured. It should be
mentioned that contact angles were obtained directly after droplet deposition. Changes in
ink composition during drying, which could also influence the wetting behaviour, are not
considered. Based on the gained data, wetting envelopes were created. Detailed
description of the theory is found in chapter 2.2.4.2. Further surface energy and surface
tension data as well as wetting envelopes for ink-layer combinations with less challenging
wetting situation, such as ZnO ink on an ITO surface, active layer ink on a ZnO surface
and silver ink on a PEDOT:PSS surface, are provided in the appendix of this work. Figure
IV-9 shows the wetting envelopes of the different active layer materials as well as the
surface tension of the chosen PEDOT:PSS inks. The corresponding values are listed in
Table IV-6. The wetting envelopes of the polymeric active layer materials lie closely
together, indicating similar wetting behaviour. One can assume that the obtained results
can be transferred to many other polymeric OPV active layers.
The polar/dispersive surface tension coordinates of all three as-purchased PEDOT:PSS
inks are located outside the P3HT:PC60BM wetting envelope. PEDOT:PSS P VP Al 4083
clearly exhibits the highest value, and thus the most challenging wetting properties.

107
From Inks to Layers

Figure IV-9: a) Wetting envelopes (𝜃 = 0 °) of commonly used OPV active layer


materials P3HT:PC60BM and PV2000:PC70BM and surface tension data of commercially
available PEDOT:PSS inks P VP Al 4083, F HC Solar, P JET N V2. An amount of
0.1 %wt wetting agent CFS was added to the PEDOT:PSS inks to lower the surface
tension, and thus to improve wetting behaviour. Inks within the wetting envelope are
expected to show good wetting on the corresponding surface. Inks located outside are
expected to show only partial wetting or dewetting. b) Photograph (size: 1x1cm) of a dry
inkjet printed PEDOT:PSS layer from F HC Solar ink, c) F HC Solar + 0.1 %wt CFS, d)
P JET N V2 and e) P JET N V2 + 0.1 %wt CFS on a P3HT:PC60BM surface. Published in
reference 184 and reproduced with permission.

The dispersive and polar surface tension components of P VP Al 4083 are similar to
water, which is the main component of the ink. Other PEDOT:PSS inks, like P Jet NV2
and F HC Solar, have a lower surface tension. This indicates that the manufacturer
already modified the aqueous inks by the addition of solvents or additives. Nevertheless,

108
From Inks to Layers

wetting properties of all as purchased PEDOT:PSS inks are insufficient for inkjet printing
a homogeneous layer on the P3HT:PC60BM substrate. All tested pristine PEDOT:PSS
inks dewet on the active layer surface (see photographs Figure IV-9). To improve this
situation, surfactants are added to these inks. As depicted in Figure IV-9, for F HC Solar,
a small amount of 0.1 %wt fluorosurfactant addition (CFS) is sufficient to print
homogeneous layers on P3HT:PC60BM.

Table IV-6: Surface energies (𝛾𝑆 ) of commonly used OPV active layer materials and
surface tensions (𝛾𝐿 ) of PEDOT:PSS inks with and without the fluorosurfactant Capstone
FS 31 (CFS). The polar (𝛾 𝑃 ) and dispersive parts (𝛾 𝐷 ) were determined following the
Owens, Wendt, Rabel and Kaelble method (OWRK). Each pendant drop and sessile drop
measurement was repeated at least three times. The presented data are mean values. The
standard deviation is in all cases < 0.2 mN/m. Published in reference 184 and reproduced
with permission.

γS γD
S γPS
Surface (mN/m) (mN/m) (mN/m) Structure

P3HT:PC60BM 24.5 24.3 0.2


PV2000:PC70BM 26.3 26.2 0.1

γL γD
L γPL
Ink (mN/m) (mN/m) (mN/m) P3HT PC60BM

Reference H2O 74.1 19.5 54.6


P VP Al 4083 75.4 21.0 54.4

P VP Al 4083 + 0.1 %wt CFS 13.9 6.6 7.3


PEDOT PSS
P Jet N V2 26.8 22.3 4.5
P Jet N V2 + 0.1 %wt CFS 19.8 18.0 1.8
F HC Solar 26.6 19.5 7.1
F HC Solar + 0.1 %wt CFS 20.9 17.6 3.3 CFS

A relatively small change in surface tension majorly impacts film formation. This
illustrates the delicate balance between surface tension and layer quality for this system.
The addition of 0.1 % CFS to P Jet N V2 as well shifts the surface tension within the

109
From Inks to Layers

wetting envelope of P3HT:PC60BM. However, experiments show that P Jet N V2 layers


still shrink or rupture during drying. Similar results were found when testing other
wetting agents such as Dynol and Surfynol in combination with P Jet N V2 or the wetting
of P Jet N V2 on PV2000:PC70BM surfaces. The misprediction of the wetting envelope
can have multiple reasons, such as the evaporation of volatile ink components during the
layer drying process and an associated change of the surface tension. At the current stage,
the reason for the overestimation of the wetting agents efficacy remains unclear,
highlighting that the consequences of modification of the ink in order to achieve wetting
are more complex than expected from textbook predictions.
A further point of concern is the long term interaction of surfactants with the film
components. Wetting promoters typically remain within the dry layer or at its interfaces
and may change the physical properties. One such example is provided in Figure IV-10.
The addition of 0.5 %wt fluorosurfactant CFS leads to a conductivity decrease of
PEDOT:PSS F HC Solar layers by more than 10 %. Evidence that high amounts of
fluorosurfactants can adversely affect OPV efficiency was provided by Savva et al. in the
178
supplementary information to reference . By adding higher amounts of the
fluorosurfactant Zonyl FS-300 to PEDOT:PSS, they reported significantly reduced 𝐹𝐹
and 𝑃𝐶𝐸 values due to 𝑅𝑆 increase and introduction of an s-shape into the J-V-curve.

Figure IV-10: Surface tension and conductivity of PEDOT:PSS F HC Solar with different
184
amounts of fluorosurfactant CFS. Published in reference and reproduced with
permission.

110
From Inks to Layers

The results described above highlight the importance of surfactants to overcome wetting
limitations of hydrophilic PEDOT:PSS inks on top of hydrophobic surfaces. However,
the example of P Jet N V2 shows that addition of small amounts of fluorosurfactant
cannot prevent dewetting phenomena for every ink type. At the same time, larger
amounts can negatively influence the device performance. Moreover, many surfactants
are known to be toxic and harmful to the environment.185 With respect to the issues
associated with enforcing wetting by additives, an alternative ‘pinning centre’ approach
has been developed.

4.3.2. Pinning centres


The following chapter introduces a generic additive-free approach to achieve wetting
based on inkjet printed anchoring points, which enables the template-free deposition of
functional films of arbitrary shape on discretionary surfaces with resolutions down to the
184
10 µm range. The presented results are published in reference . Furthermore, the
function principle behind this newly introduced strategy to overcome wetting challenges
is submitted as provisional European Patent under the file number 17 168 806.2.

Pinning centre strategy:

The method is based on inkjet printing of anchoring points on the surface to be coated.
After drying and, if necessary, annealing, these pinning centres prevent a subsequently
deposited wet film from dewetting. In the following, the concept and implementation of
this approach are described.

1. In a first step, an array of single droplets is inkjet printed onto the hydrophobic
substrate at a resolution (typically 100 - 300 DPI) low enough to avoid merging of
the droplets. Due to the low resolution, this step is usually very fast. With the
given laboratory scale setup, several cm² of surface area are covered within 1 s.
2. Subsequently, the ink is dried, resulting in shrinkage of the droplets and formation
of small solid islands which are referred to as ‘pinning centres’. In order to
enhance adhesion of the pinning centres to the substrate, they may be annealed at
temperatures up to 140 °C.
3. Finally, a continuous layer is inkjet printed in a second pass of the same material
at higher resolution (400 - 1000 DPI). As the liquid film sticks to the previously

111
From Inks to Layers

placed pinning centres, shrinkage and layer rupturing due to dewetting are
prevented. It should be noted that the second pass does not necessarily have to be
performed by inkjet printing, but may also be performed by other coating
techniques, even 1-dimensional ones, such as blade coating or slot die coating.

Figure IV-11: Schematic illustration of the creation and use of inkjet printed pinning
centres. Published in reference 184 and reproduced with permission.

The implementation of the pinning centre concept is illustrated in Figure IV-11.


PEDOT:PSS pinning centres printed from P Jet N V2 ink on top of highly hydrophobic
P3HT:PC60BM (contact angle water 103 °) and poly(vinylidene fluoride) PVDF surfaces
(contact angle water 91 °) are presented. Figure IV-12 a) shows inkjet deposition of a
continuous P Jet N V2 wet film on top of a P3HT:PC60BM surface which is partially
covered by an array of pinning centres. While the wet film remains perfectly
homogeneous in the area covered by pinning centres, it retracts from the non-modified
area.

112
From Inks to Layers

Figure IV-12: a) Optical microscope image of a dry PEDOT:PSS P Jet N V2 layer on a


P3HT:PC60BM surface. The dashed and dotted lines mark the edges of the pinning and
printing areas, respectively. Printing in an area covered by pinning centres results in
stable film formation. Printing on a blank surface results in dewetting and shrinkage of
the layer. b) PEDOT:PSS pinning centres dried at 60 °C do not provide sufficient
adhesion to pin the subsequently applied continuous wet film. c) The same sample after
annealing at 140 °C shows stable film formation. d) Photograph of PVDF surface with
water droplets. The square shaped droplet is held in position by P Jet N V2 pinning
centres. e) Microscope image the water droplet’s pinned three phase contact line.
Published in reference 184 and reproduced with permission.

113
From Inks to Layers

The forces acting on the pinning centres can be considerable. Weakly anchored pinning
centres can be swept away by the pull of the retracting wet film (Figure IV-12 b)). This
can be easily resolved by a temperature annealing step of the pinning centres. An
annealing temperature of 140 °C is found to be sufficient for the system of
P3HT:PC60BM and P Jet N V2 (Figure IV-12 c)). Images d) and e) highlight that the
pinning centre concept is of generic character and can be transferred also to other
systems, such as poly(vinylidene fluoride) PVDF surfaces (contact angle with water
 = 91 °).

Patterning of surfaces by pinning centres:

The concept of inkjet printed pinning centres allows the deposition of patterned layers
with high resolution by pursuing one of the following two strategies:

1. Cover the complete surface by inkjet printed pinning centres and apply the
continuous wet film locally, using inkjet technology. Implementation of this
strategy is demonstrated in Figure IV-12 c).
2. Cover a discretionary and well-defined area of the surface with pinning centres.
This is easily achieved by inkjet printing the desired pattern at low resolution,
resulting in a spatial modification of the surface energy. The subsequent
application of the continuous wet film can even be done with a one-dimensional
process, such as blade or slot die coating, as the ink will only wet on the pinning
centre modified part of the surface. An example of this approach is provided in
Figure IV-13. The photographs show a P3HT:PC60BM layer, locally covered by
PEDOT:PSS pinning centres. A subsequently applied continuous layer of
PEDOT:PSS ink P Jet N V2 dewets and leaves the areas without pinning centres
uncovered, thus revealing the pattern of the ZAE Bayern research institute logo.

114
From Inks to Layers

Figure IV-13: Demonstration of locally defined pinning centres on a P3HT:PC60BM


surface for patterning applications. a) Schematic illustration of pinning centre deposition
in form of the ZAE research institute lettering as negative image and b) blade coating of
continuous PEDOT:PSS liquid film. c-h) Photographs of samples after liquid film
deposition recorded in 5 s intervals. The PEDOT:PSS film ruptures and uncovers the
surface area without pinning centres, thus creating the desired layer patterning. Samples
and images were made by L. M. Eisenhofer and are reproduced in this work with
permission. Published in reference 184 and reproduced with permission.

With the printing setup used in this work, both patterning approaches allow high spatial
resolution of ~25 µm. The precision is mainly limited by the recession of the three phase
contact line between the pinning centres located at the edge of the liquid layer (see Figure
IV-14).

115
From Inks to Layers

Figure IV-14: Optical microscope image of dry PEDOT:PSS P Jet N V2 layer pinned on
a P3HT:PC60BM surface. b) The zoom shows a 3D confocal microscope image of the
pinned PEDOT:PSS layer edge. The pinning centres printed from 30 pl droplets have an
184
average diameter of 10.2 ± 0.5 µm after drying. Published in reference and
reproduced with permission.

Clearly, the achievable patterning precision highly depends on the printer setup. In this
work, Spectra SE 128 AA printheads offering a droplet size of 30 pl are used. More
advanced printheads offer drop volumes of less than 1 pl.186 This enables the deposition
of much smaller pinning centres which can be packed in denser arrays. Consequently,
shorter distances between neighbouring pinning centres can be realized, leading to less
recession of the three phase contact line. With a 500 fl printhead and under the
assumption of the same spreading characteristics of the impacting droplet on the substrate
surface (spreading factor βmax , equation 49), a minimum pinning centre distance of
22 µm and patterning precision of 5 µm is anticipated.
Besides the question how to increase the pinning centre density and patterning precision,
there is also the question how far the pinning centre density can be decreased before the
recession becomes critical and leads to a rupture of the wet layers between the pinning
centres.

116
From Inks to Layers

Limits of pinning centre density:

For the pinning centre approach to work properly it appears clear that there are both, an
upper and a lower limit for the pinning centre density. On one hand, a too high printing
resolution will lead to merging of the ink droplets, and thus prevent the formation of a
regular pinning centre pattern. On the other hand, if the density of pinning centres is too
low, they cannot prevent dewetting. For practical purposes, the processing window of the
pinning centre method must be determined more quantitatively. While the maximum
density of pinning centres (𝑅𝑒𝑠𝑚𝑎𝑥 ) determines the maximum achievable resolution of
surface patterning (see Figure IV-13), the minimum density (𝑅𝑒𝑠𝑚𝑖𝑛 ) results in
maximised processing speed.
To find 𝑅𝑒𝑠𝑚𝑎𝑥 , the area occupied by the liquid droplets just after their arrival at the
surface has to be considered. Numerous models have been described in literature to
estimate the maximum droplet spreading diameter 𝐷𝑚𝑎𝑥 upon impact on the substrate in
dependence of the initial droplet diameter 𝐷0 , the ink surface tension, the ink viscosity
and the droplet velocity upon impact (see chapter 2.2.8).124,187 The ratio of 𝐷𝑚𝑎𝑥 to 𝐷0 is
hereby referred to as the maximum spreading factor 𝛽𝑚𝑎𝑥 (equation 49).

𝐷𝑚𝑎𝑥
𝛽𝑚𝑎𝑥 = equation 49
𝐷0

𝑚𝑚
25.4 ⌊

𝑅𝑒𝑠𝑚𝑎𝑥 [𝐷𝑃𝐼] = 𝑖𝑛𝑐ℎ equation 50
𝛽𝑚𝑎𝑥 ∗ 𝐷0

For the applied setup (Spectra SE-128 AA printheads, Driving Voltage: 90 V, Driving
Waveform: 6-6-6 µs, droplet volume: 𝑉𝑜𝑙𝐷 = 30 pl, 𝐷0 = 38.6 µm,), a maximum
spreading factor 𝛽𝑚𝑎𝑥 of 2.2 for water based PEDOT:PSS inks, such as P Jet N V2 or
FHC Solar, is determined. This corresponds to a maximum droplet diameter upon impact
on the substrate 𝐷𝑚𝑎𝑥 of 85 µm. 𝐷𝑚𝑎𝑥 is also the minimum distance at which two
neighbouring droplets can be placed on the substrate without merging together. Equation
50 translates this value into the maximum pinning centre resolution 𝑅𝑒𝑠𝑚𝑎𝑥 of 300 DPI.
Experimental verification of 𝑅𝑒𝑠𝑚𝑎𝑥 is provided in Figure IV-15. For printing resolutions
< 𝑅𝑒𝑠𝑚𝑎𝑥 , a regular pattern of pinning centres is formed, whereas for printing resolutions
> 𝑅𝑒𝑠𝑚𝑎𝑥 droplets merge together forming irregular patterns.

117
From Inks to Layers

Figure IV-15: Pinning centres printed with P Jet NV2 ink on P3HT:PC60BM layers at
different resolutions. a) 250 DPI – regular pattern of pinning centres. b) 400 DPI –
pinning centres start to merge. c) 800 DPI – droplets have merged to a wet film which
184
dewets from the surface by contracting. Published in reference and reproduced with
permission.

The minimum pinning centre resolution at which dewetting of the wet film deposited on
top is prevented, is estimated by comparing the difference of the Gibbs free energies 𝐺 of
the system in the wetted (𝑤) and dewetted (𝑑) state, 𝐺𝑤 and 𝐺𝑑 , respectively. According
to the principle of lowest energy, wetting is thermodynamically favourable, if the
difference of the Gibbs free energy (∆𝐺) from equation 51 is positive.

∆𝐺 = 𝐺𝑑 − 𝐺𝑤 > 0 equation 51

A pure liquid, a smooth and homogeneous solid surface, constant pressure 𝑃, constant
temperature 𝑇𝑒 and a constant number of molecules 𝑁𝑚 are assumed. Under this
conditions, the free surface energies 𝛾 can be expressed as the change in Gibbs free
energies (equation 52), with 𝐴 being the surface area.

𝜕𝐺
𝛾=( ) equation 52
𝜕𝐴 𝑃,𝑇𝑒,𝑁𝑚

Hence, ∆G is obtained by multiplication of the free surface energies of the interfaces


solid-gas (𝑆), solid-liquid (𝑆𝐿) and liquid-gas (𝐿) with the respective interface areas in the
wetted (w) and dewetted (d) state, according to equation 53.

∆𝐺 = (𝐴𝑆𝑑 − 𝐴𝑆𝑤 ) ∗ 𝛾𝑆 + (𝐴𝑆𝐿𝑑 − 𝐴𝑆𝐿𝑤 ) ∗ 𝛾𝑆𝐿 +


equation 53
+ (𝐴𝐿𝑑 − 𝐴𝐿𝑤 ) ∗ 𝛾𝐿

118
From Inks to Layers

The interface areas in the wetted and dewetted states are given by equation 54 - equation
59, based on the geometric considerations illustrated in Figure IV-16. Due to the
translational symmetry of the quadratic pinning centre array, it is sufficient to describe a
single unit cell of length 𝑥𝑈𝐶 , which equals the smallest distance between pinning centres.

Figure IV-16: Schematic illustration of the model used to estimate the minimum pinning
centre density: a) Substrate surface covered with a regular pattern of pinning centres and
a continuous wet film. b) If the pinning centre density is too low, dewetting results in
formation of spherical caps of ink surrounding the pinning centres. c) Zoom-in on
spherical cap. d),e) PEDOT:PSS P Jet N V2 at a P3HT:PC60BM surface with different
pinning centre densities. d) A pinning centre resolution of 100 DPI (< 𝑅𝑒𝑠𝑚𝑖𝑛 ) leads to
dewetting of the wet film printed on top. e) A pinning centre resolution of 200 DPI
(> 𝑅𝑒𝑠𝑚𝑖𝑛 ) stabilizes the wet film and prevents layer rupturing and subsequent shrinkage
during drying. Published in reference 184 and reproduced with permission.

119
From Inks to Layers

In the wetted state, the solid-vapour interface disappears, while both, the solid-liquid and
the liquid-vapour interface areas are equal to the square of the unit cell length. In the fully
dewetted state, it is assumed that a spherical cap of ink forms around every pinning centre
(𝑟𝑆𝐶 : Radius of spherical cap base circle; ℎ𝑆𝐶 : Height of spherical cap). Furthermore, it is
assumed that the pinning centres are of negligible volume compared to the wet film (solid
contents of PEDOT:PSS inks are usually ~1-2 %wt). By substituting the surface areas of
equation 53 with equation 54 - equation 59 as well as the surface energies with Young’s
law (equation 60), 𝛥𝐺 is expressed as a function of the wet layer thickness 𝑑, the unit cell
width 𝑥𝑈𝐶 and the contact angle 𝜃. Mathematical reshaping leads to an expression for the
maximum unit cell length 𝑥𝑈𝐶 (equation 61), i.e. the maximum distance between pinning
centres at which the wet film covering is stable. Equation 62 gives the corresponding
minimum pinning centre resolution 𝑅𝑒𝑠𝑚𝑖𝑛 , which increases with decreasing thickness 𝑑
of the wet film and with increasing contact angle  of the corresponding ink at the surface
to be wetted.

𝐴𝑆𝑤 = 0 equation 54

𝐴𝑆𝐿𝑤 = 𝑥𝑈𝐶 2 equation 55

𝐴𝐿𝑤 = 𝑥𝑈𝐶 2 equation 56

𝐴𝑆𝑑 = 𝑥𝑈𝐶 2 − 𝜋 ∗ 𝑟𝑆𝐶 2 equation 57

𝐴𝑆𝐿𝑑 = 𝜋 ∗ 𝑟𝑆𝐶 2 equation 58

𝐴𝐿𝑑 = 𝜋 ∗ (𝑟𝑆𝐶 2 + ℎ𝑆𝐶 2 ) equation 59

𝛾𝑆 − 𝛾𝑆𝐿
𝑐𝑜𝑠(𝜃) = equation 60
𝛾𝐿

3
1 2 equation 61
6 ∗ 𝑑 ∗ √𝜋 ∗ (1 + )
1 + 𝑐𝑜𝑠(𝜃)
𝑥𝑈𝐶 =
𝜃 𝜃
𝑡𝑎𝑛 (2) ∗ (3 + 𝑡𝑎𝑛2 (2))

𝑚𝑚
25.4 equation 62
𝑅𝑒𝑠𝑚𝑖𝑛 [𝐷𝑃𝐼] = 𝑖𝑛𝑐ℎ
𝑥𝑈𝐶

120
From Inks to Layers

The validity of the model derived above is confirmed by the experiments shown in Figure
IV-16. For P Jet N V2 (𝜃 = 32.9 °), printed with a wet layer thickness of 10 µm (460 DPI,
𝑉𝑜𝑙𝐷 = 30 pl) onto a P3HT:PC60BM surface, a minimum pinning centre resolution of
114 DPI is calculated. For resolutions smaller than the predicted value, the wet film
ruptures between the pinning centres, whereas for pinning centre resolutions higher than
the calculated number, formation of a stable PEDOT:PSS wet film is achieved.
Due to the novel pinning centre approach and clear definition of processing windows, it is
now possible to manufacture layers from inks with extremely challenging wetting
behaviour, such as hydrophilic PEDOT:PSS ink P JET N, onto an otherwise non-wettable
hydrophobic P3HT:PC60BM active layer (see Figure IV-9). The beneficial application of
this new strategy to printed electronics is demonstrated in chapter 5.2 by fabricating well
working organic solar cells.

4.4. Conclusion
In this chapter of the thesis inkjet inks were developed which simultaneously fulfil the
following critical requirements that were previously specified in chapter 1.5:
(i) Eco-friendly materials (avoidance of environmentally highly critical halogenated
solvents)
(ii) Formation of well-defined spherical droplets
(iii) Sufficient ink stability and nozzle open time to ensure a robust and scalable
printing process
(iv) Printing of homogeneous layers which do not suffer from dewetting

A summary of the achievements is presented in Table IV-7. To lay the foundation for
development of OPV inkjet inks that fulfil the above-mentioned aspects, the Ohnesoge
theory was applied as guideline. Suitable solvents which are, according to theory, located
in the region of stable drop formation were selected to modify well known and well
working OPV inks, thereby making them applicable for inkjet processing. This comprised
the addition of high boiling point solvents, such as PG, EG and 1-pentanol, to IPA based
ZnO and AgNW inks. Besides the suppression of satellites, ink modification resulted in
prolonging of the nozzle open times.

121
From Inks to Layers

Table IV-7: Overview of OPV inks applied in this work and rating based on the critical
requirements for a robust inkjet printing process. Underlined inks were developed or
modified in this work to fulfil the processing requirements.

Ink Eco-friendly Drop Ink stability Layer


materials formation nozzle open time homogeneity

P3HT: x CB+BrAni x satellites x fast clogging 


PC60BM x o-DCB   
() o-X x no drops x gelling 
() o-X:THN   

PV2000: () o-X:THN   


PC70BM

PEDOT:PSS  H2O   pinning


P Jet N V2 centres

PEDOT:PSS  H2O   CFS


F HC Solar addition

ZnO NPs  IPA x satellites x fast clogging 


N-10  1-pentanol   

AgNW  IPA x satellites x fast clogging 


ClearOhm  1-pentanol   

Ag NPs  DGME   
I50DM106

The stable drop formation of the 1-pentanol based AgNW ink allows the conclusion that
digital printing of nanowires with an average length of ~30 µm with industrial printheads
having nozzle diameters in the same size range is possible. Furthermore, a solvent blend
consisting of o-X and THN mixed in a ratio of 1:1 (vol.) was identified for printing the
active materials P3HT:PC60BM and PV2000:PC70BM. By using a higher THN content
than suggested in literature (typically < 25 %vol),85,93 fast gelling of the P3HT:PC60BM
inks at RT is suppressed, making this solvent blend a suitable alternative to
environmentally critical halogenated organic solvents, such as o-DCB.

