Surface Functionalization with Amino-based Self-Assembled Monolayers

Tailoring Electrode/Molecule Interfaces for Organic Electronics

Von der Fakultät für Mathematik, Informatik und Naturwissenschaften der RWTH Aachen University zur Erlangung des akademischen Grades eines Doktors der Naturwissenschaften genehmigte Dissertation

vorgelegt von

Dipl.-Phys.

Dominik Meyer

aus Geilenkirchen, Deutschland

Berichter: Univ.-Prof. Dr. rer. nat. Matthias Wuttig Jun.-Prof. Dr. rer. nat. Riccardo Mazzarello

Tag der mündlichen Prüfung: 12.06.2015

Diese Dissertation ist auf den Internetseiten der Hochschulbibliothek online verfügbar.

Abstract

Control of the energetics at interfaces is one of the key steps that improve the efficiency of light-emitting and energy-conversion devices based on organic materials. In this context, the present work is embedded in the field of organic electronics with a special focus on the elec- trode/semiconductor interface. Efficient electron injection and extraction is still a crucial mech- anism that deserves improvement. Electron transport through this interface is dominated by the energetic injection barrier ∆e. In order to gain control of this barrier, polar molecules have been used in this thesis to tailor the work function of the electrode. Three topics are addressed: (i) Un- derstanding the chemical and electrical interface of N,N-dialkyl dithiocarbamates (CnDTC) on the noble metals Au, Ag and Cu. (ii) Work function tuning using tailor-made dithiocarbamates. (iii) Introduction of first results to generate low work functions of the transparent conductive indium tin oxide (ITO) based on a modified chemical structure of DTCs. To study these systems, a combina- tion of X-ray and ultra-violet photoemission spectroscopy (XPS/UPS) with additional results from density functional theory (DFT) has been employed. Charge transport in electronic devices has been investigated using organic thin film transistors based on the n-type material PTCDI-C13.

(i) Prior to this thesis, the class of dithiocarbamates (DTC) have been reported by our research group to produce very low work functions of gold down to 3.2 eV. This result marks the start- ing point for the present work. The assembly of N,N-dialkyl dithiocarbamates on the three noble metals Au, Ag and Cu has been analyzed with XPS. Therein, densely packed monolayers are iden- tified on Au and Ag, while a disordered layer is observed on Cu. Nonetheless, a common work function of 3.5 eV is revealed for CnDTC with n 4 on all metals with UPS. Only those molecules ≥ with the shortest alkyl chains, i.e. n {1,2}, were able to achieve 3.2 eV on Au and Ag. The final = work function of the modified metals results from a complex interplay of molecular arrangement (surface coverage, orientation), interface dipole formation and the energy level alignment at the interface. The detailed comparison of C2DTC as a model system on the three noble metals gen- erates a deeper insight into the chemical and electronic structure. Based on experimental and theoretical results, a common hybridization of the frontier S 3p orbital with those of the frontier

iii Abstract

metal d-states is identified to generate a metal/molecule interface state, that remains irrespective of the initial metal work function.

(ii) A new set of dithiocarbamates are introduced and the effect of different polar groups attached to the molecular backbone is discussed. The molecular structure of dithiocarbamates has been modified using, among others, aromatic and fluorinated substitutions to induce different dipole moments. These directly affect the resulting work function of a modified electrode and enables the generation of a broad range of available work function values for noble metals using differently substituted dithiocarbamates.

(iii) At last, the step towards modification of transparent conductive oxides based on the molec- ular structure of dithiocarbamates is explored for the commonly used indium tin oxide (ITO) as substrate. For metal oxides, a different linker group is required to bind to this kind of surface. Therefore, amine based carboxylic acids have been investigated as first candidates to lower the work function of ITO. A reduction down to Φ 4.0eV is demonstrated. This result may serve as ≈ guideline to create new surface modifiers that should generate even lower work functions similar to those of dithiocarbamates on noble metals.

iv Kurzfassung

Kontrolle über die energetischen Prozesse an Grenzflächen ist einer der zentralen Schritte zur Effizienzsteigerung von Licht-emittierenden und -absorbierenden Bauelementen zur Energieer- zeugung auf Basis der organischen Materialklasse. In diesem Kontext gliedert sich die vorlie- gende Arbeit in das Gebiet der organischen Elektronik ein mit besonderem Fokus auf die Elek- trode/Halbleiter Grenzfläche. Effiziente Injektion und Extraktion von Elektronen ist heutzutage immer noch ein kritischer Mechanismus, welcher optimiert werden muss. Der Transport von Elek- tronen durch diese Grenzfläche wird durch die energetische Injektionsbarriere ∆e dominiert. Um die Kontrolle über diese Barriere zu gewinnen, wurden polare Moleküle in dieser Arbeit verwen- det, die eine gezielte Anpassung der Austrittsarbeit der Elektrode erlauben. Dabei wurden drei zentrale Themen behandelt: (i) Verständnis der chemischen und elektronischen Grenzfläche von

N,N-dialkyl Dithiocarbamaten (CnDTC) auf den Edelmetallen Au, Ag und Cu. (ii) Einstellung der Austrittsarbeit mittels maßgeschneiderten Dithiocarbamaten. (iii) Vorstellung von ersten Er- gebnissen zur Generation von niedrigen Austrittsarbeiten von transparentem, leitfähigem Indium Zinn Oxid (ITO) auf Basis einer modifizierten chemischen Struktur des DTC. Um diese Systeme zu studieren, wurde eine Kombination von Photoelektronenspektroskopie (XPS/UPS) mit Simu- lationsrechnungen basierend auf Dichtefunktionaltheorie (DFT) verwendet. Ladungstransport in elektronischen Bauelementen wurde anhand organischer Dünnschichttransistoren untersucht, wel- che mit dem n-typ Material PTCDI-C13 betrieben wurden.

(i) Vor Anfertigung dieser Arbeit wurde von unserer Forschungsgruppe berichtet, dass die Klasse der Dithiocarbamate (DTC) sehr niedrige Austrittsarbeiten von Gold bis zu 3.2 eV produzieren. Diese Ergebnisse markieren den Startpunkt für diese Arbeit. Die Bindung von N,N-dialkyl Di- thiocarbamaten auf den drei Edelmetallen Au, Ag und Cu wurde mittels XPS untersucht. Dabei wurden dicht gepackte Monolagen auf Au und Ag identifiziert, während eine ungeordnete La- ge auf Cu beobachtet wird. Trotzdem wird eine identische Austrittsarbeit von 3.5 eV für CnDTC mit n 4 auf allen Metallen mit UPS beobachtet. Nur Moleküle mit den kürzesten Alkylket- ≥ ten, d.h. n {1,2}, ermöglichen Werte von 3.2 eV auf Au und Ag. Die finale Austrittsarbeit der =

v Kurzfassung

modifizierten Metalle resultiert von einem komplexen Zusammenspiel von molekularer Anord- nung (Bedeckungsdichte, Orientierung), Grenzflächendipol und Angleichung der Energielevel an der Grenzfläche. Ein detaillierter Vergleich des C2DTC als Modellsystem auf den drei Metallen ermöglicht einen tieferen Einblick in die chemische und elektronische Struktur. Basierend auf ex- perimentellen und theoretischen Ergebnissen kann eine gleiche Hybridisierung des molekularen S 3p Orbitals mit denen der energetisch höchstbesetzten, metallischen d-Bänder identifiziert wer- den. Dies resultiert in einem Metall/Molekül Grenzflächenzustand, welcher unabhängig von der ursprünglichen Austrittsarbeit des Metalls ist.

(ii) Ein neuer Satz von Dithiocarbamaten wird vorgestellt und der Effekt von verschiedenen, pola- ren Endgruppen diskutiert. Die molekulare Struktur von Dithiocarbamaten wurde unter anderem mit aromatischen und fluorierten Substitutionen modifiziert, um unterschiedlich polare Momente des Moleküls zu erzeugen. Diese haben direkten Einfluss auf die resultierende Austrittsarbeit der modifizierten Elektrode und ermöglichen ein breites Fenster von verfügbaren Austrittsarbeiten für Edelmetalle auf Basis modifizierter Dithiocarbamate.

(iii) Zuletzt wird der Schritt zur Modifikation von transparenten, leitfähigen Oxiden auf Basis der molekularen Struktur der Dithiocarbamate vollzogen. Dabei wird das häufig verwendete Indium Zinn Oxid (ITO) als Substrat genutzt. Zur Modifikation von Oxiden ist eine alternative, moleku- lare Bindungsgruppe nötig, um eine chemische Reaktion mit der Oberfläche einzugehen. Deshalb wurden hier Amino-substituierte Carboxylsäuren zur Modifikation verwendet. Damit konnte eine Reduzierung auf Werte bis zu 4.0 eV demonstriert werden. Diese ersten Resultate dienen als Weg- weiser zur Identifikation neuer Moleküle, um noch geringere Austrittsarbeiten im Bereich derer von Dithiocarbamaten auf Edelmetallen zu generieren.

vi Contents

Table of Contents ix

1 Introduction1

I What Governs Electronic Processes at Interfaces to Organic Materials? A Theoretical Introduction5

2 Materials and Theoretical Background7 2.1 Molecular Materials...... 7 2.2 Fundamental Processes at Metal/Molecule Interfaces...... 10 2.2.1 Interface Regimes as a Function of Charge Transfer...... 10 2.2.2 Vacuum Level Alignment in the Schottky-Mott Limit...... 13 2.2.3 Fermi Level Pinning...... 17 2.2.4 Molecule-Metal Interaction Featuring Hybridization...... 21 2.2.5 Chemisorption featuring Bond Formation...... 23 2.3 Charge Transport in Organic Transistors...... 26 2.3.1 Charge Transfer through Metal/Molecule Interfaces...... 26 2.3.2 Organic Field-Effect Transistors...... 28 2.3.3 Contact Resistance in OFETs...... 30

3 Photoemission Spectroscopy 33 3.1 Photoemission from Surfaces...... 33 3.2 Theory of Photoemission...... 34 3.3 Electron Energetics at Surfaces...... 38 3.4 Quantitative Analysis of Chemical Composition...... 41 3.5 Surface Coverage of Organic Layers using XPS...... 43 3.5.1 Determining Surface Coverage according to Schreiber...... 43

vii Contents

3.5.2 Determining Surface Coverage after Bramblett et al...... 45 3.6 Experimental Setup...... 49

4 Density Functional Theory 51 4.1 Basic Description of an Interacting Many-Body System with Electrons and Nuclei 52 4.2 The Hohenberg-Kohn Theorems...... 54 4.3 The Kohn-Sham Ansatz...... 55 4.4 Exchange and Correlation...... 56 4.5 Van-der-Waals Forces in DFT...... 58 4.6 Repeated Slab Approach...... 60 4.7 Bond Dipole Analysis...... 61 4.8 Computational Details...... 64

II Dithiocarbamate Modifications of Noble Metal Surfaces 67

5 Dithiocarbamates for Low Work Function Noble Metals 69 5.1 N,N-dialkyl Dithiocarbamates on Gold...... 70 5.2 Chemical Binding of N,N-dialkyl Dithiocarbamates on Noble Metals...... 73 5.2.1 XPS of C2DTC bound to Au(111)...... 74 5.2.2 XPS of C2DTC bound to Ag(111)...... 76 5.2.3 XPS of C2DTC bound to Cu(111)...... 79 5.2.4 Effect of Alkyl Chain Length on Monolayer Formation...... 83 5.3 Electronic Interface of N,N-dialkyl Dithiocarbamate Monolayers...... 87 5.3.1 Work Functions of N,N-diethyl Dithiocarbamates on Noble Metals..... 87 5.3.2 Work Functions of N,N-dialky Dithiocarbamates on Noble Metals..... 90 5.3.3 Valence Structure of N,N-diethyl Dithiocarbamate Monolayers...... 92 5.3.4 Bond Induced Pinning...... 98 5.4 Electrode Modification in PTCDI-C13 OTFTs...... 103 5.4.1 Contact Resistance in OTFT with CnDTC Modified Electrodes...... 103 5.4.2 In-Situ Evolution of Charge Transport using DTC Modified Electrodes.. 105

6 Work Function Tuning with Tailor-Made Dithiocarbamates 113 6.1 Reference Systems: Thiols...... 114 6.2 Novel Dithiocarbamates...... 117 6.2.1 Odd vs. Even Effect: Dimethyl-Dithiocarbamate...... 117 6.2.2 Aromatic Modification: Pyrrolyl-Dithiocarbamate...... 119 6.2.3 Cyanoethyl-Dithiocarbamate...... 121 6.2.4 Fluorinated Dithiocarbamates...... 123

viii 6.3 A New Window of Work Functions for Au, Ag, Cu...... 124

III Beyond Metal Surfaces: Modifications of Metal Oxides 127

7 Amino-based SAMs for Metal Oxides 129 7.1 Chemical Structure of ITO...... 130 7.2 Surface Modifications of ITO...... 133 7.2.1 4-Dimethylamino Benczoic Acid...... 134 7.2.2 Lysine...... 138

8 Conclusion and Outlook 143

IV Appendix 149

A Supporting InformationI A.1 Experimental Details...... I A.2 Electronic Structure: Projected DOS of C2DTC on metals...... III A.3 Binding sites of Dithiocarbamates on Noble Metals...... IV A.4 PTCDI-C13 OFETs with Modified Copper Electrodes...... IV A.5 Molecular Orbitals of Pyridine-based Phenylene Molecules...... V

B Supporting XPS Measurements VII B.1 Nitrogen on Ag(111)...... VII B.2 Carbon Disulfide on Cu(111)...... VIII B.3 Reference Systems: Thiols...... IX B.4 Dimethyl-Dithiocarbamate...... XI B.5 Aromatic Modification: Pyrrolyl-Dithiocarbamate...... XII B.6 Cyanoethyl-Dithiocarbamate...... XIII B.7 Fluorinated Dithiocarbamate...... XIV

Bibliography XVII

List of Figures XLIII

List of Tables XLVII

Acknowledgments XLIX

ix

Chapter 1

Introduction

Since the discovery of electroluminescence in molecular compounds by A. Bernanose in 1955 [1], it is generally known that the class of organic materials incorporate the fundamental properties for energy-conversion and light-emitting devices. Additionally, Melvin Calvin could explain the photovoltaic effect and photoconductivity in organic system 1958 [2] and received the Nobel Prize in Chemistry for his explanation of photosynthesis in plants 1961 [3]. However, these reports could not yet initialize an intense research of organic electronics. Meanwhile, William Shockley, John Bardeen and Walter Brattain have been awarded with with Nobel Prize in Physics 1956 “for their researches on semiconductors and their discovery of the transistor effect” [4]. Their research marks one of the most important inventions of the 20th century for the development of electrical devices and depicts a milestone for the present semiconductor technology. Indeed, the semiconductor industry association ISA reports a continuous growth of the global semiconductor industry and expects world wide sales to reach $333.2 billion for 2014 [5]. Still, this industry is dominated by inorganic materials.

It took more than 30 years after the report of Bernanose, that a general interest in organic elec- tronics has been initialized by the work of Ching W. Tang. At first, he developed the first two layered organic photovoltaic cells in 1986 [6]. While this cell reached only a power conversion efficiency of 1 %, the prospect of a sustainable energy-generation based on organic materials has pushed the development in recent years [7]. Nowadays, the combination of novel architectures for buildings and solar cells based on organic materials can be envisioned to lead to self-sustaining living conditions, as demonstrated in the project “the District of Tomorrow”1 [9].

1Created by Zuyd University in Heerlen, NL. Building in progress. Funded by the INTERREG project ORGANEXT [8].

1 Chapter 1: Introduction

E E

Φ Φ E hv

E

Figure 1.1: Schematic energy diagram of an OLED stack. All energy levels are referenced to the vacuum level Evac. In a common OLED, the anode (cathode) usually serves as hole (electron) injecting electrode. Thus, a high (low) work function Φa (Φc) is required for efficient charge transfer into the frontier molecular levels. These critical interfaces are marked with red circles.

Only one year after his report on an organic solar cell, Tang and van Slyke also reported a light emitting diode based on organic materials [10]. In parallel, the improvement of electrical proper- ties of molecular semiconductors led to the first reports of organic field-effect transistors in 1987 [11, 12]. Indeed, the development of new conductive molecular materials has been awarded with the Nobel Prize for Alan Jay Heeger, Alan Graham MacDiarmid and Hideki Shirakawa in 2000 “for their discovery of electrically conductive polymers” [13]. Therewith, the foundation for ef- ficient organic electronics has been set. In fact, the field of organic electronics experiences a constant industrial importance [14] and has developed from simple two-colored displays in small electronic devices to large area HD-TVs. Recently, the first curved OLED TVs have entered the market [15, 16]. It is indeed an advantage over common inorganic semiconductors, that molecular materials can be processed on large area (“roll-to-roll” [17, 18]) and cost-efficient at low temper- atures. Moreover, the deposition of active layers on bendable substrates enables flexible displays and solar cells [19, 20]. Indeed, first rumors indicate that LG will announce a new with a truly flexible display at the 2015 Consumer Electronics Show (CES) in January [21].

Despite the recent success of OLED technology, there still remains the need to further improve each step within the device [22]: From charge injection at the electrode, charge transport to the light-emitting layer and efficient light generation, as schematically represented in figure 1.1. In- deed, has faced such serious difficulties that the company has announced to decrease its efforts regarding the development of new OLED-TVs [23]. On the other hand, LG still pushes its investments into the next generation of OLED based TVs and displays [24].

2 Section

The present work tries to support the latest development of novel organic electronics with the fo- cus on improvement of electron injection at the electrode/molecule interface. Therein, the work function Φ plays a central role: It defines the energy, which is required to lift an electron from the 2 Fermi level EF up to the vacuum level Evac . This is also illustrated in figure 1.1, which schemat- ically depicts an energy diagram of a common OLED stack. Therein, the transparent electrode, usually ITO3 coated glass, is employed as the anode to inject holes. This combination is reason- able, since ITO exhibits a moderately high work function [25]. A high work function Φa is desired for hole injection, since the Fermi level of the anode should ideally match the energetic position 4 of the molecular HOMO level. Consequently, a low work function material (Φc) is required as cathode on the other side of the stack to match the molecular LUMO5 for efficient electron in- jection. Among other processes within the OLED stack, these electrode/molecule contacts mark critical interfaces of the full device: The best light emitting material becomes useless if charges are inefficiently injected at the electrodes. In particular, low work function electrodes remain a key challenge in modern organic optoelectronics. Natural low work function materials are usually highly reactive to the organic layers and ambient atmosphere, which results in defective . One opportunity, that has established, is the use of a thin LiF layer in combination with an Al or Ag cathode [26]. Lithium generates a low work function at the surface of the metal, but a careful encapsulation is required. Hence, an alternate combination is desired.

In this context, the current thesis investigates how electrode/molecule interfaces can be modified with organic materials in order to optimize the charge transfer at these contacts. Therein, the following question has to be answered:

Why is there a need for interface modifications in electrode/molecule junctions?

At this point, the present work exploits the recent progress in the field of dithiocarbamates (DTCs) and introduces another approach: The surface modification of electrode materials with amino- based self-assembled monolayers. As has been recently reported, N,N-dialkyl dithiocarbamates generate extremely low work functions when arranged in a monolayer on gold [27]. The question arises to what extend the metal species influences monolayer adsorption and electronic structure of the modified surface. In particular: Can dithiocarbamates be universally used to produce low work functions of noble metals? Moreover, gold is too expensive to be of interest for industrial applications. Therefore, the next steps are taken in this thesis. First, the assembly of dithiocar- bamates on the noble metals Au, Ag and Cu is investigated and their effect on the work function

2The vacuum represents the energetic level, at which the electron is not affected by the surface anymore, and is therefore a reference level for the energetic alignment of electronic states of solids. 3Indium Tin Oxide, i.e. tin doped indium oxide with a high tin dopant concentration. This results in an electrically conductive, but optically transparent compound. 4Highest Occupied Molecular Orbital 5Lowest Unoccupied Molecular Orbital

3 Chapter 1: Introduction

∆Φ

Φ

Figure 1.2: Schematic work function shifts upon DTC modification. is discussed. The comparison of these molecules assembled on all three metals generates an in- sightful analysis on the effects that govern the energy level alignment at metal/molecule interfaces featuring chemisorption. In particular, a strong linear trend of the work function shift with respect to the metal species is observed, as depicted in figure 1.2, and will be discussed in chapter 5. Sec- ond, the prospect of a work function tailoring of noble metals is presented on the basis of novel dithiocarbamates with functionalized substitutions in chapter 6. Third and last, a set of first guess molecules for the functionalization of ITO is presented in chapter 7. These molecules are based on the molecular structure of dithiocarbamates, but with different linker groups that bind to the oxide surface. While these results are more preliminary, they illustrate the opportunity to take advantage of the polar amine group in a more general approach to generate low work function electrodes from metals to oxides. Then, the OLED stack as depicted in figure 1.1 can be reversed and ITO can be used as the cathode, while an air-stable noble metal, such as silver, can be employed as the anode on top without the need for encapsulation.

4 Part I

What Governs Electronic Processes at Interfaces to Organic Materials? A Theoretical Introduction

Chapter 2

Materials and Theoretical Background

Charge transport at and through interfaces is one of the crucial mechanisms that determines the efficiency of organic electronic devices. Therefore, the energy level alignment between the Fermi level of the (metallic) electrode and the frontier molecular orbitals is essential for the charge trans- fer through the metal/organic interface. Optimizing the electrical interface can be achieved by the application of polar molecules that alter the surface energies of the conductive electrodes. Thus, after a brief presentation of the class of organic materials, this chapter introduces the physical background that is essential for the following experimental and theoretical results.

2.1 Molecular Materials

In general, organic molecules can be summarized as the class of hydrocarbons. Since this formu- lation generates a potentially infinite number of compounds, additional classifications are made, such as the small molecules and polymers. Furthermore, conjugated and aromatic molecules have established in the field of optoelectronic applications due to their unique optical properties as well as charge transport characteristics. Aromaticity is the result of a conjugated coordination of carbon in a ring system as a result of six sp2-hybridized carbon atoms. Such an aromatic configuration, also known as a resonant bond, generates delocalized electrons which dominate the frontier elec- tronic structure. One example of a small, aromatic molecule is N,N’-ditridecylperylene-3,4,9,10- tetracarboxylic diimide, which will be abbreviated with PTCDI-C13 in the following. PTCDI-C13 is known as a high mobility n-type semiconducting organic material and will be of use in this thesis as well [28]. An illustration of this compound is given in figure 2.1: It constitutes of an aromatic

7 Chapter 2: Materials and Theoretical Background

Figure 2.1: Molecular structure of N,N’-ditridecylperylene-3,4,9,10-tetracarboxylic diimide, ab- breviated with PTCDI-C13 (top). Carbon is colored in gray, hydrogen in white, nitrogen in blue and oxygen in red. Representations of the isosurfaces ( ψ 0.03 ) of the HOMO and LUMO lev- els are drawn based on DFT calculations using the B3LYP| = functional| with 6-311+g(d,p) basis set [30]. Both frontier molecular orbitals can be identified as π-type orbitals, which are delocalized over the whole aromatic core of PTCDI-C13.

core based on perylene (C20H12) with two imides (R1-CO-NR2-CO-R3) as functional groups at- tached to it. These imides are further substituted with an alkyl chain with 13 methyl units. Such a substitution generally affects the growth and diffusion of these molecules, but leaves the frontier molecular states unaffected [29]. This can be understood immediately by examining the HOMO and LUMO representations of PTCDI-C13, which are depicted in figure 2.1 using isosurfaces of these molecular states. Both orbitals can be identified as π-type with a delocalized nature across the aromatic core of the molecule, but without any probability distribution of electrons at the alkyl chains.

In contrast to PTCDI-C13, which is a perfect candidate for n-type transistor devices, the class of dithiocarbamates are non-aromatic small molecules [31]. However, these molecules provides means to effectively alter surface properties of (noble) metals and, thus, play a central role in this thesis. A set of molecules based on this ligand will be introduced and investigated in the course of this work. Here, a brief overview of N,N-dialkyl dithiocarbamates, abbreviated with CnDTC (n indicates the number of methyl units in the alkyl chain), will be given. Dithiocarbamates closely resemble dithiols (R-SSH), thus originate from the class of alkanethiols (R-SH) and can be equally employed for surface passivation of noble metals [32], too. It is well known, that thiols establish

8 Section 2.1: Molecular Materials

Figure 2.2: Schematic formation of dithiocarbamate monolayer. The dithiocarbamate compound is synthesized in-situ in a solution of ethanol. Therein, a secondary amine (here: diethylamine (CH3CH2)2NH) and carbon disulfide (CS2) are immersed in a molar ratio of 2:1 (50 mM and 25 mM, respectively). The reaction to form the dithiocarbamate then takes place initially. By addition of a noble metal, the reactive CS2 group chemically binds to the surface and a self- assembled monolayer is formed. very strong chemical bonds of the thiolate linker group to noble metals [32, 33]. Thereafter, such molecules attach to the metal surface until a dense monolayer is formed, which is then referred to self-assembled monolayer. In contrast to dithiols, N,N-dialkyl dithiocarbamates exhibit a nitrogen atom that interconnects the CS2 linker group to the alkyl chains (R1,R2-N-CS2).

While PTCDI-C13 can be directly purchased from common suppliers for chemical reagents such as Sigma-Aldrich, the dithiocarbamates in this thesis have to be synthesized. This process is rather simple and takes place spontaneously as soon as the reactants are brought together in a solution of ethanol. Therefore, the best option to modify a freshly prepared noble metal surface with a dithiocarbamate layer is an in-sit processu, i.e. the metal substrate is immersed into the same solution that serves for the synthesis itself. The general routine is a solution of ethanol with a 2:1 molar ratio of carbon disulfide and a secondary amine (50 mM and 25 mM, respectively). This reaction is illustrated in figure 2.2 and should be carried out in an inert atmosphere to minimize the incorporation of impurities. Especially oxygen is an unwanted species, since an oxidized metal surface is less attractive for the monolayer formation and will lead to disorder and a decreased layer density. In addition, oxidized sulfur in the molecule is less likely to react with the metal surface. 10 However, employing a glovebox attached to an ultra-high vacuum system (p 1 10− bar) at ≈ · the I. Institute of Physics (IA), all samples could be prepared without any contact to ambient conditions at any step of the preparation.

9 Chapter 2: Materials and Theoretical Background

2.2 Fundamental Processes at Metal/Molecule Interfaces

In the course of this thesis, derivatives of the dithiocarbamate ligand are introduced and their ef- fect on surface properties of noble metals is analyzed. The focus of this research is placed on the chemical and energetic interaction of these molecular adsorbents on conductive substrates. Conse- quently, the fundamental processes at the metal/molecule interface will be introduced on a theoret- ical level, first. The discussion in this section is mainly based on the textbook The Molecule-Metal Interface edited by Norbert Koch, Nobuo Ueno and Andrew Wee [34] as well as the review article Energy-Level Alignment at Organic/Metal and Organic/Organic Interfaces written by Slawomir Braun, William Salaneck and Mats Fahlman [35].

2.2.1 Interface Regimes as a Function of Charge Transfer

Up to this day, there does not exist a unified model that is capable to describe and predict the processes that emerge when a molecule approaches any kind of surface, such as reactive or non- reactive metals, oxides or organic materials. Due to a variety of effects that can occur during the process of adsorption, it has been proven useful to separate the problem into a set of regimes that are classified by the strength of interaction. Thus, different models have emerged to describe the different interaction types. In principle, the interaction strength is correlated to the amount of charge that is rearranged upon interface formation. Any molecular adsorbate will inevitably affect the electrostatic potential of the underlying substrate surface. For a clean metal surface, the situation is illustrated in figure 2.3A. As depicted, the difference of the chemical potential inside of the solid µbulk and the electrostatic potential φout across the metal surface determines the magnitude of the work function Φ. Since the bulk chemical potential is a characteristic quantity of the solid, atomic and molecular modifications of the surface will affect the electrostatic potential. This correlation can be best understood by the visualization of the distribution of positive and negative charges at the metal surface in figure 2.3B. While the nuclei of the solid represent a sharp surface by an abrupt drop to zero of density p(x) at x x , electrons are able to spill out into = 0 the vacuum region, hence smearing out the electronic density at the transition across the surface.

The electronic density n(x) decays exponentially from the surface at x0, but still can reach several Angstrom into the vacuum [36]. As a consequence for the metal solid to remain uncharged, the excess of electronic density outside of the solid must be compensated by an electron deficiency inside the solid adjacent to the surface. Accordingly, this charge inhomogeneity results in a surface dipole ∆φ and is directly correlated to the difference in the evolution of the electrostatic potential from the bulk across the surface ∆φ φ φ . Hence, the work function Φ can be calculated = in − out by the difference between the bulk chemical potential µbulk and the total surface dipole ∆φ:

10 Section 2.2: Fundamental Processes at Metal/Molecule Interfaces

ϕ Φ Δϕ μ μ ϕ

Figure 2.3: Simplified scheme of the potential energy (A) and charge density (B) of a metal sur- face. The bulk chemical potential µbulk and the evolution of the electrostatic potential φ(x) across the surface define the magnitude of the work function Φ. A total surface dipole of ∆φ originates from the unequal distribution of negative and positive charges, n(x) and p(x), respectively. The nuclei define the sharp surface by an abrupt drop to zero of p(x) at x x0. Yet, electronic wave- functions can spill out into the vacuum region for several Angstrom, hence= effectively creating a positive charge (inside the solid).