122
From Inks to Layers

Finally, the layer formation after droplet impact on the substrate was investigated,
exposing problematic wetting behaviour of hydrophilic PEDOT:PSS inks on top of
hydrophobic BHJ active layers, thus preventing deposition of homogenous layers.
Investigation of the ink’s surface tension and substrate’s surface energy as well as visual
inspection of the layer quality led to the conclusion that fluorosurfactant addition cannot
resolve the wetting problems for every tested PEDOT:PSS inkjet ink on the active layer
surface.
As alternative, a novel, surfactant-free strategy to coat homogeneous, defect-free layers
on surfaces with challenging wetting behaviour was presented. This was achieved by
inkjet printing single droplets, which form anchoring points upon drying. Dewetting of
the subsequently coated continuous layer of ink is prevented, as the anchoring points pin
the wet film to the desired position. Combining the advantages of digital printing and
dewetting, surface coatings can be patterned in arbitrary shapes at resolutions down to a
few tens of micrometres without employing masks. This is possible by printing the
pinning centres in the desired pattern prior to wet film deposition. Applying this strategy,
a unprecedented film quality of aqueous PEDOT:PSS inks on hydrophobic
P3HT:PC60BM layers was presented. Calculations based on the Free Surface Energy of
ink and solid substrate combined with simple geometric considerations yield minimum
and maximum printing resolutions required for suppressing the dewetting. Furthermore,
the highest achievable patterning resolution was determined. With the current generation
of 30 pl heads a patterning precision of 21 µm was presented. By further reduction of the
droplet size to 500 fl, a resolution to below 6 µm is anticipated. Registration is no longer
limited by the alignment of the coating or printhead, but instead by the pinning centre
pattern and the non wetting properties of the ink and surface. The quantitative
investigation of the limitations of the pinning centre strategy provides a clear definition of
the processing window and thus allows the precise processing of functional inks on non-
wetting surfaces. The (i) film quality, (ii) resolution and (iii) registration of this method is
only limited by the precision of the pinning centre pattern and the wetting properties of
ink and surface.
The proposed pinning centre strategy also provides the solution to a major problem of
printing functional inks. In many applications, like printed electronics, the use of ink
additives and surface modifiers is limited, because any change of ink formulation has
direct influence on the properties and therefore functionality of the printed film. Pinning
of the wet film obviates the need for such modifications, leaving the ink and layer

123
From Inks to Layers

properties unaffected. The successful deposition of the pinning centres on a highly


hydrophobic PVDF surface, allows the conclusion that the newly introduced pinning
centre strategy is of highly generic character. It can thus be regarded as versatile tool to
the field of surface engineering going far beyond OPV applications. The beneficial
application of this new strategy to printed (opto-)electronics is demonstrated in chapter
5.2.

124
CHAPTER V: From Layers to Cells

Abstract:

In this chapter, step-by-step transition from a well working blade coated reference solar
cell to a fully inkjet printed device is demonstrated. Hereby the focus is set on the critical
aspects for an industrially applicable OPV inkjet process as defined in chapter 1.5 (v)
layer functionality and (vi) high efficiency. In contrast to many literature reports, a loss-
free transfer from blade coating to inkjet printing is achieved for ZnO, PEDOT:PSS,
P3HT:PC60BM and PV2000:PC70BM layers. As replacement for the sputtered
semitransparent ITO base electrode of the reference device inkjet printed silver grid
fingers with optimised dimensions are applied. Introduction of this electrode type results
in a 𝑃𝐶𝐸 loss of ~15 %, which is mainly caused by shading effects and parasitic
absorption of light in the PEDOT:PSS layer. As alternative to the optically conspicuous
silver finger structures a digitally printed AgNW mesh electrode is introduced. This is
possible by applying the novel alcohol based inkjet ink reported in the previous chapter.
The AgNWs offer an excellent balance of sheet resistance (< 20 Ω/sq) and transmittance
(> 90 %) as well as a unique visual appearance. Fully inkjet printed devices with AgNW
electrodes achieve up to 4.3 % efficiency, which is a literature record.

Parts of this chapter have been published in:


P. Maisch, K.C. Tam, L. Lucera, H.-J. Egelhaaf, H. Scheiber, E. Maier, et al., Inkjet printed silver
nanowire percolation networks as electrodes for highly efficient semitransparent organic solar cells, Org.
Electron. (2016). doi:10.1016/j.orgel.2016.08.006, reproduced with permission from Elsevier.
P. Maisch, L.M. Eisenhofer, K.C. Tam, A. Distler, M.M. Voigt, C.J. Brabec, H.-J. Egelhaaf, A generic
surfactant-free approach to overcome wetting limitations and its application to improve inkjet-printed
P3HT:non-fullerene acceptor PV, J. Mater. Chem. A. (submitted)
P. Maisch, K.C. Tam, F.W. Fecher, H.-J. Egelhaaf, H. Scheiber, E. Maier, et al. Inkjet printing of highly
conductive nanoparticle dispersions for organic electronics. Molded Interconnect Devices (MID), 12th
International Congress (2016). doi:10.1109/ICMID.2016.7738932, reproduced with permission from IEEE.
Only parts of the publication text authored by P. Maisch are reproduced in subchapters 5.2 and 5.3 of this
work. Contributors and specific contributions in form of SEM images to this chapter are clearly stated in the
figure captions of subchapter 5.3. Additionally, permission to utilise the whole content of the publications
as part of this thesis was given from all co-authors involved through author ownership declarations.

125
From Layers to Cells

5.1. Inkjet printed active layers


It is well known that the maximum achievable efficiency of an organic solar cell is highly
dependent on the morphology of the active layer BHJ (see chapter 2.1.1.). In solution
processed OPV, BHJ optimisation typically focuses on the solvent system,61 processing
additives,36 ink/coating temperature188 and annealing.189 However, the inkjet process
limits the use of these optimisation paths. The ink formulation is mainly dictated by
considerations of printability and drop formation. At the same time, the temperature of
ink, printhead and substrate is typically kept low to minimise evaporation at the nozzle
plate, which leads to short open times. Ideally, industrial printing should be done at room
temperature. Considering these boundary conditions, it is not surprising that most
literature works report significant changes in BHJ morphology and performance losses
when switching from a well working reference process, such as spin- or blade coating, to
inkjet printing of active layers.129,131 Minimising these efficiency losses for inkjet printing
of OPV active layers is the main target of this chapter.
In order to distinguish between efficiency losses related to ink formulation and losses
related to the printing procedure, the inkjet inks developed in the previous chapter were
first applied in blade coated solar cells. The results in Figure V-1 and Table V-1 show
similar efficiencies of ~2.8 % for the o-DCB and o-X:THN 1:1 (vol.) inks. Compared to
the highly optimised reference device blade coated from CB + 5 %vol BrAni ink, an
efficiency drop of ~10 % is observed, which is mainly caused by decreased FF. To obtain
further information about the loss mechanism, the J-V curves were fitted with a one-diode
equivalent circuit approach under assumption of bimolecular recombination (see chapter
2.1.2). Excellent agreement between experimental and simulated curves is found by
variation of the charge carrier mobility µ (µCB+BrAni: 1.7*10-4 cm²/(V*s), µo-DCB:
1.4*10-4 cm²/(V*s), µo-X:THN: 1.3*10-4 cm²/(V*s)). Furthermore, UV-Vis absorption
spectra reveal more expressed vibronic peaks at ~500 nm and ~600 nm for layers
fabricated from CB + 5%vol BrAni ink compared to o-DCB and o-X:THN 1:1 (vol.) inks.
This can be attributed to increased crystallinity due to the interchain π−π stacking of the
P3HT backbone.190 The observations from UV-Vis and J-V curve analysis are strong
indicators that the performance loss is caused by different BHJ morphology and resulting
charge transport properties. This fits well with literature reports about the influence of
solvent formulation and processing additives, such as BrAni, on the active layer
microstructure.191 Detailed investigations of the blade coated active layer BHJ and

126
From Layers to Cells

correlation to device efficiency go beyond the scope of this thesis and can be found
elsewhere.36,192

Figure V-1: Representative J-V characteristics of solar cells with doctor bladed (DB)
P3HT:PC60BM active layers from different ink formulations. The inset shows the J-V
characteristics measured in the dark. b) Normalised UV-Vis absorbance spectra of
P3HT:PC60BM layers blade coated on a glass substrate.

Table V-1: P3HT:PC60BM solar cell key performance values of representative devices
with active layers blade coated (DB) or inkjet printed (IJ) from different ink formulations.
The charge carrier mobility µ and the bimolecular recombination coefficient 𝑘𝑟𝑒𝑐,𝐵𝑖𝑚𝑜𝑙
are extracted from simulated fit curves using a single diode model as described in chapter
2.1.2.

Sample 𝑃𝐶𝐸 𝐹𝐹 𝐽𝑆𝐶 𝑉𝑂𝐶 𝜇 𝑘𝑟𝑒𝑐,𝐵𝑖𝑚𝑜𝑙


(%) (%) (mA/cm²) (V) (cm²/(V*s)) (cm³/s)

DB CB + 5 %vol BrAni 3.20 67.24 8.75 0.56 1.7*10-4 6.0*10-13

IJ CB + 5 %vol BrAni 3.26 68.80 8.15 0.56 2.1*10-4 6.0*10-13

DB o-DCB 2.89 62.05 8.32 0.56 1.4*10-4 6.5*10-13

IJ o-DCB ‘dot’-pattern 1.54 46.28 5.94 0.55 3.3*10-5 6.1*10-13

IJ o-DCB 2.73 57.89 8.54 0.56 1.1*10-4 6.5*10-13

DB o-X:THN 1:1 (vol.) 2.75 60.76 8.39 0.56 1.3*10-4 6.0*10-13

IJ o-X:THN 1:1 (vol.) 2.94 61.65 8.66 0.55 1.3*10-4 6.0*10-13

127
From Layers to Cells

After identification of ink formulation related efficiency losses, the influence of the inkjet
printing process on the device performance is investigated. To mimic the parameters used
for the blade coating reference process, first inkjet printing tests were performed with
o-DCB ink at elevated substrate table temperatures of 60 °C. This results in a distinct
‘dot’-pattern shown in Figure V-2.

Figure V-2: a)-c) Microscope images of inkjet printed P3HT:PC60BM layers from o-DCB
ink. a) Substrate temperature 60 °C, inset: Confocal microscope height profile measured
along the red line. b) Substrate temperature 50 °C. c) Substrate temperature RT. All
layers were printed with a speed of 100 mm/s and QF 6. d) AFM image of inkjet printed
P3HT:PC60BM layer from o-DCB (substrate temperature 60 °C – dot pattern. e) AFM
image of P3HT:PC60BM reference layer from CB + 5 %vol BrAni ink.

With the used printer settings (speed 100 mm/s, QF = 6, see chapter 3.2.2 for explanation
of printing parameters) the single droplets have sufficient time to dry on the substrate
before overlapping neighbouring droplets impact on the substrate. An AFM image of the
surface shows a coarse pattern of the droplet shape and a microstructure with feature sizes
that appear smaller than for the well working P3HT:PC60BM layer from CB + 5 %vol
BrAni ink. UV-Vis spectra of the ‘dot’-patterned layers show very little expressed

128
From Layers to Cells

shoulder formation at ~550 nm and ~600 nm (Figure V-3 b)). These observations indicate
that the small droplet volume of ~30 pl in combination with elevated substrate
temperature leads to too fast evaporation of solvent causing poor crystallinity. It is known
from literature that P3HT:PC60BM layers with such microstructure typically yield low
solar cell efficiency.190 This is experimentally confirmed by manufacturing solar cells.
Devices with the distinct ‘dot’-pattern suffer from low 𝑃𝐶𝐸 of only 1.5 %. Simulation of
the J-V characteristics with the one-diode equivalent circuit approach reveals low charge
carrier mobility of only 3.3*10-5 cm²/(V*s). By lowering the substrate table temperature
to 50 °C, the o-DCB ink droplets start merging before they are completely dried.
However, the droplet pattern is still visible. Keeping the substrate temperature at RT, the
droplets have sufficient time to merge completely, thus forming a homogeneous wet film
(Figure V-2 c)). Subsequent drying on a 70 °C hot plate results in P3HT:PC60BM layers
with similar appearance to those manufactured by blade coating. UV-Vis and optical
microscope analysis do not show any notable difference between the inkjet printed (RT +
subsequent drying) and blade coated layers. The J-V characteristics (Figure V-3) of solar
cells with inkjet printed P3HT:PC60BM layers are highly comparable to the blade coated
devices.

Figure V-3: a) Representative J-V characteristics of solar cells with inkjet printed (IJ)
P3HT:PC60BM active layers from different ink formulations and simulated fit curves used
to extract the charge carrier mobility. The inset shows the J-V characteristics measured
in the dark. b) Normalised UV-Vis absorbance spectra of P3HT:PC60BM layers printed
on glass substrates.

129
From Layers to Cells

The same behaviour is also observed when printing P3HT:PC60BM from o-X:THN
1:1 (vol.) or CB + 5 %vol BrAni inks. Please note that the latter formulation is poorly
suited for inkjet purpose because the low viscosity prevents formation of stable round
shaped droplets (see chapter 4.1.1). Nevertheless, complete merging of the droplets on the
substrate enables printing of homogeneous layers, albeit with poor edge definition due to
satellites. The statistical evaluation provided in Figure V-4 confirms that the change of
ink formulation from CB + 5 %vol BrAni to o-X:THN in order to achieve stable drop
formation with green solvents results in an efficiency drop of ~10 %. The difference in
efficiency between blade coated and inkjet printed layers using the same ink is almost
negligible.

Figure V-4: Box plot 𝑃𝐶𝐸, FF, 𝐽𝑆𝐶 and 𝑉𝑂𝐶 comparison of P3HT:PC60BM solar cells
with doctor bladed reference (black) and inkjet printed (red) active layer from
a) CB + 5 %vol BrAni ink and b) o-X:THN 1:1 (vol.) ink. Each box contains data of at
least 5 solar cells.

130
From Layers to Cells

The strategy of inkjet printing the active ink at RT and subsequent drying can also be
transferred to donor:acceptor blends of higher efficiencies, such as PV2000:PC70BM. To
gain information about the influence of the solvent formulation and printing procedure on
the solar cell efficiency, well working reference devices according to the recommended
processing recipe from Raynergy Tek were fabricated. With the blade coated devices
from pristine o-X ink efficiencies of 6 - 7 % are reached. However, due to limited ink
stability at room temperature (gelling) and satellite formation, pristine o-X inks are poorly
suited for inkjet printing (see chapter 4.1 Ink formulation). Experiments reveal that, in
analogy to P3HT:PC60BM, THN is a useful co-solvent to improve printability. As shown
in Figure V-5, the addition of 50 %vol THN does not negatively influence the device
performance. Even solar cells blade coated from pristine THN show only a slight
decrease in performance, reaching 𝑃𝐶𝐸𝑠 of almost 6 %.

Figure V-5: a) Representative J-V characteristics of PV2000:PC70BM solar cells blade


coated and inkjet printed from different ink formulations. The inset shows the J-V
characteristics measured in the dark. b) Normalised UV-Vis absorbance spectra of
PV2000:PC70BM layers on a glass substrate.

The normalised absorbance spectra of the PV2000:PC70BM layers coated from pristine
o-X, pristine THN or solvent blend inks appear similar, indicating that there is no
significant change in polymer crystallinity.

131
From Layers to Cells

Table V-2: Device performance parameters of the PV2000:PC70BM solar cells from
o-X:THN 1:1 (vol.) ink shown in Figure V-5.

Sample 𝑃𝐶𝐸 𝐹𝐹 𝐽𝑆𝐶 𝑉𝑂𝐶


(%) (%) (mA/cm²) (V)

PV2000:PC70BM reference DB 6.90 64.6 13.4 0.80

PV2000:PC70BM IJ 6.49 61.5 13.2 0.80

Solar cells with PV2000:PC70BM layers inkjet printed at room temperature and
subsequently dried at 70 °C yield similar performance as blade coated reference devices.
Hero cells reach record efficiencies of 7.3 %, slightly exceeding literature reports of
comparable inkjet printed OPV.143 The statistical evaluation (Figure V-6) proves that
loss-free transition from blade coating to inkjet printing is possible by application of
active layer inkjet inks from (i) environmentally non-critical solvents which offer (ii)
good drop formation, (iii) sufficient ink stability and nozzle open time as well as (iv)
formation of homogeneous layers.

Figure V-6: Box plot 𝑃𝐶𝐸, 𝐹𝐹, 𝐽𝑆𝐶 and 𝑉𝑂𝐶 comparison of solar cells with doctor bladed
(black) and inkjet printed (red) PV2000:PC70BM active layers (o-X:THN 1:1 vol.). Each
box contains data of six solar cells.

Therefore, the two further critical aspects, (v) layer functionality and (vi) high efficiency,
for an industrially applicable OPV inkjet printing process, as defined in chapter 1.5, have
been fulfilled for the active material. Furthermore, by separating printing and drying

132
From Layers to Cells

process, a typical R2R manufacturing scenario is mimicked, in which the web passes
printing and drying stations consecutively.

5.2. Inkjet printed buffer layers


The next step on the way to fully inkjet printed devices is to replace the blade coated EEL
(ZnO nanoparticles) and HEL (PEDOT:PSS) of the reference solar cell with an inkjet
printed equivalent.

Inkjet printed EEL

The EEL was printed with an ink formulation derived from IPA based N-10 ink from
Avantama. As described in chapter 4.1, the ink was modified with either propylene glycol
(PG) or 1-pentanol to achieve stable droplet formation and to extend the nozzle open
times. Similar to inkjet printing of the active layer, a strongly influence of the substrate
table temperature on the morphology of the printed layers is found. Figure V-7 shows
confocal microscope images of single ZnO droplets printed from a) IPA + 10 %vol PG
ink and b) 1-pentanol ink on an ITO substrate surface, revealing a coffee stain structure
for both inks. A 3D surface topography image of a ZnO layer printed at a substrate table
temperature of 40 °C (Figure V-7 c), IPA + 10 %vol PG, QF 6, printing speed 100 mm/s)
shows that droplets deposited at the surface have not enough time to merge before drying.
Single droplets overlap each other, resulting in a rough layer with spikes of more than
100 nm height.
Most of the small scale P3HT:PC60BM solar cells manufactured with such patterned ZnO
layers achieve efficiencies in the 3 % range, but some devices show high leakage current
and decreased 𝐹𝐹 and 𝑃𝐶𝐸 (see blue graph Figure V-8), which could be caused by the
ZnO spikes protruding into the active layer or even penetrating it. To avoid this effect,
inkjet printing was performed at room temperature. Thus the single inkjet droplets on the
substrate have enough time to merge and to form a homogeneous layer.

133
From Layers to Cells

Figure V-7: Confocal microscope images of inkjet printed ZnO droplets on an ITO
substrate surface printed from a) IPA + 10 %vol PG and b) 1-pentanol ink. c) 3D
confocal microscope image of a ZnO layer printed at 40 °C from IPA + 10 %vol PG ink
(QF 6, printing speed 100 mm/s). d) 3D confocal microscope image ZnO droplets printed
from 1-pentanol ink. e) Optical microscope image of ZnO layer printed at room
temperature from IPA + 10 %vol PG ink and f) 1-pentanol ink.

However, experiments show that wet films printed from IPA + 10 %vol PG based ZnO
ink suffer from extensive spreading and fingering instabilities at the three phase contact
line. This makes any precise patterning of inkjet printed layers impossible (see Figure V-7
e)). It is known from literature that such instabilities occur when liquid spreading is
forced e.g. by gravitational, centrifugal, shear or Marangoni stress.193

134
From Layers to Cells

Figure V-8: Representative J-V characteristics of P3HT:PC60BM solar cells with ZnO
nanoparticle EEL inkjet printed from 1-pentanol based ink (red) and blade coated
reference (black). Also shown is a solar cell with a ZnO layer (blue) inkjet printed at
elevated substrate temperature from IPA + 10 %vol PG, resulting in a ‘dot’-shaped
pattern as displayed in Figure V-7 c). The inset shows the J-V characteristics measured
in the dark.

Table V-3: Device performance parameters of the solar cells shown in Figure V-8.

Sample 𝑃𝐶𝐸 𝐹𝐹 𝐽𝑆𝐶 𝑉𝑂𝐶


(%) (%) (mA/cm²) (V)

ZnO reference DB 3.29 70.3 8.22 0.57

ZnO (1-pentanol) IJ 3.20 68.8 8.29 0.56

ZnO (IPA + 10 %vol PG) 2.74 63.5 7.68 0.56


‘dot’-pattern

Further investigations reveal that this effect does not appear for pristine IPA or pristine
PG, but is highly expressed for binary blends of both components, independent of the
deposition method (Figure V-9). Similar observations have repeatedly been described in
literature for binary mixtures of volatile solvents.194,195

135
From Layers to Cells

Figure V-9: Marangoni flow induced spreading and fingering instability of a droplet of
IPA + 10 %vol PG placed with a pipette tip on a on an ITO surface (40 °C). Droplet after
a) 1 s, b) 5 s, c) 10 s, d) 20 s.

The effect is explained by an increased evaporation rate at the droplet edge, which results
in a local increase of the less volatile component (in this case PG). The concentration
gradient is associated with a gradient in surface tension (𝛾𝑃𝐺 = 38 𝑚𝑁/𝑚 , 𝛾𝐼𝑃𝐴 =
22 𝑚𝑁/𝑚), causing mass transport (Marangoni effect) to the droplet edge, thus resulting
in the observed drop shape and spreading characteristics. A detailed study and FEM
modelling of the liquid flow characteristics in drying droplets goes beyond the scope of
this thesis and is found elsewhere.196 To avoid Marangoni stress induced liquid film
fingering instability, ZnO nanoparticle inkjet inks based on pristine 1-pentanol, as
introduced in chapter 4.1, are applied. Inkjet printed layers yield excellent edge definition
(Figure V-7 f)). A comparison to layers blade coated from an ink of ZnO nanoparticles in
pristine IPA (N-10) shows no significant difference in performance (Figure V-8, Figure
V-10, Table V-3).

Figure V-10: Box plot 𝑃𝐶𝐸, 𝐹𝐹, 𝐽𝑆𝐶 and 𝑉𝑂𝐶 comparison of P3HT:PC60BM solar cells
with ZnO nanoparticle EEL inkjet printed from 1-pentanol based ink (red) and blade
coated reference (black). Each box contains data of six solar cells.

136
From Layers to Cells

This example highlights the importance of optimised inks and printing parameters to
create smooth and homogeneous layers for the production of highly efficient OPV
devices.

Inkjet printed HEL

For inkjet printing of the HEL, PEDOT:PSS was selected because it is well studied in
literature, commercially available in large quantities and physical properties like the
conductivity can be tuned in a wide range. As described in chapter 4.1.1, a stable inkjet
process can be achieved for multiple commercially available PEDOT:PSS ink
formulations without any further modification. The main challenge for inkjet fabrication
of PEDOT:PSS HELs is to achieve good wetting of the highly hydrophilic inks on the
hydrophobic polymer active layers. In this work, the fluorosurfactant CFS was used to
lower the surface tension of the inks in order to be able to inkjet print homogeneous
layers of F HC Solar on top of P3HT:PC60BM or PV2000:PC70BM surfaces (see chapter
4.3.1). A comparison of solar cell performance with inkjet printed and blade coated
PEDOT:PSS F HC Solar layers shows no significant difference in performance (Figure
V-11, Figure V-12, Table V-4).

Figure V-11: Representative J-V characteristics of solar cells with inkjet printed HEL
PEDOT:PSS F HC Solar + 0.1 %wt CFS (red) and blade coated reference (black).

137
From Layers to Cells

Table V-4: Device performance parameters of the solar cells shown in Figure V-11.

Sample 𝑃𝐶𝐸 𝐹𝐹 𝐽𝑆𝐶 𝑉𝑂𝐶


(%) (%) (mA/cm²) (V)

F HC Solar reference DB 3.16 63.3 9.12 0.55

F HC Solar IJ 3.42 67.3 9.10 0.56

Figure V-12: Box plot 𝑃𝐶𝐸, 𝐹𝐹, 𝐽𝑆𝐶 and 𝑉𝑂𝐶 comparison of solar cells with inkjet printed
HEL PEDOT:PSS F HC Solar + 0.1 %wt CFS (red) and blade coated reference (black).
Each box contains data of at least six solar cells.

For other PEDOT:PSS inks, such as P Jet N V2, which is especially designed for inkjet
purposes, addition of 0.1 %wt fluorosurfactant did not result in homogeneous layer
formation. Blade coated or inkjet printed P Jet N V2 wet layers rupture and uncover the
P3HT:PC60BM surface (see chapter 4.3.1). Only upon the addition of high amounts of
fluorosurfactant (1.5 %wt) wetting can be enforced. However, such devices suffer from
severe degradation of all performance parameters (Figure V-13). This is in agreement
with literature reports by Savva et al., who found significantly reduced performance and
increased series resistance of P3HT:PC60BM solar cells when adding 2 %wt of the
fluorosurfactant Zonyl FS 300 to a PEDOT:PSS ink.178 Nevertheless, it is possible to
produce homogeneous PEDOT:PSS films on the active layer using the newly developed
pinning centre strategy as described in chapter 4.3.2. The resulting J-V characteristics
(Figure V-13) and the performance parameters (Figure V-14) demonstrate that the solar
cells manufactured with the pinning centre strategy are comparable to that of state-of-the-
art devices using similar materials and device structure.197

138
From Layers to Cells

Figure V-13: Representative J-V characteristics of solar cells with inkjet printed HEL
PEDOT:PSS from P Jet N V2 ink. The layer does not allow wetting with small amounts of
surfactant. The red curve is from a device with a PEDOT:PSS layer fabricated according
to the pinning centre strategy (see chapter 4.3.2 and inset). The black curve shows a
reference solar cell where high amounts of fluorosurfactant are applied to enforce
wetting. Published in reference 184 and reproduced with permission.

Table V-5: P3HT:PC60BM solar cell key performance values of representative devices
shown in Figure V-13. Published in reference 184 and reproduced with permission.

Sample 𝑃𝐶𝐸 𝐹𝐹 𝐽𝑆𝐶 𝑉𝑂𝐶


(%) (%) (mA/cm²) (V)

P Jet N V2 + CFS 1.5 %wt 1.51 41.2 8.11 0.45

P Jet N V2 pinning centres 3.23 63.6 9.46 0.54

It was thus demonstrated that the inkjet printed pinning centre strategy can be applied to
overcome wetting limitations even in systems in which conventionally used surfactants
fail. The given example highlights the great potential of this newly introduced method for
printed electronics production.

139
From Layers to Cells

Figure V-14: Box plot 𝑃𝐶𝐸, 𝐹𝐹, 𝐽𝑆𝐶 and 𝑉𝑂𝐶 comparison of solar cells with inkjet printed
HEL PEDOT:PSS P Jet N V2 applied by using the pinning centre strategy (red) and
reference with high amount of fluorosurfactant (black) to enforce wetting. Each box
contains data of at least six solar cells.

5.3. Inkjet printed electrodes


In the previous chapters, a transition from blade coating to inkjet printing was
demonstrated for the active and buffer layers without any notable solar cell efficiency
loss. In the following, replacement of the reference device electrodes by inkjet printed
equivalents will be discussed. An overview of commonly applied semitransparent
electrode types is presented in chapter 2.1.3.2. In this work, silver grids in combination
with HC PEDOT:PSS and AgNW meshes are chosen for implementation by inkjet
because they are suitable for upscaling and provide excellent balance between
conductivity and transmittance.

5.3.1. Silver grid electrodes


Silver grids are commonly fabricated by screen printing198 or flexographic printing199 and
have already shown great potential in OPV applications. However, most works about
inkjet printing of silver grids report performance losses of typically ~15-60 % compared
to reference devices fabricated with conventional electrodes, such as ITO or thermally
imprinted grid structures,136,137,143 leaving significant room for optimisation. The
performance losses connected to inkjet printed silver grids are often attributed to shunting
of devices, limited conductivity of silver or PEDOT:PSS, shading by the grid or parasitic
140
From Layers to Cells

light absorption in the HC PEDOT:PSS layer. Optimisation of device performance


requires careful balancing of all these optical and electrical parameters to identify a global
loss minimum.
In this work, a 2D Comsol Multiphysics FEM model of a solar cell with grid finger
electrode, which was developed by F.W. Fecher, is applied for this purpose. Details about
158
the model are found in reference . According to F.W. Fechers simulation, linear silver
fingers yield higher efficiency compared to quadratic, hexagonal or triangular silver grid
structures, and were thus chosen for experimental implementation. By applying
parametric sweeps of silver line width, length and spacing to neighbouring silver lines,
F.W. Fecher identified optimised solar cell and silver grid layouts for different sets of
layer properties, such as PEDOT:PSS and silver sheet resistance, contact resistance and
transmittance. Considering the properties of the inkjet printed layers produced in this
work as well as resolution restrictions from the inkjet machine (single silver droplets
spread on the substrate to ~100 µm diameter, thereby defining the minimum silver line
width), 5 mm long and 100 µm wide silver fingers with 1 mm spacing were selected for
printing.
A photograph of a substrate containing six PV2000:PC70BM solar cells with inkjet
printed silver grid base electrodes is shown in Figure V-15. The grid structures are printed
on a glass substrate from nanoparticle silver ink I50T13 at a resolution of 500 DPI, dried
at 100 °C for 5 min and additionally annealed at 200 °C for 15 min. This results in a
silver sheet resistance of 0.22 Ω/sq (measured with 4-point probe). A PEDOT:PSS (F HC
Solar) layer of 100 nm thickness is applied on top of the grid structure to achieve the
necessary lateral conductivity and to planarise the printed silver.137 Compared to the
reference with ITO base electrode, the cells with inkjet printed silver grid reveal a 𝑃𝐶𝐸
reduction of ~15 %. This efficiency loss is mainly caused by decreased 𝐽𝑆𝐶 of ~10 %,
while the 𝐹𝐹 drop of ~4 % is relatively small.

141
From Layers to Cells

Figure V-15: a) Photograph of PV2000:PC70BM solar cells with inkjet printed silver grid
base electrode and evaporated silver grid top electrode. b) Confocal microscope 3D
image of an inkjet printed silver finger on a glass substrate ending in a conductive trace
Published in reference 200 and reproduced with permission from IEEE. c) Representative
J-V characteristics of solar cells with inkjet printed silver grid base electrode (red) and
ITO reference (black). Additionally, a simulated J-V curve of a solar cell with silver grid
base electrode (blue) is shown. The inset shows the J-V characteristics measured in the
dark.

142
From Layers to Cells

Table V-6: Device performance parameters of the solar cells shown in Figure V-15.

Sample 𝑃𝐶𝐸 𝐹𝐹 𝐽𝑆𝐶 𝑉𝑂𝐶


(%) (%) (mA/cm²) (V)

Reference ITO 6.72 58.4 14.25 0.81

Ag grid nanoparticles IJ 5.66 56.26 12.85 0.79

Ag grid simulated 5.85 56.53 12.79 0.81

Figure V-16: Box plot 𝑃𝐶𝐸, 𝐹𝐹, 𝐽𝑆𝐶 and 𝑉𝑂𝐶 comparison of PV2000:PC70BM solar cells
with inkjet printed silver grid base electrode and ITO reference. Same device structure
and measurement parameters as described in Figure V-15 apply.

To gain further insights into the main loss mechanisms, the FEM model, developed by
F.W. Fecher, was adopted with permission and modified to fit the exact specifications of
the inkjet printed solar cells. A schematic drawing of the modified model and list of
simulation parameters is presented in the experimental section (chapter 3.4.2). Besides the
sheet resistances of the printed layers, also the specific contact resistance 𝜌𝐶 at the
PEDOT:PSS/silver interface was taken into account. To gain information about this
parameter, transmission line structures as described in chapter 3.3.2.4 were inkjet printed.
From the transfer length 𝐿𝑇 and contact resistance 𝑅𝐶 , which are extracted from the x-
and y-axis intercepts of the interpolated transfer line graph (Figure V-17 c)), the geometry
independent specific contact resistance 𝜌𝐶 is determined. The obtained value of
0.11 Ω*cm² is within the range reported in literature for similar PEDOT-silver
contacts.157,201 Furthermore, the transmittance of the electrodes was considered in the

143
From Layers to Cells

simulation. To gain information about optical losses, UV-Vis transmission curves of the
grid electrodes were recorded. A comparison to the ITO reference electrode is shown in
Figure V-17. The simulated J-V characteristics of the solar cells with inkjet printed silver
grid are presented in Figure V-15 and show excellent agreement with experiment.

Figure V-17: a) Normalised 𝑃𝐶𝐸 of solar cells with ITO-, inkjet printed silver grid- and
simulated silver grid electrode. b) Analysis of 𝑃𝐶𝐸 loss mechanisms derived by FEM
simulation. c) Transmission line measurement of HC PEDOT:PSS (F HC Solar) and
inkjet printed silver nanoparticles (ink I50DM106) revealing a contact resistance of
0.11 Ω*cm². d) Comparison of transmission spectra of glass, ITO coated glass and glass
covered with 100 nm PEDOT:PSS + silver grid according to the above specified grid
layout (10 % area loss by shading).

According to the simulation, shading by the silver grid (54.1 %) and parasitic absorption
within the PEDOT:PSS layer (36.0 %) are the dominating 𝑃𝐶𝐸 loss contributors.
Compared to this, resistive losses caused by the PEDOT:PSS sheet resistance, silver line

144
From Layers to Cells

resistance and PEDOT:PSS-Ag contact resistance are with 1.2 %, 0.2 % and 8.5 % much
lower. This fits well with the experimentally derived J-V curves, which confirm that the
𝑃𝐶𝐸 loss is mainly caused by decreased 𝐽𝑆𝐶 .
Compared to most literature reports of organic solar cells with inkjet printed grid
electrodes (see literature summary Table II-5), efficiencies achieved in this work are
higher. This is attributed to the high efficiency of the reference device and application of
an optimised grid layout, accounting for the electrical and optical properties of the inkjet
printed layers.156 Power losses of ~15 % when switching from ITO to printed silver grids
are comparable to state of the art literature.143 According to F.W. Fechers simulations,
further improvement is possible by decreasing the silver grid finger width. This can be
achieved by applying inkjet printheads with smaller droplet diameter.
A limiting factor for application of this electrode type in OPV devices is the visibility of
the silver lines. As described in chapter 1.3, new solar cell applications, like the
integration into wearables or windows, requires visually inconspicuous electrodes.
Therefore, the state of the art has to be overcome by introducing a novel shy inkjet
printable electrode.

5.3.2. Silver nanowire electrodes


As an alternative to the visually obstructive silver grids, AgNW meshes are fabricated by
inkjet printing and used as base and top electrode in OPV devices. This is demonstrated in
the following chapter, which is partly published in reference 164.
AgNW mesh electrodes have already demonstrated excellent performance in multiple
applications including OPV, OLED and perovskite PV of various geometries and
architectures.202–204 Most of the nanowire meshes reported in literature are applied by
spray coating, spin coating or slot die coating followed by an additional structuring
process. Inkjet printing of AgNW has faced multiple challenges and authors often
describe the high aspect ratio of the AgNWs to be responsible for poor printability as
jetting nozzles become blocked.141,205 Some authors attempted to address these challenges
by blending the formulation with silver nitrate solution or by AgNW scission.205,206
However, the applicability of these methods is limited as they introduce problems like
high processing temperature and reduced conductivity. Very few works, such as Lu et al.,
have demonstrated inkjet printed AgNW networks from pristine ink for OPV so far.141
Their reports reveal several limitations of printed AgNW top electrodes, like poor mesh

145
From Layers to Cells

uniformity and a relatively low conductivity compared to other coating methods. In this
work, for the first time inkjet printed AgNW mesh electrodes of comparable quality to
those manufactured by conventional techniques are reported. This is achieved by printing
AgNW layers from a novel alcohol based ink formulation. The applied ink, as described
in chapter 4.1.1 of this work and reference 164, is a modified version of the developmental
AgNW dispersion ClearOhm Y Series from Cambrios, diluted with 1-pentanol in a
volume ratio of 1:2. Due to the high boiling point alcohol a stable printing process and
nozzle open times in the range of minutes are achieved. The length of the printed AgNWs
as determined from light- and scanning electron microscopy (Figure V-18 a), b)) is in the
range specified by the manufacturer (~30 µm). Obviously, the jetting process does not
induce fracture of the nanowires, probably due to their flexibility or flow induced
alignment.205 This proves that nanowires with similar dimensions as the jetting nozzles
(35 µm) can be inkjet printed.

Figure V-18: a) Optical micrograph and b) SEM image of inkjet printed AgNW layer
164
(500 DPI, four passes). Published in reference and reproduced with permission from
Elsevier B.V.. The SEM image was recorded by K.C. Tam and is reproduced in this work
with permission. c) Optical micrograph of a liquid 10 µl AgNW ink droplet on a
microscope slide directly after deposition. d) Droplet during drying 1 min after
deposition. e) Inhomogeneous dry AgNW layer resulting from droplet drying.

146
From Layers to Cells

With the alcohol based AgNW inkjet formulation, layers were printed on glass and
optimised to reach sheet resistances of ~20 Ω/sq. These values were achieved by
repeating the printing and layer drying process four times at a resolution of 500 DPI. The
necessity to print multiple passes instead of a single pass with a higher resolution is
elaborated in Figure V-18. AgNW clustering occurs for too thick wet films independent
of the deposition method. In-situ microscope imaging of a 10 µl AgNW ink droplet on a
microscopy slide during drying (Figure V-18 c), d)) reveals a distinct convective flow
pattern inside the droplet with the appearance of Rayleigh-Bénard convection cells. It is
clearly visible that AgNWs are carried by the liquid convection and entangle with each
other, thereby forming clusters. This process is highly expressed for wet films which are
thicker than the AgNW length (e.g Resolution: 1000 DPI, Drop size: 26 pl, wet film
thickness: 40 µm). For wet films with smaller thickness than the AgNW length
(Resolution: 500 DPI, Drop size: 26 pl, wet film thickness: 10 µm) very little convective
transport of wires is observed during drying, resulting in a more homogeneous AgNW
distribution.
Besides the local AgNW density, also the network formation was investigated by SEM
image processing (Figure V-19).

Figure V-19: Percolation analysis of inkjet printed AgNW meshes by image processing of
stitched SEM pictures. a) Single layer (500 DPI - largest connected region: 99.4 %). b)
Two layers printed on top of each other (500 DPI - largest connected region: 99.4 %). c)
Four layers printed on top of each other (500 DPI - largest connected region: 98.7 %).
Connected AgNWs forming a network are marked in red colour. SEM images and
analysis were provided by Manuela Göbelt from the Max Planck Institute for the Science
of Light and is gratefully acknowledged.

147
From Layers to Cells

Even after a single printing pass with 500 DPI, 99.4 % of the AgNW covered area is
connected in a single percolation network. With four printing passes, according to image
processing software, 98.67 % of the AgNW covered area is connected. The slightly lower
value is attributed to difficulties in threshold setting for thick AgNW layers, which results
in lower analysis accuracy. Nevertheless, the obtained values are highly comparable to
19
blade coated AgNW layers reported in reference . The high number of AgNWs
connected to each other allows the conclusion that a bulk regime description of the T vs.
Rsq characteristics, as described in see chapter 2.1.3.2, can be applied.
The optical properties of the AgNW electrodes were determined by UV-Vis transmission
measurements (Figure V-20).

Figure V-20: a) Total transmittance of inkjet printed AgNW layers (1-4 printing passes)
corrected for the absorption of the glass substrate and haze of printed AgNW layer (4
printing passes). b) Specular transmittance at 550 nm vs. sheet resistance of inkjet
printed AgNW layers (1-4 printing passes) corrected for the absorption of the glass
substrate. The solid line is a fit curve according to equation 12. Published in reference 164
and reproduced with permission from Elsevier B.V..

According to ISO 9005:2003, which takes the spectral sensitivity of the human eye into
account, the transmittance of the AgNW electrode (corrected for the absorption of the
glass substrate) after four passes of the printhead is ~94 %. The haze of the AgNW
electrode was calculated by dividing the diffuse transmission of the sample by the total
transmission. Considering the sensitivity of the human eye, a mean value of 2.3 % is
obtained. This very low number enables a clear view through semitransparent devices,

148
From Layers to Cells

such as OPV or OLEDs. The correlation between sheet resistance and optical
transmittance for inkjet printed AgNW layers (Figure V-20 b)) shows excellent
agreement with those reported in literature by other coating methods.63 The obtained
values are in the same range or even better than other frequently used semitransparent
electrode systems, like ITO.207 Fitting the transmittance vs. sheet resistance curve
according to equation 12 (bulk regime - see chapter 2.1.3.2), a conductivity ratio
𝜎𝐷𝐶,𝐵 /𝜎𝑂𝑃 of ~250 is obtained. This FoM, which characterises the quality of
semitransparent electrodes, is comparable to literature reports of spin- or spray coated
AgNW mesh electrodes (typically 100 to 500), highlighting the excellent quality of the
inkjet printed layers.63,70
To demonstrate the applicability of the inkjet printed AgNW meshes as base and top
electrodes, fully inkjet printed solar cells were manufactured. The device with an active
area of 1 cm² was fabricated by consecutively printing the following layers:

 Contact pads from Ag nanoparticle ink I40TM-106, ~0.5 Ω/sq


 AgNW bottom electrode
 ZnO nanoparticle EEL
 Active layer, consisting of the bulk heterojunction (PV2000:PC70BM)
 PEDOT:PSS F HC Solar HEL
 AgNW top electrode

Figure V-21 shows the transmittance spectra and the cross-sectional SEM image of the
active area of the device. Notice that the bottom AgNW mesh appears to have blended
with the ZnO nanoparticles after deposition. The brighter and bigger nanostructures in
this layer appear to be the AgNWs as they have higher conductivity and larger average
dimensions than the ZnO nanoparticles used in this work. A detailed description of the
mechanisms behind this mixing effect and the influence on the device performance are
currently under investigation by K.C. Tam and will be part of a separate publication.
Other than that there are no signs of mixing of individual layers. The cross section reveals
that the novel alcohol based AgNW ink can be printed on top of PEDOT:PSS layers even
without the addition of cross-linking agents.

149
From Layers to Cells

Figure V-21: a) Specular transmittance of the encapsulated device. b) Photograph of


semitransparent fully inkjet printed cell. c) Cross-sectional SEM image of the whole
device structure. The thickness of each individual layer is included in the image. The
nanoparticle flakes on the active and glass layer are due to the glass cutting process. The
SEM image was recorded by K.C. Tam and is reproduced in this work with permission.
164
The images are published in reference and reproduced with permission from Elsevier
B.V..

Figure V-22: Electrical characteristics of the fully inkjet printed OPV cell. a) Current
density vs. voltage characteristics of the cell during illumination by a sun simulator
through a mask of 1 cm² area. Inset: J-V characteristics measured in the dark. b) EQE of
the device. The integrated response using a standard reference spectrum results in a 𝐽𝑆𝐶
of 10.7 mA/cm². Published in reference 164 and reproduced with permission from Elsevier
B.V..

150
From Layers to Cells

Figure V-22 shows the current-voltage characteristics of a fully inkjet printed solar cell
with AgNW electrodes. The photovoltaic key parameters are given in Table V-7. The
resulting 𝑃𝐶𝐸 of 4.3 % slightly exceeds the highest values reported so far in literature for
fully inkjet printed organic solar cells of similar structure and dimension, which used
silver grid electrodes to achieve semitransparency.143

Table V-7: Photovoltaic parameters of the fully inkjet printed organic solar cell with
164
AgNW electrodes. Published in reference and reproduced with permission from
Elsevier B.V..

𝑃𝐶𝐸 (%) 𝐹𝐹 (%) 𝐽𝑆𝐶 (mA/cm2) 𝑉𝑂𝐶 (V)

4.3 52.8 10.7 0.76

Further improvement of efficiency seems possible and is part of an ongoing work at ZAE
Bayern. Nevertheless, the given results demonstrate a big potential of inkjet printed
AgNWs mesh electrodes, with the unique optical appearance of the devices as key
advantage (Figure V-22 b) - no visually obstructive grids, low haze). This aspect
combined with the high performance and the freedom of shape offered by inkjet
technology enables the OPV technology to enter highly customised markets, such as
facade integration or automotive applications.

5.4. Conclusion
This chapter of the thesis described inkjet printing of OPV layers and their application in
small scale solar cells. Hereby the focus was set on the critical aspects (v) layer
functionality and (vi) high efficiency as defined in chapter 1.5. A summary of the
achieved results is presented in Table V-8.
Starting with a well working doctor bladed reference device, the fabrication of fully inkjet
printed solar cells by stepwise replacement of the single layers with inkjet printed
equivalents was demonstrated.
Firstly, inkjet printed P3HT:PC60BM and PV2000:PC70BM active layers were introduced.
It was found out that elevated substrate temperatures of 60 °C (similar to the blade
coating reference process) do not result in formation of homogeneous films, because the
single ink droplets do not have sufficient time to merge before drying (printing speed 100

151
From Layers to Cells

mm/s, QF = 6). UV-Vis absorbance measurements, AFM imaging and J-V curve fitting
lead to the conclusion that P3HT:PC60BM layers manufactured in such manner suffer
from low crystallinity and poor charge transport properties, resulting in decreased device
efficiency.

Table V-8: Overview of inkjet printed OPV layers and rating based on the critical
requirements for a robust inkjet printing process.

Material Layer functionality High efficiency

P3HT:PC60BM  
from CB+BrAni  Inkjet printing at RT and 3.3 %
from o-DCB subsequent drying yields 2.7 %
from o-X:THN  similar performance as 2.9 %
blade coated reference

PV2000:PC70BM  Similar performance as 7.3 %


from o-X:THN blade coated reference

PEDOT:PSS 
Wetting enforced by 3.2 % (P3HT:PC60BM)
from P Jet N V2 pinning centres does not
lower 𝑃𝐶𝐸

PEDOT:PSS 
Similar performance as 3.4 % (P3HT:PC60BM)
from F HC Solar blade coated reference

ZnO NPs 
Good edge definition 3.2 % (P3HT:PC60BM)
from N-10 compared to PG, EG inks
(1-pentanol) and similar performance as 
blade coated reference

AgNW mesh 
Homogeneous mesh 4.3 % (Fully inkjet
electrode 93 % T, 18 Ω/sq printed, base and top
from ClearOhm electrode AgNW,
(1-pentanol) PV2000:PC70BM) 

Ag grid electrode 
Ag: 0.22 Ω/sq, minimised 5.7 % (base electrode,
from I50DM106 optical and electrical PV2000:PC70BM)
losses

To improve this situation, inkjet printing was performed at RT and layers were dried
subsequently. This process mimics a typical R2R manufacturing scenario, in which the

152
From Layers to Cells

web successively passes printing and drying stations. With this new strategy, a loss-free
transition from blade coating to inkjet printing was demonstrated. Hero cells with inkjet
printed PV2000:PC70BM layers reach record efficiencies of 7.3 %, thereby slightly
exceeding literature reports of comparable inkjet printed OPV.143 Statistical evaluation
leads to the conclusion that losses related to the printing procedure of the active layer are
marginal compared to those attributed to changes in the ink formulation.
Loss-free transition from blade coating to inkjet printing was also demonstrated for the
ZnO EEL and PEDOT:PSS HEL. Inkjet application of the newly developed 1-pentanol
based ZnO nanoparticle ink resulted in layers with excellent edge definition and device
efficiencies similar to those manufactured by blade coating of pristine IPA based N-10. A
comparison of solar cells with PEDOT:PSS layers applied by doctor blade or inkjet leads
to the conclusion that the printing procedure does not negatively influence the HEL
functionality and resulting device performance. This observation even holds when exotic
methods like the newly introduced pinning centre approach are applied.
In order to replace the sputtered semitransparent ITO layer of the reference device inkjet
printed silver grids or AgNW mesh electrodes were introduced. By applying inkjet
printed silver line grids with optimised dimension as base electrodes in PV2000:PC70BM
solar cells, record device efficiencies higher than those reported in literature (see
summary Table II-5) of 5.66 % were achieved. The increased performance is attributed to
the high efficiency of the reference device. Power losses of ~15 % when switching from
ITO to printed silver grids are comparable to state of the art literature (see summary Table
II-5). An FEM simulation leads to the conclusion that losses are mainly caused by
shading of the silver finger (54.1 %) and parasitic absorption within the PEDOT:PSS
layer (36.0 %). According to FEM predictions further improvement is possible by
decreasing the silver grid width.156 This can be achieved with inkjet printheads that offer
smaller droplet diameters. However, a limiting factor for application of this electrode type
in OPV devices is the visibility of the silver lines.
As alternative, AgNW mesh electrodes are inkjet printed. This is possible by applying a
newly developed alcohol based inkjet ink. Characterisation of printed AgNW meshes
shows a strong correlation between the thickness of the applied wet film and mesh
uniformity after drying. In-situ microscopy of a drying AgNW droplet reveals that
AgNWs are transported by liquid convection and entangle with each other, thereby
forming clusters. To avoid this effect, printing parameters were adjusted to result in wet
film thicknesses smaller than the AgNW length (Resolution: 500 DPI, Drop size: 26 pl,

153
From Layers to Cells

wet film thickness: 10 µm). The resulting excellent AgNW mesh uniformity leads to the
conclusion that restriction of convective flow of AgNWs in the liquid film is a key aspect
for controlling the AgNW distribution, and thus electrical and optical properties of the
layer. AgNW meshes inkjet printed with multiple passes have an excellent balance
between conductivity and transmittance of > 90 % T @ < 20 Ω/sq, which is comparable
to layers fabricated by conventional methods, such as slot die or spray coating. A
percolation analysis reveals that even after a single printing pass (500 DPI) > 99 % of the
AgNW covered area is connected in a single network. This leads to the conclusion that a
bulk regime description of the transmittance vs. sheet resistance curve according to
reference 63 is a suitable assumption (see chapter 2.1.3.2). A conductivity ratio 𝜎𝐷𝐶,𝐵 /𝜎𝑂𝑃
of ~250, which is comparable to literature reports of spin- or spray coated AgNW mesh
electrodes, is found, highlighting the excellent quality of the printed layers. Diffuse
transmission measurements also show that inkjet printed AgNW meshes have a very low
haze of only 2.3 %. This allows a clear view through semitransparent devices, such as
OLEDs or solar cells, in which they can be used. Compared to silver grids, inkjet printed
silver nanowires do not introduce visually obstructive patterns into the solar cells. This
enables the OPV technology to meet the high standards of visually demanding
applications like facade integration (BIPV) or automotive applications. In order to
demonstrate the potential of inkjet printed AgNW electrodes in the field of printed
electronics, they were applied as cathode and anode in a fully inkjet printed solar cell
stack. The resulting 𝑃𝐶𝐸 of 4.3 % is the highest efficiency for fully inkjet printed OPV
cells reported so far.