Φ ∆φ µ . (2.1) = − bulk

The chemical potential µbulk is a central potential energy in thermodynamics and can be derived from the fundamental Gibbs equation of the inner energy of a system: dU T dS p dV = − + P i µi dni . For any system, µ can be interpreted as the amount of energy that is necessary to add one further particle (i.e., electrons for the free electron gas) to the system (i.e., the bulk solid). Hence at T 0K, the chemical potential is equal to the Fermi energy µ(T 0K) E . According = = = F to the picture drawn in figure 2.3, it becomes now feasible to classify different regimes of interac- tion by the modifications of the electrostatic potential due to charge rearrangements [37]. Starting form the bare metal surface, figure 2.4A, the effect of a non-reactive adsorbent like the noble gases neon or xenon can be considered as shown in figure 2.4B. Having a closed shell, rare-gas atoms tend to only weakly physisorb on any surface by van-der-Waals forces. Still, this soft phys- ical adsorption affects the electrostatic potential φ at the surface and therefore may even change the work function of the substrate [38]. Regardless of potential charge transfer or formation of a permanent dipole, any atom influences the tail of the electronic density that reaches out into vacuum. Even closed-shell atoms like neon or xenon will effectively push back those electrons of the metal surface that partially spill out into vacuum, as indicated in figure 2.4B. This effect can be easily understood by the Pauli repulsion for electrons of the same spin [39]. Therefore, this

11 Chapter 2: Materials and Theoretical Background

Δϕ Δϕ Δϕ Δϕ Δϕ

Figure 2.4: Illustration of the surface and interface effects of different adsorbates on a metal sur- face. Different regimes of interaction can be classified by the type of charge rearrangement. (A) depicts the bare metal surface with a relative strong surface dipole due to the spilling of electrons into the vacuum region, as discussed at figure 2.3.(B) Soft physical adsorption without charge transfer is shown for the case of e.g. the physisorption of noble gas atoms. The electronic tail of the metal surface is pushed back into the bulk region, thus reducing the surface dipole ∆φ and consequently the work function Φ.(C) Introducing a permanent dipole D to the surface (in this case without charge transfer) gives rise for an additional upward shift of the− electrostatic potential and results in an increase of the work function. (D) and (E) illustrate the effect of direct charge transfer between adsorbate and surface. Charge rearrangement across the surface plane results in a dipolar momentum, referred to as bond dipole BD that raises (lowers) the electrostatic potential when an electron is withdrawn from (donated to) the metal surface. Combinations of a permanent dipolar contribution with charge transfer due to bond formation (e.g. in a self-assembled mono- layer) become more difficult to describe (see section 2.2.5), but are still possible and can add up or compensate each other depending on the direction of each dipole moment. For all effects, the potential shift as a result of surface modifications usually remains within a range of maximal 2eV. ±

12 Section 2.2: Fundamental Processes at Metal/Molecule Interfaces

reaction is known as push-back or pillow effect and does not only apply for non-reactive rare-gas species, but also takes place for aromatic molecules on noble metals [40]. Consequently, this kind of repulsion will always act as an initial reaction of the electronic distribution at the surface on any physical or chemical adsorption. As indicated in figure 2.4, pushing the electrons back into the surface reduces the intrinsic surface dipole and thus the work function.

If the non-reactive rare-gas atom is replaced by a layer of permanent dipoles, still excluding of charge transfer across the surface, the electrostatic potential experiences an additional shift. If the direction of the permanent dipole acts in the same the direction as the surface dipole, then both contributions will add up as illustrated in figure 2.4C. Consequently, a permanent dipole oriented in the opposite direction will counteract the surface dipole, hence (partially) compensating the latter. Introducing either a reactive substrate or adsorbate eventually results in a charge transfer across the surface plane. In analogy to the incorporation of a permanent dipole, the direction of charge flow directly affects the electrostatic potential. A charge transfer across the surface alters the shape of the electrostatic potential and can be represented by an additional interface dipole. This will be denoted as bond dipole BD, since bond formation of small molecules to surfaces is frequently discussed in this thesis. Hence, electron donation to (withdrawal from) the adsorbate to the metal lowers (raises) the work function of the surface (figure 2.4D&E). Certainly, charge transfer accompanying a permanent dipole yields several further combinations, which are not depicted in figure 2.4. In fact, such combinations are the most accurate description in this simple picture for self-assembled monolayers, which have been characterized within this thesis. Yet, these systems become more difficult to analyze and will be discussed in detail in section 2.2.5. Furthermore as also discussed by Ishii et al., there are further factors affecting the interfacial dipole formation, such as mirror forces, surface reconstructions and interface states beyond chemical bonding [37].

Nevertheless, figure 2.4 allows to classify different regimes of interaction from weak physisorp- tion without electron transfer to complex systems exhibiting permanent and bond induced dipolar contributions. Yet up to this point, two items remain unclear: First, to what extend are charges able to rearrange; and second, what are the limits of changing the electrostatic potential and therefore the work function?

2.2.2 Vacuum Level Alignment in the Schottky-Mott Limit

In the simplest case of the metal-molecule interface, charge rearrangements are negligible, i.e. permanent dipoles and charge transfer are absent. In this scenario, the concept of vacuum level alignment holds for the description of the energetic interface. Thus, the vacuum level Evac is the only reference for the energetic level of both the metal and molecule. Before any interaction takes

13 Chapter 2: Materials and Theoretical Background

R RR RR R R

Φ Φ Φ Φ Φ

Δ Δ

Δ Δ

Figure 2.5: Schematic representation of the energetic structure at the metal/organic interface for the regime of vacuum level alignment. (A) illustrates the situation before any interaction takes place. The vacuum level Evac is the reference level for both the metallic substrate and organic semiconductor. The work function of the metal ΦM is the difference between Evac and the Fermi level EF. The positions of HOMO and LUMO are defined by the ionization potential IP and the electron affinity EA, respectively. (B) For IP Φ EA, the vacuum levels can simply align > > if there is no charge rearrangement. Then the electron (hole) injection barrier ∆e (∆h) is given directly by the distance of the LUMO (HOMO) level to EF.(C) If charges rearrange without charge transfer across the surface, an interface dipole BD will shift the molecular levels with respect to the metal Fermi level EF. This is realistic for weak interaction, e.g. facing no more than the “pillow effect”.

place (figure 2.5A), the work function Φ of the metal is defined as the energetic distance between

Evac and the Fermi level EF, Φ E E . (2.2) = vac − F In analogy, the position of the molecular LUMO (HOMO) is calculated by the difference between

Evac and the electron affinity EA (ionization potential IP). The Schottky-Mott limit essentially states that the semiconductor has to remain entirely free of gap states in the process of forming a metal-semiconductor interface. Otherwise, charge transfer between metal and gap states affects the energetic interface, as will be discussed in section 2.2.3. In addition, the Fermi level EF of the metal is required to lie within the band gap between HOMO and LUMO of the semiconductor: IP Φ > > EA. Then, the vacuum levels directly match and all energy levels simply align without any shift. This scenario is depicted in figure 2.5B. If the work function of the substrate is continuously varied under the assumption that IP Φ EA remains valid, then the charge injection barrier for > >

14 Section 2.2: Fundamental Processes at Metal/Molecule Interfaces

electrons ∆e (holes ∆h) is given by the difference between EA (IP) and EF via

∆ Φ EA , (2.3) e = − ∆ IP Φ . (2.4) h = −

Instead of the injection barrier, it is also common to relate the work function of the modified1 substrate ΦM/ORG to the initial work function of the bare metal Φ, as will also be discussed in section 2.2.3. In the regime of vacuum level alignment, ΦM/ORG linearly scales with Φ and, in an ideal scenario, exactly equals Φ [35]. As illustrated in figure 2.5C, there are examples where the interaction strength is extremely weak, while small charge rearrangements still induce an interface dipole BD. This may be a realistic example for the aforementioned discussion of rare-gases on clean metal surfaces that face no more than the “pillow effect”, figure 2.4B. While there is an interface dipole incorporated in the electronic structure that shifts the metal and organic levels with respect to each other, one may identify this example as a special case of the transition between vacuum level alignment and more complex models including chemisorption. Anyhow, the expressions for charge injection barriers have to be adjusted when an interface dipole (BD 0) > is formed at the metal/molecule contact:

∆ Φ EA BD , (2.5) e = − + ∆ IP Φ BD . (2.6) h = − −

In any case of forming a metal/molecule contact, band bending might be considered in analogy to the common metal/semiconductor junction. This can also occur for molecular semiconductors, even though the mechanism can differ due to the different mechanism of charge transport in or- ganic solids. At a distance sufficiently far away from the metal, the work function is dominated by the work function of the organic layer ΦORG, which may differ from the work function ΦM/ORG of the first layer. Figure 2.6 illustrates the electronic levels before and after contact. Without the for- mation of an interface dipole, the charge injection barriers ∆e and ∆h are determined as discussed in analogy to figure 2.5B. With increasing layer thickness of the organic semiconductor, the po- sition of Evac and the molecular levels can shift with respect to the Fermi level of the metal EF.

Still, EF has to remain in the band gap of the semiconductor at all time. The work function is then defined by the organic material ΦORG and may differ from ΦM. Hence, a build-in potential Vbi drops across the organic layer to shift the molecular orbitals with respect to the metal Fermi level

EF. Still, EA and IP remain constant at any point. For such a system, in which an organic material is deposited onto a (metal) substrate, the work function of the organic material is not always a unique quantity, but may be affected by the substrate material even at a large layer thickness in the

1In this work, a modified substrate generally refers to a system of a metal substrate, which is covered by a molecular monolayer.

15 Chapter 2: Materials and Theoretical Background

k kkkk

Φ

Φ Φ

Δ

Δ

k

Figure 2.6: Band bending in an organic semiconductor for a metal/organic interface. (A) The situation before contact is illustrated. Again, the Fermi level EF of the metal is positioned some- where in the band gap between HOMO and LUMO of the molecular semiconductor. (B) In a contact without an interface dipole, i.e. in the regime of vacuum level alignment, the injection barriers for electrons (holes) are represented by ∆e (∆h). Upon further deposition of the molecu- lar material, band bending might occur. Far away from the metal substrate, the work function is defined by the organic semiconductor ΦORG and may differ from ΦM. Hence, a build-in potential Vbi drops across the organic layer to shift the molecular orbitals with respect to the metal Fermi level EF. Yet, EA and IP remain constant at any point. order of several tens of nanometers. This has been observed and discussed by Ishii et al. for the material TPD on different metallic substrates [37]. Thus, the build-in potential Vbi may be also affected by the choice of the substrate.

In addition, gap states will also affect the energy level alignment at the interface as well as the process of band bending [34]. This very important aspect of the influence of gap states on the electronic interface will be discussed in the following section, since the limits of vacuum level alignment have not been addressed yet. As the Fermi level EF marks the highest ground-state energy of the solid (and therefore of the interface), it is impossible that EF moves above (below) the molecular LUMO (HOMO) at the interface. At these points, the Fermi level is said to get pinned.

16 Section 2.2: Fundamental Processes at Metal/Molecule Interfaces

2.2.3 Fermi Level Pinning

In the regime of vacuum level alignment, the charge injection barriers at the interface are directly given by the difference between the metal Fermi level EF and the frontier molecular orbitals. For extreme cases, in which the work function ΦM is larger (smaller) than the molecular ionization potential IP (electron affinity EA), the vacuum levels cannot align without forming an interface dipole. In thermodynamic equilibrium, the Fermi level EF marks the highest occupied electronic level of the solid. By forming a metal/semiconductor interface, this condition must remain valid for the total system. Hence, if Φ IP before contact, the occupied HOMO level of the organic M > material lies above EF. This situation cannot persist when these materials are brought into contact. Consequently, charges have to rearrange across the interface to induce an interface dipole. Fig- ure 2.7 schematically correlates the energy level diagram of a metal/molecule interface (A) with the corresponding potential barriers for charge injection (B) as a function of ΦM. As indicated in figure 2.7A, there exists a level P1 (P2) in the band gap below (above) the molecular LUMO

(HOMO), that the Fermi level EF cannot pass: The Fermi level is pinned to this energetic level. Hence, there will always remain a minimal energy barrier for charge injection, irrespective of the substrate’s work function [41]. This is illustrated in figure 2.7B: At very low work function val- ues ΦM, EF is pinned below the LUMO level to P1 yielding a minimal (maximal) injection barrier ∆ (∆ ) for electrons (holes). For IP Φ EA, the regime of vacuum level alignment applies e h > M > and the injection barriers linearly scale with the metal work function. This can be expressed by a slope parameter S, d∆e d∆ S h , (2.7) = dΦM = −dΦM which is unity when the Schottky-Mott limit holds. Even if an interface dipole is formed, S equals one if the dipole is independent on the metal substrate. Otherwise, values in between unity and zero will be observed. For the scenario in which EF is pinned to either P1 or P2 S yields zero.

Now, the following question arises: What is the origin of the pinning levels P1 and P2? There have been two models established to explain the origin of Fermi level pinning, which apply to different regimes of charge transfer. The first model was developed by William Salaneck, Mats Fahlman and coworkers [35, 42, 43]: Therein, an integer number of electrons is considered to travel across the interface between the first organic layers and the metal substrate. In another approach favored by Antoine Kahn and Norbert Koch, the attention is put on the density of states that reach into the gap beyond the HOMO and LUMO edges [34, 44]. Both models will be discussed in the following.

The Integer Charge-Transfer (ICT) model applies for interfaces in the absence of pronounced hybridization of molecular orbitals with the substrate. Thus, ICT is often observed when molecules are deposited onto passivated substrates with extremely high or low work functions. An example

17 Chapter 2: Materials and Theoretical Background

p

Φ

Δ Δ

Δ

p Δ

pp Φ

Figure 2.7: Fermi level pinning as the limiting factor of vacuum level alignment for weak inter- actions at metal/molecule interfaces. (A) Energy level diagram of the metal/molecule interface. Upon variation of the metal work function ΦM, the electron (hole) injection barrier ∆e (∆h) can be tuned, but will never completely vanish, since EF cannot move above (below) the LUMO (HOMO) level of the semiconductor. Instead, EF gets pinned to either P1 or P2.(B) Schematic illustration of the evolution of the injection barriers as a function of the metal work function ΦM. The regime, in which the slope parameter S equals unity, corresponds to vacuum level alignment. If Φ IP (Φ EA), then S 0 and E is pinned at the HOMO (LUMO). At this point, the M ≥ M ≤ = F charge injection barrier ∆h (∆e) remains unaffected on a further increase (decrease) of ΦM.

is given by the group of Norbert Koch that has identified an integer charge transfer in the energy level alignment of C60 on a silver surface, which has been passivated with a monolayer of α- sexithiophene (α-6T) [45]. Passivation of the (metal) substrate decouples the frontier π-electrons from the metallic bands. Thus, hybridization is suppressed and a partial charge transfer cannot take place. Still, electron transfer across the passivating layer may occur if the energy difference is sufficiently large to promote the formation of a charged state in the organic layer. Consequently, the integer charge-transfer state EICT (EICT ) is defined by the energy required to withdraw (add) + − an electron from (to) the molecule to produce a fully relaxed state. Since organic materials can induce a significant geometrical reorganization when the molecule is charged, a self-localized state is formed and referred to as polaron (single charge) or bipolaron (double charge). These new states appear in the energy diagram in the band gap of the molecule, as depicted in figure 2.8A. For the following discussion, it is assumed that Φ IP, i.e. the Fermi level E is located beneath the M > F molecular HOMO level before contact. Upon interface formation (figure 2.8B), charges move from the organic semiconductor to the metal surface in order to establish an interface dipole BD, that aligns EF to EICT , figure 2.8C. Consequently, the direction of the charge transfer would + be reversed once EF gets pinned to EICT . As illustrated, the resulting work function ΦM/ORG −

18 Section 2.2: Fundamental Processes at Metal/Molecule Interfaces

/ / ///

Φ Φ Φ Φ

/ / ////

Φ Φ Φ Φ

Figure 2.8: Fermi level pinning: Comparison of integer charge transfer model (ICT) versus gap states at the example of pinning close to the HOMO level. (A-C) represents the processes accord- ing to the ICT model [35]: If ΦM IP before contact, charges will travel from the semiconductor to the metal, inducing a strong interface> dipole. Still, hybridization or chemical bonding are ne- glected. This scenario is realistic for passivated surfaces with high work functions. The Fermi level of the metal ΦM is pinned to the EICT state, which is close to the (former) HOMO level. + (D-F) illustrates the pinning of EF to gap states according to Kahn and Koch [34]: Filled gap states adjacent to the HOMO level can get emptied in the process of interface formation. The molecular orbitals then remain unchanged and EF gets pinned to the position of gap states.

19 Chapter 2: Materials and Theoretical Background

is independent on the initial work function ΦM of the metal when EF is pinned to an ICT-state. The alignment of the energetic levels at the interface is thus dominated by the ICT-states of the molecular semiconductor. Note that the position of HOMO and LUMO in the energy scheme in figure 2.8B andC illustrate the former frontier molecular orbitals prior to charge transfer. After relaxation of the charged molecule, the ICT-state effectively represents the new HOMO level. Thus, charges can principally be transferred across the interface between the metal and molecule without energy gain or loss if EF is pinned to an ICT-state. Still, the difference between

EICT and the HOMO level remains to be the minimal energy barrier ∆h for charge transfer, since + the subsequent molecular layer will align to the first layer without charge transfer (vacuum level alignment). Hence, a minimum energy of ∆h has still to be spent for charge injection into the bulk semiconductor, as unambiguously demonstrated by Greiner et al. [41].

In contrast to the ICT model, Antoine Kahn and Norbert Koch have often considered the effect of gap states on the energy alignment of a metal/molecule interface [34, 44]. The condition that the level mismatch before contact has to generate enough energy to form a polaron, is not always met. Still, Fermi level pinning can be observed such that the origin must be of different nature. As calculated by Kalb et al., even highly purified single crystals of molecular materials exhibit a non- negligible amount of trap states in the forbidden band gap [46, 47]. The existence of occupied trap states in the band gap near the HOMO is schematically represented in figure 2.8D. These states can then be emptied when the semiconductor is brought into contact with the metal substrate, figure 2.8E, which will induce an interface dipole in analogy to the ICT model. The final situation at equilibrium is illustrated in figure 2.8F: The Fermi level EF is now pinned to one of the gap states of the semiconductor. Thus, the energetic position of the HOMO level still accurately resembles the highest occupied molecular orbital, in contrary to the ICT model. A minimum charge injection barrier ∆h is directly observable from figure 2.8F, since the pinning to gap states ensures an energetic distance to the HOMO level.

Both models haven shown to be applicable to accurately describe and explain how the Fermi level can get pinned to keep the condition of thermodynamic equilibrium. Still, there is no officially accepted rule that states when the ICT model or the gap-states model has to be employed. How- ever, a few conditions can be stated: Both models require that the molecule only weakly interacts with with (metallic) substrate. Otherwise, the molecular orbitals completely alter by approaching the surface (discussed in the following sections). The question, whether an electron is transferred between a molecular orbital or gap-state crucially depends on the density and energetic position of these gap states as well as the position of the ICT-state. A prediction of such density of states is difficult and an experimental analysis is more reliable. However, if a molecule is brought into direct contact with the (metallic) substrate, the (weak) interaction may generate gap states via small deformations of the molecule and impurities play an important role. Otherwise, a passivated surface can reduce these effects and promote a pinning process according to the ICT-model [45].

20 Section 2.2: Fundamental Processes at Metal/Molecule Interfaces

As stated, both models describe interfaces with negligible hybridization of the frontier molecular orbitals with the metal bands. If a molecule is deposited on a reactive surface that has not been pas- sivated, then stronger interactions including the formation of chemical bonds can be expected.

2.2.4 Molecule-Metal Interaction Featuring Hybridization

As discussed in the previous section, charging a molecule generates a polaron, i.e. the molecule geometrically relaxes upon charge transfer of an electron. The situation becomes more complex if molecules get into direct contact with a more reactive surface. Then molecular orbitals may hybridize with metallic bands at the surface and partial charge transfer can be observed. This leads to more complicated molecular reorganization including chemical bonding as compared to the formation of a polaron. For weakly chemisorbed molecules, a model based on an induced density of interface states (IDIS) has been proposed by Flores and coworkers [48, 49]. This model is most suitable for the interfaces of molecules to clean, but non-reactive (metal) surfaces. Then, hybridization can be observed and the chemical interactions, albeit being moderate, cannot be neglected anymore. The principle of the concept is visualized in figure 2.9: Before the molecule gets into contact with the metal, its molecular orbitals can be represented by discrete energy levels, figure 2.9A. Hence, the positions of both HOMO and LUMO are clearly defined to the vacuum level Evac with IP and EA, respectively, and the regime in between is the well-known forbidden energy gap. When a molecule is brought into contact with a substrate featuring hybridization (B), these discrete energy levels broaden and a continuous density of states (DOS) is induced via interaction. The energy gap then vanishes and a new distribution of electronic states is observed by a subsequent filling of the induced DOS with electrons until the integrated induced DOS yields the number of electrons of the undisturbed molecule. Thus, the charge-neutrality level ECNL is defined as an effective Fermi level of the molecule. ECNL dominates the subsequent energy alignment through charge transport across the interface if its energetic position mismatches that of the Fermi level E of the metal. Hence, the initial energy mismatch (Φ E ) is reduced by the screening F M − CNL parameter S (0 S 1), which depends on the strength of the interaction. The resulting work < < function of the metal/molecule surface ΦM/ORG in figure 2.9C is then expressed via

Φ E S (Φ E ) . (2.8) M/ORG − CNL = · M − CNL

This can also be rearranged to express the strength of the interface dipole BD:

BD (1 S) (Φ E ) . (2.9) = − · M − CNL

The screening parameter S effectively matches the slope parameter defined in equation (2.7) and represents the strength of the interaction. For strong coupling, S 0 and the interface dipole BD →

21 Chapter 2: Materials and Theoretical Background

G GG GG G G

Φ Φ Φ Φ

Figure 2.9: Energy level diagram of a metal/molecule interface according to the model of in- duced density of states (IDIS) [48, 49]. (A) Before contact, the molecular orbitals are discrete energy levels as observed for single molecules in vacuum. (B) In contact with a reactive surface, hybridization of the frontier molecular orbitals with the metal bands occurs. Hence, these energy levels broaden and generate a continuous density of states. This induced DOS is then filled with electrons until charge neutrality is reached, thus defining the charge-neutrality level ECNL, which is located in the former energy gap of the molecule. (C) Charge transfer across the metal/molecule interface induces an interface dipole BD that aligns EF with ECNL. After charge transfer, the final charge neutrality level defines the resulting work function ΦM/ORG. completely screens the initial energy mismatch. Otherwise, for extremely weak interaction, the interface dipole fully vanishes as S 1. In addition, within the framework of the IDIS model, S → can be derived to [35, 49]:

dΦM/ORG 1 S , (2.10) = dΦ = 1 4πe2D (E )d/A M + F where D (EF) is the induced DOS of the interface molecules at the Fermi level, d is the metal-to- molecule distance and A is the average interface area of the molecule. This interface area A can be easily estimated if the geometric arrangement of the molecule is known. The distance d between substrate and molecule, however, is more difficult to obtain, since computational methods (e.g. density functional theory) usually struggle with a correct representation of weak van-der-Waals forces. Yet, this is an important parameter since it strongly affects the amount of charge transfer. In addition, a short distance of the molecule to the unpassivated metal always results in the “pillow” effect, which has been discussed in section 2.2.1. The incorporation of this effect into the IDIS model is often denoted as unified IDIS-model [49].

22 Section 2.2: Fundamental Processes at Metal/Molecule Interfaces

2.2.5 Chemisorption featuring Bond Formation – When Do Molecular Dipoles Count?

At last, highly reactive surfaces and/or molecules can strongly interact when brought into con- tact and new chemical bonds can be established. This is particularly important for the class of molecules that form a self-assembling monolayer (SAM), such as thiols on gold [32, 50]. To understand the intricate processes accompanying bond formation, usually atomistic calculations based on density functional theory (DFT) have to be employed [51, 52].

To begin with, the electronic structure of a single molecule has to be understood. Figure 2.10 helps to visualize the importance of the electrostatic potential for the further discussion. The illus- trated results represent the evolution of the electrostatic potential through a free-standing layer2 of molecules that contain an aromatic biphenly core [53]. As indicated by the horizontal lines, the energetic positions of HOMO and LUMO remain unaffected by the substitution of head and docking group, since the frontier molecular orbitals are dominated by the aromatic cores. Still, the electrostatic potential and therewith the position of the vacuum level depend strongly on the local composition of the molecule. Hence for an asymmetric molecule, one can identify two different left vacuum levels: One for the docking group on the left side of the molecular layer, Vvac , and one right for head group on the right side, Vvac . Consequently, the energy needed to inject (withdraw) an electron from the molecule is given by EAleft (IP left) for the left side, but may differ on the right side with EAright (IP right). As discussed in the previous sections, the position of the vacuum level

Evac crucially affects the energy level alignment of a metal/molecule interface. Substituting the docking group hence effectively enables to tune the position of the frontier MO’s with respect to the metal Fermi level EF. As can be observed directly from figure 2.10, asymmetric molecules face right different vacuum levels for the left and right side. The potential difference ∆E V left V vac = vac − vac correlates to the intrinsic molecular dipole moment according to the heuristic Helmholtz equation [54]: ( e)µ ∆Evac − ⊥ , (2.11) = − ²0 A with e being the (by definition positive) elementary charge, µ the dipole moment of a molecule ⊥ within the monolayer perpendicular to the surface, and the area per molecule A. Hence, an electron-withdrawing head like the cyano (-CN) group raises the vacuum level. On the con- right trary, a more electron-donating amino (-NH2) group lowers Evac and therewith the corresponding electron affinity and ionization potential. A symmetric molecule with a thiol (–SH) group at- tached as both head and docking group consequently exhibits the same vacuum level on both sides right E left E . left = left 2The concept of the free-standing monolayer will be discussed in detail in section 4.7. In brief, this layer represents one layer of oriented molecules in the absence of any surface, thus representing only the molecular properties using periodic calculations.

23 Chapter 2: Materials and Theoretical Background

Figure 2.10: Electrostatic potential energy for electrons across a free molecular layer from DFT calculations. The effect of substitution of head (A) and docking group (B) on the potential energy is illustrated at the example of molecules containing a biphenyl core. In all cases, the positions of HOMO and LUMO remain (in first approximation) unaffected by the substitution, since the frontier molecular orbitals are dominated by the aromatic cores. The diagrams demonstrate how the electrostatic potential and therewith the electron affinity EA and ionization potential IP lo- cally vary and can be tuned with different substitutions. In particular, asymmetric molecules face different potential conditions depending on the spatial orientation. Hence, vacuum level ¡ left ¢ and molecular energies are given for the left and right side of the molecule, Vvac ,EAleft, IPleft ³ right ´ and Vvac ,EAright, IPright , respectively. Reprinted with permission from [53]. Copyright 2007 American Chemical Society.

In DFT, the charge density plays a central role for the analysis of the chemical bond formation, as will be discussed in chapter 4. Therewith, DFT enables to analyze the bond formation by a direct comparison of the molecular and metal’s charge distribution before and after adsorption. The resulting charge rearrangements display a newly formed dipole moment, which is induced through bond formation. Hence, this dipole will be referred to as bond dipole BD. The sum of the vacuum level shift ∆Evac of the free molecular layer with the bond dipole hence generates the work function shift ∆Φ ∆ BD. Still, this shift remains limited to the band gap region between = vac + HOMO and LUMO of the organic semiconductor: The effect of Fermi level pinning (section 2.2.3) persists and can be reproduced also on the basis of DFT [53–56]. Its origin however, differs again from the models discussed above, since bond formation and strong hybridization of orbitals are involved. Thus, both the ICT and gap-states models are not appropriate for organic monolayers featuring bond formation. Instead, it is important to highlight the function of the binding groups of figure 2.10B: Since the polarity of this molecular group dominates the position of the vacuum left level Vvac and therewith the level alignment to the (metal) surface. In most examples, charge transfer across the interface is then spatially confined to this region [53, 56].

How does Fermi level pinning take place when chemical bonds are formed between the molecule and the metal? Figure 2.11A illustrates this situation: The broadened HOMO level is illustrated as a Gaussian-shaped state. In this case, a significant amount of occupied states are located above

24 Section 2.2: Fundamental Processes at Metal/Molecule Interfaces

Figure 2.11: Schematic illustration of Fermi level pinning in a SAM as has been often observed with DFT [53–57]. (A) Before contact, occupied states of the molecular HOMO (illustrated as broadened, Gauss-shaped level) are located above EF of the metal. (B) In contact, charges transfer from the molecule to the metal, moving the HOMO back below EF (C). The net charge transfer can be very small if charge can move along the whole aromatic molecular backbone and/or if several bonds polarize in order to achieve the necessary potential shift. Adapted with permission from [57]. Copyright 2009 American Chemical Society.

the Fermi level of the metal, EF, just before the contact is established. In contact, charge flows from the molecule to the metal (B), inducing an interface dipole BD that shift the HOMO back below EF (C). As indicated in figure 2.11C, only a small amount of the frontier molecular orbital is effectively emptied, as has been observed frequently [53–56]. Still, a little charge transfer can only generate a small dipole moment at first glance using equation (2.11). For aromatic molecules, this statement is not sufficiently accurate, since the (small) charge transfer can expand over the aromatic backbone. If the charge is transferred over a considerable distance, a sizable potential difference can be generated with a small net charge transfer [54]. Otherwise, if the net charge transfer remains minor and/or locally confined to the docking group of the molecule, a polarization of the SAM can take place: Several bonds of the molecule, irrespective of being π- or σ-type, can get slightly polarized upon adsorption, which effectively generates an interface dipole that pushes

EF back into the band gap region of the semiconductor [55, 56].

If Fermi level pinning occurs (in a SAM), the question arises to what extent one can tune the effec- tive work function of the surface or if it remains limited as soon as EF gets pinned. This question has been addressed by Egbert Zojer and coworkers for different types of aromatic adsorbates us- ing DFT techniques [58, 59]. In this work, EF is pinned to the molecular HOMO of derivatives of terpyrmidine SAM’s on gold. Since this MO is located in the molecular backbone, charge transfer can also be observed to reach this region of the molecule. Yet, to what extent is the work func- tion affected by a replacement of the binding group? Φ remains constant and the work function shift does not depend on the nature of the docking group despite being highly polarized for a free molecule. Since charge is transferred across the interface region of the docking group, the polarity of this group is compensated by the charge rearrangement. Otherwise, a variation of the head group of the SAM directly alters the work function shift: The head group remains unaffected by the charge rearrangements due to the pinning to the molecular backbone. Thus, it is still able to push the electrostatic potential and therewith the work function despite Fermi level pinning.