154
CHAPTER VI: From Cells to Modules

Abstract:

In this final results chapter, the transition from small area inkjet printed organic solar cells
to large area modules is demonstrated. This is done by monolithic series connection of
single cells. First, inkjet printed solar modules with conventional interconnection are
presented, reaching 𝑃𝐶𝐸s higher than any literature value of comparable devices.
However, these modules still show visually obstructive stripes resulting from the P2 gap
that opens the active layer for contact formation between adjacent cells. To resolve this
issue, inkjet printed silver ‘bridges’ that can replace the P2 gap are introduced. This novel
strategy allows fabrication of almost invisible interconnection regimes. The new
technology is used for first time demonstration of large area inkjet printed OPV modules
of arbitrary shape and unique appearance. Finally, the transition to a high throughput
single pass printing machine is presented. Fully inkjet printed solar modules, consisting of
only three layers, with efficiencies of op to 4.67 % on an active area of 10 cm² are
demonstrated.

Part of this chapter has been published in:


P. Maisch, K.C. Tam, P. Schilinsky, H.-J. Egelhaaf, C.J. Brabec, Shy Organic Photovoltaics: Digitally
Printed Organic Solar Modules with Hidden Interconnects, Sol. RRL 2, 1-9 (2018). doi
10.1002/solr.201800005, reproduced with permission from John Wiley and Sons.
Only parts of the publication text authored by P. Maisch are reproduced in subchapter 6.1.2 of this work.
Contributors and specific contributions in form of SEM images to this chapter are clearly stated in the
figure captions. Additionally, permission to utilise the whole content of the publication as part of this thesis
was given from all co-authors involved through author ownership declarations.

155
From Cells to Modules

6.1. Different interconnection strategies


Most solar applications require higher power and higher voltages than those provided by
a single solar cell, and thus a transition from the small laboratory scale cells to large
areas. As described in chapter 2.1.4, it is not sufficient to just increase the size of a single
solar cell because the generated current, the electrode resistance and consequently also
efficiency losses scale with the dimension. To prevent this from happening, larger areas
are typically sliced into several smaller solar cells which are monolithically connected in
series. This leads to the summation of the voltages of the single cells, while the current
and the electrode resistance stay at a low level. In the following, different strategies to
achieve such interconnection by inkjet printing are presented. For comparison, a literature
overview of solar modules fabricated by conventional stripe coating or LASER patterning
is provided in Table II-3 of the theoretical background section.

6.1.1. Conventional interconnection


Using inkjet technology, it is possible to fabricate organic solar modules with
conventional cell interconnection. This is achieved by leaving gaps in the base electrode
(P1), the active- and charge extraction layers (P2) and in the top electrode (P3). In
literature, only few approaches to fabricate OPV modules by inkjet have been
reported.83,142 In these works, only selected layers have been inkjet printed while the
remaining layers of the module were created by other deposition methods, such as rotary
screen printing, slot die coating and sputtering. Furthermore, reported efficiencies do not
exceed 2 % and 𝐺𝐹𝐹s are limited to < 80 %. In the following, solar modules with
conventional cell interconnection are demonstrated with inkjet printed active layer, EEL,
HEL and top electrode which surpass the state of the art 𝑃𝐶𝐸 and 𝐺𝐹𝐹 values. These
modules were created by applying the inks developed in Chapter IV and the processing
parameters as described in Chapter V. Figure VI-1 shows the layout and performance of
an inkjet printed solar module with the layer sequence
ITO/ZnO/P3HT:PC60BM/PEDOT:PSS/Ag-grid. The module has an active area of 1.5 cm²
with a 𝐺𝐹𝐹 of 80 % and allows contacting of each individual cell. Except ITO, all layers
are manufactured by consecutive inkjet printing. For printing of the grid top electrode, the
silver nanoparticle ink EMD 5800 was selected because layers manufactured from this
dispersion achieve sheet resistances of 2.2 ± 0.1 Ω/sq (measured with 4-point probe)

156
From Cells to Modules

already at moderate curing conditions (Hotplate 140 °C, 10 min) when printed on top of a
PEDOT:PSS surface. Higher curing temperatures or longer curing times can be applied to
further decrease the electrode sheet resistance, but this might cause damage to the
underlying solar cell.

Figure VI-1: Layout and J-V characteristics of inkjet printed P3HT:PC60BM solar
module with silver grid top electrode and single cells. a) Layout of patterned ITO base
electrode. b) Printing pattern of ZnO and P3HT:PC60BM. c) Printing pattern of HC
PEDOT:PSS F HC Solar. d) Printing pattern of silver grid top electrode and contact
pads. e) J-V characteristics of module and single cells. f) Photograph of the device.

The printed P3HT:PC60BM solar modules reach efficiencies of up to 2.07 % with a 𝐹𝐹 of


50 %. To unveil the influence of the silver finger sheet resistance on the solar cell
performance, FEM simulations of silver grid unit cells as described in chapter 3.4.2 were
conducted. According to the simulation, the high sheet resistance of the printed silver of
2.2 Ω/sq causes a significant drop in 𝐹𝐹 (6 %) and 𝑃𝐶𝐸 (0.2 %), compared to silver grids

157
From Cells to Modules

with a lower sheet resistance of 0.2 Ω/sq (see chapter 5.3.1 - silver grid base electrodes
allow harsher curing conditions than top electrodes resulting in higher conductivities).

Table VI-1: Device performance parameters of the solar cells and module shown in
Figure VI-1, NOTE: 𝑃𝐶𝐸 and 𝐽𝑆𝐶 are calculated from active device area

Sample 𝐴𝑎𝑐𝑡𝑖𝑣𝑒 𝑃𝐶𝐸 𝐹𝐹 𝐽𝑆𝐶 per cell 𝑉𝑂𝐶


(cm²) (%) (%) (mA/cm²) (V)

Single Cell 0.5 2.00 49.84 7.32 0.55

FEM Ag 2.2 Ω/sq 0.075 2.21 55.26 7.28 0.55

FEM Ag 0.2 Ω/sq 0.075 2.43 61.03 7.24 0.55

Module 1.5 2.07 50.04 7.55 1.64

However, experimentally achieved device efficiencies are lower than expected from
simulation. This leads to the conclusion that further loss mechanisms, which are harder to
assess, such as diffusion of solvent or nanoparticles into the cell, degrade the device
performance. Clearly, there is still room for improvement of the printed top electrode.
Nevertheless, the module efficiencies of up to 2.07 % achieved in this work are
significantly higher than most literature reports of partly inkjet printed OPV modules
described in Table II-5. Furthermore, it is the first time that four layers of an OPV module
were realised by inkjet.
Besides the inkjet printed silver grids, also AgNW mesh electrodes were applied in OPV
modules. Figure VI-2 shows the layout of such a device with the layer sequence
ITO/ZnO/PV2000:PC70BM/PEDOT:PSS/AgNWs. AgNW mesh electrodes with a sheet
resistance of < 20 Ω/sq were printed as described in chapter 5.3.2. Except the ITO base
electrode all layers of the module were fabricated by digital printing. The high precision
of the drop placement allows creating narrow P2 and P3 interconnection lines. Therefore,
printed solar modules with a 𝐺𝐹𝐹 of 85 % could be realised. This value is significantly
higher than in modules fabricated by other wet layer patterning methods (see literature
summary Table II-3).

158
From Cells to Modules

Figure VI-2: Layout and J-V characteristics of inkjet printed PV2000:PC70BM solar
module with AgNW top electrode. a) Layout of the device consisting of a patterned ITO
base electrode (white), ZnO/P3HT:PC60BM (red) and PEDOT:PSS/AgNWs layers
(black). b) Photograph of the device. c) J-V characteristics of module and single cells.

Table VI-2: Device performance parameters of the solar cells and module shown in
Figure VI-2. NOTE: 𝑃𝐶𝐸 and 𝐽𝑆𝐶 are calculated from active device area.

Sample 𝐴𝑎𝑐𝑡𝑖𝑣𝑒 𝑃𝐶𝐸 𝐹𝐹 𝐽𝑆𝐶 per cell 𝑉𝑂𝐶


(cm²) (%) (%) (mA/cm²) (V)

Cell 1 0.85 4.64 49.49 11.86 0.79

Cell 2 0.85 4.37 49.09 11.32 0.79

Cell 3 0.85 4.35 49.51 11.17 0.79

Cell 4 0.85 3.70 42.97 11.21 0.77

Module 3.40 4.29 47.83 11.48 3.13

159
From Cells to Modules

The first demonstration of inkjet printing the high performance polymer active layer
PV2000:PC70BM in a module structure allows pushing the efficiency limit of inkjet
printed OPV modules to 4.3 %. Achievement of such high efficiency is possible by
combination of the optimised inks and printing procedures introduced in the previous
chapters and the application of the inkjet printed AgNWs developed in this work.
Furthermore, the inkjet printed AgNW mesh electrode allows for the first time the
fabrication of inkjet printed semitransparent solar modules with inconspicuous electrodes
(see photograph Figure VI-2 b)). The only remaining optical inhomogeneity in the
module structure is a transparent stripe resulting from the P2 line. A strategy how to
further improve the visual appearance of the module by avoiding the P2 gap in the active
layer is presented in the following chapter.

6.1.2. Shy interconnection


The following chapter is part of the publication ‘Shy Organic Photovoltaics: Digitally
Printed Organic Solar Modules with Hidden Interconnects’ and reproduced with
permission from John Wiley and Sons.159

Motivation

The biggest attractions of OPV technology are the semitransparency, the flexibility and
the wide variety of colours that can be achieved. Instead of focusing on grid connected
applications, OPV consistently should concentrate on markets fitting to these strengths
and the fantastic power-to-weight ratio.208 Integration of OPV into portable solutions is
certainly today’s strongest marketing concept. Such a tight integration with products
requires an optical appearance to match with the product of concern. Digital printing of
OPV is a promising approach to address this issue because its drop-on-demand process
allows arbitrary shape deposition of materials. In the previous chapters it was
demonstrated how to print single solar cells with unique appearance by using optically
homogeneous and low haze AgNW mesh electrodes (chapter 5.3.2). However, a major
remaining challenge is to print whole modules with optically inconspicuous Z-
interconnection lines without the use of highly complex post-structuring processes, such
as LASER patterning. The last hurdle on the path to such ‘shy’ OPV technology is a
technological solution to mask the printed interconnection region without compromising
the device performance. This is difficult to achieve with the traditional way of OPV

160
From Cells to Modules

modularisation processes, which typically results in parallel stripes with mm-sized gaps in
the interconnection area. Specifically the P2 step will unavoidably introduce visually
conspicuous stripes which greatly affects the optical appearance of OPV modules (see
chapter 2.1.4, Figure II-10 a)). As a result, clearly visual artefacts from cell-to-cell
interconnections are typically observed on supposedly visually attractive, free-form inkjet
printed OPV modules. A solution to this problem, which utilises a visually seamless
interconnection method to elegantly minimise the conspicuity of the interconnection, is
presented in the following.

Design and fabrication

Invisible interconnections are realised by inkjet printing of highly conductive silver


‘bridges’, which penetrate the solar cell stack, thus forming an electrical connection
between adjacent cells. Figure VI-3 a) and b) schematically compare the device layout for
the traditional ‘gap’ and the novel ‘bridge’ interconnection proposed in this work.

Figure VI-3: a) Schematic illustration of a conventional interconnection (Ic) region using


a ‘gap’ in the EEL and AL to realise the cell-to-cell contact. b) ‘Bridge’ arrangement
proposed in this work. c) Photograph of a ‘gap’ type solar module. d) Photograph of a
‘bridge’ type solar module. e) Schematic drawing showing the 3-cell OPV module layout.
Each solar cell can be contacted in a way that allows measurement from bottom to top
electrode (cell 1: contact 1-2) or through the interconnection region (cell 1: contact 1-3).
c)-e) Are published in reference 159 and reproduced with permission from John Wiley and
Sons.

161
From Cells to Modules

The device structure ITO/ZnO/P3HT:PC60BM/AgNW:PEDOT:PSS was chosen for


experimental implementation because it offers high reliability of the printing process. The
ZnO EEL and the P3HT:PC60BM active layer were inkjet printed as described in the
previous chapters. As top electrode, a hybrid material consisting of a blend of
PEDOT:PSS and AgNWs was applied by inkjet printing. K.C. Tam et al. introduced this
material showing its excellent printability and its good performance in OPV devices
similar to PEDOT:PSS and AgNWs bilayer structures.209 Figure VI-3 c) shows a
reference module printed with conventional interconnection. Figure VI-3 d) and e) show
the device layout and implementation of three-cell modules realised with a silver ‘bridge’
for contacting of adjacent cells. The reference layout is fabricated by printing the layers
on top of each other, leaving gaps at the P2 positions. In order to manufacture the ‘bridge’
layout, conductive silver nanoparticle ink (Sicrys I40DM-106) was first inkjet printed
directly on top of a prepatterned ITO substrate next to the P1 line and dried at 80 °C. This
printing step was repeated for five times to ensure that the silver ‘bridge’ is thick enough
(~5 µm) to penetrate all subsequently printed layers. The silver structures were thermally
annealed (150 °C, 10 min) to attain a sufficiently low resistivity of < 25 µΩ*cm. All
following layers except the top electrode are printed as continuous homogeneous wet film
without interruptions by P2 gaps. Protrusion of these layers by the silver structures is
shown in Figure VI-4 g). This fact together with the high conductivity of the silver
guarantees the formation of low resistance contacts between the bottom and the top
electrodes of adjacent module cells.
The advantage of the ‘bridge’ type interconnections is twofold. On the one hand, higher
visual aesthetics is achieved than a traditional ‘gap’ type interconnection could offer. This
is clearly demonstrated by Figure VI-3 c) and d). In the ‘gap’ type module layout, the
interconnection areas introduce a visual conspicuity that can be seen even from distance.
In contrast, the interconnection area is visually much less conspicuous for the ‘bridge’
type layout. On the other hand, modules with ‘bridge’ type interconnection offer higher
𝐺𝐹𝐹 values than such with ‘gap’ type interconnection. In the ‘gap’ layout each cell has
0.8 cm2 active area and 0.2 cm2 interconnection area, which results in a 𝐺𝐹𝐹 of 80 %.
This is the highest value that could be reproduced with good statistics for the inks and
processing conditions described. A better control of the wetting behaviour of the single
layers and inks would probably allow the 𝐺𝐹𝐹 to reach values above 80 %.

162
From Cells to Modules

Figure VI-4: Schematic illustration of the a) ‘lines’ interconnection and b) ‘dots’


interconnection. c) Dark field microscope image of the ‘lines’ interconnection and d)
‘dots’ interconnection. e) SEM image of the cross-section at the edge of an
interconnection dot. f) Higher magnification SEM image of the dot edge shows all solar
cell layers as printed on top of each other. g) Higher magnification SEM image closer to
the middle of the dot shows direct contact formation of ITO, silver and the
AgNW:PEDOT:PSS hybrid top electrode. The SEM pictures were recorded by K.C. Tam,
159
are published in reference and are reproduced with permission from John Wiley and
Sons and the authors of the paper.

For the ‘bridge’ interconnection, however, only the electrodes need to be patterned while
all other layers overlap the interconnection area. As a result, it is much easier to control
the quality and shape of the printed layers because the respective surfaces are

163
From Cells to Modules

homogeneous and no alignments are required. With this design, each cell has an active
area of 0.85 cm2 and an interconnection area of 0.15 cm2, which corresponds to a 𝐺𝐹𝐹 of
85 %. Even higher 𝐺𝐹𝐹s might be possible with a smaller printhead nozzle diameter.
Two different ‘bridge’ interconnection patterns were realised to optimise not only
interconnection resistance, but also visual appearance: Line-interconnections (‘lines’) and
dot-interconnections (‘dots’). Figure VI-4 a)-d) shows the layout and dark field
microscope images of both patterns. Here, the silver dots have a diameter of 100 µm and
were printed 0.25 mm apart from each other and. The effect of these arrangements will be
discussed in more detail in the subsequent sections.

Performance of shy modules

Figure VI-5 and Table VI-3 show typical J-V-characteristics and performance parameters
of OPV modules fabricated with all three types of interconnections. Twelve devices were
fabricated with four devices for each type of interconnection.

Figure VI-5: Typical J-V characteristics of inkjet printed OPV modules with different
159
types of interconnections. Published in reference and reproduced with permission
from John Wiley and Sons.

164
From Cells to Modules

Table VI-3: Device performance parameters of the modules shown in Figure VI-5.
159
NOTE: 𝑃𝐶𝐸 and 𝐽𝑆𝐶 are calculated from active device area. Published in reference
and reproduced with permission from John Wiley and Sons.

Module area 𝐴𝑎𝑐𝑡𝑖𝑣𝑒 Type 𝑃𝐶𝐸 𝐹𝐹 𝐽𝑆𝐶 per cell 𝑉𝑂𝐶


(cm2) (cm2) (%) (%) (mA/cm2) (V)

3 2.55 dots 2.73 54.1 9.21 1.64

3 2.55 line 2.76 57.6 8.94 1.61

3 2.4 gap 2.59 54.4 8.64 1.65

The module performance for all devices of each interconnection type varies by less than
5 %, indicating excellent reproducibility for the inkjet printing process. The achieved
𝑃𝐶𝐸 of more than 2.5 % for all devices is reasonable for P3HT:PC60BM modules of the
given size and at the high end of the literature overview provided in Table II-3. The high
forward bias injection of the J-V curve shown in Figure VI-5 is the first evidence that the
silver ‘bridge’ type interconnects do not introduce additional resistive losses.
Characterisation of the interconnection area using Cross Kelvin Bridge Resistor
measurements (CKBR) was done by K.C. Tam.210 A detailed description of the
measurement technique and results is provided in the shared publication.159 K.C. Tam
determined a specific interconnection resistance of ~0.002 Ω*cm2 for the silver ‘bridge’
structure. This means that the interconnection resistance of a single 0.1 mm dot used in
the ‘dots’ type OPV module is ~22 Ω, while 40 of them in parallel (1 cm length with
0.25 mm pitch) would give ~0.56 Ω. For a 1 cm ‘line’ type interconnection the resistance
is ~0.18 Ω.

Electrical simulations

The impact of the interconnection resistance on the solar module performance was
elucidated by a 2D FEM simulation. The model mimics the experimentally implemented
module layout precisely and is described in detail in chapter 3.4.2. The J-V curves in
Figure VI-6 are simulated for the dots interconnection geometry, varying the inserted
interconnect resistance value. Simulated modules with 𝜌𝐼𝐶 = 0.002 Ω*cm² (obtained from
CKBR measurements) show almost the same J-V characteristics as obtained from solar
simulator measurements of the inkjet printed devices (see Figure VI-6).

165
From Cells to Modules

Figure VI-6: Experimental (triangles) and simulated (lines) J-V curves of a solar module
with dotted cell-to-cell interconnection. Various specific interconnection resistances are
simulated ranging from 0.0002 to 0.1 Ω*cm2. The simulated curves with
𝜌𝐼𝐶 = 0.0020 Ω*cm2 (obtained from CKBR measurements), show excellent agreement
with the measured module J-V characteristics. Published in reference 159 and reproduced
with permission from John Wiley and Sons.

The simulated 𝐹𝐹 and consequently the 𝑃𝐶𝐸 values in Table VI-4 are slightly higher than
experimentally observed (Table VI-3). Such small differences may come from
occasionally non-penetrating P2 dots or from a distribution of resistance values of the
interconnecting dots. Most importantly, the simulation correctly reproduces the
performance behaviour of modules with ‘line’ and ‘dots’ type interconnection and
confirms that the higher total interconnection resistance of the ‘dots’ architecture is
indeed the origin of the lower module 𝐹𝐹 and 𝑃𝐶𝐸 (Figure VI-5). The simulation model
also allows to take a deeper look into the in the interconnection area. Loss processes like
current crowding, inhomogeneous current distributions around the contact areas or the
voltage drop across interconnection dots, are visualised by the 2D FEM simulation. The
latter loss mechanism is of special interest.

166
From Cells to Modules

Table VI-4: Simulated device performance parameters of OPV modules with different 𝜌𝐼𝐶 .
The resulting values for 𝜌𝐼𝐶 = 0.0020 Ω*cm2 (obtained from CKBR measurements) are
159
highlighted. Published in reference and reproduced with permission from John Wiley
and Sons.

Type 𝜌𝐼𝐶 𝑃𝐶𝐸 𝐹𝐹 𝐽𝑆𝐶 per cell 𝑉𝑂𝐶


(Ω*cm2) (%) (%) (mA/cm2) (V)

line 0.0020 2.94 60.6 8.82 1.65

dot 0.0002 2.87 58.9 8.82 1.66

dot 0.0020 2.84 58.1 8.82 1.66

dot 0.0200 2.43 50.0 8.78 1.66

dot 0.0500 2.17 44.9 8.70 1.66

dot 0.1000 1.60 34.0 8.53 1.66

Figure VI-7 zooms into the simulated potential distribution at the interconnection area of
a module held at the maximum power point (MPP - 1.2 V being applied over the whole
module). The overall potential drop is equally split among the three cells. Simulation
reveals that the voltage losses across the ‘dots’ interconnection, i.e. from ITO on the one
side to the AgNW:PEDOT:PSS on the other side, is only 0.0039 V. Such a low value
suggests that the actual resistance losses in the interconnection region are negligible.
Another important mechanism supporting the observation of a well working
interconnection region is the lateral current flow (shown by the arrows in Figure VI-7).
The current spreading from the interconnection dots into the ITO electrode can parallelise
before reaching the P3. The low sheet resistance of both electrodes is a key parameter for
this effective charge transport.
Ultimately, the simulation model offers the unique opportunity to analyse the impact of
interconnection resistance on OPV module performance with ‘bridge’ type
interconnections. Such analysis would be otherwise experimentally challenging as it does
require the production of modules with different interconnection resistances. It confirms
that the interconnection resistance achieved by experiment is already at optimum, as
lower values do not increase the 𝐹𝐹 by more than 1 % (see Figure VI-6, Table VI-4).

167
From Cells to Modules

Figure VI-7: Simulated potential distribution (colour bar) and lateral current flow
(arrows) at the ‘dot’ type interconnection area when the module is operated at the
maximum power point (𝜌𝐼𝐶 = 0.0020 Ω*cm2). Different colour scales for the potential
drop in bottom and top electrode are necessary for visualisation, due to their different
resistance (bottom 15 Ω/sq, top: 10 Ω/sq). Published in reference 159 and reproduced with
permission from John Wiley and Sons.

The simulations prove that a specific interconnection resistance of 0.02 Ω*cm2 already
significantly affects the shape of the J-V curve, thus reducing the 𝐹𝐹 from 58.1 % to
50.0 %. At even higher values, i.e. 𝜌𝐼𝐶 = 0.1 Ω*cm2, the 𝐹𝐹 drops to less than 34.0 % and
the 𝑃𝐶𝐸 is reduced by half. Table VI-4 is very helpful in summarising the interconnection
resistances required to achieve minimum loss ‘dots’ type interconnections. Before
applying the ‘invisible’ interconnections in large area arbitrarily shaped inkjet printed
modules in the next chapter, it should be highlighted that the CKBR structure in
combination with 2D FEM simulation works very effectively in evaluating thin film PV
technologies with respect to their interconnection losses.

168
From Cells to Modules

6.2. Arbitrary shape modules


As described in chapter 1.3, superior aesthetics, freedom of form and scalability are some
of the USPs which make OPV technology highly interesting for integration into portable
electronic devices and BIPV applications. With the introduction of inkjet printable
AgNW meshes as visually inconspicuous electrode and the demonstration of ‘shy’
‘bridge’ type interconnections, the necessary technological background for realising
devices which can meet the high standards of such applications was provided in the
previous chapters. The scalability of these concepts and freedom of form is demonstrated
in the following by realisation of inkjet printed large area arbitrary shape OPV modules.
One demonstrator fabricated in the scope of this work is an OPV module portrait of
Professor Alan J. Heeger, who was awarded with the Nobel Prize in Chemistry (2000)
together with Alan G. MacDiarmid und Hideki Shirakawa for discovering conductive
polymers. The active area of this device is 84 cm2. It consists of the five layers
ITO/ZnO/P3HT:PC60BM/PEDOT:PSS/AgNW. All layers except the ITO bottom
electrode are manufactured by consecutive inkjet printing. The P2 interconnection lines
are realised by printing visually inconspicuous silver ‘bridges’. P1 and P3 lines were
patterned by LASER with 0.26 mm separation.
Figure VI-8 shows the photo used for the design of the OPV module as well as the
processing procedure which comprises the following steps:

 First, a binary image is extracted from the photograph by thresholding method.


 The area of the picture to be provided with the solar module is split by means of a
software program into ten equally sized solar cells which can be connected in
series. This is done by using a slicing software tool as described in 3.4.
 P1 and P2 lines are introduced to the ITO substrate.
 Layers of ZnO nanoparticles and active material P3HT:PC60BM are inkjet printed
fully covering the image contour (red area Figure VI-8 b)).
 Layers of PEDOT:PSS and AgNWs are printed in a contour pattern which is
scaled down by 5 % (blue area Figure VI-8 b)). This is necessary to avoid
overlapping of bottom and top electrodes, and thus shunting.
 The module is finished by introduction of the P3 lines, annealing and
encapsulation (additionally contours can be printed on the module or any covering
substrate to change the visual appearance).

169
From Cells to Modules

211
Figure VI-8: a) Photo of Alan Heeger used for the design. Obtained from with
permission. b) Printing pattern for ZnO and P3HT:PC60BM layer in red colour,
PEDOT:PSS and AgNW layer in blue colour. The white background is the ITO substrate.
Black lines indicate the interconnection region. c) Additional AgNP layer printed on the
encapsulation glass to change visual appearance. d) The completed OPV module portrait
159
of Prof. Alan J. Heeger. Published in reference and reproduced with permission of
John Wiley and Sons.

Thanks to the silver ‘bridge’ technology the interconnection areas are hardly visible at all.
The device performance parameters of the demonstrator are summarised in Table VI-5.
The values presented are the highest achieved for inkjet printed OPV modules of similar
size. Moreover, this is the first report on an inkjet printed OPV module with irregular
shape. Despite the encouraging performance, further studies need to address the few loss
mechanisms inherent to this module technology. As the lengths of individual cells are not
the same, the 𝑅𝑆 of each cell will be different. Furthermore, as the sheet resistance of the
top electrode is relatively high (~15 Ω/sq), the 𝑅𝑆 losses for the first and last cell will be
the highest, considering their relatively large cell lengths. Finally, the P3HT:PC60BM is
intentionally printed slightly thinner than optimum to enhance the visual impression, thus
achieving better optical transparency and improving the aesthetics of the module.
Nevertheless, the achieved module efficiency is more than sufficient for low power
applications, such as driving a digital clock. The combination of a radio clock and the
portrait module was realised in form of a demonstrator unit, which was handed over to
Prof. A.-J. Heeger as a present for his 80th birthday (Figure VI-9).

170
From Cells to Modules

Figure VI-9: J-V characteristics of inkjet printed 84 cm² OPV portrait module. The inset
shows the integration of the module into a radio clock demonstrator where it provides the
necessary power for operation under room light (stray light – no direct illumination).
Published in reference 159 and reproduced with permission of John Wiley and Sons.

Within the frame of an industrial cooperation with the large format inkjet machine
manufacturer Durst Phototechnik, several more arbitrarily shaped OPV modules were
printed. A selection of them is presented in Figure VI-10.

Table VI-5: Key device performance parameters of free-form OPV modules.


Demonstrator 𝐴𝑎𝑐𝑡𝑖𝑣𝑒 𝑃𝐶𝐸 FF 𝐽𝑆𝐶 per cell 𝑉𝑂𝐶
(cm²) (%) (%) (mA/cm²) (V)

Portrait A. Heeger 84 1.60 47.0 6.80 5.50

Module Durst ‘Crystal’ 98 1.4 41.7 6.0 4.0


building

Module Durst Logo 150 2.5 41.3 8.1 5.3

171
From Cells to Modules

Figure VI-10: Inkjet printed arbitrary shape OPV devices. a) P3HT:O-IDTBR* (same
printing parameters as for P3HT:PC60BM) OPV module in shape of the Durst
Phototechnik ‘Crystal’ building. b) Photograph of the building obtained from ck-
projekt.it. c) P3HT:O-IDTBR OPV module in shape of the Durst Phototechnik company
logo integrated into a digital clock. d) P3HT:PC60BM solar cell portrait of Prof. C.-J.
Brabec.
*(5Z,5'Z)-5,5'-((7,7'-(4,4,9,9-tetraoctyl-4,9-dihydro-s-indaceno[1,2-b:5,6-b']dithiophen-2,7-
diyl)bis(benzo[c][1,2,5]thiadiazole-7,4-diyl))bis(methanylylidene))bis(3-ethyl-2-thioxothiazolidin-4-one)

The versatility of design, colour and voltage of the realised demonstrators highlights the
big potential that the combination of digital printing and OPV technology has to offer. It
is the author’s strong believe that these key factors have to be focused on in order to
transfer OPV from the niche of lighthouse projects to a bigger consumer driven market.

6.3. Large scale production


In this last chapter of the modules section, insights about the transfer from laboratory to
industrial scale inkjet fabrication of OPV devices are presented.
To provide a background for the reader what ‘large scale’ means, the term should briefly
be elucidated. Most publications about OPV or solar cell research focus on new materials,

172
From Cells to Modules

exotic structures and record efficiencies. Experimental cells are typically realised on a
very small scale of only a few mm². This strategy keeps the resistive losses within the
electrodes as well as the probability for defects in the layers low and enables processing
by spin-coating. Regarding such reference, literature considers solar cells with a size of
more than 1 cm² often already as ‘large’. The arbitrarily shaped modules presented in the
last chapter, which were printed with a laboratory inkjet machine PiXDRO LP 50, can
consequently already be labelled as ‘very large’ (up to 150 cm²). However, to
demonstrate the possibility of inkjet printing OPV devices on industrial scale, this is still
not sufficient. A printing system is required that allows the deposition of even bigger
layers in a homogeneous, highly reproducible and very fast manner. In the following, this
is realised by transferring the inkjet processes described in the previous chapters to a high
throughput single pass machine which was set up in the framework of this thesis. A
detailed description of the machine, which is equipped with 4 printing stations containing
in total 16 SAMBA printheads, an ink circulation system and multiple drying stations, is
provided in chapter 3.2.3. In analogy to the previous chapters, the transfer from the
reference process to the single pass inkjet machine is conducted by stepwise replacement
of the single layers. The specifications of the printheads (chapter 3.2.2, 3.2.3) hereby
allow application of the same PEDOT:PSS, ZnO and AgNW ink formulations as
described in chapter IV.

Single pass EEL and HEL

Figure VI-11 a) and b) show photographs of the single pass inkjet printing station and a
continuous PEDOT:PSS film printed on top of a P3HT:PC60BM active layer. The layer is
deposited while the moving web passes the inkjet unit with approximately 1 mm distance
to the nozzle plate. With this setup, homogeneous layers of PEDOT:PSS and also ZnO
nanoparticles from 1-pentanol ink can be printed at high web speeds of up to 40 m/min.
This corresponds to throughput speeds of 6.88 m²/min when the print station is fully
equipped with four printheads (print width 17.2 cm). For the given setup, the speed is
limited by the accuracy of the web movement. In principle, the SAMBA technology
(firing frequency up to 100 kHz, resolution 1200 DPI) would allow even higher print
speeds of up to 127 m/min. The buffer layers printed with the single pass machine were
first applied in small scale reference solar cells with the layer sequence
ITO/ZnO/P3HT:PC60BM/PEDOT:PSS/Ag. Figure VI-11 c) and Figure VI-12 show a

173
From Cells to Modules

comparison of the devices with ZnO and PEDOT:PSS layers printed with the high
throughput loop coater setup and a doctor bladed reference. The resulting solar cell
characteristics show highly comparable behaviour.

Figure VI-11: a) Single pass prototype inkjet integrated in a quasi-R2R machine, which
runs a 7 m long PET substrate in an endless loop. b) Homogeneous PEDOT:PSS F HC
Solar layer printed on top of a P3HT:PC60BM surface with the single pass inkjet system.
The white arrows show the web movement direction. c) Representative J-V characteristics
of P3HT:PC60BM solar cells with ZnO (red) and PEDOT:PSS (F HC Solar) layers inkjet
printed with the single pass machine. The inset shows the J-V characteristics measured in
the dark.

Table VI-6: Device performance parameters of the solar cells shown in Figure VI-11.

Sample 𝑃𝐶𝐸 𝐹𝐹 𝐽𝑆𝐶 𝑉𝑂𝐶


(%) (%) (mA/cm²) (V)

Reference DB 3.17 61.5 9.66 0.53

ZnO IJ loop coater 3.26 61.2 9.83 0.54

PEDOT:PSS IJ loop coater 3.22 59.9 10.0 0.54

Furthermore, investigation of the printed layers by means of optical and confocal


microscopy shows no visible printing pattern and a layer thickness comparable to the

174
From Cells to Modules

reference (ZnO ~50 nm, PEDOT:PSS ~100 nm). This proves direct and loss-free
transferability of inkjet printed buffer layers to a high throughput industrial inkjet system.

Figure VI-12: Box plot 𝑃𝐶𝐸, 𝐹𝐹, 𝐽𝑆𝐶 and 𝑉𝑂𝐶 comparison of P3HT:PC60BM solar cells
with ZnO (N10 - red) and PEDOT:PSS (F HC Solar - blue) layers inkjet printed at the
single pass machine as well as doctor bladed reference (black).

Single pass active layer

Attempts to print the PV2000:PC70BM active layer from o-X:THN 1:1 (vol.) based ink
with the single pass inkjet machine resulted in printing defects and non-functional layers
(Figure VI-13 b)). It appears that the print pattern is caused by clogging of the SAMBA
printhead nozzles (nozzle diameter 10 µm). Interestingly, as demonstrated in chapter 5.1,
the same ink can be applied with the PiXDRO LP 50 using Spectra SE 128 AA printheads
(nozzle diameter 35 µm), forming homogeneous films. At the current stage, the reasons
for the nozzle failure of the SAMBA printheads as well as the question whether this effect
is related to the nozzle size remain unclear and are part of an ongoing investigation.
Further experiments reveal that homogeneous layers of PV2000:PC70BM can only be
printed with the single pass inkjet when decreasing the o-X amount to 0. Printing of the
active layer from pristine THN ink resulted in homogeneous films (Figure VI-13 a)),
which yield a 𝑃𝐶𝐸 of ~5.2 % when applied in the solar cell architecture
(ITO/ZnO/PV2000:PC70BM/PEDOT:PSS/Ag).

175
From Cells to Modules

Figure VI-13: a) Printing process of homogeneous PV2000:PC70BM layers from THN ink
with the single pass inkjet. b) Dry non-functional PV2000:PC70BM layer showing
printing defects from nozzle malfunction. c) Representative J-V characteristics of solar
cells with PV2000:PC70BM active layers doctor bladed (black) and printed at the single
pass loop coater (red) from THN based ink. The inset shows the J-V characteristics
measured in the dark. d) Box plot 𝑃𝐶𝐸, FF, 𝐽𝑆𝐶 and 𝑉𝑂𝐶 comparison of PV2000:PC70BM
solar cells with doctor bladed (black) and inkjet printed (red) active layer from THN ink.
Each box contains data of at least five solar cells.

Table VI-7: Device performance parameters of the solar cells shown in Figure VI-13 c).
Sample 𝑃𝐶𝐸 𝐹𝐹 𝐽𝑆𝐶 𝑉𝑂𝐶
(%) (%) (mA/cm²) (V)

PV2000:PC70BM THN DB 5.91 60.20 12.46 0.79

PV2000:PC70BM THN IJ 5.08 58.80 10.85 0.80


loop coater

176
From Cells to Modules

The efficiency of cells with active layers manufactured from pure THN ink is slightly
lower compared to o-X:THN 1:1 (vol.) ink (~6.5 %, reported in chapter 5.1). This effect
could be related to a non-optimised BHJ microstructure caused by the high boiling nature
of the solvent in combination with the drying conditions (hot air oven) at the loop coater.
Similar effects were described by Hoth et al. for inkjet printed polymeric active layers
from THN ink.129 Strategies to further improve of the printing process and device
efficiency by changing to a small molecule donor material or nanoparticle active layer
inks as proposed by Sankaran et al.,145 are part of an ongoing investigation by co-workers.
Nevertheless, the above described results clarify that it is possible to manufacture highly
efficient organic solar cells entirely with the single pass high throughput inkjet machine.

Single pass cells and modules

The final missing step on the way to an industrially relevant printing process is to
demonstrate high throughput production of large area highly efficient and fully inkjet
printed solar cells and modules. To make the process as easy and robust as possible, a
device structure consisting of only three layers was applied. This was realised by using
hybrid electrodes consisting of AgNW meshes blended with ZnO nanoparticles or
PEDOT:PSS. These material combinations and their printability were developed and
investigated by K.C. Tam.209 Compared to typical solar modules consisting of five to six
layers, the three layer setup drastically reduces the manufacturing steps, thus increasing
reproducibility and processing yield. A provisional patent (US62/521,696) was filed
jointly with the AgNW supplier Cambrios Advanced Materials Co. to protect this idea.212
The devices were manufactured by consecutive inkjet printing of
AgNW:ZnO/PV2000:PC70BM/AgNW:PEDOT:PSS layers on a flexible PET substrate.
Both hybrid electrodes were inkjet printed using the single pass loop coater machine. For
printing of the AgNW blends, the on-head filters of the SAMBA printhead were removed.
The successful printing process proves that AgNWs with lengths of several 10s of
micrometers can even be inkjet printed with a nozzle size of only 10 µm. Printing was
repeated three times by circling the web around the loop coater to achieve a sheet
resistance < 15 Ω/sq for both electrodes.
The PV2000:PC70BM active layer was either fabricated at the single pass inkjet machine
from THN based ink or at the PiXDRO LP 50 from o-X:THN 1:1 (vol.) ink.

177
From Cells to Modules

The layout of a fully inkjet printed module with an active area of 10 cm², consisting of
three single solar cells, which are monolithically connected in series, is presented in
Figure VI-14 a). For realising the interconnection, the regular P2 ‘gap’ layout as
described in chapter 6.1.1 was chosen. Processing of the silver ‘bridge’ P2 strategy as
described in 6.1.2 was not possible because of limited accuracy of the machines web
movement. In order to maximise the efficiency, the active layer was printed from
o-X:THN ink using the PiXDRO LP 50.

Figure VI-14: a) Layout of fully inkjet printed 10 cm² three-layer modules, black:
AgNW:ZnO base electrode, green: PV2000:PC70BM active layer, red:
AgNW:PEDOT:PSS top electrode. b) Photograph of modules in production. c) J-V
characteristics of PV2000:PC70BM three-layer solar module and single module cells. d)
SEM cross section image of an inkjet printed AgNW:PEDOT:PSS blend electrode and e)
inkjet printed AgNW:ZnO blend electrode. The red arrows mark AgNWs, which appear
brighter compared to the matrix material. The SEM images were recorded by K.C. Tam
and are reproduced in this work with permission.

The cross section SEM images (Figure VI-14 d), e)) reveal that the AgNWs are
embedded in the EEL/HEL material, showing no signs of demixing, aggregations or
spikes, which could cause shunting of the device. The performance of the fully inkjet
printed modules is reported in Figure VI-14 and Table VI-8.

178
From Cells to Modules

Table VI-8: Device performance parameters of the solar cells shown in Figure VI-14.
Sample 𝐴𝑎𝑐𝑡𝑖𝑣𝑒 𝑃𝐶𝐸 𝐹𝐹 𝐽𝑆𝐶 per cell 𝑉𝑂𝐶
(cm²) (%) (%) (mA/cm²) (V)

Cell 1 3.3 4.57 46.15 12.48 0.79

Cell 2 3.3 4.49 47.50 11.87 0.80

Cell 3 3.3 4.50 46.45 12.05 0.80

Module 10.0 4.67 46.99 12.57 2.37

Solar simulator measurements of the modules show full voltage and current of all three
cells. The achieved efficiencies of up to 4.7 % are even higher than the 4.3 % presented
for the single cell with nanowires electrodes (chapter 5.3.2), which is up to now the
highest value for fully inkjet printed organic solar cells published in literature. The results
prove the robustness of the three layer approach and scalability by inkjet technology.

To further demonstrate that the core strength of inkjet technology, namely to print any
arbitrary design, is not lost when switching to a high-throughput single pass process, a set
of demonstrator cells with the above described three-layer structure was fabricated. The
devices show pictograms of the main renewable energy sources, solar power, wind
power, hydro power and power from regrowing biomass in the AgNW:PEDOT:PSS
electrode (Figure VI-15 a), b)). All layers of the devices are fabricated with the single
pass inkjet machine. Compared to other PV2000:PC70BM devices presented in this work,
the performance of the single pass inkjet printed demonstrator cells is lower. This is
mainly attributed to the big cell size of 3.5 cm² and the irregular shape of the top
electrode used to achieve the image pattern. Nevertheless, the devices are a successful
first time proof that a process relying entirely on single pass inkjet printing can be used to
manufacture arbitrarily shaped OPV at high throughput rates.

179
From Cells to Modules

Figure VI-15: Fully single pass inkjet printed flexible three-layer OPV demonstrator
consisting of four 3.5 cm² solar cells showing pictograms of the renewable energy
sources solar power, wind power, hydro power and power from regrowing biomass. a)
Layout of the device, black: AgNW:ZnO base electrode, green: PV2000:PC70BM, red:
AgNW:PEDOT:PSS top electrode. b) Photograph of the demonstrator before
encapsulation. c) During encapsulation the UV-curable epoxy glue (Katiobond LP 655)
dissolves the PC70BM in areas which are not protected by the top electrode, thereby
changing the active layer colour in the semitransparent devices from brown to green.
d) J-V characteristics of the 3.5 cm² PV2000:PC70BM solar cells.

Table VI-9: Device performance parameters of the solar cells shown in Figure VI-15.
Sample 𝐴𝑎𝑐𝑡𝑖𝑣𝑒 𝑃𝐶𝐸 𝐹𝐹 𝐽𝑆𝐶 𝑉𝑂𝐶
(cm²) (%) (%) (mA/cm²) (V)

Cell 1 3.5 2.89 39.19 9.94 0.74

Cell 2 3.5 2.56 36.98 9.48 0.74

Cell 3 3.5 2.88 39.87 9.71 0.74

Cell 4 3.5 2.78 37.00 10.0 0.75

Furthermore, it should be highlighted that the printed cells are flexible, semitransparent
and all layers are manufactured in ambient atmosphere. All these properties are of high
importance to bring inkjet printed OPV closer to market readiness.

180
From Cells to Modules

6.4. Conclusion
In this chapter the transition from inkjet printed single solar cells to complete solar
modules was presented. Hereby the focus was set on the last two critical aspects for the
successful realisation of an industrial OPV inkjet printing process as defined in chapter
1.5. These are (vii) large area and (viii) superior design. A summary of the achieved
results is presented in Table VI-10.

Table VI-10: Overview of inkjet printed OPV layers and rating based on the critical
requirements for a robust inkjet printing process.

Material/ Large area Superior design


Technology

P3HT:PC60BM  Applied in 150 cm² modules 


Red colour,
Applied in free-form
modules

PV2000:PC70BM  Applied in 10 cm² modules, 


Green to brown colour,
Successfully printed at high Applied in free-form
throughput single pass IJ cells

PEDOT:PSS  Applied in 84 cm² modules, 


‘Invisible’ in device,
Successfully printed at high Applied in free-form
throughput single pass IJ modules

ZnO NPs  Applied in 150 cm² modules, 


‘Invisible’ in device,
Successfully printed at high Applied in free-form
throughput single pass IJ modules

AgNW mesh +  Applied in 150 cm² modules, 


‘Invisible’ in device,
hybrid electrodes Successfully printed at high Applied in free-form
(blends with ZnO throughput single pass IJ modules
or PEDOT:PSS)

Ag grid electrode  Printable at high throughput x Visually obstructive


single pass IJ

Conventional ‘gap’  Applied in 10 cm² modules at x Visually obstructive


interconnection high throughput single pass IJ

Novel ‘bridge’ ()Applied in 150 cm² modules, 


‘Invisible’ in device,
interconnection Demonstration of single pass Applied in free-form
printing to be done modules

181
From Cells to Modules

First, solar modules with conventional interconnection comprising of P1, P2 and P3


structuring lines were realised. For the first time inkjet printing of a whole module
structure comprising of four printed layers was demonstrated. Furthermore, the achieved
𝐺𝐹𝐹 of up to 85 % and module 𝑃𝐶𝐸 of up to 4.3 % are significantly higher than any
values reported in literature about solar modules that were partly realised by inkjet
printing.83,142
In a second step, an innovative technology to process solar modules with literally
‘invisible’ cell-to-cell interconnections was introduced. By replacing the P2 line with an
inkjet printed silver interconnection ‘bridge’, all other layers can be printed over the
whole area without additional patterning, thereby greatly reducing the visual conspicuity
of the interconnection area. Besides visual improvement, the design also increases the
𝐺𝐹𝐹 of OPV modules and reduces the number of alignment steps required. Visual and
electrical investigations prove the excellent connection between cells. Furthermore, a
FEM model of the solar module was developed to analyse the influence of the
interconnection resistance on the solar module performance. The simulations lead to the
conclusion that the achieved interconnection resistance of 0.002 Ω*cm2 is already at
optimum, as lower values do not increase 𝐹𝐹 by more than 1 %.
The application of this novel ‘shy’ interconnection strategy in combination with the
printed AgNW electrodes enables the fabrication of large area arbitrarily shaped OPV
module demonstrators of unique appearance. To find suitable printing layouts a software
tool was developed which slices the arbitrary module shape into several smaller cell areas
that are monolithically connected in series. The fabricated demonstrators with active areas
of up to 150 cm² are the first time demonstration of inkjet printed arbitrary shape OPV
modules. Furthermore, it is the first time realisation of efficient inkjet printed OPV
devices on such large area, confirming the robustness and scalability of the technology
developed in this work.
Finally, the transfer of the printing processes to a high throughput single pass inkjet
printer (2.4 pl SAMBA printheads; 5 m²/min printing speed) is demonstrated.
P3HT:PC60BM solar cells with ZnO and PEDOT:PSS buffer layers fabricated at the
single pass machine yield the same performance as well working blade coated reference
devices, which leads to the conclusion that loss-free upscaling of the processes developed
in this work is possible. AgNWs with a length of several 10s of µm were printed using
SAMBA print heads (10 µm nozzle diameter) without on-head filters. Applying the

182
From Cells to Modules

AgNWs in form of ZnO:AgNW as well as PEDOT:PSS:AgNW hybrid electrodes, fully


inkjet printed 10 cm² solar modules were manufactured. Their 𝑃𝐶𝐸s of up to 4.7 % are
even higher compared to the single cells with AgNW electrodes (4.3 %) that held the
efficiency record for fully inkjet printed organic solar cells. Furthermore, to demonstrate
that the core strength of inkjet technology, namely to print any arbitrary design, is not lost
when switching to a high-throughput single pass process, a set of flexible,
semitransparent demonstrator cells with a unique layout was fabricated.

183
CHAPTER VII: Summary and Outlook

Abstract:

This chapter summarises and rates the main results and achievements of this work,
including:

- The first time demonstration of inkjet printed AgNWs with a length of ~30 µm.
- A novel and generic approach to achieve wetting based on inkjet printed
anchoring points.
- A new strategy to print solar modules with almost invisible interconnection.

Combination of all these achievements enables the production of optically homogeneous


large area and arbitrarily shaped digitally printed OPV modules. This technological
progress brings OPV as well as other types of printable electronics closer to commercial
applicability, which will be addressed in the outlook chapter.

184
Summary and Outook

7.1. Summary
This work aimed to develop and demonstrate digital printing of OPV at an industrially
relevant level in order to lead printed PV and similar technologies closer to market
readiness. Based on a literature overview several critical aspects for achieving this goal
have been identified at the beginning of the thesis and categorised into ink-, layer-, and
device aspects. In the following, the main achievements of this work in these three sectors
are summarised.

Ink requirements:

(i) Eco-friendly raw materials:

Electrode- and buffer layer inkjet inks were developed using environmentally friendly
solvents or dispersion media. This comprises 1-pentanol based ZnO nanoparticle and
AgNW inks as well as water based PEDOT:PSS:AgNW blends. Furthermore, active
layer inkjet inks free of halogenated aromatic compounds were created. However, further
research is necessary to exchange THN and o-X with even less environmentally critical
solvents.

(ii) Formation of well-defined spherical droplets:

The creation of electronic inkjet inks which offer a stable and satellite-free printing
process and extended nozzle open times was presented using the Ohnesorge theory as
formulation guideline. A solvent system consisting of o-X and THN mixed in a ratio of
1:1 (vol.) was identified as most suitable for inkjet printing blends of P3HT:PC60BM or
PV2000:PC70BM. Even though the 𝑂ℎ-number of the pristine solvents is low
(𝑂ℎ𝑜-𝑋 = 0.03, 𝑂ℎTHN = 0.06), addition of the polymers results in an increase in viscosity.
The associated 𝑂ℎ-number gain to a value of 0.14 allows a stable drop formation. By
changing the solvent in commercially available ZnO nanoparticle- and AgNW inks from
IPA to 1-pentanol, an 𝑂ℎ-number increase from 0.09 to 0.16 and suppression of satellites
was achieved. Using the 1-pentanol based AgNW ink, digital printing of nanowires with
an average length ~30 µm through inkjet nozzles of similar dimension (10-35 µm) was
demonstrated for the first time.