25 Chapter 2: Materials and Theoretical Background

2.3 Charge Transport in Organic Transistors

Tuning the work function of a substrate aims an optimized charge injection and extraction in or- ganic optical and electronic devices for light emission and power conversion devices, i.e. organic light emitting diodes (OLED) and organic solar cells (OSC). While the realization of stable OLED and OSC devices is a technically challenging task, organic thin film transistors are more versatile to investigate the effect of an interface barrier. Nonetheless, the incorporation of DTC based mono- layers in novel architecture OLEDs has just started in cooperation with the group of Prof. Vescan from the department of electrical engineering.

2.3.1 Charge Transfer through Metal/Molecule Interfaces

For any optoelectronic device that can be envisioned with molecular materials, charge carriers have to be either injected, e.g. in light emitting devices, or extracted for power generation from light absorption in a solar cell. The basic principles of charge injection from a metal electrode into an organic semiconductor is briefly discussed in this section based on the discussion on the metal/organic interface. In principle, two charge injection mechanisms are frequently discussed: Thermionic injection and field emission tunneling [60]. However, as discussed by J. Campbell Scott [61], in any real experiment neither of the two ideal mechanisms is sufficiently accurate to describe charge injection into an organic semiconductor. Still, it is noteworthy to introduce both models since they are the foundation of any discussion on charge injection. The basic principles are illustrated in figure 2.12.

To reach the molecular LUMO from the Fermi level EF of the metal, electrons have to overcome the injection barrier ∆e (this discussion works equally for the injection of holes into the HOMO level with ∆h). If no other than thermal energy is provided to the system, the emission current J of electrons depends only on the barrier height ∆e and the temperature T according to the Richardson formula: ? µ ¶ µ ¶ em 2 ∆e ? 2 ∆e J 3 (kBT ) exp A T exp , (2.12) = 2π~ · −kBT = · −kBT ? ? where A is known as the “effective Richardson constant”, kB is the Boltzmann constant and m is the effective mass in the semiconductor. This scenario of thermionic charge injection is depicted in figure 2.12A, but it makes rigorous simplifications. Originally derived for the emission of elec- trons from a metal into vacuum, the adjustments in equation (2.12) (work function Φ replaced ? by injection barrier ∆e and electron mass me replaced by effective mass m ) imply that the elec- tron are still freely propagating in the semiconductor. This, however, is not the case for organic semiconductors [61, 62].

26 Section 2.3: Charge Transport in Organic Transistors

Δ

Figure 2.12: Charge injection processes at a metal/organic interface. (A) illustrates the regime of plain thermionic injection, in which charge carriers can move from the (metal) electrode into the organic semiconductor if the barrier ∆e can be surmounted by thermal energy. (B) At a sufficiently high voltage drop across the device, charge carrier tunneling through the forbidden band gap area is possible to reach the transport level of the molecule. (C) depicts the realistic scenario that trap states are localized within the band gap region [34, 47]. Charge carriers can use these states to reach the molecular transport level in a multi-step injection process. For a real device, a combination of (A-C) will be most accurate to describe the injection process. Adapted from section 2.4 of [60].

In addition to thermionic injection, tunneling from the metal into the transport level of the semi- conductor can take place if the applied voltage is sufficiently high, figure 2.12B. This process is denoted as field emission tunneling (also known as Fowler-Nordheim tunneling), in which an exponential term describes the reduction of the barrier height by the applied voltage according to [61, 63]: · µ eV ¶ ¸ J exp 1 . (2.13) ∝ kBT − In general, both thermionic injection and field emission simultaneously take place in a real device if a voltage is applied. Still, the localized nature of the electronic states is not incorporated ac- curately. This can be achieved by solving the drift-diffusion equation in the potential across the depletion zone [61, 63]:

µ ¶· µ ¶ ¸ ∆e eV J N0evD exp exp 1 , (2.14) = −kBT kBT − where v is the effective drift velocity, which reduces to v µE for small electric fields E, and D D = N0 is the electron density of states. The key point of the discussion up to here is the fact that the emission current J of equation (2.14) is still dominated by the barrier height ∆e even though it can be lowered by an applied electric field. An additional injection mechanism that takes place is a multi-step tunneling process through localized gap states that are known to appear even in highly

27 Chapter 2: Materials and Theoretical Background

Figure 2.13: Scheme of an OFET with electrode modifications as has been employed in this work. Left depicts the complete layer stack with source and drain electrodes (here: gold) on a doped silicon substrate that serves as gate electrode with a SiO2 as dielectric layer. The n-type material PTCDI-C13 has been employed as organic semiconductor. The inset in the middle highlights the modified interface of the metal electrode to achieve an improved electrical contact for electron injection. On the right, the molecule C2DTC is illustrated in detail, which has been taken for the electrode modification. purified materials [34, 47], as sketched in figure 2.12C. However, an analytical description then becomes difficult and the reader is referred to the discussion of J. Campbell Scott in [61] and the references therein.

In conclusion, the control over the barrier height ∆e remains a crucial step on the path to optimized charge transfer at metal/organic interfaces.

2.3.2 Organic Field-Effect Transistors

Figure 2.13 illustrates a common organic field-effect transistor (OFET) on the left side and high- lights the molecular electrode modification as has been employed in this work. All devices have been fabricated in a bottom-gate bottom-contact geometry: A highly doped silicon layer serves both as gate electrode and substrate. On top, a thermal oxide layer (400 nm) of SiO2 is grown to electrically isolate the active layer, source and drain from the gate electrode. A bottom-contact geometry has been chosen since the source and drain electrodes are coated with a DTC-based monolayer prior to the deposition of the n-type semiconducting material PTCDI-C13.

Figure 2.14 graphically illustrates the functionality of an OFET. The simplified sketch basically displays the principal effects when voltages are applied to gate and drain electrode (the source electrode is always connected to ground potential). Figure 2.14A depicts the situation before any voltage is applied. The (localized) positive and negative charges are randomly, but homogeneously distributed in the organic semiconductor (SC). Applying only a drain voltage without charge accu- mulation at the dielectric generates only negligible charge flow, since the intrinsic mobility of most

28 Section 2.3: Charge Transport in Organic Transistors

Figure 2.14: Schematic illustration of charge flow in an OFET. (A) Without any voltage applied to the device, all charge carriers are homogeneously distributed in the organic semiconductor. (B) Charge accumulation at the dielectric is generated for electrons (holes) as soon as a positive (negative) gate voltage VG is applied. (C) When a drain voltage VD is set, the charges accumulated near the dielectric surface can move almost freely from source to drain. molecular solids is too low [64, 65]. Hence, charge carriers have to be accumulated by applying a 3 gate voltage VG, as sketched in figure 2.14B for electrons . Finally, by raising the drain voltage

VD, figure 2.14C, a significant charge flow can be observed.

During operation, one can classify different regimes of an OFET and it is thus useful to quantify the charge distribution [66, 67]. The density of accumulated, free charges q(x) in the channel at position x can be expressed as

q(x) n(x)ed C (V V V (x)) . (2.15) = = i · G − T − with the electron density n(x), elemental charge e and the thickness d of the conducting chan- nel. Equation (2.15) takes advantage of the geometry of the device as a capacitor with an area- normalized capacitance of the insulating oxide Ci. VT represents the so-called threshold voltage that indicates, at which gate voltage the device turns from the OFF to the ON-state. Using the combination of the current density j σE and the conductivity σ µn(x)e with equation (2.15) = = leads to: µ dV µCi dV j q(x) (VG VT V (x)) . (2.16) = d dx = d · − − · dx Now, this expression can be integrated over the complete channel of length L and width W . Using j ID , the drain current I is given as = W d D

" 2 # W VD ID µCi (VG VT) VD . (2.17) = L − · − 2

Equation (2.17) accurately describes the drain current in the channel of an OFET in the so-called

3The working principle works equally for hole accumulation with applied voltages of the opposite sign.

29 Chapter 2: Materials and Theoretical Background

linear regime. For the case that (V V ) V , the second term in the brackets of equation (2.17) G − T À D becomes negligible and the influence of VD on the distribution of free charge carriers almost van- ishes, i.e. all charges are uniformly distributed within the channel. The situation changes when V reaches the same order of magnitude as the difference (V V ). This transition regime leads D G − T to the pinch point: The channel is said to get pinched-off at (V V ) V . At this point, there G − T = D is no potential difference between the gate and the channel region near the drain electrode, which results in a region absent of free carriers. Thus, a further increase of VD cannot raise the drain current, which is denoted as saturation regime with

2 W (VG VT) ID µCi − . (2.18) = L 2

The pinch point, at which (V V ) V , is a useful point in a measurement that can be employed G − T = D to estimate the threshold voltage VT and the field-effect mobility µ. Plotting pID against VG yields a linear trend that can be easily fitted with a straight line. From the slope of the fit, the mobility µ can be calculated and the intercept at the x-axis can be used to extract VT. This is a versatile and very common technique to quantify the characteristics of an OFET. Still, it should be noted that the field-effect mobility µ is field-dependent on VG and therewith strongly affected by the device geometry, temperature, traps states and magnitude of the applied voltage [67–72]. It is thus not uncommon to find a broad range of mobility values for one material.

2.3.3 Contact Resistance in OFETs

In order to quantify the influence of an electrode surface modification in an OFET, the transfer line method (TLM) can be employed to examine the contact resistance Rc of the device [70, 73]. This method is restricted to the linear regime of an OFET and should only be used for devices with a coplanar stacking geometry [71, 73]. It is based on the assumption that the channel path from source to drain can be represented by a series of resistances: A contact resistance Rc at the electrode/semiconductor interface and a channel resistance Rch. The TLM is furthermore based on the assumption that Rc is independent on L, but Rch raises linearly with L [70]. Then, the drain current can be written as

VD VD VD I , (2.19) D ¡ ¢ 1 = RON = Rc Rch = R W/L µC V V − + c + · i | B − T| where the linear-regime channel conductance has been used to estimate Rch [70, 73]. The ON-state channel resistance RON can be directly calculated by measuring both ID and VD:

VD L RON Rc . (2.20) = I = + W µC V V D i | G − T|

30 Section 2.3: Charge Transport in Organic Transistors

It is useful to rearrange equation (2.20) by multiplying the expression with W to normalize the contact resistance to the channel width:

L RON W Rc W . (2.21) × = × + µC V V i | G − T| This expression can be exploited to obtain a value for R by plotting the product of R W c ON × versus the channel length L. A linear fit then yields R W at the intercept of the y-axis. Hence, c × several devices with different channel lengths have to be analyzed in order to obtain a value for

RON. This can turn out to be a crucial point, since this analysis is strongly dependent on accurate channel lengths. In addition, device characterization under ambient conditions may affect the result when the diffusion of molecules like oxygen or water creates trap states in the channel and at the electrode/semiconductor interface.

31

Chapter 3

Photoemission Spectroscopy

Since the explanation of the photoelectric effect by Albert Einstein [74], photoelectron spec- troscopy (PES) has developed to become one of the commonly employed methods to probe the electronic and chemical structure of materials. Therefore, this technique plays a key role in this thesis to analyze and understand chemical bonding of molecules on metal surfaces. In the fol- lowing sections, the fundamental principle of photoemission, the processes accompanying a PES measurement and experimental details will be introduced and briefly discussed. In particular, the discussion of the fundamental processes is based the textbooks Photoemission in Solids, edited by Cardona and Ley, and Photoelectron Spectroscopy written by Stefan Hüfner [75, 76].

3.1 Photoemission from Surfaces

In 1887 Heinrich Hertz reported the external photoelectric effect using illumination of ultra-violet light on a metal plate [77]. Yet, it was Albert Einstein’s ground-breaking assumption of localized light quantums in 1905 that finally could explain this effect and enabled the prediction of further interactions of light with matter on a theoretical level [74]. Upon absorption of a single photon, an electron may get liberated from the nucleus if the photon energy is sufficiently large to compensate the binding energy of the electron. After leaving the surface, the electron moves with a kinetic energy EK through the vacuum according to

E hν E Φ , (3.1) K = − B − where the binding energy EB is referred to the Fermi level EF. Hence, there exists a minimum energy barrier Φ, also known as work function of a surface, which has to be overcome. If the

33 Chapter 3: Photoemission Spectroscopy

kinetic energy EK of the electron is less than this threshold value Φ, the electron will always remain in the surface. Generally, PES is roughly categorized into two major sections depending on the excitation energy of the incident light: Ultraviolet spectroscopy (UPS) usually probes the valence structure using excitation energies in the range of 10 eV to 100 eV, while X-ray photoelectron spectroscopy (XPS) employs higher energies up to several keV to excite and analyze strongly bound core electrons.

At this point, it is important to point out that PES is very surface sensitive technique. The experi- mental information depth is equal to the electron escape depth, which is strongly dependent on the inelastic mean free path λ (IMFP) of electrons moving through a material to reach the surface. As analyzed by Seah and Dench, the electron’s IMFP depends on its kinetic energy EK as well as the material species, in which the electrons travel [78]. This defines the characteristic path length that dominates the electron’s probability decay given by

I I exp( x/λcos(θ)) , (3.2) = 0 − where x is the distance that the electron has traveled from its origin. With a moving direction at an angle θ to the surface normal, the effective escape depth is then reduced to λ cos(θ). With · values for the IMFP λ in the range of a few Å for UPS and up to a few tens of Å for XPS, it is easy to conclude that the experimental information depth also remains confined to the first view nanometers using PE techniques.

3.2 Theory of Photoemission

Photoemission is usually explained in the so-called three-step model, which has been pushed for- ward by Berglund and Spicer [79, 80]. In spite of being purely phenomenological, it has been proven to successfully explain the fundamental principles of photoemission by dividing the com- plete process into three parts. Nevertheless, peak intensities are incorrectly calculated caused by some poor approximations (e.g. the IMFP of electrons is inaccurately incorporated in any of the steps). A correct, yet less descriptive approach is the one-step model, in which the electron is accurately described by a Bloch wave that decays inside the solid, but moves freely in vacuum [76]. For a detailed description of the one-step model of photoemission, the reader is referred to the review article of Braun [81]. For the phenomenological explanation of the complex process of photoemission, the three-step model is discussed in the following:

1. Optical excitation of an electron in the solid

2. Transport of the excited electron to the surface of the solid

3. Escape of this electron through the surface into vacuum

34 Section 3.2: Theory of Photoemission

Figure 3.1: Impact of crystallinity and roughness on the generation of secondary electrons in a UPS measurement of Ag(111). (A) illustrates the valence structure of a highly crystalline and tex- tured Ag(111) surface with extremely low roughness grown on a mica substrate. (B) depicts the UPS spectrum of a polycrystalline Ag(111) surface, which has been sputter-cleaned several times in order to remove unwanted adsorbents. Sputtering, however, increases the surface roughness.

The total amount of electrons I ext that escape from the solid are a sum of primary and secondary electrons: ³ ´ ³ ´ ³ ´ I ext E ,hν,~k I ext E ,hν,~k I ext E ,hν,~k . (3.3) K = prim K + sec K Primary electrons have succeeded the complete photoemission process without any energy loss after excitation, whereas secondary electrons are mainly generated by inelastic scattering. The amount of secondary electrons is influenced by many parameters such as crystallinity of the solid material, surface roughness, temperature (i.e. phonon scattering), contamination of the surface, material species and more. A demonstrative illustration of the effect of secondary electrons on a spectrum is visualized in figure 3.1 using a comparison of UPS spectra of a clean polycrystalline Ag(111) sample versus a clean and highly textured (monocrystalline-like) Ag(111) thin film. The increased amount of secondary electrons are a direct result of to scattering due to increased surface roughness as well as a less optimal texture. Thus, any kind of impurities, (organic) surface layers and also 2d-ordered SAMs serve as additional scattering objects, which also raise the amount of secondary electrons. In the following, however, only the processes accompanying primary electrons will be considered. An expression for secondary electrons has also been developed by Berglund and Spicer [79, 80].

Coming back to the external emission current I ext of the three-step model, one can express I ext de- pending on kinetic and photon energies EK and hν, respectively, as a product of the three processes by ³ ´ ³ ´ ³ ´ I ext E ,hν,~k P int E ,hν,~k T (E ) X E ,~k . (3.4) K = K · K · K

35 Chapter 3: Photoemission Spectroscopy

int Here, P denotes the internal distribution of excited electrons, T (EK) the transport function, and ³ ´ X EK,~k the escape function.

To generate an expression for the distribution of excited electrons inside the solid, the transition probability W for photoexcitation of an electron from an initial state Ψi to a final state Ψj can be expressed using Fermi’s golden rule:

³ ´ 2π ¯­ ¯ ˆ ¯ ®¯2 ~ ~ W ¯ Ψj ¯ H 0 ¯ Ψi ¯ δ Ej(k) Ei(k) hν , (3.5) = ~ · − − which was originally derived by Dirac [82], but Fermi named it as one of the golden rules of quantum mechanics. In equation (3.5), the initial and final states i and j have the energies Ei(~k) ­ ¯ ¯ ® and Ej(~k), respectively. The perturbation operator Hˆ 0 of the matrix element Mi j Ψj ¯ Hˆ 0 ¯ Ψi → = can be expressed as performed in textbook quantum-mechanics for the interaction of light with matter while energy conservation is ensured by the delta-function. Summing over all initial and final states then yields an estimation for the internal distribution of photoexcited electrons [76]:

int ³ ´ X¯ ¯2 ³ ´ ³ h i´ P EK,hν,~k ¯Mi j¯ δ Ej(~k) Ei(~k) hν δ EK Ej(~k) (EF Φ) . (3.6) ∝ i,j → − − − − +

As indicated by the second delta-function in equation (3.6), an electron will be detected with a kinetic energy EK that is equal to the energy difference between the final state Ej(~k) and the vacuum level E E Φ. vac = F +

In the second step, the excited electrons travel through the material and may reach the surface. As discussed above, the IMFP of electrons is the dominating parameter describing scattering effects. In order to incorporate this effect, the transmission function T (E) has been classically derived by Berglund and Spencer as the fraction of total number of photoelectrons within one mean free path [76, 79]: αλ(EK) T (EK) , (3.7) =∼ 1 αλ(E ) + K with the optical absorption coefficient α of the incident light. Hence, in the limit of αλ 1, in ¿ which the IMFP of the electrons is much smaller that the penetration depth of the photon (which is a valid assumption in many cases), the transmission function converges to T (E) αλ. →

Finally in the third step, the transmission of an electron through the surface into vacuum is consid- ered. In normal direction to the surface, the electron must have a minimal energy to just surmount the energy barrier E Φ [75, 83]. With an angle θ to the surface normal, the following condition F + can be stated: µ ¶1/2 k EF Φ cos(θ) ⊥ + , (3.8) = k ≥ EK

36 Section 3.2: Theory of Photoemission

θ

θ ┴

θ

Figure 3.2: Escape condition for the photoexcited electron defining the angular region, in which electrons are able to leave the surface. On the bottom, the internal escape cone is highlighted by the segment with the thick line. Due to the condition that the photoelectron must have a minimum kinetic energy and, therefore, a minimum wave vector component k in the direction ⊥ perpendicular to the surface, a maximal escape angle θmax can be defined. Outside of the solid, the wave vector is defined by k0 p/~. Scheme adapted from [76]. =

2 2 with the kinetic energy inside the solid EK ~ k /2m. Equation (3.8) effectively defines an escape = cone with an opening angle θ depicted in figure 3.2, illustrating that the k -component outside of ⊥ the solid is always smaller than inside, while the parallel component k remains conserved. Only ∥ those photoexcited electrons, which propagate within the angular region θmax of the escape cone are able to leave the surface. Under the assumption that all generated photoelectrons move with equal probability in all directions, the fraction of electrons that actually escapes from the surface is given by:

" µ ¶1/2# 1 EF Φ X (E) 1 + EK EF Φ (3.9) = 2 − EK > + 0 elsewhere. (3.10) =

Putting all processes of the three-step model from equation (3.4) together generates an expression of the external current of photoexcited electrons. Therein, the transport and escape functions

T (EK) and X (EK,~k) describe processes of photoelectrons, but not their generation. Instead, T (EK) and X (EK,~k) lower the total intensity by reducing the current of primary photoelectrons, which results in a smearing of the spectrum. Hence, Fermi’s golden rule from equation (3.5) effectively dominates the spectral shape of a measured PE spectrum: Due to the delta-function that represents all initial and final states, a PE spectrum projects the density of states (DOS(Ei)) which can be

37 Chapter 3: Photoemission Spectroscopy

assessed with a photon energy of hν. Electrons with different angular momenta will contribute differently to intensity, because the angular momentum of the initial state is stored in the matrix ¯ ¯2 element ¯Mi j¯ . Taking advantage of angle-resolved PE measurements then offers the opportunity → to investigate the band structure of a sample.

3.3 Electron Energetics at Surfaces

A key objective of this work using photoelectron spec- E E troscopy experiments is the quantification of the work function Φ, i.e. the determination of the minimal energy, E ∞ which is necessary to liberate an electron from a sam- ϕ ple’s surface. At this point, it is important to examine E the concept of the vacuum level: Ishii et al. have pointed out the fact that the effective work function of an ele- mental crystal may depend on the (hkl) surface plane [37, 84, 85]. As an example, the work functions of dif- Figure 3.3: Vacuum level close to the sur- face, Evac(s), and at large distance ap- ferent surfaces of a tungsten single crystal can be consid- proaching infinity, E ( ). The work vac ∞ ered: Φ 4.63eV, Φ 5.25eV and Φ 4.47eV function Φ is the energy difference be- 100 = 110 = 111 = [84]. Since the Fermi level is an invariant energy of the tween Fermi level EF of the sample and its local vacuum level. solid, independent on any crystal orientation, the differ- ences of the observed work function values have to originate from the surface itself. This, however, raises the question if there is a unique vacuum level that may serve as an appropriate reference level. An electron at rest in vacuum with infinite distance to any surface resides at the vacuum level E ( ), which is also an invariant energy. Yet, this energy level usually differs from the vac ∞ vacuum level close to a surface and will be denoted as Evac(s). The difference between Evac(s) and E ( ) is correlated to electrostatic effects that affect the electron’s energy in a local environment vac ∞ to the surface. Figure 3.3 illustrates the different vacuum levels associated with a surface: In the vicinity of the surface, one has to invest a minimal energy ∆E E (s) to carry an electron from = vac the Fermi level EF to the vacuum level of the surface Evac(s). This minimal energy difference is known as the work function Φs of a surface. However, at large distance from the surface, the ref- erence level of the electron is the vacuum level at infinity E ( ). The electron may gain kinetic vac ∞ energy due to loss of potential energy when leaving the surface, thus conserving total energy. The energy difference between these two vacuum levels originates from contributions such as surface dipoles or charges at the surface. For metallic surfaces, surface dipoles usually originate from the spilling of electronic charge out of the surface, which leaves a positive charge inside of the solid and raising the vacuum level E (s) compared to E ( ), as depicted in figure 3.3. vac vac ∞

38 Section 3.3: Electron Energetics at Surfaces

E

E E Δ Figure 3.4: Schematic illustra- E tion of the energy landscape of ∞ ϕ Δ vacuum levels of a crystal with ϕ different (hkl) facets. No energy E is gained or lost when transfer- ring an electron from the (100) to the (110) or any other (hkl) surface. Explanation see text.

In an educational review article, Cahen and Kahn have discussed the consequences of the different vacuum levels [86]. Using a gedanken experiment, they conclude that it is not possible to build a perpetuum mobile by subsequently ejecting electrons from a low work function surface and injecting them through a high work function surface. At the example of tungsten, a minimal energy of 4.63 eV is required to excite an electron from the Fermi level EF just above the (100) surface. A subsequent movement of this electron above the (110) surface requires an energy investment to lift the electron to the higher vacuum level. Thereafter, the electron gains energy of 5.25 eV upon entering the solid again. The energy discrepancy equals the difference between the vacuum levels of the two surfaces, i.e. 0.62 eV for this example. This energy difference turns out to be exactly the energy gain, one would expect from the difference of the two work functions. Therefore, no energy is spent or gained in this experiment. Alternatively, the electron could be moved far away from the (100) surface to reach the vacuum level at infinity and to subsequently approach the (110) surface. Still, there would be no energy gain, as illustrated in figure 3.4. In the beginning, the energy

Φ100 has to be invested to excite an electron from the Fermi level to the vacuum level Evac(100) of the (100) surface plane. At large distance to this surface, the electron relaxes to the universal vacuum level E ( ) and gains an unknown energy value of ∆ . There again, by approaching vac ∞ 1 from infinity to the (110) surface, one needs to invest the energy ∆ ∆ (Φ Φ ). Thus 2 = 1 + 110 − 100 in total, the energy difference ∆1 cancels, therefore total energy remains conserved. In summary, there will be no energy gain or loss when moving an electron from one crystal surface to another through the vacuum.

Hence, the vacuum level at infinity E ( ) is an invariant reference level, which is experimen- vac ∞ tally not accessible. However, it is also an experimentally irrelevant parameter, since the minimal energy to extract an electron from a surface is the difference between the Fermi level EF and the lo- cal vacuum level of the surface E (s), and not at infinity E ( ). The aforementioned gedanken vac vac ∞ experiment can be applied to a real experiment as well. Therein, electrons are liberated from a sample’s surface and measured via dropping onto the surface of a detector. Since the detector is a solid with its own work function Φd, electrons move a similar way from sample to detector

39 Chapter 3: Photoemission Spectroscopy

E e- e-

- EK,max Evac(s) e Figure 3.5: Energy diagram of EK,min a sample and detector at equi- Evac(d) librium during photoexcitation. Evac(∞) Electrons are excited with hv hv ϕs and travel to the detector, at ϕd which a characteristic kinetic energy range is measured. EF EK,min and EK,max depend only on the energy level of the sample. Without an external voltage, the Fermi energy EF sample detector remains a constant of the total system, adapted from [86]. as depicted in figure 3.4. Again, the vacuum level at infinity E ( ) remains irrelevant for the vac ∞ experimental work function. A more detailed scheme of this process is depicted in figure 3.5. Upon photoexcitation, electrons travel from the sample to the detector with a characteristic kinetic energy EK, which depends on the initial electron binding energy EB and the energy of the irradia- 1 tion hν. Under the assumption , that the vacuum level of the detector Evac(d) is below the one of the sample Evac(s), electrons that had just enough energy to leave the surface of the sample, will always arrive at the detector with a minimal kinetic energy EK,min. This minimal energy corre- sponds to the onset of photoionization in a measurement and remains independent on the position of E ( ), but can be directly related to the energy difference Φ Φ . Electronic states can vac ∞ s − d therefore be characterized for binding energies E E hν Φ , as schematically represented B ≤ F − + s in figure 3.6. Primary electrons, which are located relatively close to the Fermi level EF, profit from a weak background noise in a measured spectrum. Since every electron has a non-zero prob- ability of scattering during (and after) excitation, secondary electrons are generated especially at lower kinetic energies. Thus, an increasing background noise towards the ionization threshold is measured with the detector at EK,min. Deep core level states with high binding energies are not available with UPS. Finally, the full spectrum of electronic states, as measured at the detector, can be exploited to calculate the work function Φs of the sample by

Φ E E E hν E . (3.11) s = max − K,max = K,min + − K,max

1This is in fact a reasonable condition: In a usual UPS measurement, a constant voltage is applied to lower the Fermi level of the detector with respect to the Fermi level of the sample. Thus, it is ensured that a measured photoionization threshold is a unique feature of the sample and not an artifact of the work function of the detector.

40 Section 3.4: Quantitative Analysis of Chemical Composition

Emax = Evac(s) + hv ϕs

EK,max = EF + hv

hν Evac(s) EK,min

ϕs Evac(d) ϕd EF EF

valence states hν

core level states

Sample Detector

Figure 3.6: Schematic representation of the energetic processes in a photoelectron measurement. Upon photoexcitation, electrons from valence and core level states gain energy and can escape from the sample if the final energy is at least equal to the vacuum level Es of the sample. The grey shaded energy area represents the region, which is not accessible in a measurement, since electrons in this area will not be able to escape from the surface. All other electrons travel to the detector, at which the kinetic energy of the impinging electrons is measured. Kinetic energy can be calculated back into binding energies via E E Φ hν. B = K + d −

3.4 Quantitative Analysis of Chemical Composition

In this thesis, XPS measurements have been conducted with two goals: First, the position, i.e. the binding energy EB, of core level electrons is characteristic for the chemical surrounding of the liberated electron. This allows to identify atomic species and chemical bonds. Second, a quantitative evaluation can thus be performed in order to estimate the chemical stoichiometry of the composition.

The binding energy of an electron is strongly dependent on its chemical environment. Still in most examples, a measured binding energy EB can be traced back to one elemental species. The com- bination of being both sensitive to the surface and the elemental species is therefore an advantage of a PES measurement when investigating a surface composition. Moreover, chemical shifts of core level arise for different chemical surroundings, i.e. bonds to different elements. Depending

41 Chapter 3: Photoemission Spectroscopy

S S

SS SS

Figure 3.7: Core level spectra of carbon upon fluorination. A comparison of a hydrogenated decanethiol monolayer on gold (A) with that of the fluorinated counterpart 1H,1H,2H,2H- perfluorodecanethiol (B). All peaks refer to the carbon C 1s orbital. The shifts in (B) from the C-C/C-H state at 285.2 eV are attributed to the different chemical environment of carbon with bonds to either hydrogen or different amounts of fluorine. Since fluorine is strongly electron- withdrawing, carbon is partially electron deficient. Thus the binding energy of the remaining electrons in the corresponding carbon atom is increased.

on the nature and strength of participating chemical bonds, these shifts can range from several meV up to 10 eV [75]. The effect of fluorination of an alkyl chain is illustrated in figure 3.7 for the carbon C 1s core level in a monolayer of 1H,1H,2H,2H-perfluorodecanethiol on gold. Fluorine is the most electronegative element in the periodic table of elements. When fluorine is bound to carbon, it causes a shift of the C 1s core level to higher binding energies. Electrons, which are incorporated in this chemical bond, are effectively withdrawn from carbon by fluorine. Thus, the remaining core electrons are more strongly bound to the carbon core. In general, unknown binding energies are usually compared to published reference values for similar compositions involving the same elements.