185
Summary and Outlook

(iii) Sufficient stability and nozzle open time to ensure robust and scalable printing:

An almost 10-fold increase in nozzle open time was achieved by modification of the ZnO
and AgNW inks with 1-pentanol. Furthermore, it was found out that introduction of THN
to the o-X based P3HT:PC60BM or PV2000:PC70BM active layer inks, suppresses fast
gelling. This allows a stable printing process with nozzle open times of almost 1 h and
makes the solvent blend a suitable alternative to o-DCB.

Layer requirements:

(iv) Printing of homogeneous films:

The wetting behaviour of different water based PEDOT:PSS inks on highly hydrophobic
BHJ active layers was investigated. Experiments revealed that addition of
fluorosurfactants does not always result in good wetting and homogeneous layer
formation. As alternative, a surfactant-free approach to achieve wetting based on inkjet
printed anchoring points was introduced. These so-called pinning centres are printed onto
the surface prior to wet layer deposition and can prevent shrinkage or rupturing of the
liquid film. The proposed pinning centre strategy does not only allow wet film
structuring with a precision in the 10 µm range, but also provides the solution to a major
problem of printing functional inks. In many applications, like printed electronics, the
use of ink additives and surface modifiers is limited, because any change of ink
formulation has direct influence on the properties and therefore functionality of the
printed film. Pinning of the wet film obviates the need for such modifications, leaving
the ink and layer properties unaffected.

(v) Electrical and optical functionality of the dry layers:

The functionality of the inkjet printed layers was demonstrated by manufacturing of


devices and comparison to well working blade coated reference solar cells. For ZnO
(EEL) and PEDOT:PSS (HEL) layers as well as P3HT:PC60BM and PV2000:PC70BM
active layers, loss-free transition from blade coating to inkjet printing was achieved. In
order to replace the semitransparent ITO electrode, inkjet printed silver grids or AgNW
mesh electrodes were applied. The inkjet printed and visually non-obstructive mesh
electrodes offer an excellent balance of sheet resistance and transmittance of
𝑅𝑠𝑞 < 20 Ω/sq @ T > 90 % as well as a very low haze of only 2.3 % and conductivity

186
Summary and Outook

ratio 𝜎𝐷𝐶,𝐵 /𝜎𝑂𝑃 of ~250. These values are highly comparable to literature reports of
layers fabricated by conventional methods, such as slot die- or spray coating,
highlighting their excellent quality.

Device requirements:

(vi) High device efficiency:

Hero cells with inkjet printed PV2000:PC70BM layers reach record efficiencies of 7.3 %,
thereby slightly exceeding literature reports of comparable inkjet printed OPV.
Introduction of an optimised line grid bottom electrode resulted in 𝑃𝐶𝐸𝑠 of up to 5.66 %,
which is ~15 % lower than the ITO reference device, but significantly higher than record
device efficiencies presented in literature (see summary Table II-5). An FEM simulation
reveals that the efficiency losses are mainly caused by shading of the silver finger
(54.1 %) and parasitic absorption within the PEDOT:PSS layer (36.0 %). The results of
solar cells with inkjet printed AgNWs are even better. Fully inkjet printed
PV2000:PC70BM solar cells and modules with AgNW base and top electrodes with
𝑃𝐶𝐸𝑠 up to 4.7 % were demonstrated. This is the highest value reported for fully inkjet
printed OPV so far.

(vii) Large area:

The transition from small scale single solar cells to large area inkjet printed solar
modules was demonstrated. For the first time OPV modules with four inkjet printed
layers (𝐺𝐹𝐹: 85 %, 𝑃𝐶𝐸: 4.3 %, size: 3.4 cm²) were realised. Furthermore, solar module
demonstrators with active areas of up to 150 cm² and a 𝑃𝐶𝐸 of 2.5 % were fabricated,
thus proving the robustness and scalability of the printing processes developed in this
work. Finally, the transfer to a high throughput quasi-R2R single pass inkjet machine
was addressed. Applying this printer, ZnO and PEDOT:PSS buffer layers were fabricated
at throughput speeds of more than 5 m²/min. Resulting P3HT:PC60BM solar cell 𝑃𝐶𝐸s of
> 3 % confirm the excellent quality of the printed layers. Even, AgNWs as well as
ZnO:AgNW and PEDOT:PSS:AgNW blends were printed with the high throughput
single pass machine and applied as cathode and anode in the 10 cm² hero modules.

187
Summary and Outlook

(viii) Superior design:

A novel approach to avoid the visually obstructive gap in the active layer (P2 line),
which is necessary for monolithical series connection of adjacent cells, was developed.
By introducing inkjet printed silver ‘bridges’, modules with almost invisible cell
interconnection and very low interconnection resistance of only 0.002 Ω*cm2 were
manufactured. This new technology in combination with visually inconspicuous inkjet
printed AgNW mesh electrodes enables the fabrication of solar modules with unique
appearance. Choosing different active layer materials, free-form solar modules in the
shape of buildings, portraits and company logos of different colour (red: P3HT:PC60BM,
green/brown: PV2000:PC60BM/PV2000:PC70BM, purple: P3HT:O-IDTBR) were
realised, showcasing the strength of combining OPV and inkjet technology.

In order to rate these achievements, the processes established in this work have to be
compared to an ideal OPV fabrication scenario as introduced by Krebs (see chapter
2.1.5).88

‘The ideal process should involve:


- solution processing of all layers
 Fully inkjet printed solar cells and modules were demonstrated
- on flexible substrates
 PET foil was used as substrate in the single pass prototype machine
- by the combination of as few coating and printing steps as possible.
 Fully printed solar cells consisting of only 3 layers were demonstrated
- The process should be free from costly indium,
 Inkjet printed AgNW mesh and silver grid electrodes were introduced
- toxic solvents as well as chemicals and
 Electrodes and buffer layers were printed from alcohol or water based inks.
Active layer inkjet inks free of halogenated aromatic compounds were developed.
- the final polymer solar cell product should have a low environmental impact and
a high degree of recyclability’
 The fully inkjet printed and flexible devices demonstrated in this work can be
disposed in an almost waste-free incineration process.

In conclusion, the technology introduced in this work perfectly matches the ideal OPV
fabrication scenario. Therefore, it is possible to answer a question that was raised already
188
Summary and Outook

in the year 2008, when first reports about inkjet printed OPV were published (see chapter
2.3).88

It is difficult at the current stage to determine whether inkjet printing will play an
important role in the high volume fabrication of polymer solar cells. The advantages
offered by the technique and the good literature examples of its use makes it likely that it
will become of use for creating complex patterns and perhaps devices with a small outline
of the active area.

Since formulation of this question, substantial progress has been made in OPV and inkjet
technology. Modern printers can reach throughput speeds exceeding 1000 m²/h,2 thereby
breaking the limitation to use inkjet technology only on small areas. At the same time,
OPV efficiencies of more than 13 % (current record: 17.3 %) have been realised.1,213,214
These developments and the significant progress in digitally printed OPV (see literature
summary Table II-5) combined with the achievements of this work prove the feasibility
of industrial scale OPV manufacturing by inkjet. It is the authors believe that the
combination of inkjet and OPV can become a major success story.

7.2. Outlook
In this final paragraph, an outlook on possible applications for the inkjet technology
developed in this work is provided. An illustrative summary is shown in Figure VII-1.
As described in paragraph 1.3 the USPs of OPV are superior aesthetics (i), flexibility (ii),
the tunability of colour by band gap adjustment (iii), the light weight (iv), the freedom of
form (v) and scalability (vi). These factors make OPV technology highly interesting for
BIPV applications. Hereby the integration into facade-, shading elements or windows
requires a cost efficient fabrication method at varying batch size that allows quick and
easy adaption of shape and colour to specifications defined by the architect. The inkjet
processes developed in this work provide the best possible match to these requirements.
The main research topic that has to be addressed in order to realise such BIPV
applications is the lifetime of the OPV devices, which has to be further improved, e.g. by
advanced encapsulation strategies and barrier foils.
Another even bigger potential market for OPV is the integration into wearables and
textiles. OPV might be used as power source for smartphones, watches, fitness trackers or
sensors. For the tight integration into products, superior aesthetics and flexibility are the

189
Summary and Outlook

key aspects. Especially visually homogeneous AgNW mesh electrodes and the ‘shy’
interconnection strategy introduced in this work could contribute to realisation of such
applications. A prototype OPV wrist band for powering a fitness tracker, which was
realised in a side project together with co-workers, is shown in Figure VII-1.

Figure VII-1: Potential application scenarios for the inkjet processes developed in this
work, including integration into wearables, BIPV, OLED lightning and display
215
applications, photodetectors (image reproduced with permission from ), solar parks
216
(image reproduced with permission from ) and perovskite PV (image reproduced with
permission from 204)

Low power devices, such as sensors, can already be operated with integrated OPV.
However, further research in pushing the record efficiencies of the active layer materials
is necessary to also power consumers electronics that require higher power, such as smart
watches or smart phones, with small area integrated OPV.

190
Summary and Outook

Besides OPV, many of the inkjet processes and materials presented in this work can
directly be transferred to other printed electronics technologies. By replacing the active
layer of the solar cell stack with Livilux SPG-01T, OLEDs can be produced. Figure VII-1
shows such a device with the layer sequence ITO/ZnO/LiviluxSPG-01T/PEDOT:PSS/Ag
made by a co-worker K.C. Tam. All layers except the electrodes were manufactured with
the inkjet processes described in this work. By using inkjet printed AgNW meshes to
replace the opaque silver, even semitransparent devices with homogeneous illumination
from both sides were demonstrated. Printed OLEDs could for example be used for
lightning or displays applications. In this context the company Kateeva should be
mentioned, which is very active in the field.
Furthermore, inkjet printed electrodes and buffer layers presented in this work have
already proven their applicability in photo detectors. Fully inkjet printed devices
consisting of an Ag finger structure covered by a layer of ZnO N-10 nanoparticles and
PbS nanocrystal active active material were presented by A. Amin et al..215
Finally, the possibility to apply inkjet printing for large area energy harvesting
applications should be mentioned. Even though inkjet technology offers the high
throughput and material saving manufacturing process required for such application, it is
highly unlikely that OPV will compete with the market dominating silicon PV in the near
or mid future. However, new emerging technologies, such as perovskite PV, could be
able to step into this field. Very cheap raw materials, record efficiencies of more than
22 % and high throughput R2R production from solution could make this technology a
real game changer.6,217 Problems of perovskite PV that have to be solved before
application are the high content of toxic lead and limited lifetime of the solar cells. 218,219
If scientists are able to overcome these boundaries, high throughput fabrication by means
of inkjet seems an appealing strategy, which is already focused by research institutes,
universities and even start-ups, such as Saule Technologies.220 The inkjet printed ZnO and
PEDOT:PSS buffer layers as well as electrodes introduced in this work could be applied.
In summary, the detailed description of the complete process chain from ink formulation
to large area device application makes this work a ‘tool box’ for everybody who is
working on printed electronics. The large variety of potential applications for the
developed inks and printing procedures presented in the outline provides confidence that
one or more of these applications could make the step from science to industry. This
motivates further work in the field.

191
Bibliography
(1) Meng, L.; Zhang, Y.; Wan, X.; Li, C.; Zhang, X.; Wang, Y.; Ke, X.; Xiao, Z.;
Ding, L.; Xia, R.; et al. Organic and Solution-Processed Tandem Solar Cells with
17.3% Efficiency. Science (80-. ). 2018.

(2) Durst Phototechnik. Rho 1330 Specifications https://www.durst-group.com/


(accessed Mar 19, 2018).

(3) US Energy Information Administration. International Energy Outlook 2016


https://www.eia.gov/outlooks/ieo/ (accessed Sep 9, 2017).

(4) Statistisches Bundesamt - Bruttostromerzeugung in Deutschland für 2015 bis 2017


https://www.destatis.de/DE/ZahlenFakten/Wirtschaftsbereiche/Energie/Energie.ht
ml (accessed Mar 12, 2018).

(5) Ecofys GmbH 2007 - in Zusammenarbeit mit dem Ministerium für Wirtschaft,
Mittelstand und Energie NRW https://www.ecofys.com/de/ (accessed Mar 13,
2018).

(6) Green, M. A.; Hishikawa, Y.; Dunlop, E. D.; Levi, D. H.; Hohl-Ebinger, J.; Ho-
Baillie, A. W. Y. Solar Cell Efficiency Tables (Version 51 ). Prog. Photovoltaics
Res. Appl. 2018, 26 (1), 3–12.

(7) Schneider, K.; Wirth, H. Fraunhofer ISE - Recent Facts about Photovoltaics in
Germany
https://www.ise.fraunhofer.de/content/dam/ise/en/documents/publications/studies/r
ecent-facts-about-photovoltaics-in-germany.pdf (accessed Mar 13, 2018).

(8) Fraunhofer Institute for Solar Energy Systems, ISE - Photovoltaics report
https://www.ise.fraunhofer.de/content/dam/ise/de/documents/publications/studies/
Photovoltaics-Report.pdf (accessed Mar 13, 2018).

(9) Lee, T. D.; Ebong, A. U. A Review of Thin Fi Lm Solar Cell Technologies and
Challenges. Renew. Sustain. Energy Rev. 2017, 70 (November 2016), 1286–1297.

(10) Zhang, S.; Yang, X.; Numata, Y.; Han, L. Highly Efficient Dye-Sensitized Solar
Cells: Progress and Future Challenges. Energy Environ. Sci. 2013, 6, 1443–1464.

(11) Sun, Y.; Welch, G. C.; Leong, W. L.; Takacs, C. J.; Bazan, G. C.; Heeger, A. J.
Solution-Processed Small-Molecule Solar Cells with 6.7 % Efficiency. Nat. Mater.
2011, 11 (1), 44–48.

(12) Liu, Y.; Zhao, J.; Li, Z.; Mu, C.; Ma, W.; Hu, H.; Jiang, K.; Lin, H.; Ade, H.; Yan,
H. Aggregation and Morphology Control Enables Multiple Cases of High-
Efficiency Polymer Solar Cells. Nat. Commun. 2014, 1–8.

(13) Lee, M. M.; Teuscher, J.; Miyasaka, T.; Murakami N, T.; Snaith, H. J. Efficient
Hybrid Solar Cells Based on Meso-Superstructured Organometal Halide
Perovskites. Science (80-. ). 2012, 338 (6107), 643–647.

192
Bibliography

(14) Gevorgyan, S. A.; Espinosa, N.; Ciammaruchi, L.; Roth, B.; Livi, F.; Tsopanidis,
S.; Züfle, S.; Queirós, S.; Gregori, A.; Alves, G.; et al. Baselines for Lifetime of
Organic Solar Cells. Adv. Energy Mater. 2016, 6, 1–9.

(15) Machui, F.; Hösel, M.; Li, N.; Spyropoulos, G. D.; Ameri, T.; Søndergaard, R. R.;
Jørgensen, M.; Scheel, A.; Gaiser, D.; Kreul, K.; et al. Cost Analysis of Roll-to-
Roll Fabricated ITO Free Single and Tandem Organic Solar Modules Based on
Data from Manufacture. Energy Environ. Sci. 2014, No. 9, 2792–2802.

(16) Espinosa, N.; Hösel, M.; Angmo, D.; Krebs, F. C. Solar Cells with One-Day
Energy Payback for the Factories of the Future. Energy Environ. Sci. 2012, 5,
5117–5132.

(17) Lucera, L.; Kubis, P.; Fecher, F. W.; Bronnbauer, C.; Turbiez, M.; Forberich, K.;
Ameri, T.; Egelhaaf, H.-J.; Brabec, C. J. Guidelines for Closing the Efficiency Gap
between Hero Solar Cells and Roll-To-Roll Printed Modules. Energy Technol.
2015, 3 (4), 373–384.

(18) Brabec, C. J. Organic Photovoltaics : Technology and Market. Sol. Energy Mater.
Sol. Cells 2004, 83, 273–292.

(19) Lucera, L. Doctoral Thesis: Closing the Efficiency Gap between Lab-Produced
Organic Solar Cells and Roll-to-Roll Printed Modules, Friedrich-Alexander-
Universität Erlangen-Nürnberg, 2016.

(20) van der Wiel, B.; Egelhaaf, H.-J.; Issa, H.; Roos, M.; Henze, N. Market Readiness
of Organic Photovoltaics for Building Integration. MRS Online Proc. Libr. 2014,
1639.

(21) OPV Solar Trees at EXPO 2015 http://www.opvius.com/opv-solar-trees-at-expo-


2015.html (accessed Aug 16, 2017).

(22) HeLi-on - The World’s Most Compact Solar Charger


https://infinitypv.com/products/heli-on (accessed Mar 15, 2018).

(23) Machui, F. EnCN, ZAE Bayern, Semitransparent Organic Photovoltaic (OPV)


Modules in Greenhouse Applications. 2017.

(24) Emmott, C. J. M.; Jason, A. R.; Kirchartz, T.; Urbina, A.; Ekins-daukes, N. J.;
Nelson, J. Organic Photovoltaic Greenhouses: A Unique Application for Semi-
Transparent PV. 2015, 8, 1317–1328.

(25) Fujifilm Samba G3L


http://www.fujifilmusa.com/products/industrial_inkjet_printheads/print-
products/printheads/high-performance/samba_g3l/index.html (accessed Aug 5,
2018).

(26) LLC, BCC Research: Global Inkjet Market Inking Double-Digit Growth Rates
Through 2021 https://www.bccresearch.com/pressroom/avm/global-inkjet-market-
inking-double-digit-growth-rates-through-2021 (accessed Dec 31, 2017).

(27) Paolo Samorí; Palermo, V.; Feng, X.; Maisch, P.; Lucera, L.; Brabec, C. J.;

193
Bibliography

Egelhaaf, H. Flexible Carbon Based Electronics - Chapter: Flexible Solar Cells,


ISBN: 978-3-527-34191-7. In Flexible Carbon Based Electronics - Chapter:
Flexible Solar Cells; Copyright Wiley-VCH Verlag GmbH & Co. KGaA, 2018; p
480.

(28) Krebs, F. C. Polymeric Solar Cells; Materials, Design, Manufacture. Lancaster, PA


DEStech Publ. Inc. 2010, 229.

(29) Martin Pope; Swenberg, C. E. Electronic Processes in Organic Crystals and


Polymers - 2nd Edition, ISBN: 9780195129632; 1999.

(30) Wellmann, P. Materialien Der Elektronik Und Energietechnik: Halbleiter,


Graphen, Funktionale Materialien, ISBN: 978-3-658-14006-9; 2016.

(31) Deibel, C.; Dyakonov, V. Polymer – Fullerene Bulk Heterojunction Solar Cells.
Reports Prog. Phys. 2010, 73 (9), 096401/1–096401/39.

(32) Torabi, S.; Jahani, F.; Severen, I. Van; Kanimozhi, C.; Patil, S.; Havenith, R. W.
A.; Chiechi, R. C.; Lutsen, L.; Vanderzande, D. J. M.; Cleij, T. J.; et al. Strategy
for Enhancing the Dielectric Constant of Organic Semiconductors Without
Sacrificing Charge Carrier Mobility and Solubility. Adv. Energy Mater. 2015, 25
(1), 150–157.

(33) Tang, C. W. Two-Layer Organic Photovoltaic Cell. Appl. Phys. Lett. 1986, 48
(183).

(34) Hoppe, H.; Sariciftci, N. S. Morphology of Polymer / Fullerene Bulk


Heterojunction Solar Cells. J. Mater. Chem. 2005, 16 (1), 45–61.

(35) Yu, G.; Gao, J.; Hummelen, J. C.; Wudl, F.; Heeger, A. J. Polymer Photovoltaic
Cells : Enhanced Efficiencies via a Network of Internal Donor-Acceptor
Heterojunctions. Science (80-. ). 1995, 270.

(36) Machui, F.; Maisch, P.; Langner, S.; Krantz, J.; Ameri, T.; Brabec, C. J.
Classification of Additives for Organic Photovoltaic Devices. ChemPhysChem
2015, 16, 1275–1280.

(37) Vandewal, K.; Tvingstedt, K.; Gadisa, A.; Inganäs, O.; Manca, J. V. Relating the
Open-Circuit Voltage to Interface Molecular Properties of Donor : Acceptor Bulk
Heterojunction Solar Cells. Phys. Rev. B 2010, 81, 125204–1 – 8.

(38) Vandewal, K.; Tvingstedt, K.; Gadisa, A.; Inganäs, O.; Manca, J. V. On the Origin
of the Open-Circuit Voltage of Polymer-Fullerene Solar Cells. Nat. Mater. 2009, 8
(11), 904–909.

(39) Bartesaghi, D.; del Carmen Pérez, I.; Kniepert, J.; Roland, S.; Turbiez, M.; Neher,
D.; Koster, L. J. A. Competition between Recombination and Extraction of Free
Charges Determines the Fill Factor of Organic Solar Cells. Nat. Commun. 2015, 6:
7083.

(40) Shockley, W.; Queisser, H. J. Detailed Balance Limit of Efficiency of P-N


Junction Solar Cells. J. Appl. Phys. 1961, 32, 510–519.

194
Bibliography

(41) Shockley, W. The Theory of P-N Junctions in Semiconductors and P-N Junction
Transistors. Bell Labs Tech. J. 1949.

(42) Mertens, K. Photovoltaik: Lehrbuch Zu Grundlagen, Technologie Und Praxis,


ISBN: 978-3-446-442323-0; 2015.

(43) Waldauf, C.; Scharber, M. C.; Schilinsky, P.; Hauch, J. A.; Brabec, C. J. Physics of
Organic Bulk Heterojunction Devices for Photovoltaic Applications. J. Appl. Phys.
2006, 99, 104503.

(44) Laquai, F.; Andrienko, D.; Deibel, C.; Neher, D. Elementary Processes in Organic
Photovoltaics - Charge Carrier Generation , Recombination , and Extraction in
Polymer – Fullerene Bulk Heterojunction Organic Solar Cells, ISBN:
9783319283388; 2016.

(45) Cowan, S. R.; Roy, A.; Heeger, A. J. Recombination in Polymer-Fullerene Bulk


Heterojunction Solar Cells. Phys. Rev. B 2010, 82, 245207.

(46) Spanggaard, H.; Krebs, F. C. A Brief History of the Development of Organic and
Polymeric Photovoltaics. Sol. Energy Mater. Sol. Cells 2004, 83 (2-3), 125–146.

(47) Dou, L.; You, J.; Hong, Z.; Xu, Z.; Li, G.; Street, R. A.; Yang, Y. 25th
Anniversary Article : A Decade of Organic / Polymeric Photovoltaic Research.
Adv. Mater. 2013, 5 (46), 6642–6671.

(48) Distler, A. Doctoral Thesis: The Role of Fullerenes in the Photo-Degradation of


Organic Solar Cells, Friedrich-Alexander-Universität Erlangen-Nürnberg, 2015.

(49) Li, G.; Zhu, R.; Yang, Y. Polymer Solar Cells. Nat. Photonics 2012, 6, 153–161.

(50) Scharber, B. M. C.; Mühlbacher, D.; Koppe, M.; Denk, P.; Waldauf, C.; Heeger,
A. J.; Brabec, C. J. Design Rules for Donors in Bulk-Heterojunction Solar Cells —
Towards 10 % Energy-Conversion Efficiency. 2006, 5090 (502783), 789–794.

(51) Dou, L.; Liu, Y.; Hong, Z.; Li, G.; Yang, Y. Low-Bandgap Near-IR Conjugated
Polymers / Molecules for Organic Electronics. Chem. Rev. 2015, 115 (23), 12633–
12665.

(52) Grey, J. Organic Photovoltaics: Strong Absorption in Stiff Polymers. Nat. Maerials
2016, 15, 705–706.

(53) Nelson, J. Polymer : Fullerene Bulk Heterojunction Solar Cells. Mater. Today
2011, 14 (10), 462–470.

(54) Vezie, M. S.; Few, S.; Meager, I.; Pieridou, G.; Dörling, B.; Ashraf, R. S.; Goñi, A.
R.; Bronstein, H.; Mcculloch, I.; Hayes, S. C.; et al. Exploring the Origin of High
Optical Absorption in Conjugated Polymers. Nat. Mater. 2016, 15, 746–753.

(55) Li, G.; Shrotriya, V.; Huang, J.; Yao, Y.; Moriarty, T.; Emery, K.; Yang, Y. High-
Efficiency Solution Processable Self-Organization of Polymer Blends. Nat. Mater.
2005, 4 (11), 864–868.

195
Bibliography

(56) Liang, Y.; Xu, Z.; Xia, J.; Tsai, S.; Wu, Y.; Li, G.; Ray, C.; Yu, L. For the Bright
Future — Bulk Heterojunction Polymer Solar Cells with Power Conversion
Efficiency of 7.4 %. Adv. Energy Mater. 2010, 22 (20), 135–138.