A quantitative analysis of core level spectra is in principle a complex task [75, 76], since there are many effects involving the absolute intensity of a photoelectron peak. Additionally, many of them cannot be easily calculated. Nevertheless, when taking the intensity ratio of different core levels in order to evaluate the elemental composition, many of these factors cancel in a good approximation. The most dominant factors are the absolute peak intensity area Ax of element x and the corresponding cross section for photoionization σx, calculated by Scofield [87]. The relative

42 Section 3.5: Surface Coverage of Organic Layers using XPS

intensity nx can be estimated by the quotient

Ax nx . (3.12) = σx

In an ideal analysis, the Scofield cross sections σx are replaced by true atomic sensitivity factors, which incorporate all effects during the process of photoionization (see alsosection 3.2). The relative intensities from equation (3.12) can be further used to calculate the percentage Cx of a specific element x in a homogeneous sample according to

A nx x/σx Cx . (3.13) P P A = i ni = x i/σi

3.5 Surface Coverage of Organic Layers using XPS

Determining the surface density of molecular monolayers is a key step to characterize the surface properties upon SAM formation. Hence, it is mandatory to find an accurate way of estimating surface coverages of molecules on different (metal) surfaces using the information acquired with XPS. As shown in previous studies, the molecular density and packing motif are important char- acteristics of an organic monolayer [32, 88–90]. The determination of surface coverage in this work is based on two publications, which both have established the estimation of molecular den- sities of molecules in a monolayer. The first method is based on the work of Frank Schreiber, who analyzed the crystal structure, packing and orientation of thiolate SAMs on gold with various experiments [50]. The second model has been developed by Bramblett et al., who compared two different experimental methods to validate their results [91]. Both publications served in this work as guidelines to estimate the surface density of molecules on surfaces. Thus, these methods will be discussed in the following.

3.5.1 Determining Surface Coverage according to Schreiber

In a comprehensive work, Frank Schreiber has reviewed growth and structural phases of self– assembled monolayers [50]. The focus of his study is concentrated on alkyl thiols on gold surfaces as an archetypal system to investigate the process of self assembly, different phases (high versus low molecular coverage), effects of different molecular backbones, temperature dependencies and more. In particular, the manifold investigation of decanethiol (C10H21SH) on gold with various techniques allows to conclude on the molecular density from a single XPS measurement. It is gen- erally accepted, that the XPS sulfur to gold intensity ratio can be employed to estimate a molecular coverage [27, 92]. Decanethiol arranged in a hexagonally close–packed (hcp) monolayer occupies 2 14 an area of 21.6 Å , which corresponds to a surface density of 4.6 10 molecules/cm2. In addition, ·

43 Chapter 3: Photoemission Spectroscopy

the sulfur to gold intensity ratio χ measured with XPS yields a value of approximately 0.05. This combination can be exploited to estimate the surface coverage Θ of dithiol or dithiocarbamate (DTC) covered gold surfaces. Since a DTC molecule exhibits two sulfur atoms (compared to one for thiols)this factor of 2 leads to the relation:

IS2p σAu4f 14Thiols 2 χThiol 0.05 Θ 4.6 10 /cm− (3.14) = IAu4f · σS2p = ⇐⇒ = · IS2p σAu4f 14DTCs 2 χDTC 0.05 Θ 2.3 10 /cm− . (3.15) = IAu4f · σS2p = ⇐⇒ = ·

The validity of equation (3.15) can be verified by comparing results for diethyl–dithiocarbamate monolayers (abbreviated by C2DTC) on gold published by Schulz et al. with those of Morf and coworkers. [27, 93] In the latter, Morf et al. investigated the supramolecular structures of dithiocarbamates on gold using STM imaging and obtained an average area per C2DTC on gold 2 14 of approximately 45 Å , which corresponds to a molecular surface density of 2.2 10 molecules/cm2. · 14 In parallel, Schulz et al. have calculated a value of 2.5 10 molecules/cm2 from XPS measurements · employing equation (3.15), which only slightly overestimates the value reported by Morf et al.

However, equation (3.15) is only applicable for thiolate SAMs on a gold surface. In this work, monolayers formed on silver and copper are also investigated to further examine monolayer for- mation. To quantify molecular surface densities, equation (3.15) has to be modified to account for the signal intensities of different core level in a XPS measurement. Hence, a quantitative descrip- tion for X–ray photoelectron intensities from pure elements has to be found. An analysis of such intensities has been performed by Seah et al. in a combined experimental and numerical work

[94]. Therein, the X–ray photoelectron intensity IXi of electrons in a core level i of element X is expressed by I (theo,hν) σ (hν)sec(α)N Q ¡E ¢λ ¡E ¢ , (3.16) Xi = Xi X X K,Xi X K,Xi where σXi is the cross section for photoionization at a photon energy hν (taken from Scofield

[87]), α is the angle of of the incident X–ray beam from the surface normal, NX is the atomic ¡ ¢ concentration per unit volume, QX EK,Xi is the intensity reduction due to elastic scattering effects ¡ ¢ of electrons with kinetic energy EK,Xi (after Jablonski [95]), and λX EK,Xi is the inelastic mean free path (IMFP) for electrons with kinetic energy EK,Xi (after Seah and Dench [78]). In addition, photoelectrons excited from a metal core level have to move through an organic layer before entering the analyzer. Therefore, these electrons experience an additional exponential decay due ¡ ¢ to scattering while passing the organic material of thickness t with exp t/λ [78]. Here, the − XiM IMFP λXiM refers to the movement of electrons through an organic layer as reported by Seah and Dench [78]. Finally, working with a hemispherical analyzer causes a dependency of the measured intensity with the kinetic energy of the photoelectrons by I 1/E [96]. Taking all these factors ∝ K,Xi into account, the intensity of core level i from element X can be set into relation to the intensity of

44 Section 3.5: Surface Coverage of Organic Layers using XPS

the Au 4f core level via

¡ ¢ ¡ t ¢ IXi σXi NXi QXi EK,Xi λXi EK,Au4f exp /λXiM ¡ ¢ ¡ − ¢ . (3.17) IAu = σAu · NAu · Q E · λAu · EK,X · exp t/λAu M 4f 4f 4f Au4f K,Au4f 4f i − 4f This quotient can now be applied to equation (3.15) for the examination of surface densities of DTC molecules on surfaces other than gold,

I ¡ ¢ ¡ t ¢ S2p σXi NXi QXi EK,Xi λXi EK,Au4f exp /λXiM χDTC ¡ ¢ ¡ − ¢ 0.05 (3.18) = IX · σS · NAu · Q E · λAu · EK,X · exp t/λAu M = i 2p 4f Au4f K,Au4f 4f i − 4f 14 2 Θ 2.3 10 DTC molecules/cm− . ⇐⇒ = ·

The latter expression has been employed as a first estimate to calculate molecular densities of DTC monolayers on different noble metal surfaces based on X–ray photoelectron spectroscopy measurements. Nonetheless, this method may have systematic drawbacks, since this compari- son is originally based exclusively for alkanethiols on a gold surface. As discussed by Ulman, alkanethiols arrange differently on silver as compared to gold: Albeit an almost identical nearest- neighbor distance in the metal (111) surface, different tilt angles of the alkyl backbone with respect to the surface normal are observed [32]. However, this seems to remain a minor source of error for the application of equation (3.18): Since the sulfur atoms are directly adjacent to the metal atoms of the substrate, photoelectrons excited from the respective elements experience almost the same attenuation length due to the overlying backbone of the alkanethiolate SAM.

3.5.2 Determining Surface Coverage after Bramblett et al.

While the original work of Frank Schreiber is based on the structure of thiols on gold, a more generic approach to estimate molecular densities is desirable. Bramblett and coworkers have in- vestigated how porphyrin derivatives form monolayers on gold surfaces using two different tech- niques: Ultraviolet visible absorption and x-ray photoelectron spectroscopy [91]. Porphyrins are useful in this study, since derivatives of this molecular class are often colorful dyes with strong ab- sorption coefficients in the visible and ultraviolet spectral region [97, 98]. To achieve an analysis using XPS, the authors have derived a theoretical formalism to determine the molecular density in a monolayer. Even though they discuss their analysis only for molecules on gold, this framework can easily be applied to any other (metal) surface as will be discussed in the following.

The starting point is a simple equation that describes the molecular density in a monolayer Γ, see equation (3.19). Dividing the density ρ, given in g/cm3, by the molecular mass MW in g/mol yields the total number of molecules in a unit volume. Since the real molecular layer is supposed to be a monolayer of thickness t, one can derive the number of molecules in this layer by multiplying

45 Chapter 3: Photoemission Spectroscopy

t to the total number of molecules in a unit volume, thus leading to equation (3.19):

3 7 t (nm) ρ (g/cm ) 10− (cm/nm) NAv (molecules/mol) molecules 2 ΓSAM ( /cm− ) · · · . (3.19) = MW (g/mol)

Yet, the monolayer thickness t cannot be defined as an universal parameter that only depends on the molecular geometry. As has been shown for alkyl thiolate SAMs on gold and silver, the effec- tive monolayer thickness depends on the tilt angle of the alkyl chains with respect to the substrate’s surface [88]. Since the sulfur atoms of the thiolate molecules occupy different binding sites on dif- ferent metal surfaces, the alkyl backbone of the molecules tilt differently in order to maximize the intermolecular binding energy via van-der-Waals interactions [32]. Thus, the monolayer thickness t remains an uncertainty that has to be determined experimentally. Fortunately, the attenuation length of photoelectrons is very sensitive to the thickness of a medium that the electrons are pass- ing through. Assuming an organic overlayer, indicated by M, on a gold surface2, the number of M photoelectrons from gold that pass the overlayer, IAu, are a fraction of the original number of elec- pure trons that have been excited from the gold substrate, IAu , and decreases exponentially with rising layer thickness: µ ¶ M pure t IAu IAu exp −¡ ¢ . (3.20) = · λAuM EK,Au ¡ ¢ λAuM EK,Au represents the inelastic mean free path of a photoelectron that has been excited from gold with a kinetic energy EK,Au and passes through the organic layer M. If a nonzero take–off angle θ is considered, then the photoelectron will have to pass an effective distance t t/cos(θ). eff = Thus, to generalize equation (3.20) t is replaced by t/cos(θ):

µ t ¶ I M I pure exp − . (3.21) Au = Au · λ ¡E ¢ cos(θ) AuM K,Au · Accordingly, the monolayer thickness t can be calculated by rearranging equation (3.21). The number of excited photoelectrons for a bare and a modified gold metal surface can be directly detected with XPS by measuring both a plain reference metal sample and a SAM–modified sub- strate. This leads to the first model of Bramblett et al. to estimate the surface density of an organic monolayer with equation (3.19) using

à ! I M t λ (E ) cos(θ) ln Au . (3.22) = − AuM K,Au · · pure IAu

While this first model is simple and easily applied, there are sources of error that can hinder an accurate estimation of the molecular density. First, the density may differ between the molecu-

2For the current discussion, a gold substrate is assumed. Yet, the concept can be directly transferred to other metals by accounting for the different intensities, cross sections and inelastic mean free paths of photoelectrons from other core level.

46 Section 3.5: Surface Coverage of Organic Layers using XPS

lar solid and a monolayer. In addition, the densities of organic materials under investigation in this thesis are not known for all compounds and have been estimated. Second, the total intensity of a photoelectron measurement is crucially dependent on the position of the sample under the analyzer. This is especially challenging when different samples have to be compared. Small dif- ferences in sample height will strongly affect the intensity ratio in equation (3.22) and therefore impact the layer thickness t and finally surface density Γ. At last, surface contaminants will also directly affect the intensity of gold core level photoelectrons, since photoelectrons are equally at- tenuated by the organic monolayer as well as unwanted contaminants. In the end, every attenuation of the photoelectrons will be interpreted as an effective layer thickness independent on the origin of attenuation. A more precise analysis would therefore include a comparison of core levels of the organic monolayer with the gold core level of one sample, similar to the model of Schreiber (sec- tion 3.5.1). In order to achieve a more sophisticated approach, the absolute photoelectron intensity that is measured in an experiment has to be expressed on a more fundamental basis.

At first, a bare surface without any adsorbent is considered. The intensity of photoelectrons excited from core levels depends on several parameters: The spectrometer constant β0, in which the pass energy Epass contributes; the initial X-ray photocurrent J0 impinging the sample; the cross–section

σXi for photoionization of an electron from an orbital i of atom X; the kinetic energy EK,X of an electron from element X; the average atomic concentration NX of element X; the mean free path

λXY of an electron from element X passing through medium Y and the take-off angle θ. Considering a plain gold surface, the intensity can be expressed in the following form:

Z · ¸ pure β0σAu J0 ∞ z IAu NAu(z)exp − dz (3.23) = EK,Au 0 λAuAu cos(θ) β0σAu J0 © ª NAu[ λAuAu cos(θ)] exp( ) exp(0) (3.24) = EK,Au − −∞ − β0σAu J0 NAuλAuAu cos(θ). (3.25) = EK,Au

This expression can be utilized to consider a photoelectron from a gold core level passing through an organic overlayer by inserting equation (3.25) into equation (3.21):

µ ¶ M pure t IAu IAu exp − = λAuM cos(θ) µ ¶ β0σAu J0 t NAuλAuAu cos(θ)exp − . (3.26) = EK,Au λAuM cos(θ)

In analogy to the derivation of equation (3.25), the same considerations are performed for photo- electrons from elements of an uniform organic overlayer. Since any element incorporated in the monolayer can be employed for this calculation, this method is more versatile compared to the analysis by Schreiber, which is limited to the estimation of the sulfur to gold ratio, 3.5.1. Accord-

47 Chapter 3: Photoemission Spectroscopy

M ing to equation (3.23), the intensity IX of an photoelectron from element X passing through the organic layer M can be written as

Z t · ¸ M β0σX J0 z IX NX(z)exp − dz (3.27) = EK,X 0 λXM cos(θ) ½ · ¸ ¾ β0σX J0 t NX [ λXM cos(θ)] exp − exp(0) (3.28) = EK,X − λXM cos(θ) − ½ · ¸¾ β0σX J0 t NXλX,M cos(θ) 1 exp − . (3.29) = EK,X − λXM cos(θ)

M Au Now, the intensity ratio of IX /IX can be expressed as

β0σX J0 n ³ t ´o I M NXλXM cos(θ) 1 exp − X EK,X − λXM cos(θ) ³ ´ (3.30) M = β0σAu J0 t IAu NAuλAuAu cos(θ) exp − EK,Au · λAuM cos(θ) n ³ t ´o σXEK,AuNXλXM 1 exp λ −cos(θ) − XM . (3.31) = ³ t ´ σAuEXNAuλAuAu exp − · λAuM cos(θ) Equation (3.31) represents the intensity ratio as can be obtained in a single measurement. Still, the thickness t of the organic layer remains an unknown parameter, so that equation (3.31) has to be rearranged and solved for t. To simplify, let γ include all prefactors, such that

M I σAuEK,XNAuλAuAu γ X . (3.32) = M IAuσXEK,AuNXλXM

Then, equation (3.31) can be rearranged to

µ t ¶ µ t ¶ γexp − 1 exp − λAuM cos(θ) = − λXM cos(θ) µ t ¶ µ t ¶ 0 1 exp − γ exp − , (3.33) = − λXM cos(θ) − · λAuM cos(θ) which has to be iteratively solved for t. Subsequently, this estimation of the thickness t can be inserted again into equation (3.19) to quantify the molecular surface coverage Γ. With this latter method, each elemental species can be employed to calculate Γ, which is a great advantage to the previous approaches. In an ideal case, all extracted surface coverages yield the same value within error bars. However, when the elemental distribution in a molecular monolayer is asymmetric, as apparent in dithiocarbamate SAMs, systematic errors may occur in the latter analysis. In DTC and thiolate SAMs, the photoelectrons excited from sulfur will systematically move through a longer molecular layer than e.g. photoelectrons from carbon or nitrogen. Still, these systematic deviations are often minor within the error bars since the monolayer quality may differ from sample to sample due to different amounts of surface defects and/or impurities of any kind.

48 Section 3.6: Experimental Setup

3.6 Experimental Setup

All photoelectron measurements present in this work have been conducted at the ORPHEUS cluster system at the I. Institute of Physics (IA). ORPHEUS is a custom-build ultra-high vacuum (UHV) setup for the preparation and characterization of complex metal-organic thin films and structures. The main analysis chamber is based on a Specs PHOIBOS 100 system, which is schematically de- picted in figure 3.8. The working principle is based on collection and separation of photoelectrons employing a hemispherical energy analyzer (HSA) [96]. After photoexcitation, electrons can exit from the surface as described by the processes in section 3.2. At a working distance of 40 mm, an acceptance angle of usually less than 10 ◦ is allowed for these electrons to enter the HSA. If other electrons than those ejected from the surface normal, as depicted in figure 3.8, are of interest, then the sample can be tilted, thus enabling the characterization of band structures. Once entered into the HSA, these photoelectrons are focused by a system of electronic lenses, which fulfill several functions. First, the lenses are focused on the surface of the sample to directly collect excited photoelectrons. Subsequently, these electrons are focused to the entrance slit of the hemisphere.

The lens systems retards the kinetic energy EK of the electrons by an energy value Eret:

E E E , (3.34) pass = K − ret with the pass energy Epass. Hence, the lens systems enables the energy separation and resolution in combination with the hemisphere of the HSA: Only photoelectrons with a kinetic energy equal to

Epass are able to pass the hemisphere without deflection. The central trajectory in between the inner and outer sphere of the HSA is defined by an electric field E~ that deflects the electrons according to the Lorentz force F~ q E~. Typically, the setup is employed in the fixed analyzer transmission = · (FAT) mode, in which the pass energy is kept constant. Hence, the magnitude of retardation Eret is continuously varied during a measurement in order to collect all electrons within the available energy range hν Φ E 0. − ≤ K ≤ Having passed the exit slit of the HSA, the photoelectrons are detected using a PHOIBOS CCD3 detector [99]. This kind of 2d detector is based on two micro-channel plates (MCP) that serve as electron multiplier and a phosphor screen. Upon impingement of electrons, this screen locally generates light, which is then recorded by a highly sensitive CCD camera. Electrons are thus detected and analyzed by the intensity and location of bright pixels on the screen.

For photoexcitation, there are two sources attached to the analysis chamber: A high-energy X-ray source based on Al Kα with a photon energy of hν 1486.74eV is employed to excite strongly = bound electrons, which are located near the atomic nuclei. These electrons transport information such as elemental species incorporated in the thin film as well as chemical environment [100, 101].

3CCD = Charge Couple Device

49 Chapter 3: Photoemission Spectroscopy

Valence band spectra are taken using the high intensity UV irradiation of a helium discharge lamp [102]. This measurement is based on the characteristic He I spectral line at 21.22 eV.

II

I

I I

I

I

hv

Figure 3.8: Scheme of a state-of-the-art hemispherical energy analyzer (HSA) of a photoemis- sion spectrometer. After excitation by monochromatic electromagnetic radiation, photoelectrons escape from the surface at various angles. Only electrons inside a small acceptance angle are collected and retarded by an electrostatic lens systems of the analyzer. Thereby, only electrons whose kinetic energy matches the pass energy at the entrance slit of the hemisphere pass the HSA without deflection. After passing the exit slit of the hemisphere, these electrons are photomulti- plied by micro-channel plates (MCP) and detected by a combination of a phosphor screen and a sensitive CCD camera. Scheme after [96, 99].

50 Chapter 4

Density Functional Theory

“ There is an oral tradition that, shortly after Schrödinger’s equation for the electronic wave-function Ψ had been put forward and spectacularly validated for simple small

systems like He and H2, P.M. Dirac declared that chemistry had come to an end – its contents was entirely contained in that powerful equation. Too bad, he said to have added, that in almost all cases, this equation was far too complex to allow solution. ” From Walter Kohn’s Nobel lecture [103]

At the end of the 20th century, computational chemistry has finally established in the field of theoretical chemistry and physics. For their leading role in the development of density functional theory (DFT), Walter Kohn and John Pople have been awarded with the Nobel prize in Chemistry 1998. As expressed in the quote of Walter Kohn’s Nobel lecture, it was Schrödinger’s ground- breaking work in 1926 that has dramatically changed the view on atomic physics and modern quantum mechanics [104]. Yet, Schrödinger’s powerful equation can not be solved analytically for a many-body system with interacting particles. While many approximations have been developed to solve or rather avoid the problems accompanying many-particle systems, density functional theory has become one of the most versatile methods to generate fundamental insights into solid state physics on the basis of Schrödinger’s work [103]. The success of density functional theory has certainly been facilitated by the emerging power of modern computational techniques. Still, the original idea to take advantage of a density functional to describe quantum systems has been proposed shortly after Schrödinger’s equation by Thomas [105] and Fermi [106] in 1927. The success of their methods, however, remained rather insignificant due to extensive approximations that resulted in insufficient accuracy. This finally changed when Hohenberg and Kohn developed a formalism that allowed density functional theory to remain an exact theory of many correlated particles [107]. In the following, the fundamental theorems and equations of Kohn’s work will

51 Chapter 4: Density Functional Theory

be sketched. A comprehensive introduction to DFT is given by Richard Martin in his textbook Electronic Structure [108], which also served as guideline for the following discussion.

4.1 Basic Description of an Interacting Many-Body System with Electrons and Nuclei

The fundamental equation governing a system with interacting electrons and nuclei in the non- relativistic limit is the time-dependent Schrödinger equation

dΨ¡{~r },{R~ },t¢ ˆ ¡ ¢ i I H Ψ {~ri},{R~I},t i~ , (4.1) = dt in which the spatial distribution of all particles involved in the system is stored in the quantum- ¡ ¢ mechanical wave-function Ψ {~ri},{R~I},t . Here, electrons are denoted by a lower case annotation

~ri and ions by an upper case R~I. The key parameter governing Schrödinger’s equation is the Hamil- tonian Hˆ that acts on the wave-function Ψ. Within this Hamiltonian, all energetic contributions1 are stored in form of operators of kinetic and potential energies of the electrons with mass me and nuclei with atomic number ZI of mass MI, such that

2 Z e2 1 e2 2 1 Z Z e2 ˆ ~ X 2 X I X X ~ 2 X I J H i ¯ ¯ ¯ ¯ I ¯ ¯ . (4.2) = −2me ∇ − ¯~r R~ ¯ +2 ¯~ri ~rj¯ − 2MI ∇ +2 ¯R~ R~ ¯ i i,I i I i j − I I J I J | {z } | {z− } | 6= {z } | {z } | 6= {z − } ˆ ˆ ˆ E Te Vext Vˆint TI II

Here, Tˆe and TˆI denote the kinetic energy operators for electrons and ions, respectively. The external potential Vˆext represents the Coulomb interaction between electrons and nuclei, while

Vˆint incorporates all electron-electron interactions. Finally, the last term EII contains the classical interactions of the ion cores with each other and potentially any other static energy contribution that is not relevant to describe the electrons.

As mentioned by Dirac, the multiple electron-electron and electron-nuclei interactions in equa- tion (4.2) generate a mathematical problem that is analytically not solvable. Thus, simplifications are required: First, the kinetic energy of the nuclei TˆI is a rather small contribution in the ground state due to the inverse mass of the heavy ions. Compared to the mass of the electrons me, the ionic mass MI can be approximated to infinity, thus leading to a vanishing kinetic energy term

TˆI. In parallel, this argument directly leads to the Born-Oppenheimer approximation [110], also known as Adiabatic approximation of quantum mechanics: Due to the large difference in mass,

1For the discussion here, relativistic effects, magnetic fields and quantum electrodynamics are not included, but also not mandatory for the systems under investigation in this thesis. For more information on such systems, the inter- ested reader is referred to the textbooks Electronic Structure [108] and Computational Chemistry [109].

52 Section 4.1: Basic Description of an Interacting Many-Body System with Electrons and Nuclei electronic processes can be separated from those of the nuclei. Therefore, it is possible to split the total wave-function Ψ into two parts, i.e. a product of two wave-functions via

Ψ¡{~r },{R~ }¢ Ψ ¡{~r },{R~ }¢ Ψ ¡{R~ }¢, (4.3) i I = e i I · I I where Ψe denotes the wave-function for the electrons, while ΨI represents the ions. Additionally, this attempt leads to a separation of the Schrödinger equation, i.e. one for the electrons and one for the nuclei.

The second simplification of Schrödinger’s equation (4.1) can be achieved under the condition of orthonormal eigenstates ( Ψ Ψ 1). Using the Rayleigh-Ritz principle, it can be deduced 〈 | 〉 = that the possible wave-functions of the ground state remain stationary solutions [108]. Thus, the Schrödinger equation for electrons and nuclei can be written in the time-independent form:

Hˆ Ψ¡{~r },{R~ }¢ E Ψ¡{~r },{R~ }¢, (4.4) i I = i I with the simplified expression of the Hamiltonian

Hˆ Tˆ Vˆ Vˆ E . (4.5) = e + ext + int + II

In the scope of density functional theory, the density of electrons n is a fundamental value and will become important in the following sections. In general, one can write the density of particles P as the expectation value of the density operator nˆ (~r ) i 1..N δ(~r ~ri). Under the constraint of = = − orthonormal eigenstates Ψ for possible spins σ1 corresponding to position ~r1, n (~r ) then yields

R 3 3 P 2 d r2 d rN σ Ψ(~r ,~r2,~r3, ,~rN) n (~r ) Ψ nˆ (~r ) Ψ N ··· 1 | ··· | . (4.6) = 〈 | | 〉 = R d3r d3r d3r Ψ(~r ,~r ,~r , ,~r ) 2 1 2 ··· N| 1 2 3 ··· N | Analogously, the total energy of the system can be expressed as the expectation value of the Hamiltonian of equation (4.5). In addition, the energy contribution due to the external potential of the nuclei acting on the electrons can now be written in terms of the electron density, leading to Z ­ ¯ ¯ ® ­ ® ­ ® ­ ® 3 E Ψ ¯ Hˆ ¯ Ψ Hˆ Tˆ Vˆ d rV (~r )n (~r ) E . (4.7) = ≡ = e + int + ext + II

The expression of the total energy for a system of electrons in an external potential of nuclei, as formulated in equation (4.7), is still an exact representation without major simplifications. In par- ticular, the complex correlation effects of electrons, both kinetic and potential, are still explicitly ­ ® ­ ® incorporated in Tˆe and Vˆint . The following section will discuss the way how Walter Kohn and coworkers further reformulated the total energy in terms of the electron density to achieve a formalism that is capable of solving Schrödinger’s equation (4.4).

53 Chapter 4: Density Functional Theory

4.2 The Hohenberg-Kohn Theorems

The first important step to solve the many-particle Schrödinger equation (4.4) has been the work of Hohenberg and Kohn [107]. They proved, that the total energy according to equation (4.7) can be reformulated as an energy functional depending solely on the electron density n (~r ) without losing any information of the correlated, self-interacting system. The well-known Hohenberg- Kohn Theorems can be expressed as follows (adapted from section 6.2 of Electronic Structure, [108]):

• Theorem I: For any system of interacting particles in an external potential Vext (~r ), the potential is determined uniquely, except for a constant, by the ground state particle density

n0 (~r ). Corollary I: Since the Hamiltonian is thus fully determined, except for a constant shift of the energy, it follows that the many-body wave functions for all states (ground and excited) are determined. Therefore all properties of the system are completely determined given only

the ground state density n0 (~r ).

• Theorem II: A universal functional for the energy E[n] in terms of the density n (~r ) can be

defined, valid for any external potential Vext (~r ). For any particular Vext (~r ), the exact ground state energy of the system is the global minimum value of this functional, and the density

n (~r ) that minimizes the functional is the exact ground state density n0 (~r ). Corollary II: The functional E[n] alone is sufficient to determine the exact ground state energy and density. (...)

It is important to point out the fact, that only the ground state density n0 (~r ) determines the exact ground state energy of the system. Hence, the use of variational techniques allows to find the global energy minimum of the system and, in principle, to determine all properties of the system. Yet, as corollary II states, the energy functional E[n] is only capable of determining the exact ground state energy and density, but a direct conclusion on excited states cannot be made. This becomes particularly important for systems, in which the electron density significantly differs from that of the ground state. In fact, this is the case for ground and excited states of organic molecules, which is basically described by the Franck-Condon principle [111]. Basic DFT is generally known as a ground state theory.

Following the Hohenberg-Kohn approach, the total energy from equation (4.7) can be rewritten in the form of energy functionals depending on the electron density: Z E [n] T [n] E [n] d3rV (~r )n (~r ) E HK = e + int + ext + II Z F [n] d3rV (~r )n (~r ) E . (4.8) ≡ HK + ext + II

54 Section 4.3: The Kohn-Sham Ansatz

The difference to equation (4.7) is now, that also all internal energies can be directly expressed as density functionals. Thus, equation (4.8) defines the Hohenberg-Kohn functional FHK [n], that incorporates all kinetic and potential energy contributions of the self-interacting electron density,

F [n] T [n] E [n] . (4.9) HK = e + int It is an interesting point, that one can describe a system with all its properties exclusively based on the ground state of the particle density. However, there is actually no gain for the objective to find a solution for the interacting many-particle Schrödinger equation. To master this crucial step, the idea of Kohn and Sham is required and will be introduced in the following.

4.3 The Kohn-Sham Ansatz

In 1965, Kohn and Sham proposed an attempt to find a workaround of solving the complex many- particle Schrödinger equation (4.4)[112]. This is basically the foundation of today’s successful application of density functional theory in the field of ground state computational chemistry. The ansatz postulated by Kohn and Sham is both simple and spectacular: The exact ground state density of the interacting particles can be represented by the ground state density of an aux- iliary system consisting of non-interacting particles. According to this attempt, the complex Schrödinger equation of a single wave function, which contains all electronic properties of N particles, can be split into N separate equations of independent particle wave functions, i.e. via ¡© ª¢ Ψ ~r ψi 1..N (~r ): ⇒ = (H ² )ψ (~r ) 0 , (4.10) KS − i i = with the eigenvalue ²i for each electron i.

To retain the physical information of the interacting particles, the complex contributions have to be incorporated with additional terms. This is the genius of the Kohn-Sham approach: While the auxiliary system consists of independent particles, an interacting density is assumed. Thus, all complex interactions of the many-particle system are put into a so-called exchange-correlation functional of the density. Consequently, the Hamiltonian of equation (4.5) has to be rewritten for the auxiliary Kohn-Sham system:

H (~r ) T V (~r ) . (4.11) KS = s + KS

Here, Tˆs represents the kinetic energy of the single, independent particles. Hence, it is essential to define the new Kohn-Sham potential VKS (~r ) of equation (4.11) in order to accurately describe the system. Therefore, the internal energy functional of the self-interacting electron from equa- tion (4.9) is partitioned into two contributions. First, the self-interaction energy due to Coulomb

55 Chapter 4: Density Functional Theory

forces of a classical charge density is introduced in the form of the Hartree energy as

Z 1 3 3 n(~r )n(~r 0) EHartree [n] d r d r 0 . (4.12) = 2 ~r ~r | − 0| Second, the missing quantum-mechanical interactions, such as the Pauli exclusion principle [39], are described with the exchange-correlation functional Exc [n]. There are different ways to imple- ment those effects and depending on the system under investigation, one has to determine which approximation is most suited to describe the system. This will be briefly discussed in section 4.4.

Employing the Hartree energy functional of equation (4.12) and a exchange-correlation functional

Exc [n] to be defined later, one can rewrite the Hohenberg-Kohn expression of the exact ground state energy functional from equation (4.8) in the form Z E [n] T [n] d3rV (~r )n (~r ) E [n] E E [n] . (4.13) KS = s + ext + Hartree + II + xc

Using this Kohn-Sham energy functional, it is now possible to express the Kohn-Sham potential

VKS (~r ) from equation (4.11) as

∂EHartree ∂Exc VKS (~r ) Vext (~r ) = + ∂n (~r ) + ∂n (~r ) V (~r ) V (~r ) V (~r ) . (4.14) = ext + Hartree + xc

Under the assumption of an approximate estimation for the exchange-correlation functional Exc [n], the expression of equation (4.14) can be used within the Schrödinger-like Kohn-Sham equa- tion (4.10) with the formulation of the Kohn-Sham Hamiltonian of equation (4.11). It is then possible to iteratively solve these equations after an initial guess has been made, as illustrated in

figure 4.1. A good first guess for the independent-particle density n(~r ) and wave functions ψi(~r ) are usually based on a linear combination of atomic orbitals [113]. Self-consistency is then vali- dated by minimal differences in total energy and forces on the nuclei. The latter can be identified employing the Hellman-Feynman Theorem, also known as Force Theorem, which correlates the force on a nucleus explicitly with the charge density [114].

4.4 Exchange and Correlation

In the end, the whole approach of Kohn and Sham crucially depends on an accurate modeling of the exchange-correlation functional Exc[n]. Comparing the exact formalism of the total energy functional from Hohenberg-Kohn (equation (4.8)) with that of the auxiliary system proposed by Kohn-Sham in equation (4.13) helps to gain an appealing insight into the nature of the effects of exchange and correlation. Since the particle density n of the two different systems is equal by

56 Section 4.4: Exchange and Correlation

initial final charge guess n density |² i E Ψ [ | n Σ ] =

n

ψKS VKS

(H 0 K = S - εi) Ψi

Figure 4.1: Schematic illustration to reach self-consistency in a DFT calculation. After an ini- tial guess has been made for the independent particle density n(~r ) and wave function Ψi(~r ), the Schrödinger-like Kohn-Sham equation can be iteratively solved.

definition of the Kohn-Sham ansatz, the exchange-correlation functional Exc[n] can be rewritten as

E [n] F [n] (T [n] E [n]) xc = HK − s + Hartree T [n] T [n] E [n] E [n] . (4.15) = e − s + int − Hartree

Thus, Exc[n] is explicitly the difference of the kinetic and internal interaction energies of the exact system with interacting particles from those of the auxiliary system with electron-electron interac- tions described by the classical Hartree energy. In this form, the genius of this approach becomes apparent: Via separation of independent-particle kinetic energy Ts and long-range self-interaction energy EHartree from exchange and correlation effects, Exc[n] exhibits mainly short-range interac- tions. It can therefore be interpreted as the modeling of an exchange-correlation hole surrounding each electron in order to effectively describe the consequences of the Pauli exclusion principle [39] and other quantum-mechanical interactions. The correlation term in particular incorporates effects for electrons with opposite spins, i.e. interaction effects that are not affected by Pauli exclu- sion. Therefore, understanding and modeling this exchange-correlation hole is the key objective in developing high quality approximations for Exc[n].

The most commonly employed models are the local density approximation (LDA) and the gener- alized gradient approximation (GGA). For the former, Exc[n] is approximated to be a functional depending only on the absolute electron density, which is assumed to be only slowly varying. This leads to a description of exchange and correlation energy of a uniform electron gas of density n,

57 Chapter 4: Density Functional Theory

as discussed by Kohn and Sham [112], and can be expressed as Z E LDA[n] d3r n (~r )² (n) . (4.16) xc = xc

Here, ²xc is the energy per electron that depends on the density n. Apparently, this model is best suited to describe systems, in which the homogeneous electron gas is a valid approximation, especially for metals. The idea to not only model exchange and correlation via the local electron density, but additionally its gradient n was also first described by Kohn and Sham [112] and then ∇ carried out in detail by Herman [115], Z E GGA[n] d3r f (n (~r ), n (~r )) . (4.17) xc = ∇

In comparison with the LDA attempt, generalized gradient approximations improve total and atomic energies [116, 117], especially for systems, which cannot be described as a homogeneous electron gas. Among all GGA’s, the model of Perdew et al. is one of the most frequently employed and has been of use in this thesis as well [118]. In particular, GGA’s are known to accurately describe not only molecules and metals in single systems, but also molecule/metal interfaces as investigated in this thesis [27, 51–53, 117].

4.5 Van-der-Waals Forces in DFT

Any realistic solution of Schrödinger’s equation using Density Functional Theory is based on an accurate description of exchange and correlation effects, as discussed in the previous section. In standard DFT calculations, the commonly employed exchange-correlation functionals (LDA and GGA) are local or, in the case of GGA’s, semi-local approximations. This may lead to major dif- ferences to the experimentally observed systems if non-local correlation effects are important, e.g. van-der-Waals forces in organic crystals. Yet, an accurate modeling of van-der-Waals forces based on first-principles is challenging, since non-local correlations generate many additional computa- tional steps to implement the interactions of atoms and electrons between adjacent molecules.

In this study, van-der-Waals interactions affect the molecular arrangement in a monolayer, in par- ticular for those with longer alkyl chains. Van-der-Waals forces can be interpreted as the attractive interaction between two spontaneous dipole moments or the interaction between one spontaneous and one induced dipole. The physical origin of van-der-Waals forces, however, is based on the same basics as the formation of chemical bonds: The solution of the time-independent Schrödinger equation (4.4), as discussed in references [119, 120]. Thus, these attractive forces of the charge density occur at any position in each solid, but they are much weaker than ionic or covalent bonds. Yet, for separate alkyl chains, which are oriented parallel to each other in an oriented monolayer,

58 Section 4.5: Van-der-Waals Forces in DFT

van-der-Waals forces dominate the orientation and packing of these molecular backbones. To achieve an accurate description of such systems, an additional energy correlation term has to be incorporated that accounts for these long-range interactions. In general, only dipole interactions are considered in these systems and all higher order multipole interactions are neglected due to the fact that their influence is much less compared to the dipole component.

QUANTUM ESPRESSO allows to implement an approximation of van-der-Waals forces using the semi-empirical description of F. London [121]. This energy term is based on quantum-mechanical calculations using perturbation theory and is derived as

i,j X C6 E , (4.18) disp = − 6 i,j Rij

¯ ¯ with the distance Rij ¯R~i R~j¯ between the two atom i and j. In the above equation, the parameter i,j = − C6 is a material constant based on the atoms involved and correlates to the polarizability. It can be directly observed from equation (4.18) that an attractive interaction is effectively confined to short distances as a result of the sixth power of the denominator. The dispersion energy based on the

London-type of interaction only slightly increases the computational cost as Edisp does not depend on the electron density n. Thus, the whole contribution of Edisp can be derived only as a result of the positions of the nuclei at the beginning of each self-consistent step.

Within the framework of QUANTUM ESPRESSO, these London-type interactions are slightly modified to incorporate them in the framework of DFT. This can then be expressed as [122]

ij XX C6 ¡ ¢ ¡ ¢ 1 Edisp s6 fdamp Rij with fdamp Rij ³ h i´ , (4.19) = − 6 · = Rij i j i Rij 1 exp d 1 > + − · Rr − with a parameter Rr, which is given as the sum of the van-der-Waals radii of the two atoms i and j, and a scaling constants d 20 and s6 based on semi-empiric considerations. The general form of = ¡ ¢ the expression from London is still present, but an additional damping term fdamp Rij has been added. For large distances R , this term vanishes with f ¡R ¢ 1 and the full dispersion ij damp ij → ∞ → energy E from equation (4.19) is dominated by the denominator to zero, E ¡R ¢ 0. disp disp ij → ∞ → Hence, the damping factor fdamp becomes important at shorter distances and effectively confines the attractive interaction within the distance of the van-der-Waals radii. For shorter distances, the original London-type interaction of equation (4.18) would strongly increase, which is counteracted by the atoms if closed shells of the separate atoms start to overlap. Moreover, at shorter distances in the regime of chemical bonds, standard DFT methods accurately describe the distribution of electrons and nuclei without the consideration of van-der-Waals forces. It should be noted here, that London-type interactions have only been incorporated for the non-metallic atomic species to accurately describe the interaction between two molecules with interacting alkyl chains.

59 Chapter 4: Density Functional Theory

4.6 Repeated Slab Approach

In analogy to the publications of Georg Heimel and coworkers studying thiolate monolayers on gold [51, 53, 54, 123], the theoretical characterization in this thesis is mainly focused on molecule/metal interfaces of dithiocarbamates on noble metals. As mentioned in the previous section, a GGA functional has been chosen to accurately describe these complex systems. Fur- thermore, a plane wave basis set needs to be applied for the calculations if the interface energies of a metal are involved. Local models based on linear combinations of Gauss-like atomic functions in order to describe the wave functions Ψi (~r ) can be best applied to non-periodic calculations of single molecules [27, 30, 124, 125]. Yet, these models are not suited to accurately describe a system based on a molecule attached to a metal surface, since the metal will only be poorly de- scribed. It is in particular impossible to describe a homogeneous electron gas, if the metal is only modeled by a small cluster. Hence, a periodic approach is needed for larger systems and can be ³ ´ best implemented employing plane waves in the form of Ψ (~r ) exp i~k ~r . i = i

In this work, the open-source package of QUANTUM ESPRESSO has been used to gain in- sights into the interface physics2 of molecular adsorbates on metal surfaces [113]. QUANTUM ESPRESSO is exclusively based on a plane-wave approach, thus enabling calculations of periodic systems as investigated here. Yet, employing plane waves to simulate a (modified) surface of a metal requires special boundary conditions, since the periodicity is broken in the direction of the surface normal. Hence, a workaround has to be chosen in order to accurately describe a surface while keeping the periodicity valid all in three dimensions [54]: A repeated slab approach, as il- lustrated in figure 4.2, consists of the topmost atomic layers of the metal (and a molecular layer attached to the surface) with accurate periodicity in the directions parallel to the surface. For the direction perpendicular to the surface, the value of the unit cell is chosen to be arbitrarily large, e.g. up to a length of 30 to 50 Å, depending on the size of the molecular layer. In the end, the resulting surface layers are (periodically) kept apart sufficiently far, so that these layers are not affected by each other. In order to compensate for long-range polar interactions along the surface normal, QUANTUM ESPRESSO allows to incorporate a dipole layer in form of a homogeneous electric field within the vacuum layer. Its field strength can be used to cancel the total dipole mo- ment of the unit cell [54, 113]. According to figure 4.2C, the final system contains of a metal slab of several layers, a molecular layer of reacted molecules and a dipole layer in the vacuum re- gion. The top four atomic layers of metal substrate are allowed to relax during optimization while the bottom two layers are kept fixed to represent the solid. Further information on computational details relevant for this work can also be found in the master thesis of Tobias Schäfer [126].

2In this context, the following aspects are particularly important: Geometrical orientation of molecules, electron density, dipole analysis and density of states.

60 Section 4.7: Bond Dipole Analysis

Figure 4.2: Illustration of a slab for modeling molecule-metal systems at the example of a C2DTC 14 layer on a Au(111) surface at a low coverage of 0.8 10 mol/cm2.(A) depicts a single chemisorbed C2DTC molecule. (B) illustrates the top view of the· surface slab, indicating the unit of repetition (black box). (C) is the side-view of the system, illustrating the long c-axis perpendicular to the surface and incorporating a dipole layer to avoid electrostatic interaction of the adjacent slabs. The unit cell is again marked by a black rectangular box.

4.7 Bond Dipole Analysis

After a system as depicted in figure 4.2 has been fully relaxed and optimized, further information can be extracted from nuclear positions and the electronic density of the system. The analysis of the key factors that govern the work function changes upon molecular modifications is one objec- tive in this thesis. In order to decompose these effects, the evolution of the classical electrostatic potential V can be analyzed exploiting the Poisson equation [51, 54],

n (~r ) 2V . (4.20) ∇ = − ²0

In a periodic slab approach, only differences of the electrostatic potential are relevant, since the bottom part of each slab (fixed metal atoms) is rather meaningless - it just represents the solid metal. Since the zero point energy in QUANTUM ESPRESSO is arbitrarily chosen, one can reset this zero point to match the Fermi energy EF. Then, the energetic difference between the Fermi level and the electrostatic potential in the vacuum region at sufficiently large distance from the surface directly yields the work function E Φ, as illustrated in figure 4.3. At this point, the vac =

61 Chapter 4: Density Functional Theory

Figure 4.3: Evolution of electrostatic po- tential (black curve) of a complete sur- face slab containing several metal lay- ers and a thiolate monolayer. ∆Φ de- notes the work function change due to mod modification, ΦAu is the absolute work function of the modified surface. HOPS (LUPS) represent the highest occupied (lowest unoccupied) π-state of the aro- matic molecule. Reprinted with permis- sion from [53]. Copyright 2007, American Chemical Society. importance of the dipole layer of figure 4.2 becomes apparent as it effectively decouples the elec- trostatic potential between adjacent slabs. The work function in the vacuum layer on top, i.e. above the molecular layer, represents a realistic value, which should equal the value determined exper- imentally. Therefore, it is essential to avoid interference effects with the electrostatic potential at the bottom side of the next metal slab.

The analysis can be put one step forward to decompose the interplay of dipole interactions ac- companying monolayer formation. After a full system containing a metal and a molecular layer has been successfully simulated, one can cut the total system into its subsystems. These subsys- tems can then be optimized separately in order to extract an expression of the bond dipole BD. The bond dipole is a direct result of charge rearrangements that accompany chemical bonding of molecules to surfaces. However, an estimation of its magnitude from the total system remains dif- ficult to determine. Therefore, Heimel and coworkers have established this approach of exploiting the subsystems [51, 53, 54, 123]. Figure (4.4) illustrates the result of such a decomposition: The total system (D) comprises of the metal layer (A) as substrate, a molecular layer (C) and the dipole layer BD due to bond formation (B). The electrostatic potential along the surface normal of the total system Vtot is consequently the addition of the subsystems,

V V V V . (4.21) tot = metal + BD + mol

While the vacuum shift caused by the bond dipole BD cannot be simulated directly, one can rearrange equation (4.21) to develop an expression for VBD. Since the charge density n is a direct product of a DFT simulation, equation (4.21) can be restated with Poisson’s equation to

n n n n . (4.22) tot = metal + BD + mol

2 The dipole shift due to bonding can be directly assessed using V nBD(~r )/² . One crucial ∇ BD = − 0 point is the simulation of the free molecular layer of figure 4.4(C), as discussed also in [126, 127]. It is mandatory for obvious reasons that the geometry and orientation of the molecular layer from

62 Section 4.7: Bond Dipole Analysis

Figure 4.4: Complete representation of the electrostatic potential upon SAM formation of a thi- olate SAM on a gold layer, including the schematic evolution of the subsystems (A-C) of the complete slab D. Figure (A) represents the bare metal layer with atomic position as optimized with the molecule. The free molecular layer exhibits an intrinsic vacuum level shift ∆Vvac that results in different electron affinities EA and ionization potentials IP on the left and right side of the molecule. In order to establish a chemical bond to the metal surface, charges rearrange from the free molecular scenario, which is represented by the bond dipole BD (B). Note that the mag- nitude of BD is increased for visual clarity. Summing up the subsystems ((A-C)) results in the complete system (D). Reprinted with permission from [53]. Copyright 2007, American Chemical Society. the full system has to be conserved in order to accurately describe the effects of this layer in a subsystem. The difficulty arises that, in analogy to thiolate SAMs [127], one hydrogen atom of the neutral, isolated molecule is missing in a chemisorbed monolayer. Upon reaction with the metal surface, the hydrogenated sulfur usually replaces the R-S-H bond to form a metal-sulfur bond in the form R-S-metal. In an experiment, the hydrogen atom may form a H2 -molecule with another split hydrogen. The difficult choice for simulating this free molecular layer is whether to hydrogenate these molecules again or keep them in a radical state. Both models may have their validity depending on the scientific question, but a careful consideration is required. The molecular layer in a radical form may either crucially differ from that in the total system on an electronic basis, when the positions of the nuclei remain fixed. In consequence, the molecular layer is kept apart from its energetic minimum. Otherwise, it may crucially differ on a structural level if the radical molecules are allowed to relax into their semi-stable radical ground states. Both scenarios may inherently alter the physical properties of the molecular layer, as also explicitly stated by Wang et al. [127]. In most cases, especially for thiolate monolayers as investigated in this work, the hydrogenated molecular layer is to be preferred. As a result, the molecular ground state is mainly conserved and the chemical integrity remains for comparison. Regarding equation (4.22), the charge density due to the hydrogen atom needs to be subtracted to match the total system, i.e. with nmol nH mol nH = − − ntot nmetal nBD nH mol nH . (4.23) = + + − −

63 Chapter 4: Density Functional Theory

4.8 Computational Details

In the scope of the master thesis of Tobias Schäfer and within a subsequent collaboration, DFT simulations have been performed to understand the interface between DTC monolayers and noble metals [126, 128, 129]. Table 4.1 lists the unit cell parameters that have been employed for 4 4 × cells. These cells consist of six metal layers with a (111) surface, which is represented by a 4 4 × construction in a and b direction. The c-axis is directed parallel to the surface normal and chosen to be sufficiently large in order to avoid interference effects. In addition, 2 2 and 2 3 cells × × containing each one molecule can simply be constructed using the corresponding fraction for the unit cell axes a and b. The corresponding cells are depicted in figure 4.5. £ ¤ £ ¤ £ ¤ a Å b Å c Å α[◦] β[◦] γ[◦] gold 11.80 11.80 40.0 90.0 90.0 60.0 silver 11.72 11.72 60.0 90.0 90.0 60.0 copper 10.184 10.184 60.0 90.0 90.0 60.0

Table 4.1: Unit cell parameters of 4 4 cells containing C2DTC monolayers on noble metals as illustrated in figure 4.2. The amount× of molecules can be varied, thus representing different surface densities. In addition, 2 2 and 2 3 cells containing one molecule have been simulated. × × These cells only vary in the a and b axis from the values listed above.

For all calculations, ultra-soft pseudo-potentials as developed by Vanderbilt [130] have been em- ployed in combination with the PBE functional [118]. Semi-empirical London-like van an der Waals interactions were applied as implemented by Grimme in order to improve the alignment of alkyl-chains of the N,N-dialkyl-dithiocarbamates [121, 122]. For the 4 4 cells of the Au(111) × surface, a k-point integration was carried out using a 6 6 1 Monkhorst-Pack grid, while a 8 8 1 × × × × grid has been employed for the smaller cells [131]. Finally, the Methfessel-Paxton smearing [132] has been used with a smearing parameter of 0.2 eV and kinetic energy cutoffs for wave functions and densities have been set to 30 Ry 300 Ry, respectively. To validate the DFT results, various starting positions of the molecules with respect to the top metal layer have been tested, including different combinations of top, bridge and hollow sites for the DTC sulfur atoms (see also figure A.2 for the illustration of the most realistic binding positions of DTCs on Au, Ag, Cu).

64 Section 4.8: Computational Details

Figure 4.5: Initial surface slab geometries for DFT calculations as performed by T. Schäfer [126].

65

Part II

Dithiocarbamate Modifications of Noble Metal Surfaces

Chapter 5

Dithiocarbamates for Low Work Function Noble Metals

In the past ten years, the class of dithiocarbamates (DTC) have appeared as efficient surface mod- ifiers for gold nanoparticles [31, 133]. Prior to the application in this field of research, dithiocar- bamates have a long history in the organo-sulfur chemistry, biological research, agriculture and medicine [134, 135]. The versatility of dithiocarbamates in the field of molecular electronic junc- tions has just recently been reported [92]. Our research group has aimed to understand, how dithio- carbamates bind on a gold surface as a first step towards a new generation of surface modifications [27, 126, 136]. In this chapter, a further step forward towards a more fundamental investigation is taken, in which the following questions will be addressed: To what extent can dithiocarbamates be employed to achieve low work function surfaces of noble metals that are more important in indus- trial applications, i.e. copper and silver? How can the chemical binding as well as the electronic

69 Chapter 5: Dithiocarbamates for Low Work Function Noble Metals

interface be characterized? In the end, the comparison of the binding of dithiocarbamates to differ- ent metallic surfaces eventually reveals new insights into the understanding of the metal/molecule interface.

5.1 N,N-dialkyl Dithiocarbamates on Gold

The assembly of N,N-dialkyl dithiocarbamates (abbreviated with CnDTC) on gold as has been investigated by P. Schulz and coworkers in our group and will be briefly recapped. These results have also been published in Advanced Materials Interfaces in 2014 [27].

Dithiocarbamates have been reported to exhibit superior electronic properties in molecular junc- tions compared to thiolate monolayers with similar molecular backbones [92]. This improved charge transport in metal/molecule/metal devices originates from a shift of the frontier molecu- lar states towards the Fermi level EF. This shift reduces the potential barrier for charge injection and extraction. In fact, the carbodithioate linker group (-R-CS2) has been reported to generate delocalized metal-molecule interface states (MMIS) that are located close to the Fermi level EF [137]. The same observation can be made for the calculated density of states (DOS) of dithio- carbamate monolayers in figure 5.1. By comparison of the electronic structure between single molecular levels and the DOS of the full metal/molecule system, significant contributions to the valence region can be identified: Figure 5.1A shows the projected DOS of the C nDTC SAMs in the top-top configuration on gold. The gray areas refer to projections of the PDOS on different structural units of the C2DTC SAM. The contributions to the electronic structure at higher binding energies E 4eV can be correlated to the molecular backbone consisting of the dialkyl amine, B ≥ while the electronic states near the Fermi level at 4eV E E can be traced back to the fron- ≤ B ≤ F tier molecular orbitals located at the CS2-group. Hence, when raising the alkyl chain length of the molecular backbone, these frontier electronic states remain almost identical, while the higher energy contributions raise with the increasing amount of σ-type bonds. In combination with the molecular states from single molecule calculations (figure 5.1B), one can identify the onset of the frontier MMIS at 1 eV in figure 5.1A to originate from the 3 p sulfur lone pair that forms a bond to the underlying gold atoms. In addition, C2DTC reveals two π-type states located within the CS2 linker group to correspond to HOMO-1 and HOMO-2, as shown with the isosurfaces of the single molecule on the left side of figure 5.1.

Yet, when an electrode is brought into contact with an organic semiconductor as present in OLED and OSC devices, the electronic structure of the metal/SAM interface is only one important factor for efficient charge transfer. As elucidated in section 2.2, the energy level alignment of an elec- trode/semiconductor interface is also crucially dependent on the work function of the electrode material: Efficient electron transfer can only take place when the Fermi level EF is located close

70 Section 5.1: N,N-dialkyl Dithiocarbamates on Gold

Figure 5.1: Electronic structure of DTC monolayers on gold (A) and single molecules (B) from DFT calculations. (A) Plots of the total PDOS of CnDTC monolayers on gold are drawn with the full PDOS as solid lines. The gray areas are projections of the total PDOS on different structural units of the C2DTC SAM, illustrating that the onset of molecular states at 1 eV are dominated by the CS2 linker group. Increasing chain length in the molecular backbone raises the amount of σ-states at higher binding energies, but leaves the valence band region EB 4eV unaffected. (B) Binding energies of the molecular states from the isolated, hydrogen-terminated≤ molecules reveal the MMIS to originate from the linker group with the sulfur 3 p lone pair to be the HOMO level (see isosurface on the left). HOMO-1 and HOMO-2 are π-type orbitals and indicate the delocalized nature of the electron in the linker group. Reprinted with permission from [27].

to the LUMO of the organic semiconductor. This sets the need for stable low work function elec- trodes. Elements with low work function like lithium, potassium or sodium are generally highly reactive materials. Hence, these materials have to thoroughly encapsulated to avoid a reaction with ambient molecules, which usually cause electrode degradation. Instead, an air-stable passi- vation layer for a noble metal surface, which can shift the work function to lower values is desired. Such passivations have been targeted with small molecules as well as polymers [138, 139]. Yet, molecules that form stable SAMs with low work functions have not been until very recently: A work function Φ as low as 3.2 eV can be achieved when C2DTC forms a close-packed monolayer with a density Θ 2.5 1014 molecules/cm2 on gold [27, 93]. This corresponds to an effective = × work function reduction by 2 eV (Φ 5.2eV). For longer alkyl chains lengths in the molecu- Au111 ≈ lar backbone (CnDTC with n {4,6,8,10}), the work functions are found to lie at Φ 3.5 0.1eV ∈ = ± irrespective of the number of methyl units [27].

71 Chapter 5: Dithiocarbamates for Low Work Function Noble Metals

V V V V V V V

V V V

VV VV V V V V V V V V V V V

V V V VV VV V V V V

Figure 5.2: Decomposition of a C2DTC SAM on Au(111) (A-D) that yields a work function of 3.2 eV in comparison with an ethanethiol SAM (E-H) at comparable surface coverage. Plots of the electrostatic potential (integrated over a plane parallel to the gold surface) serve to identify dipole contributions of each unit of the cell, which generate the work function change. (A) and (E) present the clean, unmodified metal with a work function of 5.2 eV. (B) The large intrinsic dipole moment of the free standing, hydrogenated C2DTC layer yields a potential shift of 1.3 eV. (F) For the hydrogenated ethanehiol layer, a small intrinsic dipole result only in a potential shift of 0.3 eV. (C) and (G) In addition, the charge rearrangement upon adsorption of both the C2DTC and ethanethiol molecules generate a potential shift of 0.7eV (at equal surface density). (D) and (H) present the total systems. Calculations performed by Tobias Schäfer [126].

Therefore, the question arises how these very low work functions can be explained. The work function of the full molecule/metal system can be reasonably reproduced with DFT simulations for an adequate molecular density [27, 126]. Using DFT, the full system can be partitioned into its building blocks, i.e. the bare metal surface and the molecular layer (section 4.7). Fig- ure 5.2 illustrates the results of such an approach for two systems with a surface coverage of 14 2 2.5 10 molecules/cm :AC2DTC SAM in comparison with a ethanethiol SAM. The full × C2DTC system in (D) reduces the work function by ∆Φ 2.0eV down to an absolute value of = Φ 3.2eV. With an initial work function Φ 5.2eV (A and E), the reduction of the free mod = Au = standing, hydrogenated molecular layer (FSM) accounts to ∆Φ 1.3eV (B). This large shift FSM = reflects the strong dipole moment of the C2DTC molecule according to equation (4.20). The miss- ing 0.7 eV are thus generated by the interface dipole due to charge rearrangements induced via

72 Section 5.2: Chemical Binding of N,N-dialkyl Dithiocarbamates on Noble Metals

adsorption and bond formation (C). This value is in good comparison with the interface dipole of the thiolate SAM (G). However, the small dipole moment of the thiolate backbone only generates an additional shift of 0.3 eV (F). Deviations from literature values for the thiolate SAM most prob- ably originate from a different surface coverage, since thiols are able to pack more densely than dithiocarbamates [53, 54, 90]. It is noteworthy at this point to highlight the fact, that the charge rearrangements are strongly localized at the metal/molecule interface. They do not extend to the polar backbone of the molecules as can be identified from the development of the electrostatic po- tential in figure 5.2C. Thus, the polar unit does not get compensated by the charge rearrangements, but remains conserved after monolayer formation.

In conclusion, a high intrinsic dipole moment in combination with a moderate interface dipole in a DTC monolayer add up to the full work function shift up to ∆Φ 2.0eV. The question ≈ arises, if such a large shift can also be observed for different noble metals. Therefore, a detailed discussion of the assembly of N,N-dialkyl dithiocarbamates on the coinage metals gold, silver and copper follows. The focus is expressed by the following question: What are the differences in the chemical and electronic structure when replacing the gold surface with another noble metal?

5.2 Chemical Binding of N,N-dialkyl Dithiocarbamates on Noble Metals

As discussed in the previous section and reference [27], the different number of methyl groups in CnDTC molecules results in rather small variations of the work function when grown on a Au surface. This also accounts for the binding scenario observed with XPS [136]: Peak shifts in the order of 0.4 eV are only observed for the carbon C 1s level, which has been attributed to a higher coordination symmetry1as well as molecular stabilization due to interchain interactions of the diamine units with longer alkyl chains. In order to examine and highlight significant differences between the adsorption of N,N-dialkyl dithiocarbamates on the noble metals Au, Ag and Cu, the molecule C2DTC will serve as a model SAM in the following discussion. Changes from C2DTC to the higher order substitution with n 2 are discussed in section 5.2.4. >

1For small alkyl chain length, the C 1s peak equally comprises contributions of C-H/C-C and C-N bonds. These two configurations are located at slightly different binding energies, as will be discussed with the help of figure 5.3. An increasing number of methyl units generates significantly more C-H/C-C coordinations, while those of C-N remain constant. Consequently, the C 1s peak shape becomes more dominated by the alkyl chains with increasing methyl units.