(57) Lee, J. K.; Ma, W. L.; Brabec, C. J.; Yuen, J.; Moon, J. S.; Kim, J. Y.; Lee, K.;
Bazan, G. C.; Heeger, A. J. Processing Additives for Improved Efficiency from
Bulk Heterojunction Solar Cells. J. Am. Chem. Soc. 2008, 130 (11), 3619–3623.

(58) He, Z.; Zhong, C.; Huang, X.; Wong, W.; Wu, H.; Chen, L.; Su, S.; Cao, Y.
Simultaneous Enhancement of Open-Circuit Voltage , Short-Circuit Current
Density , and Fill Factor in Polymer Solar Cells. Adv. Mater. 2011, 23 (40), 4636–
4643.

(59) Raynergytek. Raynergy Tek Sets World Record Solution Processed Single
Junction OPV http://www.raynergytek.com/news.asp (accessed Jul 1, 2017).

(60) Berny, S.; Blouin, N.; Distler, A.; Egelhaaf, H.; Krompiec, M.; Lohr, A.; Lozman,
O. R.; Morse, G. E.; Nanson, L.; Pron, A.; et al. Solar Trees : First Large-Scale
Demonstration of Fully Solution Coated , Semitransparent , Flexible Organic
Photovoltaic Modules. Adv. Sci. 2016, 3 (5).

(61) Zhao, J.; Li, Y.; Yang, G.; Jiang, K.; Lin, H.; Ade, H.; Ma, W.; Yan, H. Efficient
Organic Solar Cells Processed from Hydrocarbon Solvents. Nat. Energy 2016.

(62) Guillén, C.; Herrero, J. ITO / Metal / ITO Multilayer Structures Based on Ag and
Cu Metal Films for High-Performance Transparent Electrodes. Sol. Energy Mater.
Sol. Cells 2008, 92, 938–941.

(63) De, S.; Coleman, J. N. The Effects of Percolation in Nanostructured Transparent


Conductors. MRS Bull. 2011, 36 (10), 774–781.

(64) Rathmell, A. R.; Wiley, B. J. The Synthesis and Coating of Long, Thin Copper
Nanowires to Make Flexible, Transparent Conducting Films on Plastic Substrates.
Adv. Mater. 2011, 23 (41), 4798–4803.

(65) 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 (6),
1076–1081.

(66) Galagan, Y.; Rubingh, J. J. M.; Andriessen, R.; Fan, C.; Blom, P. W. M.; Veenstra,
S. C.; Kroon, J. M. ITO-Free Flexible Organic Solar Cells with Printed Current
Collecting Grids. Sol. Energy Mater. Sol. Cells 2011, 95, 1339–1343.

(67) Choi, K.; Jeong, J.; Kang, J.; Kim, D.; Kuk, J.; Na, S.; Kim, D.; Kim, S.; Kim, H.
Characteristics of Flexible Indium Tin Oxide Electrode Grown by Continuous
Roll-to-Roll Sputtering Process for Flexible Organic Solar Cells. Sol. Energy
Mater. Sol. Cells 2009, 93 (8), 1248–1255.

(68) Tenent, R. C.; Barnes, T. M.; Bergeson, J. D.; Ferguson, J.; To, B.; Gedvilas, L.
M.; Heben, M. J.; Blackburn, J. L. Transparent Single-Walled-Carbon-Nanotube
Films for Photovoltaics Produced by Ultrasonic Spraying. Adv. Mater. 2009, 21

196
Bibliography

(31), 3210–3216.

(69) Bae, S.; Kim, H.; Lee, Y.; Xu, X.; Park, J.; Zheng, Y.; Balakrishnan, J.; Lei, T.;
Kim, H. R.; Song, Y. Il; et al. Roll-to-Roll Production of 30-Inch Graphene Films
for Transparent Electrodes. Nat. Nanotechnol. 2010, 5, 574–578.

(70) De, S.; Higgins, T. M.; Lyons, P. E.; Doherty, E. M.; Nirmalraj, P. N.; Blau, W. J.;
Boland, J. J.; Coleman, J. N. Silver Nanowire Networks as Flexible, Transparent,
Conducting Films: Extremely High DC to Optical Conductivity Ratios. ACS Nano
2009, 3 (7), 1767–1774.

(71) Kubis, P. Doctoral Thesis: Design and Development of Ultra-Fast Laser Patterning
Processes for the Production of Organic Photovoltaic Modules with High
Geometric Fill Factor, 2014.

(72) Krebs, F. C.; Spanggard, H.; Kjær, T.; Biancardo, M.; Alstrup, J. Large Area
Plastic Solar Cell Modules. Mater. Sci. Eng. B. 2007, 138 (2), 106–111.

(73) Krebs, F. C. Polymer Solar Cell Modules Prepared Using Roll-to-Roll Methods :
Knife-over-Edge Coating , Slot-Die Coating and Screen Printing. Sol. Energy
Mater. Sol. Cells 2009, 93, 465–475.

(74) Krebs, F. C.; Gevorgyan, S. A.; Gholamkhass, B.; Holdcroft, S.; Schlenker, C.;
Thompson, M. E.; Thompson, B. C.; Olson, D.; Ginley, D. S.; Shaheen, S. E.; et al.
A Round Robin Study of Flexible Large-Area Roll-to-Roll Processed Polymer
Solar Cell Modules. Sol. Energy Mater. Sol. Cells 2009, 93, 1968–1977.

(75) Bundgaard, E.; Krebs, F. C. Large Area Modules Based on Low Band Gap
Polymers. Photovolt. Spec. Conf. (PVSC), 2010 35th IEEE 2010, 1064–1067.

(76) Krebs, F. C.; Tromholt, T.; Jørgensen, M. Upscaling of Polymer Solar Cell
Fabrication Using Full Roll-to-Roll Processing. Nanoscale 2010, 873–886.

(77) Krebs, F. C.; Fyenbo, J.; Jørgensen, M. Product Integration of Compact Roll-to-
Roll Processed Polymer Solar Cell Modules : Methods and Manufacture Using
Flexographic Printing , Slot-Die Coating and Rotary Screen Printing. J. Mater.
Chem. 2010, No. 41, 8994–9001.

(78) Galagan, Y.; de Vries, I. G.; Langen, A. P.; Andriessen, R.; Verhees, W. J. H.;
Veenstra, S. C.; Kroon, J. M. Technology Development for Roll-to-Roll
Production of Organic Photovoltaics. Chem. Eng. Process. Process Intensif. 2011,
50 (5-6), 454–461.

(79) Manceau, M.; Angmo, D.; Jørgensen, M.; Krebs, F. C. ITO-Free Flexible Polymer
Solar Cells : From Small Model Devices to Roll-to-Roll Processed Large Modules.
Org. Electron. 2011, 12 (4), 566–574.

(80) Kopola, P.; Aernouts, T.; Sliz, R.; Guillerez, S.; Ylikunnari, M.; Cheyns, D.;
Välimäki, M.; Tuomikoski, M.; Hast, J.; Jabbour, G.; et al. Gravure Printed
Flexible Organic Photovoltaic Modules. Sol. Energy Mater. Sol. Cells 2011, 95,
1344–1347.

197
Bibliography

(81) Angmo, D.; Markus, H.; Krebs, F. C. All Solution Processing of ITO-Free Organic
Solar Cell Modules Directly on Barrier Foil. Sol. Energy Mater. Sol. Cells 2012,
107, 329–336.

(82) Angmo, D.; Gevorgyan, S. A.; Larsen-olsen, T. T.; Søndergaard, R. R.; Hösel, M.;
Jørgensen, M.; Gupta, R.; Kulkarni, G. U.; Krebs, F. C. Scalability and Stability of
Very Thin , Roll-to-Roll Processed , Large Area , Indium-Tin-Oxide Free Polymer
Solar Cell Modules. Org. Electron. 2013, 14 (3), 984–994.

(83) Angmo, D.; Larsen-Olsen, T. T.; Jørgensen, M.; Søndergaard, R. R.; Krebs, F. C.
Roll-to-Roll Inkjet Printing and Photonic Sintering of Electrodes for ITO Free
Polymer Solar Cell Modules and Facile Product Integration. Adv. Energy Mater.
2013, 3 (2), 172–175.

(84) Hyungcheol Back; Kong, J.; Kang, H.; Kim, J.; Kim, J.-R.; Lee, K. Flexible
Polymer Solar Cell Modules with Patterned Vanadium Suboxide Layers Deposited
by an Electro-Spray Printing Method. Sol. Energy Mater. Sol. Cells 2014, 130,
555–560.

(85) Lucera, L.; Machui, F.; Kubis, P.; Schmidt, H. D.; Adams, J.; Strohm, S.; Ahmad,
T.; Forberich, K.; Egelhaaf, H.-J.; Brabec, C. J. Highly Efficient, Large Area, Roll
Coated Flexible and Rigid OPV Modules with Geometric Fill Factors up to 98.5%
Processed with Commercially Available Materials. Energy Environ. Sci. 2016, 9,
89–94.

(86) Spyropoulos, G. D.; Kubis, P.; Li, N.; Baran, D.; Lucera, L.; Salvador, M.; Ameri,
T.; Voigt, M. M.; Krebs, F. C.; Brabec, C. J. Flexible Organic Tandem Solar
Modules with 6 % Effi Ciency : Combining Roll-to-Roll Compatible Processing
with High Geometric Fill Factors. Energy Environ. Sci. 2014, No. 10, 3284–3290.

(87) Kubis, P.; Lucera, L.; Machui, F.; Spyropoulos, G.; Cordero, J.; Frey, A.; Kaschta,
J.; Voigt, M. M.; Matt, G. J.; Zeira, E.; et al. High Precision Processing of Flexible
P3HT/PCBM Modules with Geometric Fill Factor over 95%. Org. Electron.
physics, Mater. Appl. 2014, 15 (10), 2256–2263.

(88) Krebs, F. C. Fabrication and Processing of Polymer Solar Cells : A Review of


Printing and Coating Techniques. Sol. Energy Mater. Sol. Cells 2009, 93, 394–412.

(89) Zhao, W.; Qian, D.; Zhang, S.; Li, S.; Inganäs, O.; Gao, F.; Hou, J. Fullerene-Free
Polymer Solar Cells with over 11 % Efficiency and Excellent Thermal Stability.
Adv. Mater. 2016, 28, 4734–4739.

(90) Schilinsky, B. P.; Waldauf, C.; Brabec, C. J. Performance Analysis of Printed Bulk
Heterojunction Solar Cells. Adv. Funct. Mater. 2006, 16 (13), 1669–1672.

(91) Chang, Y.; Tseng, S.; Chen, C.; Meng, H.; Chen, E.; Horng, S.; Hsu, C. Polymer
Solar Cell by Blade Coating. Org. Electron. 2009, 10 (5), 741–746.

(92) Mariani, P.; Vesce, L.; Carlo, A. Di. The Role of Printing Techniques for Large-
Area Dye Sensitized Solar Cells. Semicond. Sci. Technol. 2015, 30 (10), 1–16.

(93) Machui, F.; Lucera, L.; Spyropoulos, G. D.; Cordero, J.; Ali, A. S.; Kubis, P.;

198
Bibliography

Ameri, T.; Voigt, M. M.; Brabec, C. J. Large Area Slot-Die Coated Organic Solar
Cells on Flexible Substrates with Non-Halogenated Solution Formulations. Sol.
Energy Mater. Sol. Cells 2014, 128, 441–446.

(94) Krebs, F. C.; Espinosa, N.; Hösel, M.; Søndergaard, R. R.; Jørgensen, M. 25th
Anniversary Article : Rise to Power – OPV-Based Solar Parks. Adv. Mater. 2014,
26 (1), 29–39.

(95) Ding, J. M.; De, A.; Vornbrock, F.; Ting, C.; Subramanian, V. Patternable Polymer
Bulk Heterojunction Photovoltaic Cells on Plastic by Rotogravure Printing. Sol.
Energy Mater. Sol. Cells 2009, 93, 459–464.

(96) Euler, L. Principes Généraux Du Mouvement Des Fluides. Mem. Acad. Sci. Berlin
1755.

(97) Stokes, G. G. On the Theories of the Internal Friction of Fluids in Motion, and of
the Equilibrium and Motion of Elastic Solids. Cambridge Philos. Soc 1849, 287–
319.

(98) Wijshoff, H. The Dynamics of the Piezo Inkjet Printhead Operation. Phys. Rep.
2010, 491 (4-5), 77–177.

(99) Plateau, J. A. F. Mémoire Sur Les Phénomenes Que Présente Une Masse Liquide
sans Pesanteur. Mém. Acad. Sci. Bruxelles 1843.

(100) Lord Rayleigh. On the Instability of a Cylinder of Viscous Liquid under Capillary
Force. Philos. Mag. 1892, 34.

(101) Lord Rayleigh. On The Instability Of Jets. Proc. London Math. Soc. 1878, 1-10
(1), 4–13.

(102) Young, T. An Essay on the Cohesion of Fluids. Philos. Trans. R. Soc. 1805, 95,
65–87.

(103) Laplace, P.-S. Mechanique Celeste, Supplement Au X Libre. Cour. Paris 1805.

(104) Hoath, S. D. Fundamentals of Inkjet Printing: The Science of Inkjet and Droplets,
ISBN: 9783527337859; Hoath, S. D., Ed.; 2015.

(105) Döring, M. Ink-jet printing


http://www.extra.research.philips.com/hera/people/aarts/_Philips Bound
Archive/PTechReview/PTechReview-40-1982-192.pdf (accessed Mar 22, 2018).

(106) Hutchings, I. M.; Martin, G. D. Inkjet Technology for Digital Fabrication, ISBN:
978-0-470-68198-5; 2012.

(107) Le, H. P. Progress and Trends in Ink-Jet Printing Technology. J. Imaging Sci.
Technol. 1998, 49–62.

(108) Stokes, S. G. G. On the Effect of the Internal Friction of Fluids on the Motion of
Pendulums. Trans. Cambridge Philos. Soc. 1851, 9, 8–106.

199
Bibliography

(109) Reynolds, O. F. R. S. XXIX. An Experimental Investigation of the Circumstances


Which Determine Whether the Motion of Water Shall He Direct or Sinuous , and
of the Law of Resistance in Parallel Channels. Philos. Trans. R. Soc. London 1884,
174, 935–982.

(110) Graebel, W. Advanced Fluid Mechanics 1st Edition, ISBN: 9780080549088; 2007.

(111) Derby, B. Inkjet Printing of Functional and Structural Materials : Fluid Property
Requirements , Feature Stability , and Resolution. Annu. Rev. Mater. Res. 2010, 40
(1), 395–414.

(112) McKinley, G. H.; Renardy, M. Wolfgang von Ohnesorge. Phys. Fluids 2011, 23
(12).

(113) Jang, D.; Kim, D.; Moon, J. Influence of Fluid Physical Properties on Ink-Jet
Printability. Langmuir 2009, 25 (5), 2629–2635.

(114) Abbott, S. Ohnesorge for Inkjet https://www.stevenabbott.co.uk/practical-


coatings/ohnesorge.php (accessed Aug 9, 2017).

(115) Bonn, D.; Eggers, J.; Indekeu, J.; Meunier, J.; Rolley, E. Wetting and Spreading.
Rev. Mod. Phys. 2009, 81 (2), 739–805.

(116) Fowkes, F. M. Attractive Forces at Interfaces. Ind. Eng. Chem. 1964, 56 (12), 40–
52.

(117) Owens, Wendt, Rabel and Kaelble (OWRK) method


https://www.kruss.de/services/education-theory/glossary/owens-wendt-rabel-and-
kaelble-owrk-method/ (accessed Aug 15, 2017).

(118) Handbook of Digital Imaging, ISBN: 978-0-470-51059-9; Kriss, M., Ed.

(119) Khalate, A. A.; Bombois, X.; Ye, S.; Babuska, R.; Koekebakker, S. Minimization
of Cross-Talk in a Piezo Inkjet Printhead Based on System Identification and
Feedforward Control. J. Micromechanics Microengineering 2012.

(120) Kwon, K.-S. Waveform Design Methods for Piezo Inkjet Dispensers Based on
Measured Meniscus Motion. J. MICROELECTROMECHANICAL Syst. 2009, 18
(5), 1118–1125.

(121) Tanguy, L.; Liang, D.; Zengerle, R.; Koltay, P. arXiv:1210.4078: Droplet Break-up
with Negative Momentum Fluid Dynamics Videos. Phys. Fluid Dyn. 2012.

(122) Rioboo, R.; Marengo, M.; Tropea, C. Time Evolution of Liquid Drop Impact onto
Solid, Dry Surfaces. 2002, 33, 112–124.

(123) Rioboo, R.; Marengo, M.; Tropea, C. Outcomes from a Drop Impact on Solid
Surfaces. At. Sprays 2001, 11, 155–165.

(124) Clanet, C.; Béguin, C.; Richard, D.; Quéré, D. Maximal Deformation of an
Impacting Drop. J. Fluid Mech. 2004, 517, 199–208.

200
Bibliography

(125) Shah, V.; Wallace, D. B. Low-Cost Solar Cell Fabrication by Drop-on-Demand


Ink-Jet Printing. Proc. IMAPS 37th Annu. Int. Symp. Microelectron. 2004.

(126) Hoth, C. N.; Choulis, S. A.; Schilinsky, P.; Brabec, C. J. High Photovoltaic
Performance of Inkjet Printed Polymer:Fullerene Blends. Adv. Mater. 2007, 19
(22), 3973–3978.

(127) Hoth, C. N.; Schilinsky, P.; Choulis, S. a; Brabec, C. J. Printing Highly Efficient
Organic Solar Cells. Nano Lett. 2008, 8 (9), 2806–2813.

(128) Aernouts, T.; Aleksandrov, T.; Girotto, C.; Genoe, J.; Poortmans, J. Polymer Based
Organic Solar Cells Using Ink-Jet Printed Active Layers. Appl. Phys. Lett. 2008,
92.

(129) Hoth, C. N.; Choulis, S. A.; Schilinsky, P.; Brabec, C. J. On the Effect of poly(3-
Hexylthiophene) Regioregularity on Inkjet Printed Organic Solar Cells. J. Mater.
Chem. 2009, 19 (30), 5398.

(130) Eom, S. H.; Senthilarasu, S.; Uthirakumar, P.; Yoon, S. C.; Lim, J.; Lee, C.; Lim,
H. S.; Lee, J.; Lee, S. H. Polymer Solar Cells Based on Inkjet-Printed PEDOT:PSS
Layer. Org. Electron. 2009, 10, 536–542.

(131) Steirer, K. X.; Berry, J. J.; Reese, M. O.; Hest, M. F. A. M. Van; Miedaner, A.;
Liberatore, M. W.; Collins, R. T.; Ginley, D. S. Ultrasonically Sprayed and Inkjet
Printed Thin Film Electrodes for Organic Solar Cells. Thin Solid Films 2009, 517,
2781–2786.

(132) Lange, A.; Wegener, M.; Boeffel, C.; Fischer, B.; Wedel, A.; Neher, D. A New
Approach to the Solvent System for Inkjet-Printed P3HT:PCBM Solar Cells and
Its Use in Devices with Printed Passive and Active Layers. Sol. Energy Mater. Sol.
Cells 2010, 94, 1816–1821.

(133) Eom, S. H.; Park, H.; Mujawar, S. H.; Yoon, S. C.; Kim, S.-S.; Na, S.-I.; Kang, S.-
J.; Khim, D.; Kim, D.-Y.; Lee, S.-H. High Efficiency Polymer Solar Cells via
Sequential Inkjet-Printing of PEDOT:PSS and P3HT:PCBM Inks with Additives.
Org. Electron. 2010, 11, 1516–1522.

(134) Jeong, J.; Lee, J.; Kim, H.; Kim, H.; Na, S. Ink-Jet Printed Transparent Electrode
Using Nano-Size Indium Tin Oxide Particles for Organic Photovoltaics. Sol.
Energy Mater. Sol. Cells 2010, 94, 1840–1844.

(135) Teichler, A.; Eckardt, R.; Hoeppener, S.; Friebe, C.; Perelaer, J.; Senes, A.;
Morana, M.; Brabec, C. J.; Schubert, U. S. Combinatorial Screening of Polymer :
Fullerene Blends for Organic Solar Cells by Inkjet Printing. Adv. Energy Mater.
2011, 1, 105–114.

(136) Yu, J.-S.; Kim, I.; Kim, J.-S.; Jo, J.; Larsen-Olsen, T. T.; Søndergaard, R. R.;
Hösel, M.; Angmo, D.; Jørgensen, M.; Krebs, F. C. Silver Front Electrode Grids
for ITO-Free All Printed Polymer Solar Cells with Embedded and Raised
Topographies, Prepared by Thermal Imprint, Flexographic and Inkjet Roll-to-Roll
Processes. Nanoscale 2012, 4, 6032–6040.

201
Bibliography

(137) Galagan, Y.; Coenen, E. W. C.; Sabik, S.; Gorter, H. H.; Barink, M.; Veenstra, S.
C.; Kroon, J. M.; Andriessen, R.; Blom, P. W. M. Evaluation of Ink-Jet Printed
Current Collecting Grids and Busbars for ITO-Free Organic Solar Cells. Sol.
Energy Mater. Sol. Cells 2012, 104, 32–38.

(138) Angmo, D.; Sweelssen, J.; Andriessen, R.; Galagan, Y.; Krebs, F. C. Inkjet
Printing of Back Electrodes for Inverted Polymer Solar Cells. Adv. Energy Mater.
2013, 3 (9), 1230–1237.

(139) Lange, A.; Schindler, W.; Wegener, M.; Fostiropoulos, K.; Janietz, S. Inkjet
Printed Solar Cell Active Layers Prepared from Chlorine-Free Solvent Systems.
Sol. Energy Mater. Sol. Cells 2013, 109, 104–110.

(140) Jung, S.; Sou, A.; Banger, K.; Ko, D.-H.; Chow, P. C. Y.; McNeill, C. R.;
Sirringhaus, H. All-Inkjet-Printed, All-Air-Processed Solar Cells. Adv. Energy
Mater. 2014, 4 (14).

(141) Lu, H.; Lin, J.; Wu, N.; Nie, S.; Luo, Q.; Ma, C.-Q.; Cui, Z. Inkjet Printed Silver
Nanowire Network as Top Electrode for Semi-Transparent Organic Photovoltaic
Devices. Appl. Phys. Lett. 2015, 106, 093302.

(142) Eggenhuisen, T. M.; Galagan, Y.; Coenen, E. W. C.; Voorthuijzen, W. P.; Slaats,
M. W. L.; Kommeren, S. A.; Shanmuganam, S.; Coenen, M. J. J.; Andriessen, R.;
Groen, W. A. Digital Fabrication of Organic Solar Cells by Inkjet Printing Using
Non-Halogenated Solvents. Sol. Energy Mater. Sol. Cells 2015, 134, 364–372.

(143) Eggenhuisen, T. M.; Galagan, Y.; Biezemans, A. F. K. V.; Slaats, T. M. W. L.;


Voorthuijzen, W. P.; Kommeren, S.; Shanmugam, S.; Teunissen, J. P.; Hadipour,
A.; Verhees, W. J. H.; et al. High Efficiency, Fully Inkjet Printed Organic Solar
Cells with Freedom of Design. J. Mater. Chem. A 2015, 3, 7255–7262.

(144) Singh, A.; Gupta, S. K.; Grag, A. Inverted Polymer Bulk Heterojunction Solar
Cells with Ink-Jet Printed Electron Transport and Active Layers. Org. Electron.
2016, 35, 118–127.

(145) Sankaran, S.; Glaser, K.; Gaertner, S.; Roedlmeier, T.; Sudau, K.; Hernandez-Sosa,
G.; Colsmann, A. Fabrication of Polymer Solar Cells from Organic Nanoparticle
Dispersions by Doctor Blading or Ink-Jet Printing. Org. Electron. 2016, 28, 118–
122.

(146) Kommeren, S.; Coenen, M. J. J.; Eggenhuisen, T. M.; Slaats, T. W. L.; Gorter, H.;
Groen, P. Combining Solvents and Surfactants for Inkjet Printing PEDOT:PSS on
P3HT/PCBM in Organic Solar Cells. Org. Electron. 2018, 61, 282–288.

(147) Georgiou, E.; Choulis, S. A.; Hermerschmidt, F.; Pozov, S. M.; Christodoulou, C.;
Schider, G.; Kreissl, S.; Ward, R.; List-kratochvil, E. J. W.; Boeffel, C. Printed
Copper Nanoparticle Metal Grids for Cost-Effective ITO-Free Solution Processed
Solar Cells. Sol. RRL 2018, 2 (3), 1–8.

(148) PiXDRO B.V. User manual / Documentation PiXDRO Inkjet Printing systems
https://www.meyerburger.com/de/de/meyer-burger/produkte-und-
systeme/detail/pixdro-lp50/ (accessed Apr 18, 2018).
202
Bibliography

(149) Fujifilm print heads


http://www.fujifilmusa.com/products/industrial_inkjet_printheads/print-
products/printheads/index.html (accessed Aug 27, 2017).

(150) Shang, J.; Flury, M.; Harsh, J. B.; Zollars, R. L. Comparison of Different Methods
to Measure Contact Angles of Soil Colloids. J. Colloid Interface Sci. 2008, 328 (2),
299–307.