73 Chapter 5: Dithiocarbamates for Low Work Function Noble Metals

5.2.1 XPS of C2DTC bound to Au(111)

Starting with the binding of C2DTC on Au(111), an overview of the core level spectra is displayed in figure 5.3. The Au(111) layer has been grown on a mica substrate and yields a high crystallinity, texture and a flat surface [126, 140]. This scenario represents almost ideal conditions as given by a single crystal surface in (111) direction and therefore improves the comparison between experiments and DFT simulations. A survey scan of C2DTC on gold is plotted in figure 5.3A and reveals the prominent elemental species on the surface. As expected, the survey scan is dominated by the gold substrate, but also indicates the presence of the organic monolayer. However, elements with low atomic concentration like nitrogen are hardly noticeable and high resolution scans have to be performed (B-D). Nevertheless, a significant incorporation of oxygen or the formation of an oxide layer can be directly excluded with the survey scan due to the absence of an oxygen peak at E 532eV. A high resolution scan of the O 1s region also revealed only a minor concentration B ≈ of oxygen.

Nitrogen is found with a single peak at E 399.9eV. This binding energy fits well to a configura- B = tion of nitrogen in a sp2-hybridized state with carbon (E 400eV) rather than a sp3-hybridization B = (E 398.3eV)[141, 142]. This finding leads to the conclusion that the delocalized electrons of B = the molecular linker group are not only located at the direct metal/molecule interface but also up to the central nitrogen atom. The weak signal-to-noise ratio for the N 1s signal originates from the rather small cross-section for photoionization [87] in combination with a low atomic concentration within the SAM (one nitrogen per molecule).

For carbon, the C 1s level is measured and a slightly asymmetric peak is observed (figure 5.3C). The main peak at 286 eV is the sum of bonds to hydrogen, carbon, nitrogen and sulfur, while the smaller peak at E 288.3eV can be assigned to a minimal amount of carbon species bound to B = oxygen. This may be caused e.g. by contaminants within the DTC layer or physisorbed residues on top of the molecular layer. Carbon located in a relatively neutral coordination (e.g. for C-C, C-H and probably C-S) generates a peak at E 285.5eV. The asymmetry of the main C 1s peak is B = induced by the additional bonds to nitrogen that produce a slightly positive charge on the adjacent carbon atoms and results in a shift to higher binding energies at E 286.2eV. The sum of the fits B = accurately reproduces the measured peak. Nevertheless, it should be noticed that the areas of the fitted peaks not always reproduce the actual quantity of these bonds. When peaks are located too closely to each other for to the fitting algorithm, a distinct peak separation cannot be resolved.

The analysis of the sulfur S 2p level is the most crucial step towards the validation of a proper SAM formation. In contrast to the N 1s and C 1s core level spectra, S 2p exhibits a spin doublet due to spin-orbit coupling [143, 144]. This splitting between the S 2p3/2 and S 2p1/2 level is constant at ∆E 1.18eV. Therefore, the discussion of the S 2p level will only refer to the S 2p3/2 in the =

74 Section 5.2: Chemical Binding of N,N-dialkyl Dithiocarbamates on Noble Metals

A Au Survey spectrum C2DTC on Au(111) S 4f5/2 2p Au4f3/2

S1s C1s

Au4d3/2 Au4d1/2 300 250 200 150 Au Intensity [arb. u.] 4p3/2 N Au Au4p1/2 1s S Au5s 4s O1s C1s S2s 2p

800 700 600 500 400 300 200 100 0 Binding Energy [eV] B C D XPS N 1s XPS C 1s XPS S 2p C-C S 2p1/2 C-N S 2p3/2 C-O Intensity [arb. u.] Intensity [arb. u.] Intensity [arb. u.]

404 402 400 398 396 294 292 290 288 286 284 282 170 168 166 164 162 160 Binding energy [eV] Binding Energy [eV] Binding energy [eV]

Figure 5.3: XPS spectra of C2DTC on Au(111). The survey spectrum (A) confirms the presence off all expected elements as well as the absence of impurities. Only photoemission peaks of gold, carbon, nitrogen and sulfur are clearly observed. Oxygen as the most prominent type of impurity is not detected in in the survey spectrum, indicating that (almost) no random water, ethanol or oxygen molecules have contaminated the surface. (B-D) High resolution scans of the N 1s, C 1s and S 2p core levels. (B) The N 1s spectrum centered at EB 399.9eV fits well to nitrogen in an aromatic coordination. (C) Carbon reveals a slight splitting= of the C 1s level, which can be denoted to C-C and C-H bonds (E 285.5eV), C-N bonds (E 286.2eV) and B = B = little amount of unexpected C-O configurations (EB 288.3eV).(D) The S 2p3/2 level at (EB = = 161.9eV) corresponds to the bidentate binding state on gold.

following. Nevertheless, the S 2p1/2 level is always incorporated in the analysis of the measured S 2p spectra with the constant peak shift given above. In addition, the intensity ratio is constant with area(S2p ) : area(S2p ) 1.958 : 1 1 : 0.5107 . (5.1) 3/2 1/2 = = Moreover, an equal full width at half maximum (FWHM) of both peaks is assumed as it reflects the same chemical and instrumental peak broadening. The result of this fitting procedure is illustrated in figure 5.3D: The measured spectrum can be accurately fitted with one peak doublet, of which the S 2p3/2 is centered at EB 161.9eV. This suggests a bidentate binding of sulfur, in which = both atoms are symmetrically incorporated in the chemical binding to the gold surface [92, 145, 146]. It is interesting to note, that the S 2p level is slightly shifted (∆E 0.2eV) towards lower ≈

75 Chapter 5: Dithiocarbamates for Low Work Function Noble Metals

N C S O measured 9.4 % 63.8 % 21.9 % 4.9 % theoretical 12.5 % 62.5 % 25.0 % 0 %

Table 5.1: Stoichiometry of C2DTC monolayer on Au(111). binding energies compared to simple thiols [92, 145]. This shift can be attributed to the delocalized electronic nature of the CS2 linker group, which effectively raises the electron density at the sulfur atoms. Furthermore, the S 2p region reveals no traces of unbound sulfur in the form R-SH, which – would be located at E 163.5eV [146], nor in the charged form R-S , which was reported at B ≈ E 161.1eV [147]. Thus, both sulfur atoms of each molecule can be expected to be fully reacted B = on gold. Importantly, oxidized sulfur is also not detected at higher binding energies (E 167.0eV) B = [148].

A reasonable stoichiometry is found for C2DTC on gold with deviations resulting from oxygen contamination (less than 5 %), as listed in table 5.1. Oxygen species together with an increased amount of carbon indicates to the adsorption of some ethanol molecules on the surface and/or within the DTC layer. Nonetheless, using the S/Au-ratio yields a high packing density of C2DTC on gold with Θ (2.6 0.4) 1014 molecules/cm2 in good accordance to [27, 93]. = ± × In summary, the XPS spectra are almost free of unwanted species with only minor carbon-oxygen contributions, e.g. due to physisorbed ethanol, and non-reacted molecules. Instead, a densely packed monolayer with a surface coverage in good comparison to literature values has been ob- served [27, 93]. Additionally, core levels of nitrogen and sulfur indicate that a delocalized elec- tronic state spreads over the linker group and the nitrogen in the molecular backbone. Based on these observations, the binding of C2DTC on silver will be addressed in the next section.

5.2.2 XPS of C2DTC bound to Ag(111)

In principle, the XPS measurements of C2DTC on silver are comparable to the results on gold. This is in line with results of thiols, which are well known to form similar monolayers on silver and gold [32]. The nearest-neighbor distance in the Ag(111) plane of 2.89 Å is almost identical to that of gold with 2.88 Å. Therefore, the sulfur-sulfur spacing in a DTC molecule of 2.9 Å also perfectly matches the surface spacing of Ag atoms. These conditions suitable for monolayer formation are supported by the XPS measurements illustrated in figure 5.4: The survey scan (A) confirms the presence of expected elemental species of the DTC layer together with a variety of core levels originating from the silver substrate. Moreover, the amount of oxygen in this sample is even less than for the C2DTC layer on gold.

76 Section 5.2: Chemical Binding of N,N-dialkyl Dithiocarbamates on Noble Metals

A Survey spectrum C2DTC on Ag(111) Ag4s Ag3d3/2 Ag S2p 3d1/2 C1s S1s

300 200 100 Ag3p3/2 Ag3p1/2 Intensity [arb. u.] N Ag3s 1s Ag O1s 4p C1s S2s S2p Ag4s

800 700 600 500 400 300 200 100 0 Binding Energy [eV] B C D XPS N 1s XPS C 1s XPS S 2p C-C S2p(b) C-N S2p(ub) Intensity [arb. u.] Intensity [arb. u.] Intensity [arb. u.]

404 402 400 398 396 294 292 290 288 286 284 282 170 168 166 164 162 160 158 Binding energy [eV] Binding Energy [eV] Binding energy [eV]

Figure 5.4: XPS spectra of C2DTC on Ag(111). The survey spectrum (A) confirms the presence of all expected elements as well as the absence of impurities. Only photoemission peaks of silver, carbon, nitrogen and sulfur are clearly observed. (B-D) High resolution scans of the N 1s, C 1s and S 2p core levels. (B) The N 1s spectrum centered at E 399.9eV fits well to nitrogen in an B = aromatic coordination. (C) Carbon reveals C-C and C-H bonds at E 285.5eV and C-N bonds B = at EB 286.2eV. Traces of C-O as observed on gold are not detected. (D) The S 2p3/2 level with = a main component at E 161.8eV corresponds to the preferred bidentate binding state with little B = amount of unbound sulfur with E 160.6eV. B =

Nitrogen within the DTC layer is found again to be centered at E 399.9eV (figure 5.4B), which B = is exactly the binding energy found for the N 1s level in the C2DTC on gold. This can be inter- preted as a first indication that the molecular layer resembles the one on gold. Again, a delocalized electronic state, which originates from the linker group, extends the nitrogen atom in the molecular backbone. Yet, an increased signal-to-noise ratio is apparent compared to figure 5.3B. On silver, the N 1s level is located directly on top of a satellite peak of the Ag 3d core level (see figure B.1). With the use of standard background subtractions, the N 1s level cannot be analyzed with suffi- cient accuracy since the contributions of the satellite peak will always be incorporated in the area calculation. Therefore, it has been proven useful for this task to subtract a scan of this satellite peak from the N 1s scan of the DTC sample as a first background subtraction. Then, the com- plex background is removed and a peak fit can be performed as illustrated in (B). Nevertheless, this procedure lowers the signal-to-noise ratio since the peak shape of the Ag 3d satellite varies

77 Chapter 5: Dithiocarbamates for Low Work Function Noble Metals

XX

XX

Figure 5.5: Proposed binding scenario of C2DTC on Ag(111) from XPS measurements (left). A predominant bidentate binding, in which both sulfur atoms of one molecule are bound to Ag, is expected from the S 2p core level (right) with some amount of unbound sulfur ( 20%). The two coordination of sulfur are colored in green for bound species at 161.8 eV and in red< for an unbound configuration of 160.8 eV. The sulfur species in the illustration (left) are colored accordingly. slightly when adsorbates are present on the surface.

Carbon, however, can be analyzed with the same routine as performed for the C2DTC layer on gold. Since the amount of oxygen in this sample is almost not traceable, the C 1s level is also absent of the oxidized C-O component (figure 5.4C). Instead, the same characteristics can be observed for the main peak: It can be best reproduced again using two single peaks, where one represents the slightly charged carbon C-N species at E 286.2eV and the other the neutral B = configuration of C-C and C-H bonds at E 285.5eV. These values match the binding energies B = observed for C2DTC on gold, which also supports the identification of a similarly adsorbed mono- layer.

Finally, the sulfur S 2p region reveals a new feature as shown in figure 5.4D: A new peak at lower binding energies appears in the spectrum. It is evident that a single S 2p doublet is not able to re- produce this peak shape. The main S 2p3/2 component is centered at EB 161.8eV and only sightly = shifted towards lower binding energies compared to the formation on gold. In fact, this result fits well with literature values for thiols on silver, which has been reported to lie just above 162.0eV [149]. Keeping in mind the shift towards lower binding energies on gold of ∆E 0.2eV when thi- ≈ ols and dithiocarbamates are compared, this result seems reasonable to refer to sulfur in a covalent bond to silver. Thus, the S 2p3/2 level located at EB 160.6eV most probably corresponds to un- = bound sulfur in the absence of a hydrogen bond [147]. Such a scenario is highlighted in figure 5.5. Apparently, every one out of six sulfur atoms cannot form a bond to the underlying silver, i.e. every third DTC molecule can not establish a symmetric binding position. The origin of this asymmetry might be attributed to the actual differences between the gold and silver surfaces despite a common nearest neighbor spacing in the (111) plane, as also discussed by Abraham Ulman [32]: Silver is more reactive than gold and can spontaneously form an oxide, which, however, is not detected

78 Section 5.2: Chemical Binding of N,N-dialkyl Dithiocarbamates on Noble Metals

N C S O measured 9.5 % 63.1 % 27.4 % < 0.1 % theoretical 12.5 % 62.5 % 25.0 % 0 %

Table 5.2: Stoichiometry of C2DTC monolayer on Ag(111). for this sample. Still, a higher reactivity may also result in a stronger surface reorganization of the top silver atoms [150]. In addition, even for atomically flat surfaces the energetic landscape is heterogeneous due to relativistic effects [32, 151]. For Ag(111), the binding energy surface is less rough compared to the respective Au(111) surface. This leads to competing binding sites for the molecules on silver and might result in higher disorder in the film. However, such scenarios could not be reproduced with DFT simulations, in which a symmetric binding scenario has been observed. DTC molecules move from the top position, which is favored on gold, in a symmetric bridge position. It is noteworthy at this point, that massive surface relaxations are generally not reproduced by standard DFT techniques. For gold, a more sophisticated approach incorporating MD-DFT simulations have been proven to accurately describe the surface rearrangement of gold upon thiol modification [152]. For Ag, such an investigation has not been reported so far.

Altogether, C2DTC forms a closed-packed monolayer on Ag(111) similar to the adsorption on Au(111). A predominant bidentate binding geometry is identified with little amount of loosely bound sulfur atoms, which is attributed to surface induced disorder. The surface concentration is calculated from the S 2p to Ag 3d area ratio to Θ (2.2 0.5) 1014 molecules/cm2 using the = ± × method of Bramblett and to Θ (2.0 0.5) 1014 molecules/cm2 after the modified method of = ± × Schreiber. Both values match the packing density of C2DTC on gold. A reasonable molecular stoichiometry is found with a small underestimation of nitrogen, which should be traced back to the complex background of the N 1s level located on top of the Ag 3d satellite peak.

5.2.3 XPS of C2DTC bound to Cu(111)

Finally, the binding chemistry of C2DTC on copper is investigated and the corresponding XPS spectra are shown in figure 5.6. Copper is the most reactive of the noble metals investigated in this work. Nonetheless, a passivation layer of alkanethiolate monolayers have been reported to prevent the metal from oxidation [153]. After modification with C2DTC, the surface is found be free of unwanted species (A), especially oxygen is absent again. This finding also confirms the good working conditions at the ORPHEUS cluster tool, in which ambient conditions can be excluded at every stage of the sample preparation.

The high resolution measurements of the N 1s and the C 1s levels (B&C, respectively) strongly resemble the core levels found on silver and gold. Nitrogen is present with an improved signal-

79 Chapter 5: Dithiocarbamates for Low Work Function Noble Metals

A Survey spectrum C2DTC on Cu(111) Cu3s

Cu2p3/2 S2p C1s S1s Cu2p1/2

CuLMM 300 200 100 Cu2s

Intensity [arb. u.] Cu3p Cu3s O1s N1s C1s S2s S2p

1000 800 600 400 200 0 Binding Energy in eV B C D XPS N 1s XPS C 1s XPS S 2p C-C S2p(top) C-N S2p(hcp) Intensity [arb. u.] Intensity [arb. u.] Intensity [arb. u.]

404 402 400 398 396 294 292 290 288 286 284 282 170 168 166 164 162 160 158 Binding energy [eV] Binding Energy [eV] Binding energy [eV]

Figure 5.6: XPS spectra of C2DTC on Cu(111). The survey spectrum (A) confirms the presence off all expected elements as well as the absence of impurities. Only photoemission peaks of cop- per, carbon, nitrogen and sulfur are clearly discernible. CuLMM denotes the Auger photoemission peak. (B-D) High resolution scans of the N 1s, C 1s and S 2p core levels. (B) The N 1s spectrum centered at E 399.9eV fits well to nitrogen in an aromatic coordination. (C) Carbon reveals B = C-C and C-H bonds at E 285.0eV, C-N bonds (E 286.0eV). Traces of C-O are not trace- B = B = able. (D) The S 2p3/2 level reveals a more complex peak shape compared to the binding on gold and silver. Two core level doublets can be fitted to the full S 2p peak with the 2p3/2 components located at E 162.3eV and E 161.5eV, indicating an asymmetric binding scenario. B = B = to-noise ratio compared to silver, and can thus be fitted with a higher accuracy. The peak is centered again at E 399.9eV, indicating that DTC molecules chemisorbed on copper exhibit a B = similar electronic structure with nitrogen in a delocalized bond. Carbon is again found free of any oxidized states and can be accurately reproduced using two peaks representing C-C coordination at E 285.0eV and C-N at E 286.0eV. Compared to DTCs on gold and silver, a broadened C 1s B = B = peak shape is observed. This correlates to a shift of ∆E 0.5eV towards lower binding energies ≈ for the C-C configuration. This could point to a different charge distribution induced through the higher reactivity of copper. In particular, the carbon in the linker group reveals a different charging using the Bader charge analysis of a simulated monolayer [154]: While for C2DTC on Au(111) and Ag(111) a negative charge of q 1.22e and q 1.10e is observed, this carbon atom is Au = − Ag = − effectively neutral on Cu(111) (q 0.04e). Cu = −

80 Section 5.2: Chemical Binding of N,N-dialkyl Dithiocarbamates on Noble Metals

Figure 5.7: Proposed binding configuration of C2DTC on Cu(111). (A) The fixed sulfur-to-sulfur distance of 2.9 Å of the CS2 linker group does not match with the nearest-neighbor distance of copper (dNN 2.56 Å). Thus a asymmetric top-hcp coordination is observed by DFT for C2DTC = on copper. Such a binding scenario is also expected for residues of CS2 bound to copper, which is supposed to be incorporated in the DTC layer (B).

The sulfur S 2p core level yields a more complex structure than observed so far (D). Similar to the peak shape of S 2p on silver (figure 5.4D), two doublets are necessary to fit the measured data with sufficient accuracy. At first sight, a binding scenario can be envisioned, in which only one S atom of each DTC molecule binds to the copper surface yielding a binding energy of E 162.3eV. The B = other atom then appears to be less bound to the surface with E 161.5eV in a similar binding B = scenario as proposed for silver. In comparison, alkanethiols do not show such a split coordination of the sulfur core level on copper [149, 153, 155]. The reported binding energies for the S 2p3/2 component of alkanethiols (E 162.0 0.2eV) are in good agreement with the DTC sulfur located B ≈ ± at 162.3 eV. The small shift to higher binding energies might be induced by the asymmetric charge distribution between the two sulfur atoms of the DTC molecule.

What is the origin of the second S 2p3/2 species at E 161.5eV? Unbound sulfur in DTC B = molecules on gold has been reported to have even lower binding energies of E 161.1eV [147]. B = On silver, unbound sulfur has been identified at E 160.8eV (section 5.2.2). The energy differ- B = ence of almost 0.5 eV is too large to be associated with the same chemical coordination. Elemental sulfur in a CuS coordination has been reported at E 161.3eV after thermal treatment has decom- B = posed the C-S bonds in a methanethiol layer on Cu(100) [33, 156]. However, this finding implies that the DTC molecule or CS2 would break its covalent C-S bonds in order to form a copper sul- fide. So far, there are no reports of spontaneous decomposition of DTC molecules [145]. In order to get a deeper insight into this binding scenario, DFT calculations of C2DTC on Cu(111) have been performed by T. Schäfer [126]. Therein, the C2DTC molecules are always found to arrange in a asymmetric coordination with one sulfur placed on top of a copper surface atom, while the other is located in a quasi hollow (hcp) position figure 5.7A. This asymmetry can be traced back to the different nearest-neighbor distance of the Cu(111) plane (dNN = 2.56 Å) compared to gold and silver (2.88 Å and 2.89 Å, respectively). With a fixed sulfur-to-sulfur distance of 2.9 Å, the metal

81 Chapter 5: Dithiocarbamates for Low Work Function Noble Metals

N C S O measured 6.4 % 62.7 % 30.8 % < 0.1 % theoretical 12.5 % 62.5 % 25.0 % 0 %

Table 5.3: Stoichiometry of C2DTC monolayer on Cu(111). surface supplies favorable adsorption sites for both sulfur atoms of one DTC molecule. On copper on the contrary, a symmetric coordination in the form of top-top or hcp-hcp is not available and the C2DTC is forced to rearrange. Atomic charges calculated with the Bader scheme for C2DTC reveal a clear difference between the two sulfur atoms within one molecule [154]. For S in a top position q 0.90e due to the chemical bond to the metal, while S in hcp is slightly negatively top = + charged with q 0.26e. Such a behavior is not observed for C2DTC on gold nor silver in hcp = − DFT calculations, on which the DTC molecules relax only in symmetric binding positions. On Au(111), both sulfur atoms prefer a binding site directly atop a surface gold atom ( figure A.3A), which also leads to a symmetric charge distribution of the two sulfur atoms calculated with the Bader scheme (q 1.01e and q 1.05e). On Ag(111), a symmetric bridge position is always = + = + observed (figure A.3B), irrespective of the initial starting positions of the calculations (top, bridge, mixed, number of molecules). Similar to gold, a symmetric charge distribution is observed at the sulfur atoms (q 0.88e for both S atoms). Therefore, the measured S 2p peak shape on copper = + is expected to originate from this charge inequality.

A less than optimal molecular stoichiometry is found for C2DTC on copper (table 5.3). Although oxygen is absent, nitrogen is low by a factor of two from the ideal value, while sulfur is overly present. The discrepancy to an ideal monolayer formation is also observed in the examination of the surface coverage: Using the S/Cu area ratio, a value of Θ 4.7 1014 molecules/cm2 is ≈ × calculated. This large value reflects the high surface density of sulfur, but it seems to overestimate the surface coverage of DTC molecules. For tightly packed alkanethiols, a maximum of Θ = 5.9 1014 thiols/cm2 is expected on copper [149, 157]. From the current data, the most probable × scenario is the formation of a heterogeneous SAM, in which mainly C2DTC molecules are mixed with carbon disulfide (CS2), which has not reacted with the amines. Such a scenario is illustrated in figure 5.7B. This would explain the increased sulfur concentration in combination with an understoichiometric amount of nitrogen while carbon adapts an ideal value. Yet, both sulfur atoms of carbon disulfide are expected to bind in a similar coordination as proposed for the full DTC molecule. In fact, if only CS2 is allowed to bind to a Cu(111) surface, the S 2p level strongly resembles that of the DTC (figure B.2). Thus, estimating a molecular density from the S/Cu area ratio is a crucial attempt when different sulfur containing adsorbents are present on the surface. Nitrogen can be best employed for an estimate of the DTC surface coverage, which then yields Θ 1.6 1014 molecules/cm2. This value is slightly less compared to the layer formation DTC/Cu ≈ × on gold and silver, but reflects the presence of additional CS2 on the surface.

82 Section 5.2: Chemical Binding of N,N-dialkyl Dithiocarbamates on Noble Metals

5.2.4 Effect of Alkyl Chain Length on Monolayer Formation

At this point, longer alkyl chain substitutions are briefly discussed to highlight their influence on the chemical structure of CnDTC and the formation of monolayers on Ag and Cu compared to Au. For all monolayers, surface densities of the CnDTC SAMs have been examined and listed in table 5.4. For monolayers on Ag(111), both methods of Schreiber and Bramblett (section 3.5) yield comparable results for the S/Ag intensity ratio for all chain lengths. The resulting surface density is of the order of Θ 2.0 1014 molecules/cm2. The constant offset between the two models = × is ascribed to systematic errors, which are made in both methods via approximations. Statistical errors for each value are relatively small ( 0.2 1014 mol./cm2) based on the good sample-to- ± × sample reproducibility, while systematic errors can reach larger values up to 50 % as discussed in [91]. Nevertheless, the good comparability to the monolayer formation on Au(111) is striking and large errors up to 50 % are not expected. Similar to the observation for CnDTC on gold [27], the highest packing density is found for the shortest alkyl chain. For higher order alkyl chains n 2, > the surface packing seems to remain constant at Θ (1.8 0.1) 1014 mol./cm2. ≈ ± × The examination of the N 1s core level for DTC layers on silver is a crucial step due to its location atop of the Ag 3d satellite peak. This is also directly reflected in the extracted molecular densities listen in table 5.4. Figure 5.8 gives an overview of all important core levels of CnDTC on Ag(111), including the measurements collected for the Ag 3d core levels (A). Therein, the core levels of ni- trogen (B) reveal a low signal-to-noise ratio that further decreases with increasing methyl units in the SAM. Since nitrogen is located beneath the alkyl chains, its total intensity decreases with increasing chain length. Hence, the (naturally) low intensity of the N 1s level is convoluted with an increased noise ratio upon subtraction of the Ag 3d satellite peak as background, such that the N 1s core level diminishes further. All other core levels illustrated in figure 5.8 reveal ac- ceptable signal-to-noise ratios. The evolution of C 1s core level (C) is clearly dominated by the rising amount of C-C and C-H bonds, since the amount of C-N bonds remains independent on the alkyl chain length. Thus, a peak shift of 0.4 eV to lower binding energies can be identified from C2DTC to C4DTC. The higher order SAMs then reveal an almost constant peak center with a weak evolution back to higher energies of 0.1eV, which is most probably attributed to stabi- ≈ lization via van-der-Waals forces between molecules with long alkyl chains. Finally, the S 2p core levels of CnDTC on Ag(111) in figure 5.8D presents all features as has been previously discussed for C2DTC. A dominant doublet with its S 2p3/2 component is centered at 161.8 eV for all chain lengths, while a varying contribution of unbound sulfur at 160.8 eV is found, which never exceeds 20 %. The overall reduction of intensity for longer chains is caused by the increased scattering of photoelectrons when moving through these alkyl chains, since the constant atomic concentration of sulfur is buried beneath this layer of hydrocarbons.

For C2DTC on copper, the high calculated surface coverage using the S/Cu intensity ratio has

83 Chapter 5: Dithiocarbamates for Low Work Function Noble Metals

A B

C10DTC C10DTC Ag3d N1s C8DTC C8DTC C6DTC C6DTC C4DTC C4DTC C2DTC C2DTC Intensity [arb. u.] Intensity [arb. u.]

380 375 370 365 404 402 400 398 396 Binding energy [eV] Binding energy [eV]

C D

C10DTC C10DTC C1s S2p C8DTC C8DTC C6DTC C6DTC C4DTC C4DTC C2DTC C2DTC Intensity [arb. u.] Intensity [arb. u.]

290 288 286 284 282 170 168 166 164 162 160 158 Binding energy [eV] Binding energy [eV]

Figure 5.8: Core level spectra of N,N-dialkyl DTC on Ag(111) normalized to the Ag 3d level and vertically offset for comparison. Solid lines, except for Ag 3d, represent the sum of fits to the data point (open circles). (A) Core level spectra of the Ag 3d level of the metallic substrate illustrating the 3d5/2 to 3d3/2 doublet splitting due to spin-orbit coupling. (B) N 1s core level region revealing a low signal-to-noise ratio due to background subtraction of the Ag 3d satellite peak. Still, a constant binding energy of EB 399.9eV is observed for CnDTC with n 8.(C) C 1s = ≤ core level with rising contributions of C-C and C-H states at E 285.5eV for increasing number B = of methyl units. (D) S 2p core level region with a predominant binding energy of E 161.8eV B = and minor contribution of unbound sulfur at E 160.8eV. B =

84 Section 5.2: Chemical Binding of N,N-dialkyl Dithiocarbamates on Noble Metals

A B

C10DTC C10DTC Cu2p N1s C8DTC C8DTC C6DTC C6DTC C4DTC C4DTC C2DTC C2DTC Intensity [arb. u.] Intensity [arb. u.]

960 955 950 945 940 935 930 925 404 402 400 398 396 Binding energy [eV] Binding energy [eV]

C D

C10DTC C10DTC C1s S2p C8DTC C8DTC C6DTC C6DTC C4DTC C4DTC C2DTC C2DTC Intensity [arb. u.] Intensity [arb. u.]

290 288 286 284 282 170 168 166 164 162 160 158 Binding energy [eV] Binding energy [eV]

Figure 5.9: Core level spectra of N,N-dialkyl DTC on Cu(111) normalized to the Cu 2p level and vertically offset for comparison. Solid lines, except for Cu 2p, represent the sum of fits to the data point (open circles). (A) Core level spectra of the Cu 2p level of the metallic substrate illustrating the 2p3/2 to 2p1/2 doublet splitting due to spin-orbit coupling. (B) N 1s core level centered at E 399.9eV. Signal intensity reduces with longer chain length due to scattering B = of photoelectrons. (C) C 1s core level with rising contributions of C-C and C-H states at E B = 285.5eV due to increasing number of methyl units. (D) S 2p core level with two configurations of S 2p3/2 located at EB(top) 162.3eV and EB(hcp) 161.5eV. = =

85 Chapter 5: Dithiocarbamates for Low Work Function Noble Metals

14 2 10 mol./cm C2DTC C4DTC C6DTC C8DTC C10DTC × Bramblett: N/Ag 1.3 0.6 1.1 0.9 0.6 silver Bramblett: S/Ag 2.2 1.7 1.9 1.6 1.9 Schreiber: S/Ag 2.0 1.5 1.7 1.4 1.8 Bramblett: N/Cu 1.5 1.3 1.6 1.4 1.3 copper Bramblett: S/Cu 4.7 4.1 4.0 3.5 3.3 Schreiber: S/Cu 4.3 4.1 4.1 3.6 3.5 gold# Schreiber: S/Au 2.5 2.4 2.1 1.8 1.8

Table 5.4: Surface density of N,N-alkyl dithiocarbamates on Ag(111), Cu(111) and Au(111). Values are listed in units of 1014 mol./cm2. The models of Schreiber and Bramblett (section 3.5) have been applied to the sulfur-to-metal core level intensity ratios for Ag and Cu. On copper, only the calculation using the nitrogen core level yields a reasonable estimation for the surface # coverage of DTC molecules, as discussed in section 5.2.3. Values for CnDTC on gold adapted from [27] for comparison. been addressed in the previous section. In principle, this observation persists for the longer alkyl chain substitutions: Very large values are extracted with both methods, as listed in table 5.4. Even though the calculated densities decrease with rising alkyl chain length, all values remain larger than 3 1014 molecules/cm2 and therewith significantly higher than observed for a densely packed × monolayer on gold [27, 93]. In addition, this decrease may be induced to the methods themselves in this case: As elucidated in section 3.5, the molecular weight is incorporated in the analysis.

Yet, sulfur atoms are incorporated in both CnDTC and CS2 species. While the molecular weight raises with increasing alkyl units, that of CS2 remains constant. Thus, even for a constant atomic concentration of sulfur on the surface, the calculated density alters when the alkyl group of the

CnDTC is modified. This statement is furthermore supported by the examination of the DTC sur- face coverage using the N 1s core level. Nitrogen is exclusively incorporated in dithiocarbamates and this leads to a relative constant value of Θ (1.5 0.1) 1014 DTC molecules/cm2. = ± ×

Figure 5.9 shows the core level of CnDTC chemisorbed on Cu(111). Basically, the same effects are observed as previously discussed for the formation on gold and silver. The core level of

Cu 2p (A) is plotted for reference. The distinct splitting of the 2p level into its 2p3/2 and 2p1/2 species is induced via spin-orbit coupling. Nitrogen (B) displays a reasonable signal-to-noise ratio comparable to the measurement of C2DTC on Au(111). N 1s core level are centered at EB = 399.9eV for all SAMs with a decreasing signal intensity for rising chain length. The observation of this core level remains a clear evidence that CnDTC are bound on the copper surface. For carbon (C), a broadened C 1s peak of C2DTC has been attributed to the existence of C-N and C- C/C-H coordinated carbon atoms. Therein, the C-C/C-H states are slightly shifted (∆E 0.5eV) ≈ towards lower binding energies (E 285.0eV), which is most probably attributed to a different B = charge distribution associated with the asymmetric adsorption geometry on Cu(111). This shift

86 Section 5.3: Electronic Interface of N,N-dialkyl Dithiocarbamate Monolayers

is compensated with increasing number of methyl units, which leads to a C 1s level of C10DTC centered at E 285.5eV. Finally, the S 2p core level preserves its complex shape for all chain B = length (D). However, a good reproducibility of the peak position of the two configurations can be observed with S 2p3/2 located at EB(top) (162.3 0.1)eV and EB(hcp) (161.5 0.1)eV. = ± = ± Altogether, monolayer formation can be identified for all chain length substitutions investigated in this study. The details of the binding chemistry discussed for C2DTC on Au, Ag and Cu are mainly preserved with increasing chain length of the molecular backbone. On Ag(111), densely packed monolayers are observed for all CnDTC similar to the adsorption on Au(111). An incorporation of CS2 within the monolayers of CnDTC on Cu(111) has to be considered, which leads to a heterogeneous SAM. A synthesis, in which the adsorption of non-reacted CS2 molecules can be suppressed, would be an interesting point. This could be useful to further understand the mechanism that leads to the incorporation of CS2 on the Cu surface, which is not observed on Au and Ag. Additionally, to confirm of such a scenario, experiments based on STM or LEEM imaging should be considered. Nonetheless, these results set the foundation to generate low work function noble metals similar to those of CnDTC on Au(111). Hence, the electronic structure of the valence band region is discussed in the following.

5.3 Electronic Interface of N,N-dialkyl Dithiocarbamate Monolayers

In line with the previous section, the molecule C2DTC serves as a model system to examine and understand the electronic structure of N,N-dialkyl dithiocarbamates on the coinage metals Au, Ag and Cu. In particular, results from DFT calculations are limited to the systems of C2DTC layers. Work function changes by C2DTC are first discussed based on experimental and theoretical results. UPS measurements have also been conducted for all CnDTC modified metals and are evaluated in section 5.3.2.

5.3.1 Work Functions of N,N-diethyl Dithiocarbamates on Noble Metals

Again, the initial discussion is focused on C2DTC: Figure 5.10 presents the valence band region of C2DTC on the three metals as measured with UPS using He I radiation (hν 21.22eV). In = figure 5.10A, the ionization threshold of the secondary electrons is highlighted. This onset of photoionization is directly connected to the work function Φ of the surface. With the use of Φ hν E , where E marks the secondary electron cutoff (SEC), work function values of = − sec sec 3.2 eV, 3.1 eV and 3.5 eV are observed for C2DTC on Au, Ag and Cu, respectively. Recalling the work functions of the bare, unmodified metal (111) surfaces (table 5.5), significant shifts of

87 Chapter 5: Dithiocarbamates for Low Work Function Noble Metals

A B C

Au C2DTC Ag C2DTC Cu C2DTC Intensity [arb. u.]

19 18 17 12 10 8 6 4 2 0 4 2 0 Binding energy [eV]

Figure 5.10: UPS spectra of C2DTC on Au, Ag, Cu using He I irradiation (hν 21.22eV). (A) Secondary electron cutoff (SEC) that marks the threshold for photoionization and= has been employed to evaluate the work function via Φ hν E . Work functions of 3.2 eV, 3.1 eV and = − sec 3.5 eV are observed for C2DTC on Au, Ag and Cu, respectively. (B) Valence band region of C2DTC on Au, Ag and Cu. A common electronic structure is clearly observed for molecular states at 6eV E 12eV, which originate from σ-bonds within the molecule. (C) At lower < B < binding energies (0eV EB 6eV), the spectra comprise a convolution of molecular states and metal d-bands. The spectra< have< been vertically offset for clearance. the work function in the order of 1-2 eV are observed for all metals. The work functions listed in table 5.5 are among the lowest reported so far: For gold and silver, only the polymeric materials PEI and PEIE have produced comparable results of (3.6–3.9) eV, measured with Kelvin Probe in air [139]. On copper, a maximal shift of 2.5 eV using phenyl ispcyanide only persists at cryogenic temperature [158], while benzenethiol on Cu(100) has been reported to reach a work function of 3.7 eV [159].

The valence band region of C2DTC on Au, Ag and Cu is plotted in figure 5.10B. At higher binding energies, 6eV E 12eV, the spectra are dominated by molecular states. These originate from < B < σ-bonds within the molecule, which are clearly observable for all three samples. At lower binding energies (C), 0eV E 6eV, the spectra are less comparable and display different features. < B < These originate from a convolution of molecular states with the d-bands of the underlying metals. A more detailed discussion of this region will be given below.

Work function Φ Au(111) Ag(111) Cu(111) Bare metal 5.3 eV 4.4 eV 4.8 eV C2DTC 3.2 eV 3.1 eV 3.5 eV

Table 5.5: Work functions of C2DTC monolayer on Au, Ag, Cu.

88 Section 5.3: Electronic Interface of N,N-dialkyl Dithiocarbamate Monolayers

A Gold B Silver C Copper 3.0 bond dipole intrinsic dipole experiment 2.0 [eV] ∆Φ - 1.0

0.0 0 1 2 3 4 0 1 2 3 4 0 1 2 3 4 Surface coverage Θ [1014 cm-2]

Figure 5.11: Theoretical work function shift against surface coverage for CnDTC on Au, Ag, Cu from DFT calculations. The total shift ∆Φ is the sum of the contributions of the bond dipole (light blue) and the intrinsic dipole (dark blue) of the free, hydrogenated molecular layer. Differences in the molecular packing geometry generate alternate intrinsic dipoles for C2DTC on the three sub- strates even if the surface coverage is comparable. Calculated work function shifts are reproduced with good agreement to the experimental values. Calculations performed by T. Schäfer.

First, the following questions are central for the subsequent discussion: What is the origin of these low work functions? And: What are the differences for C2DTC on the three metals? The work functions of C2DTC on gold and silver are very similar with a difference of only 0.1 eV, while the difference to copper is more significant ( 0.3..0.4eV). Thus, DFT calculations have been per- ≈ formed to examine the interfacial dipole formation for several packing densities of C2DTC on all metal surfaces. Figure 5.11 illustrates the results of such an analysis. In each graph, the differ- ent contributions of intrinsic (dark blue) and bond dipoles (light blue) are depicted. In addition, the experimental data from XPS (surface coverage) and UPS (work function shift) are plotted for comparison. For all systems, a clear correlation between work function shift ∆Φ and molecular surface coverage Θ can be observed. On gold (A) and silver (B), the total shift can reasonably reproduce the experimental values. Furthermore, clear differences are discernible for monolayer formation on these two metals. Both a strong intrinsic dipole (1.7 eV) and bond induced dipole

(0.4 eV) generate a total work function shift of ∆Φ 2.1eV for C2DTC on gold. On silver, the sit- = uation differs, such that the overall shift is much less with ∆Φ 1.3eV as detected experimentally = with UPS. Yet, the calculations are able to reproduce this finding, as both the intrinsic (1.2 eV) and bond dipole (0.2 eV) are significantly lower than those observed on gold. Moreover, it is striking that, despite the similar monolayer formation, a different work function shift is observed. Instead, almost identical absolute work function values are observed. It is noteworthy, that Heimel and coworkers have found a similar behavior for thiolate monolayers comparing gold and silver [53].

On copper (figure 5.11C), the overall match between simulations and experiment is less convinc-

89 Chapter 5: Dithiocarbamates for Low Work Function Noble Metals

ing, which most probably has to be attributed to a larger deviation between theoretical and real layer structure. However, large work function shifts in the order of 2 eV are also observed with the calculations. In particular, relative large bond dipoles up to 1 eV are present. This finding results from a higher charge transfer across the metal/molecule interface. Here, the asymmetric adsorption geometry of C2DTC on the (111) surface of copper results in the aforementioned un- equal charge distribution within the CS2 linker group. Furthermore, the ideal representation of the Cu(111) surface may be less realistic in comparison with the experiment. As has been shown for thiols on Cu(111), a massive surface reconstruction can take place [157, 160]. Such processes are not incorporated in the current simulations. In fact, the actual reconstruction of Cu(111) upon adsorption of C2DTC remains to be validated using LEED and/or STM imaging techniques. More- over, the monolayer can be expected to be disordered with the incorporation of residues of bare

CS2 species in the SAM. All these phenomena affect the formation of a monolayer and lead to the moderate match between simulations and experiment. Nonetheless, from the DFT calculations of a homogeneous C2DTC monolayer, a work function as low as observed on silver and gold can be envisioned.

5.3.2 Work Functions of N,N-dialky Dithiocarbamates on Noble Metals

Subsequent to the discussion of the adsorption chemistry of C2DTC on Au, Cu and Ag, the effect of the alkyl chain length of the molecular backbone is briefly clarified. As observed for CnDTC on Au(111) (section 5.1 and reference [27]), the work function is almost independent of alkyl chain length. This behavior is exactly reproduced for CnDTC on Ag(111), figure 5.12A: Except for C2DTC, a common threshold for photoionization is observed. On Cu(111), a clear differ- ence between C2DTC and the higher order substitutions cannot be observed (D). In parallel to the secondary electron cutoffs, the valence structure evolves with rising alkyl chain length. The evo- lution of CnDTC on Ag(111) and Cu(111) is very similar (B&E) with an increasing amount of states located at 5eV E 12eV, which are attributed to the rising number of σ-bonds within the < B < alkyl chain. Otherwise, both the influence of the metal (d-states) and the frontier metal/molecule interface state decrease in intensity (C&F). This effect is attributed to the longer exit path for photoelectrons. Moreover, the position of the frontier interface state remains independent on the alkyl chain length. This further supports the finding that this interface state is strongly dominated by the linker group of the molecule.

Extracted work function values from the examination of the secondary electron cutoffs are high- lighted in figure 5.13A for all C nDTC on Au, Ag, Cu. Apparently, CnDTC reveal a common work function of Φ 3.5eV for n 2 on all three metal substrates. This trend is irrespective CnDTC = > of the initial work function of the substrate. A first important consequence can be identified: A variation of the work function of the metal substrate does not provide a handle to tune the work

90 Section 5.3: Electronic Interface of N,N-dialkyl Dithiocarbamate Monolayers

A B C Ag(111) C2DTC C4DTC C6DTC C8DTC C10DTC Intensity [arb. u.]

19 18 17 16 12 10 8 6 4 2 0 4 2 0 Binding energy [eV] D E F Cu(111) C2DTC C4DTC C6DTC C8DTC C10DTC Intensity [arb. u.]

19 18 17 16 12 10 8 6 4 2 0 4 2 0 Binding energy [eV]

Figure 5.12: UPS spectra of CnDTC on Ag (top row) and Cu (bottom row). For all spectra, the secondary electron cutoffs are plotted (A & D), which enables an estimation the work function via Φ 21.22eV Esec. Except for C2DTC on silver, a common ionization threshold and hence a common= work− function are observed. (B & E) The valence region of modified metals reveal an increasing amount of molecular states at 5eV E 12eV, which are attributed to the rising < B < number of σ-bonds within the alkyl chain. In parallel, both the influence of the metal, i.e. d-states, and the frontier metal/molecule interface state (C & F) decrease in intensity due to the longer exit path for photoelectrons. Spectra have been vertically offset for clearance. function of the modified surface. Instead, a linear trend is observed for the work function shift ∆Φ,

figure 5.13B. Obviously, C 2DTC does not follow this trend, since a constant offset by additional 0.3 eV is observed on gold and silver, but not on copper.

What is the origin of this common work function and why differs C2DTC? This is the key question for the following discussion. In particular, it not obvious at this point if the common level is a result from correlated effects or if this observation is only a coincidence. In general, there are several factor influencing the work function ΦSAM of a monolayer, such as depolarization within the monolayer, the factual surface coverage of molecules and the molecular orientation

For dithiocarbamates on noble metals, the binding motif of C2DTC seems to differ from those CnDTC with n 2. Therefore, different depolarizations between the molecules can be expected > [90]. On Au(111) and Ag(111), a three fold “trimeric” structure has been found to yield an optimal

91 Chapter 5: Dithiocarbamates for Low Work Function Noble Metals

A Ag Cu Au B Ag Cu Au 5.5 -0.5 C2DTC 5.0 -1.0 C4DTC 4.5 C6DTC

4.0 [eV] -1.5 C DTC [eV] 8 Φ 3.5 ∆Φ C10DTC -2.0 3.0 metals 2.5 -2.5 4.4 4.6 4.8 5.0 5.2 4.4 4.6 4.8 5.0 5.2 Φ Φ metal [eV] metal [eV]

Figure 5.13: Work functions of CnDTC on Au, Ag, Cu. (A) Absolute work functions Φ for modified and bare metal surfaces against the initial work function of the substrate Φmetal. Except for C2DTC on Au(111) and Ag(111), a common work function of ΦCnDTC 3.5eV is observed, independent on the metal species. Metal work functions plotted as gray circles= for reference. (B) Work function shift ∆Φ as a function of the initial metal Φmetal. A linear trend is observed CnDTC with n 2. > packing density with DFT simulations. These results are supported by a study of Morf et al., who performed a STM analysis of C2DTC on Au(111) [93]. Therein, C2DTC has been found to form a chiral “trimeric” structure with an asymmetric configuration of the ethyl chains. On Cu(111), however, this structure has been found be unstable in various DFT simulations [126].

In addition, the surface coverage has been shown to directly affect the total work function shift, see

figure 5.11. In fact, the highest molecular densities are found for C2DTC on Au and Ag. On Cu, an overall constant, yet lower value has been observed for all CnDTC layers. Certainly, this finding is correlated to the disorder of CnDTC on Cu(111). In addition, the “trimeric” structure seems to become disadvantageous for increasing number of methyl units. As known from thiols [88], longer alkyl chains prefer an orientation parallel to each other in order to increase intermolecular van-der-Waals forces. This behavior is also expected for N,N-dialkyl dithiocarbamates, which would counteract the high packing order of the “trimeric” structure. Apparently, this seems to be case already for CnDTC with n 4 and should explain the different work functions of C2DTC. ≥ Nevertheless, it is still not obvious why the work function reaches a common value for different metal species after modification in the first place. To address this question, the valence spectra are evaluated in the following section.

5.3.3 Valence Structure of N,N-diethyl Dithiocarbamate Monolayers

In this section, the discussion is focused on the valence band region of C2DTC on Au, Ag, and Cu. The general structure of C2DTC spectra have been introduced in figure 5.10C. As mentioned above, the electronic structure in the valence region of 0eV E 6eV displays a convolution of < B <

92 Section 5.3: Electronic Interface of N,N-dialkyl Dithiocarbamate Monolayers

the frontier molecular states with the d-bands of the metal substrates. In this region, the valence bands of CnDTC essentially matches those of C2DTC, except for a damping of these states due to the increased scattering of photoelectrons in the longer alkyl chains, figure 5.12. Thus, this discussion is performed again on the basis of the experimental and theoretical results of C2DTC monolayers. Figures 5.14, 5.15 and 5.16 present detailed graphs of this region for each system to further understand the electronic structure at the interface. Therein, each figure plots the valence band region of the bare and modified metal as measured with UPS, as well as the PDOS from DFT calculations.

Starting with C2DTC on Au(111) in figure 5.14: The valence structure of the bare metal surface is illustrated in (A) and reveals accentuated peaks in the region of 2eV E 6eV. These states < B < correspond to the 5d-bands of gold projected on to the (111) plane. In addition, the Shockley-type surface state is clearly visible at the Fermi edge [76]. All these distinct features are character- istic for a highly crystalline and textured gold surface in its (111) direction accompanied by an extremely flat surface [76]. These characteristics are otherwise only observed for a single crystal.

After adsorption of C2DTC (B), the surface state vanishes, but the presence of the Au 5d-bands is still discernible in direct comparison with the bare metal surface. From these bare spectra, it is dif- ficult to distinguish between the metal d-states and molecular orbitals. The comparability improves by performing a peak fit after subtracting the (linear) background for 0.5eV E 5.5eV (C). The < B < number of fitted peaks is chosen with the attempt to accurately reproduce the measured spectrum. Comparing these frontier metal/molecule interface states (MMIS) with the projected DOS of the sulfur 3p and gold 5d states highlights two observations: First, the region of 2eV E 5eV is < B < clearly a mixture of molecular and metallic states. A distinct separation is impossible, but the peak (colored in orange) at EB 2.5eV can be assumed to originate from the C2DTC, since the = bare gold has no sharp features at this energy (A). This state then corresponds to the HOMO-1. Second, a clear feature is observed at E 1.4eV after background subtraction (purple peak in B = figure 5.14C). Apparently, this frontier peak is a hybridized state of the S 3p orbital with compo- nents of the 5d orbital of the adjacent Au atoms (D). Such a hybridization has also been identified for thiols on gold [161]. The onset of this state reaches up to E 0.9eV and therefore marks the B ≈ highest molecular state. How will the frontier electronic states arrange in the cases of C2DTC on copper and silver, which have significantly different work functions?

Therefore, the discussion turns to C2DTC on Cu(111), whose work function is 0.5 eV lower than that of gold with Φ 4.8eV (Φ 5.3eV). Figure 5.15A illustrates the UPS spectrum Cu(111) ≈ Au(111) ≈ of the unmodified surface. Again, the most distinct peaks correspond to the d-states of the metal, which are more localized for Cu(111) compared to Au(111) and lie in a range of 2eV E 4eV. < B < Additionally, the Shockley surface state at EF is again clearly visible and verifies a metal surface of high crystallinity and a flat (111) surface plane [76, 162]. The sharp onset of 3d-states at E 2eV B = is directly recognized in the UPS spectrum of C2DTC on copper (B). As performed in figure 5.14,

93 Chapter 5: Dithiocarbamates for Low Work Function Noble Metals

A Au(111)

x10

B C2DTC on Au(111)

MMIS Binding Energy E [eV] C B sum of fits Intensity [arb. u.]

S 3p D Binding Energy EB [eV] S 3pz S 3px S 3py Au 5dz2 Au 5dzy

7 6 5 4 3 2 1 0

Binding Energy EB [eV]

Figure 5.14: Valence structure of C2DTC on Au(111). (A) UPS spectrum of bare Au(111) dis- playing accentuated 5d-states and a Shockley surface state at EF.(B) UPS spectrum of C2DTC on Au(111) with a convolution of metal d-states and molecular orbitals at 2eV E 6eV.(C) < B < UPS spectrum of C2DTC after a (linear) background subtraction for 0.5eV EB 5.5eV. Num- ber of fits chosen to reproduce the measured features. The frontier metal/molecule< < interface state located at EB 1.4eV (purple) can be identified with a projected DOS derived from DFT (D) to originate from= a hybridization of the S 3p and Au 5d orbitals of adjacent atoms. Gray dotted lines represent the remaining Au 5d spin components with minor hybridization to S 3p.

94 Section 5.3: Electronic Interface of N,N-dialkyl Dithiocarbamate Monolayers

A Cu(111) x10

B C2DTC on Cu(111)

MMIS Binding Energy E [eV] C B sum of fits Intensity [arb. u.]

S 3p D Binding Energy EB [eV] S 3pz S 3px S 3py Cu 3dz2

7 6 5 4 3 2 1 0

Binding Energy EB [eV]

Figure 5.15: Valence structure of C2DTC on Cu(111). (A) UPS spectrum of bare Cu(111) dis- playing accentuated 3d-states and a Shockley surface state at EF.(B) UPS spectrum of C2DTC on Cu(111) with a convolution of metal d-states and molecular orbitals at 2eV E 4eV.(C) < B < UPS spectrum of C2DTC after a (linear) background subtraction for 0.5eV EB 5.5eV. Num- ber of fits chosen to reproduce the measured features. The frontier metal/molecule< < interface state located at EB 1.3eV (purple) can be identified with a projected DOS (D) to originate from a hybridization= of the S 3p and Cu 3d orbitals of adjacent atoms. Gray dotted lines represent the remaining Cu 3d spin components with minor hybridization to S 3p.

95 Chapter 5: Dithiocarbamates for Low Work Function Noble Metals

a background fit reveals a more pronounced peak shape (C). And again, the convolution of metallic and molecular states is hardly separable for 2eV E 5eV. Projected DOS plots of sulfur and the < B < adjacent metal atoms also displays a superposition of d-bands with molecular states (D). Anyhow, a single feature at E 1.3eV can be correlated to a hybridized sulfur/metal state in analogy to the B = binding to gold (purple peak in figure 5.15C). This MMIS is only slightly shifted by 0.1 eV despite the different work functions of gold and copper. In addition, the frontier MMIS is more broadened on Cu(111) compared to the adsorption on Au(111). Such a broadening might be attributed to the asymmetric binding coordination of C2DTC on Cu(111) accompanied with a different charge distribution (see section 5.2.3). In parallel, this is also reflected in the PDOS of the S 3p states, which are more broadened with significant contributions at lower binding energies – in contrast to figure 5.14D for Au(111).

Finally, the lowest work function metal of this study, silver with Φ 4.4eV, is considered in Ag(111) = figure 5.16: In contrast to the metals discussed so far, silver has d-states located at higher binding energies with significant contributions to the UPS DOS at 4eV E 6.5eV. Similar to the bare < B < metal surfaces of Au(111) and Cu(111), the UPS spectrum of Ag(111) displays distinct 4d-states accompanied with a (weak) Shockley surface state at EF (A). The small height of the surface state is a common characteristic for silver and is observed for single crystals as well [45, 76, 162]. After deposition of C2DTC (B), the accentuated d-states are blurred, but the onset at EB 4eV is still = clearly visible. The frontier electronic states at 0eV E 4eV are thus less dominated by the < B < intense metallic d-bands (C), but nonetheless comprise Ag 4d- and 5s-states in combination with the molecular frontier orbitals. In contrast to C2DTC on Au and Cu, a single frontier state cannot be identified for C2DTC on Ag(111), but a set of metal/molecule peaks generate broader MMIS. Therein, a peak centered at E 1.7eV can be fitted with an onset at around 1 eV (purple peak in B = figure 5.16C). The origin of this frontier electronic structure can be found in the analysis of the

PDOS (D). Apparently, due to the binding coordination of C2DTC in a bridge position, a splitting of the S 3p orbitals can be observed in the order of 0.5 eV accompanied with a shift of 0.3 eV to higher binding energies. Nonetheless, these S 3p orbitals are hybridized with components of adjacent Ag 4d-orbitals. Hence, in spite of the different electronic structure of the UPS spectrum, a similar frontier metal/molecule interface state can still be observed.

What is key point of the examination of the valence spectra of C2DTC monolayers? Despite very different initial work functions ranging from Φ 4.4eV to Φ 5.3eV, i.e. a range of Ag(111) = Au(111) = almost 1 eV, the work functions of the modified surfaces are observed to vary much less. C2DTC has served as model system for the evaluation of the valence band region. The most important observation is the identification of a common metal/molecule interface state, which is located at

EB,MMIS (1.5 0.2)eV for C2DTC on all three metals. Based on the UPS measurements of all = ± CnDTC, this interface state is representative for all DTC monolayers. The observed deviations in the order of a view hundred meV can be expected to originate from little differences of the bond

96 Section 5.3: Electronic Interface of N,N-dialkyl Dithiocarbamate Monolayers

A Ag(111)

x10

B C2DTC on Ag(111)

MMIS Binding Energy E [eV] C B sum of fits Intensity [arb. u.]

S 3p D Binding Energy EB [eV] S 3pz S 3px S 3py Ag 4dz2 Ag 4dzy

7 6 5 4 3 2 1 0

Binding Energy EB [eV]

Figure 5.16: Valence structure of C2DTC on Ag(111). (A) UPS spectrum of bare Ag(111) dis- playing accentuated 4d-states and a Shockley surface state at EF.(B) UPS spectrum of C2DTC on Ag(111) with a mixture of metal d-states with molecular orbitals (4eV E 6.5eV) and a sep- < B < arated frontier metal/molecule interface state (1eV EB 3.5eV). (C) UPS spectrum of C2DTC < < after a (linear) background subtraction for 1eV E 4eV. Number of fits chosen to reproduce < B < the measured features. The frontier metal/molecule interface state located at EB 1.7eV (purple) can be identified with a projected DOS (D) to originate from a hybridization of the= S 3p and Ag 4d orbitals of adjacent atoms. Gray dotted lines represent the remaining Ag 4d spin components with minor hybridization to S 3p.

97 Chapter 5: Dithiocarbamates for Low Work Function Noble Metals

formation, e.g. from top to bridge binding positions. It is also possible that these discrepancies reflect a different peak broadening due to the variation of the binding geometries. Nonetheless, these deviations do not follow any trend with the work function – neither in magnitude nor di- rection. Therefore, in the picture of the energy level alignment at the metal/molecule interface (see section 2.2), the Schottky-Mott limit is clearly not applicable for the chemisorbed molecules investigated in this study. Likewise, the concept of Fermi level pinning seems inappropriate in this scenario due to the energetic position of the MMIS. In Fermi level pinning, the Fermi level usually aligns closely (within a few hundred meV) to the frontier molecular orbital [41, 56], which is not observed for dithiocarbamates in the present study.

In summary, the low work function values determined for C2DTC on Au and Ag can be traced back to a higher molecular density as compared to Cu. The common nature of the work function for all other CnDTC on Au, Ag and Cu appears together with a common interface state, which represents the chemical bond between the molecule and the metal. This observation may guide to the identification of a correlation between these properties and to an answer to the question: Is there a mechanism that explains a common work function? Based on the results in this section, a new model is proposed, discussed and evaluated in the following section: A bond-induced pinning mechanism, which is similar to Fermi Level Pinning (section 2.2.3), but dominated by the chemical bond between molecule and substrate.

5.3.4 Bond Induced Pinning

On the basis of the current results and taking into account results reported in the literature, a mechanism for strongly chemisorbed molecules is considered: It will be denoted as bond-induced pinning. Therein, it is assumed that the bond, which is established between the linker group of the molecule and the metal, dominates the position of the frontier interface state. This would directly imply a constant charge injection barrier (∆h and ∆e as defined in equation (2.6)) and potentially a constant work function. Such a type of pinning has not been reported in this way so far.

Figure 5.17 illustrates a scheme of how the energy level alignment could take place. Before a contact between the metals and molecules is established (figure 5.17A) all energy levels are referenced to Evac, as introduced in section 2.2.2. Note again, that the vacuum level of a surface is not an invariant energy level. In this illustration, the metals are therefore aligned with respect to their Fermi level EF, which leads to different vacuum levels due to their different work functions in figure 5.17A. In contact, a chemical bond is formed between the linker group of the molecules

(CS2 in the case of DTC) and the topmost atoms of the metallic surface (figure 5.17B). In the case of dithiocarbamates, the molecule/metal interface state (MMIS) has been identified to originate from a hybridization of S 3p states with metal d-states at a fixed binding energy of E (1.5 B,MMIS = ± 0.2)eV with respect to EF. At first, it is assumed that the charge rearrangement due to the formation

98 Section 5.3: Electronic Interface of N,N-dialkyl Dithiocarbamate Monolayers

Φ Φ Φ Φ Φ

Φ Φ

Figure 5.17: Proposed bond-induced pinning mechanism. (A) illustrates the situation before the metals and the molecule are brought into contact. All energy levels refer to the vacuum level Evac, as discussed in section 2.2.2. Note, that the different metals are aligned according to their Fermi level EF.(B) presents the alignment of the interface in contact with the formation of a chemical bond. The former HOMO (sulfur 3p lone pair, see figure 5.1) forms the metal/molecule interface state (MMIS) as a result of hybridized S 3p and metal d-states. As a result of the common binding energy of the MMIS with respect to EF a common work function Φeff results from different bond dipoles. of the chemical bond remains spatially confined within the linker group of the molecule2. Then, the molecular orbitals, which are not incorporated in the bond formation, are energetically equal to the free, isolated molecule and the effective work function of the modified metals Φeff is dominated by the electron affinity EA of the molecules.