(151) del Rio, O. I.; Neumann, A. W. Axisymmetric Drop Shape Analysis :


Computational Methods for the Measurement of Interfacial Properties from the
Shape and Dimensions of Pendant and Sessile Drops. J. Colloid Interface Sci.
1997, 196 (2), 136–147.

(152) Gary Tuttle, Dept. of Electrical and Computer Engineering, I. S. U. Contact


resistance and TLM measurements
http://tuttle.merc.iastate.edu/ee432/topics/metals/tlm_measurements.pdf (accessed
Aug 30, 2017).

(153) Acciani, G.; Falcone, O.; Vergura, S. Analysis of the Thermal Heating of Poly-Si
and a-Si Photovoltaic Cell by Means of Fem. Renew. Energy Power Qual. J. 2010,
1 (08), 1240–1244.

(154) Pieters, B. E. Spatial Modeling of Thin-Film Solar Modules Using the Network
Simulation Method and SPICE. IEEE J. Photovoltaics 2011, 1 (1), 93–98.

(155) Koishiyev, G. T. Doctoral Thesis: Analysis of Impact of Non-Uniformities on


Thin-Film Solar Cells and Modules with 2-D Simulations. Thesis, Colorado State
University. Libraries, 2010.

(156) Fecher, F. W. Doctoral Thesis: Simulation of Thin-Film Photovoltaic Modules: 2D


and 3D Spatially Resolved Electrical and Electrothermal Finite Element
Calculations, 2018.

(157) Slooff, L. H.; Burgers, A. R.; Bende, E. E.; Veenstra, S. C.; Kroon, J. M. Parameter
Study for Polymer Solar Modules Based on Various Cell Lengths and Light
Intensities. 28th Eur. Photovolt. Sol. Energy Conf. Exhib. 2013.

(158) Fecher, F. W.; Brabec, C. J.; Romero, A. P.; Buerhop-lutz, C. Influence of a Shunt
on the Electrical Behavior in Thin Film Photovoltaic Modules – A 2D Finite
Element Simulation Study. Sol. Energy 2014, 105, 494–504.

(159) Maisch, P.; Tam, K. C.; Schilinsky, P.; Egelhaaf, H.-J.; Brabec, C. J. Shy Organic
Photovoltaics: Digitally Printed Organic Solar Modules with Hidden Interconnects.
Sol. RRL 2018, 2 (7), 1–9.

(160) Lamont, C. A.; Eggenhuisen, T. M.; Coenen, M. J. J.; Slaats, T. W. L.; Andriessen,
R.; Groen, P. Tuning the Viscosity of Halogen Free Bulk Heterojunction Inks for
Inkjet Printed Organic Solar Cells. Org. Electron. 2015, 17, 107–114.

(161) Chueh, C.-C.; Yaoa, K.; Yip, H.-L.; Chang, C.-Y.; Xu, Y.-X.; Chen, K.-S.; Li, C.-
Z.; Liu, P.; Huang, F.; Chen, Y.; et al. Non-Halogenated Solvents for
Environmentally Friendly Processing of High-Performance Bulk-Heterojunction

203
Bibliography

Polymer Solar Cells. Energy Environ. Sci. 2013, 6, 3241–3248.

(162) Schmidt-hansberg, B.; Sanyal, M.; Grossiord, N.; Galagan, Y.; Baunach, M.;
Klein, M. F. G.; Colsmann, A.; Scharfer, P.; Lemmer, U.; Dosch, H.; et al.
Investigation of Non-Halogenated Solvent Mixtures for High Throughput
Fabrication of Polymer – Fullerene Solar Cells. Sol. Energy Mater. Sol. Cells 2012,
96, 195–201.

(163) Abbott, S.; Hansen, C. M. Hansen Solubility Parameters in Practice, ISBN: 978-0-
9551220-2-6; 2008.

(164) Maisch, P.; Tam, K. C.; Lucera, L.; Egelhaaf, H.-J.; Scheiber, H.; Maier, E.;
Brabec, C. J. Inkjet Printed Silver Nanowire Percolation Networks as Electrodes
for Highly Efficient Semitransparent Organic Solar Cells. Org. Electron. 2016.

(165) Hebner, T. R.; Wu, C. C.; Marcy, D.; Lu, M. H.; Sturm, J. C. Ink-Jet Printing of
Doped Polymers for Organic Light Emitting Devices. Appl. Phys. Lett. 1998, 72
(5), 519–521.

(166) Sirringhaus, H.; Kawase, T.; Friend, R. H.; Shimoda, T.; Inbasekaran, M.; Wu, W.;
Woo, E. P. High-Resolution Ink-Jet Printing of All-Polymer Transistor Circuits.
Science (80-. ). 2000, 290 (5499), 2123–2126.

(167) Gamerith, S.; Klug, A.; Scheiber, H.; Scherf, U.; Moderegger, E.; List, E. J. W.
Direct Ink-Jet Printing of Ag-Cu Nanoparticle and Ag-Precursor Based Electrodes
for OFET Applications. Adv. Funct. Mater. 2007, 17, 3111–3118.

(168) Chan, C.; Ko, T.; Hiraoka, H. Polymer Surface Modification by Plasmas and
Photons. Surf. Sci. Rep. 1996, 24, 1–54.

(169) Lee, S.-H.; Hwang, J. Y.; Kang, K.; Kang, H. Fabrication of Organic Light
Emitting Display Using Inkjet Printing Technology. 2009 Int. Symp.
Optomechatronic Technol. 2009, 71–76.

(170) Zhang, Z.; Zhang, X.; Xin, Z.; Deng, M.; Wen, Y.; Song, Y. Controlled Inkjetting
of a Conductive Pattern of Silver Nanoparticles Based on the Coffee-Ring Effect.
Adv. Mater. 2013, 24, 6714–6718.

(171) Bromberg, V.; Ma, S.; Singler, T. J. High-Resolution Inkjet Printing of Electrically
Conducting Lines of Silver Nanoparticles by Edge-Enhanced Twin-Line
Deposition. Appl. Phys. Lett. 2013, 102, 2141011–2141014.

(172) Cassie, A. B. D.; Baxter, S. Wettability of Porous Surfaces. Trans. Faraday Soc.
1944, No. 40, 546–551.

(173) Raj, R.; Enright, R.; Zhu, Y.; Adera, S.; Wang, E. N. Unified Model for Contact
Angle Hysteresis on Heterogeneous and Superhydrophobic Surfaces. Langmuir
2012, 28 (45), 15777–15788.

(174) Raj, R.; Adera, S.; Enright, R.; Wang, E. N. High-Resolution Liquid Patterns via
Three-Dimensional Droplet Shape Control. Nat. Commun. 2014, 5, 1–8.

204
Bibliography

(175) Wang, J. Z.; Zheng, Z. H.; Li, H. W.; Huck, W. T. S.; Sirringhaus, H. Dewetting of
Conducting Polymer Inkjet Droplets on Patterned Surfaces. Nat. Mater. 2004, 3,
171–176.

(176) Chen, S.; Su, M.; Zhang, C.; Gao, M.; Bao, B.; Yang, Q.; Su, B.; Song, Y.
Fabrication of Nanoscale Circuits on Inkjet-Printing Patterned Substrates. Adv.
Mater. 2015, 27 (26), 3928–3933.

(177) Colsmann, A.; Reinhard, M.; Kwon, T.; Kayser, C.; Nickel, F.; Czolk, J.; Lemmer,
U.; Clark, N.; Jasieniak, J.; Holmes, A. B.; et al. Inverted Semi-Transparent
Organic Solar Cells with Spray Coated , Surfactant Free Polymer Top-Electrodes.
Sol. Energy Mater. Sol. Cells 2012, 98, 118–123.

(178) Savva, A.; Neophytou, M.; Koutsides, C.; Kalli, K.; Stelios, C. A. Synergistic
Effects of Buffer Layer Processing Additives for Enhanced Hole Carrier
Selectivity in Inverted Organic Photovoltaics. Org. Electron. 2013, 14, 3123–3130.

(179) Lim, F. J.; Ananthanarayanan, K.; Lutherbc, J.; Wei Ho, G. Influence of a Novel
Fluorosurfactant Modified PEDOT : PSS Hole Transport Layer on the
Performance of Inverted Organic Solar Cells. J. Mater. Chem. 2012, No. 48.

(180) Colsmann, A.; Stenzel, F.; Balthasar, G.; Do, H.; Lemmer, U. Plasma Patterning of
Poly (3,4-Ethylenedioxythiophene): Poly( Styrenesulfonate) Anodes for Efficient
Polymer Solar Cells. Thin Solid Films 2009, 517 (5), 1744–1746.

(181) Tan, Z.; Moghaddam, R. S.; Lai, M. L.; Docampo, P.; Higler, R.; Price, M.;
Sadhanala, A.; Pazos, L. M.; Credgington, D.; Hanusch, F.; et al. Bright Light -
Emitting Diodes Based on Organometal Halide Perovskite. Nat. Nanotechnol.
2014, 9, 687–692.

(182) Jäckle, S.; Liebhaber, M.; Gersmann, C.; Mews, M.; Jäger, K.; Christiansen, S.;
Lips, K. Potential of PEDOT:PSS as a Hole Selective Front Contact for Silicon
Heterojunction Solar Cells. Nat. Sci. Reports 2017, 7, 1–8.

(183) Lövenich, W. PEDOT — Properties and Applications. Polym. Sci. Ser. C 2014, 56
(1), 135–143.

(184) Maisch, P.; Eisenhofer, L. M.; Tam, K. C.; Distler, A.; Voigt, M. M.; Brabec, C. J.;
Egelhaaf, H.-J. A Generic Surfactant-Free Approach to Overcome Wetting
Limitations and Its Application to Improve Inkjet-Printed P3HT:non-Fullerene
Acceptor PV. J. Mater. Chem. A (submitted).

(185) Giesy, J. P.; Kannan, K. Perfluorochemical Surfactants in the Environment.


Environ. Sci. Technol. 2002, 36 (7), 146A – 152A.

(186) Ozawa, K.; Usui, T.; Ide, K.; Takahashi, H.; Sakai, S. Development of a Femtoliter
Piezo Ink-Jet Head for High Resolution Printing. Int. Conf. Digit. Print. Technol.
2007, 898–901.

(187) Perelaer, J.; Smith, P. J.; Bosch, E. Van Den; Grootel, S. S. C. Van; Ketelaars, P.
H. J. M.; Schubert, U. S. The Spreading of Inkjet-Printed Droplets with Varying
Polymer Molar Mass on a Dry Solid Substrate. Macromol. Chem. Phys. 2009, 210

205
Bibliography

(6), 495–502.

(188) Zhang, J.; Xie, S.; Lu, Z.; Wu, Y.; Xiao, H.; Zhang, X.; Li, G.; Li, C.; Chen, X.;
Ma, W.; et al. Influence of Substrate Temperature on the Film Morphology and
Photovoltaic Performance of Non-Fullerene Organic Solar Cells. Sol. Energy
Mater. Sol. Cells 2018, 174, 1–6.

(189) Huang, J.; Yang, C.; Ho, Z.; Kekuda, D.; Wu, M.; Fan-Ching, C.; Chen, P.; Chu,
C.-W.; Ho, K.-C. Annealing Effect of Polymer Bulk Heterojunction Solar Cells
Based on Polyfluorene and Fullerene Blend. Org. Electron. 2009, 10, 27–33.

(190) Moulé, A. J.; Meerholz, K. Controlling Morphology in Polymer – Fullerene


Mixtures. Adv. Mater. 2008, 20, 240–245.

(191) Liu, X.; Huettner, S.; Rong, Z.; Sommer, M.; Friend, R. H. Solvent Additive
Control of Morphology and Crystallization in Semiconducting Polymer Blends.
Adv. Mater. 2012, 24, 669–674.

(192) Machui, F. Doctoral Thesis: Formulation of Semiconductor Solutions for Organic


Photovoltaic Devices, Friedrich-Alexander-Universität Erlangen-Nürnberg, 2014.

(193) Cazabat, A. M.; Heslot, F.; Troian, S. M.; Carles, P. Fingering Instability of Thin
Spreading Films Driven by Temperature Gradients. Nature 1990, 346, 824–826.

(194) Isojärvi, T. Unstable Thin Film Spreading of Volatile Binary Mixtures,University


of Jyväskylä https://jyx.jyu.fi/dspace/handle/123456789/42570 (accessed Mar 16,
2018).

(195) Pesach, D.; Marmur, A. Marangoni Effects in the Spreading of Liquid Mixtures on
a Solid. Langmuir 1987, 3 (4), 519–524.

(196) Hu, H.; Larson, R. G. Analysis of the Effects of Marangoni Stresses on the
Microflow in an Evaporating Sessile Droplet. Langmuir 2005, 21 (9), 3972–3980.

(197) Marrocchi, A.; Lanari, D.; Facchetti, A.; Vaccaro, L. Poly(3-Hexylthiophene):


Synthetic Methodologies and Properties in Bulk Heterojunction Solar Cells.
Energy Environ. Sci. 2012, 5, 8457–8474.

(198) Krebs, F. C.; Gevorgyan, S. A.; Alstrup, J. A Roll-to-Roll Process to Flexible


Polymer Solar Cells : Model Studies , Manufacture and Operational Stability
Studies. J. Mater. Chem. 2009, 19, 5442–5451.

(199) Hösel, M.; Søndergaard, R. R.; Jørgensen, M.; Krebs, F. C. Fast Inline Roll-to-Roll
Printing for Indium-Tin-Oxide- Free Polymer Solar Cells Using Automatic
Registration. Energy Technol. 2013, 1 (1), 102–107.

(200) Maisch, P.; Tam, K. C.; Fecher, F. W.; Egelhaaf, H.-J.; Scheiber, H.; Maier, E.;
Brabec, C. J. Inkjet Printing of Highly Conductive Nanoparticle Dispersions for
Organic Electronics. In Molded Interconnect Devices (MID), 2016 12th
International Congress; 2016.

(201) Hoppe, H.; Seeland, M.; Muhsin, B. Optimal Geometric Design of Monolithic

206
Bibliography

Thin-Film Solar Modules : Architecture of Polymer Solar Cells. Sol. Energy


Mater. Sol. Cells 2012, 97, 119–126.

(202) Hu, L.; Kim, H. S.; Lee, J.-Y.; Peumans, P.; Cui, Y. Scalable Coating and
Properties of Transparent, Flexible, Silver Nanowire Electrodes. ACS Nano 2010, 4
(5), 2955–2963.

(203) Yu, Z.; Zhang, Q.; Li, L.; Chen, Q.; Niu, X.; Liu, J.; Pei, Q. Highly Flexible Silver
Nanowire Electrodes for Shape-Memory Polymer Light-Emitting Diodes. Adv.
Mater. 2011, 23 (5), 664–668.

(204) Ramírez Quiroz, C. O.; Levchuk, I.; Bronnbauer, C.; Salvador, M.; Forberich, K.;
Huemueller, T.; Hou, Y.; Schweizer, P.; Spiecker, E.; Brabec, C. J. Pushing
Efficiency Limits for Semitransparent Perovskite Solar Cells. J. Mater. Chem. A
2015, 24071–24081.

(205) Finn, D. J.; Lotya, M.; Coleman, J. N. Inkjet Printing of Silver Nanowire
Networks. ACS Appl. Mater. Interfaces 2015, 7 (17), 9254–9261.

(206) Wu, J.-T.; Lien-Chung Hsu, S.; Tsai, M.-H.; Liu, Y.-F.; Hwang, W.-S. Direct Ink-
Jet Printing of Silver Nitrate–silver Nanowire Hybrid Inks to Fabricate Silver
Conductive Lines. J. Mater. Chem. 2012, 22 (31), 15599.

(207) Emmott, C. J. M.; Urbina, A.; Nelson, J. Environmental and Economic Assessment
of ITO-Free Electrodes for Organic Solar Cells. Sol. Energy Mater. Sol. Cells
2012, 97, 14–21.

(208) Gaudiana, R.; Brabec, C. Organic Materials: Fantastic Plastic. Nat. Photonics
2008, 2 (5), 287–289.

(209) Tam, K. C.; Maisch, P.; Wagner, M.; Egelhaaf, H.-J.; Brabec, C. J. Fully Solution
Processed Highly Efficient Organic Solar Modules Consisting of Only Three
Layers. Energy Environ. Sci. in preparation.

(210) Schroder, D. K. Semiconductor Material and Device Characterization, ISBN: 978-


0-471-73906-7, 3rd ed.; 2006.

(211) Shirakawa, H.; McDiarmid, A.; Heeger, A. J. Twenty-Five Years of Conducting


Polymers. Chem. Communciations 2003, 1–4.

(212) Allemand, P.-M.; Gupta, R.; Egelhaaf, H.-J.; Chen, P.-C.; Wagner, M.; Maisch, P.;
Pichler, K.; Tam, K. C.; Kubis, P. Organic Solar Module And/or Fabrication
Method, Provisional US Patent, US62/521,696, 2017.

(213) Zhao, W.; Li, S.; Yao, H.; Zhang, S.; Zhang, Y.; Yang, B.; Hou, J. Molecular
Optimization Enables over 13% Efficiency in Organic Solar Cells. J. Am. Chem.
Soc. 2017, 139 (21), 7148–7151.

(214) Heliatek. Heliatek sets new Organic Photovoltaic world record efficiency of 13.2%
http://www.heliatek.com/en/press/press-releases/details/heliatek-sets-new-organic-
photovoltaic-world-record-efficiency-of-13-2 (accessed Mar 22, 2018).

207
Bibliography

(215) Amin, A.; Sytnyk, M.; Killilea, N.; Maisch, P.; Tam, K. C.; Katharina; Poulsen;
Niehaus, J.; Heiss, W. Poster Presentation: Fully Inkjet Printed PbS Nanocrystal
Photodetectors. NaNaX 8 — Nanosci. with Nanocrystals 2017.

(216) IBC SOLAR AG, Jura-Solarpark https://www.ibc-blog.de/wp-


content/uploads/2014/02/Jura-Solarpark.jpg (accessed Mar 19, 2018).

(217) Seok, S. Il; Grätzel, M.; Park, N. Methodologies toward Highly Efficient
Perovskite Solar Cells. Small 2018, 1–17.

(218) Chen, W.; Wu, Y.; Yue, Y.; Liu, J.; Zhang, W.; Yang, X.; Chen, H.; Bi, E.;
Ashraful, I.; Graetzel, M.; et al. Efficient and Stable Large-Area Perovskite Solar
Cells with Inorganic Charge Extraction Layers. Science (80-. ). 2015, 350 (6263),
944–948.

(219) Hao, F.; Stoumpos, C. C.; Cao, D. H.; Chang, R. P. H.; Kanatzidis, M. G. Lead-
Free Solid-State Organic-Inorganic Halide Perovskite Solar Cells. Nat. Photonics
2014, 8, 489–494.

(220) Saule Technologies http://sauletech.com/ (accessed May 19, 2018).

208
Appendix

a)

b)

209
Appendix

c)

d)

Figure A1: a) Wetting envelopes (𝜃 = 0 °) showing ink-surface combinations applied in


this work. Good wetting is predicted and experimentally confirmed for a) Ag
nanoparticle- or AgNW inks on glass substrates, b) ZnO on ITO or AgNW layers, c)
active layer inks P3HT:PC60BM/PV2000:PC70BM on ZnO layers and d) Ag nanoparticle-
or AgNW inks on PEDOT:PSS layers. The challenging wetting behaviour of PEDOT:PSS
inks on the active layer is discussed in chapter 4.3.

210
List of Publications

Publications in peer-reviewed journals:


[P1] F. Machui, P. Maisch, S. Langner, J. Krantz, T. Ameri, C.J. Brabec, Classification
of Additives for Organic Photovoltaic Devices, ChemPhysChem. (2015).
doi:10.1002/cphc.201402734.
[P2] P. Maisch, K.C. Tam, L. Lucera, H.-J. Egelhaaf, H. Scheiber, E. Maier, et al.,
Inkjet printed silver nanowire percolation networks as electrodes for highly
efficient semitransparent organic solar cells, Org. Electron. (2016).
doi:10.1016/j.orgel.2016.08.006.
[P3] P. Maisch, K.C. Tam, P. Schilinsky, H.-J. Egelhaaf, C.J. Brabec, Shy Organic
Photovoltaics: Digitally Printed Organic Solar Modules with Hidden Interconnects,
Sol. RRL 2, 1-9 (2018). doi: 10.1002/solr.201800005.
[P4] P. Maisch, L.M. Eisenhofer, K.C. Tam, A. Distler, M.M. Voigt, C.J. Brabec, H.-J.
Egelhaaf, A generic surfactant-free approach to overcome wetting limitations and
its application to improve inkjet-printed P3HT:non-fullerene acceptor PV, J. Mater.
Chem. A. (submitted)
[P5] A. Karl, A.Osvet, A.Vetter, P. Maisch, N. Li, H.-J. Egelhaaf, C.J. Brabec,
Discriminating bulk versus interface shunts in organic solar cells by advanced
imaging techniques, Prog. Photovoltaics Res. Appl. 1-9 (2019).
doi:org/10.1002/pip.3121.
[P6] A.A. Yousefi-Amin, N. Killilea, M. Sytnyk, P. Maisch, K.C. Tam, et al., Fully
Printed Infrared Photodetectors from PbS Nanocrystals with Perovskite Ligands,
ACS Nano, 13, 2389–2397 (2019), doi: 10.1021/acsnano.8b09223.

[P7] K.C. Tam, P. Maisch, M. Wagner, H.-J. Egelhaaf, C.J. Brabec, Fully solution
processed highly efficient organic solar modules consisting of only three layers (in
preparation).
[P8] K.C. Tam, P. Kubis, P. Maisch, H.-J. Egelhaaf, C.J. Brabec, Fully printed organic
solar modules with bottom and top silver nanowire electrodes (in preparation).

Proceedings of conferences:

[Pr1] P. Maisch, K.C. Tam, F.W. Fecher, H.-J. Egelhaaf, H. Scheiber, E. Maier, et al.,
Inkjet printing of highly conductive nanoparticle dispersions for organic
electronics, in: Molded Interconnect Devices (MID), 2016 12th Int. Congr.,
Würzburg/GER, 2016. doi:10.1109/ICMID.2016.7738932.
Awarded with the price for best paper
[Pr 2] P. Maisch, K.C. Tam, L. Lucera, F.W. Fecher, H.-J. Egelhaaf, H. Scheiber, et al.,
Inkjet Printing of Semitransparent Electrodes for Photovoltaic Applications, in:
Proc. SPIE Vol. 9942 99420R-1, Opt. + Photonics, San Diego/USA, 2016.

211
List of Publications

Invited book chapters:

[B1] P. Maisch, L. Lucera, C.J. Brabec, H.-J. Egelhaaf, Flexible Solar Cells, in: Flex.
Carbon Based Electron., Wiley VCH, ISBN: 978-3-527-34191-7, 2018.

Oral presentations:

[O1] P. Maisch, K.C. Tam, L. Lucera, P. Kubis, H.-J. Egelhaaf, C.J. Brabec, Scalable
organic solar module fabrication– a comparison of digitally printed and LASER
patterned interconnects, in: MRS Fall Meet. Boston/USA, 2016.

[O2] P. Maisch, K.C. Tam, F.W. Fecher, H.-J. Egelhaaf, H. Scheiber, E. Maier, et al.,
Inkjet printing of highly conductive nanoparticle dispersions for organic
electronics, in: Molded Interconnect Devices (MID), 12th Int. Congr.,
Würzburg/GER, 2016.
[O3] P. Maisch, K.C. Tam, L. Lucera, F.W. Fecher, H.-J. Egelhaaf, H. Scheiber, et al.,
Inkjet Printing of Semitransparent Electrodes for Photovoltaic Applications, in:
Proc. SPIE Vol. 9942 99420R-1, Opt. + Photonics, San Diego/USA, 2016.
[O4] P. Maisch, M.M. Voigt, S. Keilwitz, C.J. Brabec, Inkjet Printing of Organic
Photovoltaics, in: Large-Area, Oranic Print. Electron. Conv. (LOPEC), Messe
München/GER, 2015.

Poster presentations:

[Po1] P. Maisch, K.C. Tam, H.-J. Egelhaaf, C.J. Brabec, Inkjet Printed Semitransparent
Organic Solar Cells and Modules, in: Next Gener. Sol. Energy Meets
Nanotechnol., Erlangen/GER, 2016.
[Po2] P. Maisch, S. Keilwitz, F.W. Fecher, M. Müller, K.C. Tam, M.M. Voigt, et al.,
Inkjet Printing as Manufacturing Method for Organic Solar Cells and Moduels, in:
11th Int. Conf. Oragnic Electron., Erlangen/GER, 2015.
[Po3] P. Maisch, S. Keilwitz, F.W. Fecher, M. Müller, K.C. Tam, M.M. Voigt, et al.,
Inkjet Printed Organic Solar Cells and Moduels, in: 4th Congr. Org. Print.
Photovoltaics, Würzbg/GER, 2015.

Patent applications:

[Pa1] P.-M. Allemand, R. Gupta, H.-J. Egelhaaf, P.-C. Chen, M. Wagner, P. Maisch, et
al., Organic Solar Module and/or Fabrication Method, provisional US patent,
US62/521,696, 2017.

212

Das könnte Ihnen auch gefallen