The proposed mechanism applies for the DTC monolayers, which have been investigated in this work. As discussed at the example of C2DTC, a common work function is only observed if the molecular density is equal on the different substrates. However, an important task is to clarify if this model also successfully applies to other systems or if there are already reports that disprove its validity.

A comparable study has been conducted by Zangmeister et al., who analyzed the binding chem- istry for a set of thiols on the noble metals Ag, Cu, Au and Pt [163]. The authors also observed a common work function for each thiolate monolayer. Thus, the final work function values again seem decoupled from the initial work functions of the unmodified metals. Moreover, the fron-

2This assumption is also valid for DTC monolayers on Au, Ag and Cu.

99 Chapter 5: Dithiocarbamates for Low Work Function Noble Metals

tier molecular states only slightly vary within a few hundred meV and appears to be independent on the underlying metal and the molecular backbone. In fact, the authors also identify “(. . . ) the one common element, the metal-thiol linkage, as the most important factor in determining energy-level alignment” (p. 392 of reference [163]). These observations are thus consistent with the modeling of the bond-induced pinning mechanism. Nonetheless, further tests with different types of metal/molecule interface states should be considered for future studies. It should be kept in mind, that both thiolate and dithiocarbamate monolayers are based on a similar chemical bond between sulfur and the respective metal atoms.

A next important question should be addressed: Which molecular orbital dominates the energy level alignment if Fermi Level Pinning competes with the bond-induced pinning? This question refers to the (hypothetical) scenarios drawn in figure 5.18. Both scenarios are unrealistic and the system will somehow avoid the LUMO from getting filled (A) or emptying the HOMO (B), since a charging of the molecule is generally unfavored. Three options can be envisioned:

1. All frontier MO including the MMIS shift with respect to EF and avoid a charging of the molecule. This would disprove the bond-induced pinning model as a universal mechanism.

2. The frontier MO except the MMIS shift with respect to EF according to avoid a charging of the molecule while the binding energy of the interface state remains constant. This would support both the Fermi Level Pinning as well as the bond-induced pinning models.

3. The molecule actually is charged to keep the MMIS at its binding energy. This would counteract the commonly observed mechanism of Fermi Level Pinning and is therefore the most unrealistic scenario.

This question can not be clarified with the current experimental and theoretical results, since Fermi Level Pinning is not observed in any DTC monolayer of this work. Hence, a new set of molecules will have to be chosen in future studies. These molecules could be comprised by a large aro- matic core to induce a LUMO at a low binding energy to compete with the interface state after chemisorption. Potential molecules could be those of the DFT studies of monolayers on gold by Hofmann and Heimel [53, 59]. In particular, one set of phenyl molecules of reference [53] seems to combine the mechanisms of Fermi Level Pinning with bond-induced pinning: Pyridines with {1,2,3} phenyl rings. Illustrations of these molecules and representations of their highest molecular orbitals are depicted in figure A.4. Therein, the nitrogen lone pair can be identified as important orbital, because this atomic orbital forms the interface state by establishing a chemical bond between nitrogen and a surface metal atom. In addition, by adding a second and third phenyl ring, the highest two MO’s interchange their character: While the HOMO comprises the nitrogen lone pair of the smallest molecule (P1), the HOMO of the higher order molecules can be identified

100 Section 5.3: Electronic Interface of N,N-dialkyl Dithiocarbamate Monolayers

Φ Φ

Φ Φ Φ Φ

Figure 5.18: Hypothetical scenarios in which Fermi level pinning competes with bond-induced pinning. (A) depicts the scenario, in which the LUMO would be located below EF after chemisorption. (B) illustrates the other possible conflict, which would lead to an emptying of the HOMO. Both scenarios are unrealistic as drawn, because a charging of the molecule is gen- erally energetically unfavored. How the molecule will react in such scenarios, is discussed in the main text. as π-type and the nitrogen lone pair is located at HOMO-1 (P2, P3)3.

In reference [53], the authors have conducted a bond dipole analysis and investigated the electronic configuration with the projected DOS (PDOS) of each monolayer using DFT. Figure 5.19 shows the resulting PDOS of the three pyridine-based monolayers on gold. These PDOS exhibit three major attributes:

1. The metal/molecule interface state (red shaded area and inset of figure 5.19) is the product of hybridized atomic orbitals of nitrogen and gold, as elucidated in the discussion above. Moreover, the interface state is located at the same binding energy for all three pyridine molecules. This can also be expected, since adding phenyl rings to the molecular backbone should not interfere with the chemical bond of the linker group, which is always a nitrogen- gold bond.

2. The order of the highest occupied molecular orbitals has changed compared to the electronic structure of the isolated molecules (see figure A.4). After adsorption, the interface state represents the highest MO for all three molecules. Hence, the π-type MO’s are now located

3The annotation P1, P2 and P3 refers to the number of phenyl units in these molecules as can also be seen at figure A.4.

101 Chapter 5: Dithiocarbamates for Low Work Function Noble Metals

Figure 5.19: Projected DOS of Pyridines on gold. Annotations P1, P2 and P3 according to the number of phenyl rings, see figure A.4. HOPS (LUPS) refers to the highest occupied (lowest unoccupied) π-state of the molecules. The red shaded area corresponds to the chemical bond between nitrogen and gold, as illustrated in the inset. A constant interface state is observed, while Fermi level pinning occurs at LUPS for P2 and P3. See text for detailed discussion. Reprinted with permission from [53]. Copyright 2007, American Chemical Society.

at higher binding energies (labeled with HOPS4 in figure 5.19). Due to bond formation, the electronic structure of the molecules have changed with only one constant state: The interface state representing the chemical bond.

3. Fermi Level Pinning takes place for the higher order molecules P2 and P3. With extend- ing the large aromatic cores, the band gap of the molecules diminishes. Yet, the lowest 5 unoccupied orbital (labeled with LUPS ) cannot be located below EF and is thus pinned to

EF. As analyzed by Heimel and coworkers, the molecule keeps the LUMO above EF by a polarization single bonds within its molecular backbone. Therewith, the energetic position of the MO’s across the aromatic backbone (including HOPS and LUPS) can be controlled (see also figure 3 of reference [53]). The interface state remains, however, unaffected by this polarization.

In summary, the evaluation of the results based on the DFT studies of Georg Heimel and coworkers suggest, that both Fermi Level Pinning can appear with a fixed interface state. Therefore, it will be most appealing to examine how these pyridine molecules bind to other noble metals such as silver, copper and platinum. The key task will be the examination and comparison of the interface states in these systems. This could finally lead to new insights of interface physics and lead to an improved understanding of the dominating processes in monolayer formation and tailoring work functions.

4HOPS: Highest Occupied Pi State 5LUPS: Lowest Unoccupied Pi State

102 Section 5.4: Electrode Modification in PTCDI-C13 OTFTs

5.4 Electrode Modification in PTCDI-C13 OTFTs

In this last section, the versatility of N,N-alkyl dithiocarbamates to support electron injection at metal/organic interfaces will be presented. For this approach, the n-type molecule PTCDI-C13 has been chosen as active material in OFET devices [28]. Moreover, the use of DTC surface modifications offers the opportunity to gain new insights into the fundamentals that govern charge transport at early stages of growth. This has been demonstrated in a joint project with Ingolf Segger in our research group and will be presented in section 5.4.2.

5.4.1 Contact Resistance in OTFT with CnDTC Modified Electrodes

The prospect of CnDTC to improve charge injection from a modified electrode into the LUMO of an (organic) semiconductor can be envisioned from the results of the previous sections. The key point to achieve an efficient charge transfer at the interface is the minimization of the interface barrier, see section 2.3.1. One important step is the optimization of the barrier height using polar molecules [164–169]. Fluorinated molecules have been employed to raise the work function of an electrode, which is desired to align the Fermi level EF of the metal to the molecular HOMO level for hole injection and extraction. However, an efficient electron injection requires a low work function to minimize the energetic barrier to the LUMO. Here, N,N-dialkyl dithiocarbamates with a constant work function of Φ 3.5eV are promising candidates to achieve this goal. =

In fact, as has been previously presented by Schulz et al. [27], CnDTC modifications of the gold source and drain electrodes in a thin film transistor are able to improve the device performance of n-type OFEts. Yet, this study is more exemplary and less quantitative. Here, a further step foward has been taken by employing the transfer line method (TLM, section 2.3.3) to OFETs with modified silver and copper source and drain electrodes. Therefore, a set of PTCDI-C13 transistors with varying channel length L have been prepared and investigated.

As many n-type charge transport materials, PTCDI-C13 is volatile under ambient conditions. Therefore, an approach of encapsulation with polyethylene has been utilized [170]. This kind of encapsulation is by far not perfect, since polyethylene is known to provide only a minor barrier for diffusion of non-polar molecules like oxygen. Still, the reproducibility of the OFET devices is significantly improved such that these devices operate stable for a few hours under ambient condi- tion. Figure 5.20A presents a set of transfer curves of P TCDI-C13 OFETs with CnDTC modified electrodes. The overall performance is excellent, which can be identified by several factors: First, only a low drain voltage of V 5V is applied. Second, a threshold voltage of virtually zero is an D = indication of (i) a small injection barrier and (ii) a small amount of traps. At this point it should be noted, that the physics and effects of the threshold voltage in organic transistors is a very complex

103 Chapter 5: Dithiocarbamates for Low Work Function Noble Metals

A B x104 10-7 30 C4DTC ] m] -1 -8 25 C6DTC Ω

Ω 10 C8DTC W [ 20 C10DTC 10-9 ON 15 10-10 C4DTC 10 -11 C6DTC 10 C DTC 5

Channel conductance [ 8 Total resistance R C10DTC 10-12 0 -10 0 10 20 30 0 50 100 150 200

Gate Voltage VG [V] Channel length L [µm]

Figure 5.20: PTCDI-C13 OFET characteristics with CnDTC modified S/D electrodes. (A) OFET transfer measurements of PTCDI-C13 transistors with modified Ag source and drain contacts revealing clear on-to-off currents at low voltages (V 5V ) and low threshold voltages of V .(B) D = T Plots of transfer line method (TLM) to extract the contact resistance Rc, which can be identified at the intercept of the linear regressions (grey dotted lines) at the y-axis. Low contact resistances in the regime of R 2..7MΩcm are observed. c = issue and care should be taken to not misinterpret its position [68, 171–173]. In particular, a low threshold voltage should not be taken as unambiguous proof for a low concentration of trap states. The actual density of traps has to be identified using a set of experiments as discussed by Kalb and Batlogg [46, 47] Third, a distinct off-to-on ratio is already achieved at low gate voltages of V 5V. G ≈ Subsequently, TLM has been applied to evaluate the width-normalized contact resistance R W c · for this set of devices according to section 2.3.3. Hence, the width-normalized resistance in the ON-state of the transistor, R W at constant V , is plotted against the channel length L and a ON · G linear fit to these data points is performed, figure 5.20B. The intercept of this fit with the y-axis then yields R W . Apparently, a linear trend for the contact resistance with increasing alkyl chain c · length of CnDTC is observed: Values of 1.9, 3.2, 4.9 and 6.6MΩcm are obtained for n 4, 6, 8 = and 10, respectively. This effect is attributed to a tunnel barrier, which increases with the length of the non-conductive alkyl chains [174–176]. In addition, the charge carrier mobility of PTCDI-C13 is calculated from a linear fit to the square root drain current in the saturation regime (section 2.3). The extracted values are in the range of 0.006 to 0.02cm2/Vs and match previously reported values [29].

In addition, PTCDI-C13 OFETs with modified Cu electrodes have been characterized as illus- trated in figure A.3. Despite a more difficult sample preparation, these devices also yield contact resistances and mobilities comparable to those with modified Ag electrodes. This is not entirely

104 Section 5.4: Electrode Modification in PTCDI-C13 OTFTs

surprising, since a common work function of the modified metals has been observed after CnDTC modification, as shown in figure 5.13. Therefore, a common contact resistance can be assumed irrespective of the choice of the S/D-electrode material (Au, Ag, Cu). Yet, not only the electronic structure is decoupled upon modification: PTCDI-C13 molecules close to the electrodes will also experience a different surface free energy, which in turn affects the growth of the organic thin film [177].

5.4.2 In-Situ Evolution of Charge Transport using DTC Modified Electrodes

The previous section has demonstrated the improved performance of PTCDI-C13 OTFTs if the source and drain electrodes are covered with a DTC layer. This offers the opportunity to further investigate the charge transport of this n-type material. As known from literature, PTCDI-C13 is a material that can show pronounced layer-by-layer growth [29]. Here, one question has always been challenging: In which layers does charge transport in an organic thin film transistor take place? It has always been assumed, and first results supported the idea, that only the first few nanometers actually participate to charge transport [178–182]. In a joint project with Ingolf Segger at the I. Institute of Physics (IA), a new characterization setup has been designed, build and utilized. This setup offers the electrical characterization of OTFTs during growth of the organic thin film, 6 i.e. in vacuum at p 1 10− mbar. Hence, the evolution of charge transport can be directly ≈ × observed by an in-situ technique. Therewith, the Bachelor students Andreas Hessler [183], Emily Hofmann [184] and Arthur Leis [185]6, could successfully generate new insights into the charge transport properties of PTCDI derivatives. In addition, a more comprehensive correlation of charge transport with optical and structural properties of organic thin films based on PTCDI derivatives can be found in the dissertation of Ingolf Segger [186]. Here, a brief overview of PTCDI-C13 is given on the basis of the results obtained by Arthur Leis [185].

Figure 5.21 shows a typical measurement of PTCDI-C13 at low deposition rate ( 0.05 Å/s) and ≈ a moderate temperature (T 40 ◦C). This sample serves as a reference for the following discus- ≈ sion. As can be seen, both the drain current ID and the gate current IG are collected during the measurement. This is important to check that the major current is actually confined in the molec- ular channel between source and drain. Otherwise, e.g. if the oxide is damaged, the applied gate voltage can directly induce a charge flow from gate to source. However, as can be observed from

figure 5.21A andB, IG is more than four orders of magnitude smaller than ID.

Both drain and gate currents are colored in order to identify the layer thickness at the beginning of each measurement, starting from dark blue (low layer thickness) to dark red (high layer thick- ness). The first measurements do not show any field-effect. Apparently, no closed paths from

6The three students have been supervised in the course of this work in collaboration with Ingolf Segger at the I. Insti- tute of Physics (IA).

105 Chapter 5: Dithiocarbamates for Low Work Function Noble Metals

− −

− −

− − k k

− −

− − kk kk kk

− −

− − kkk kkk

Figure 5.21: In-situ transfer characteristics of PTCDI-C13 during deposition at T 40 ◦C. Both ≈ drain (A) and gate currents (B), ID and IG are plotted in a color spectrum from dark blue (low layer thickness) to dark red (large layer thickness), as indicated by the colorbar. (A) shows the transfer curves at a constant drain voltage of VD 10V. The measurements with low layer thickness (dark blue) show no field-effect as long as paths= of molecules have not connected the source and drain electrodes. After this critical point is reached, the field-effect clearly evolves with rising layer thickness and rapidly saturates. These final transfer curves reveal a good reproducibility and a clear off-to-on ratio. (B) The gate current are significantly smaller than source-gate current in (A). Still, a field-effect can also be observed for IG. source to drain have formed at this early stage of deposition. As soon as a critical layer thickness

Θc is reached, the characteristic field-effect current evolves quickly with rising layer thickness and saturates with very reproducible transfer curves. The critical thickness Θc, at which the first conduction occurs, is also denoted as the point of percolation [187, 188]. For a perfect two di- mensional system with a random distribution of sites, this percolation limit is calculated to be Θ2d 0.68monolayers [189, 190]. For two dimensions, this number represents the area ratio that c = has to be covered to statistically generate the first path from one end of the system to the other. Figure 5.21 is useful to gain an overview of the sample’s characteristics, but has to be further ana- lyzed to gain a deeper insight. From figure 5.21A itself, no more than an evolution of conduction to a saturation can be observed. Indeed, saturation is reach after only a few measurements. Appar- ently, charge transport remains limited to approximately the first 3 nm. The transfer characteristics seem to be unaffected on further deposition of PTCDI-C13. This will discussed subsequently in more detail.

In analogy to section 2.3.2 the mobility and threshold voltage can be extracted from the transfer curves using equation (2.18). The results are illustrated in figure 5.22 and plotted against the layer thickness in units of a monolayer thickness ML of PTCDI-C13 (1ML 2.5nm). At first sight, ≈

106 Section 5.4: Electrode Modification in PTCDI-C13 OTFTs

A B 0.06 35 I II 0.05 30 [V] t 25 /Vs]

2 0.04 III IV V 20 [cm

µ 0.03 15 0.02 10 Mobility 0.01 Threshold voltage V 5 0 0 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 Layer thicnkess Θ [ML] Layer thickness Θ [ML]

Figure 5.22: In-situ mobility and threshold voltage of PTCDI-C13 at T 40 ◦C.(A) Mobility µ as a function of the layer thickness Θ. Five distinct regimes can be identified,≈ in which the sections I, III and V show no distinct correlation between layer thickness and mobility. A clear, almost linear increase of µ with Θ is only present within the regimes II and IV. This behavior correlates to the layer-by-layer growth mode of PTCDI-C13. The plateaus observed in I and III reflect the fact, that the current layer has not formed closed paths, that could carry charge flow [185, 186]. (B) Threshold voltage V rapidly saturates to V sat 12.5V as a function of the layer thickness. t t ≈

both the mobility µ and threshold voltage Vt in figure 5.22A andB, respectively, saturate after approximately 1 ML. While this statement can be considered true for the threshold voltage with a saturation value of V sat 12.5V, the evolution of the mobility reveals more details: For PTCDI- t = C13 at T 40 ◦C, the thickness dependent mobility can be divided into at least five regimes. At ≈ first, no field-effect mobility is observed since no closed paths are present for 0ML Θ 0.5ML. < < Secondly, an intense and almost linear increase is present for 0.5ML Θ 1.0ML, which then < < abruptly stops. In this section, the first conductive paths from source to drain are forming and enable charge transport. With rising layer thickness, more charges are able to participate to the field effect transport until this first layer is fully closed. In the third section, again a thickness- independent mobility is observed for the range of half of a monolayer. For 1.5ML Θ 2.0ML, < < a further increase of the mobility results in the final saturation mobility. This behavior correlates to the layer-by-layer growth mode of PTCDI-C13: The plateaus observed in the regimes I and III reflect the fact, that the first and second layers, respectively, have not formed closed paths yet, which could carry charge flow [185, 186]. Apparently, only the first two layers of PTCDI-C13 are involved in charge transport, which form in the regimes II and IV. Finally, in regime V, the final saturation is reached and an almost constant mobility is observed. Small deviations from a constant value are attributed to trapping of free carriers [47]. In this regime, a further deposition of PTCDI-C13 does not affect the mobility and, therefore, is irrelevant for charge transport.

At this point, it is useful to examine the derivative of the mobility dµ/dΘ, figure 5.23. Plotting

107 Chapter 5: Dithiocarbamates for Low Work Function Noble Metals

3.5 ] -3 3.0 Θ1 = 0.75 2.5 /(Vs•ML) x 10 2 2.0 [cm θ 1.5 /d µ

1.0

0.5 Θ2 = 1.70

0.0 Charge density d

0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 Layer thickness Θ [ML]

Figure 5.23: Charge carrier density dµ/dΘ of PTCDI-C13 at T 40 ◦C. The first monolayer dominates the charge transport, which forms forms closed paths≈ at a thickness range between 0.5ML Θ 1.0ML. Only two layers can be unambiguously identified to participate to charge < < transport with peak centers at Θ 0.75ML and Θ 1.70ML. 1 = 2 = dµ/dΘ against Θ also reveals the dominant role of the first monolayer. Moreover, a fit with Gaussian peaks helps to identify all layers that contribute to charge transport, since this derivative scales with the charge carrier density. It could be shown for pentacene, that this quantity depends on the deposition rate [187]. A variation of the deposition rate influences the growth mode and therewith the number of closed paths from source to drain if a mixture of layer-by-layer and island- like growth is apparent. It will remain interesting to investigate the affect of the deposition rate on the evolution of the charge transport of PTCDI derivatives.

For PTCDI-C13 at moderate temperature during growth (T 40 ◦C), two clearly separated peaks ≈ can be observed with peak centers located at Θ 0.75ML and Θ 1.70ML. This observation 1 = 2 = directly reflects the evolution of µ in figure 5.22A. Apparently, the third layer does not participate in charge transport. This might be caused by two effects: First, a change of the growth mode from layer-by-layer to a more grainy-like island growth mode can hinder charge transport in the third and subsequent layers. This has in fact been observed for PTCDI-C13 at room temperature in our research group using Atomic Force Microscopy [185, 186, 191]. Second, the molecules of the first two layers may also be able to screen the electric field of the gate electrode, such that subsequent layers are not affected by the applied gate voltage. To verify such a scenario, surface potential measurements of the OTFT channel during operation can be envisioned [192, 193]. In particular, conductive AFM and Kelvin probe measurements should be considered to solve this task [194–196].

108 Section 5.4: Electrode Modification in PTCDI-C13 OTFTs

T = 70°C T = 90°C T = 100°C T = 125°C A1 B1 C1 D1 0.4 0.4 0.4 0.4 /Vs]

2 0.3 0.3 0.3 0.3 [cm µ 0.2 0.2 0.2 0.2

Mobility 0.1 0.1 0.1 0.1

0.0 0.0 0.0 0.0 0 1 2 3 0 1 2 3 0 1 2 3 0 1 2 3 Layer thickness Θ [ML]

14 40 35 30

] A2 B2 C2 D2

-3 35 12 30 25 Θ1A Θ1A 30 Θ1 10 Θ1B 25 Θ 1A 20 25 8 20 20 Θ1B 15

/(Vs•ML) x 10 6 15

2 Θ 15 1B 10 4 Θ 10 [cm Θ2 10 2 θ Θ /d 2 Θ3 5 2 5 µ 5 Θ3 Θ Θ3 Θ Θ Θ d 4 4 2 3 0 0 0 0 0 1 2 3 0 1 2 3 0 1 2 3 0 1 2 3 Layer thickness Θ [ML]

Figure 5.24: Mobility and charge carrier density dµ/dΘ of PTCDI-C13 vs. temperature T . Top row: In-situ mobility µ as measured for different substrate temperatures. Bottom row: Cor- responding derivatives of µ against the layer thickness Θ. In general, the mobility in the top row show the well-known temperature dependency of organic semiconductors with a maximum around T 90 ◦C. Furthermore, a more pronounced evolution can be identified with up to four layers contributing≈ to charge flow in the thin film.

In the following, the influence of the substrate temperature on the in-situ mobility µ and charge carrier density dµ/dΘ is investigated. It is known for PTCDI-C13, that higher temperatures as well as annealing promotes layer-by-layer growth while lower substrate temperatures induce a poorly ordered and grainy structure [185, 186]. In figure 5.24, a set of PTCDI-C13 samples deposited at different temperatures is illustrated. The top row shows the in-situ mobility µ against the layer thickness Θ and the bottom row depicts the corresponding derivatives of µ. At first sight, the mobility shows the well-known temperature dependency for organic semiconductors [62]. At elevated temperatures, a maximal mobility is usually observed in the common models for charge

109 Chapter 5: Dithiocarbamates for Low Work Function Noble Metals

120

100 C] °

80

Temperature [ Θ 60 1A Θ 1B Θ 2 Θ 3 40 Θ 4

0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 Layer thickness Θ [ML]

Figure 5.25: Evolution of charge transport layers of PTCDI-C13 as a function of temperature.

7 transport in molecular semiconductors [64] . For PTCDI-C13, a maximum at T 90 ◦C is observed ≈ with a maximal mobility of µ 0.4cm2/Vs. For all samples, except at the highest temperature, max = the characteristic evolution of µ is apparent with a stepwise increase to a saturation. This increase of µ is always interrupted by regimes, in which the mobility remains almost independent of the layer thickness. In fact, these regimes actually reveal a slight decrease of µ which might be induced through temperature induced trapping, thermal stress and/or phonon scattering. However, at these temperatures a significant contribution of the second to fourth layer of PTCDI-C13 can be identified as a result of a more pronounced layer-by-layer growth8. Moreover, the positions of these layers shift with temperature and follow the same trend as observed for the mobility: With rising temperature up to T 90 ◦C more layers contribute to charge transport and the positions ≈ of those layers shift towards a lower thickness. Above 90 ◦C, these peaks move back towards higher thicknesses. The positions of Θ1 to Θ4 are also highlighted in figure 5.25 as a function of temperature. The turning point between 90 ◦C and 100 ◦C may be caused by an increasing disorder of the layer, e.g. due to thermally induced desorption and vibrational movement of PTCDI-C13 molecules, as well as an increasing contribution of phonon scattering. Apparently, these seem to be dominant factors for the sample at 125 ◦C, at which a distinct contribution of more than the first layer to charge transport can hardly be observed.

7(i) A modified band transport model with a significant amount of traps, i.e. the multiple trapping and release model, and (ii) a classical polaron hopping mechanism, in which charges move via frequent movements along localized states. 8A more detailed study of the correlation of growth and structure of PTCDI derivatives with temperature is also discussed in reference [186].

110 Section 5.4: Electrode Modification in PTCDI-C13 OTFTs

Additionally, a splitting of the first peak can be identified and is denoted by Θ1A and Θ1B. The origin of this splitting is not clear and needs further experiments, in particular with AFM to probe the film morphology. Nonetheless, a clear correlation between electrical properties and thin film structure can be identified: While only two molecular layers contribute to charge transport at room temperature, additional layers are also able to participate to charge flow at higher temperatures. It is known, that PTCDI-C13 forms highly ordered layers at elevated temperatures [28, 186]. Yet, at temperatures above approximately 100 ◦C, defects in the form of cracks (possibly at grain bound- aries) can occur [186], which increase scattering and lower the total conductivity.

Definitely, there are further interesting aspects that could be investigated, such as the influence of the deposition rate, substrate modifications, other electrode modifications, trap state distribution and, of course, the affect of different molecular substitutions. Some of these aspects have been addressed in recent studies at the I. Institute of Physics (IA) [184–186].

111

Chapter 6

Work Function Tuning with Tailor-Made Dithiocarbamates

F d

dd

The previous chapter has demonstrated the versatility of dithiocarbamates to functionalize surfaces of noble metal electrodes. Taking advantage of the small C2DTC as a model monolayer material, new insight into the chemical and electronic interface of dithiocarbamates on the coinage metals Au, Ag and Cu has been generated. One important finding results from the common work function of modified metal surfaces: Apparently, varying the work function of the substrate does not pro- vide a handle to tune the work function of the modified surface and, therefore, cannot be used to optimize the charge injection barriers at electrode/semiconductor interfaces. To achieve this goal,

113 Chapter 6: Work Function Tuning with Tailor-Made Dithiocarbamates

two approaches are possible: On the one hand, the energetic levels of the organic semiconductor can be shifted via attachment of functional groups to the molecular core [197]. On the other one, the energetic surface of the electrode material can be further tuned by different surface modifiers with the desired dipole moment [198]. To date, this approach has not been exploited for dithio- carbamates, even though a variety of DTCs have been reported for the functionalization of gold nanoparticles [31]. Consequently, a set of new dithiocarbamates has been investigated and will be discussed in this chapter. A first impression of these tailor-made surface modifier is given by the illustration of molecules in the chapter figure on page 113. In principle, the reaction scheme for dithiocarbamate monolayers, which is also depicted in the figure above, can be easily modified by replacing the dialkyl amine with any other secondary amine that comes to mind. Hence, this opens the door to a new generation of electrode modifications. Nonetheless, when discussing work functions of noble metals, thiolate monolayers should be included in every consideration to tune the work function – especially if high values are desired for hole injection. Therefore, this chapter begins with a short recap of thiolate monolayers. For all samples in this chapter, the discussion of XPS measurements is referring to the plots given in the appendix (section B).

6.1 Reference Systems: Thiols

On gold, alkanethiols represent one of the most intensively studied class of material that form self- assembled monolayers [50, 88, 89]. The binding chemistry, molecular orientation and packing as well as their effect on energetic properties such as the surface free energy have been addressed using various experimental and theoretical techniques [32, 143, 151, 199, 200]. In addition, thiols have been found to bind equally to different noble metals, even though some variations of the molecular arrangements have been observed as a result of the different surface structures of the (111) planes [32, 149, 153]. Hence, thiols represent valuable reference systems in the scope of this work. Depending on the constitution of the molecular backbone and head group, different work functions can be achieved. In particular, a (partial) fluorination of the alkyl or aryl unit of the molecule are commonly employed to increase the work function [164, 198, 201]. Here, two representatives of the class of thiols have been investigated as reference systems: A hydrogenated alkanethiol with ten units of hydrocarbons, i.e. decanethiol (abbreviated with C10Thiol), and its fluorinated counterpart, 1H,1H,2H,2H-Perfluorodecanethiol (FC10Thiol).

In close similarity to the ethanethiol, which has been discussed in figure 5.2, decanethiol exhibits only a small molecular dipole moment. These two molecules only differ in the length of the nonpolar alkyl backbone, i.e. two units of hydrocarbons for ethanethiol and ten for decanethiol.

For the perfluorinated FC10Thiol (see figure 3.7B), the top-most hydrocarbons are replaced by fully fluorinated units. A remaining amount of two CH2 units are necessary for structural stability

114 Section 6.1: Reference Systems: Thiols

A B C

Au C10Thiol Ag C10Thiol Cu C10Thiol Intensity [arb. u.]

18 17 16 12 10 8 6 4 2 0 4 2 0 Binding energy [eV] D E F

Au FC10Thiol Ag FC10Thiol Cu FC10Thiol Intensity [arb. u.]

18 17 16 12 10 8 6 4 2 0 4 2 0 Binding energy [eV]

Figure 6.1: UPS spectra of C10Thiol (A-C) and FC10Thiol (D-F) on Au, Ag, Cu. The secondary electron cutoffs Esec of C10Thiol (A) and FC10Thiol (D) display the different work function via Φ 21.22eV E and are listed in table 6.1. The valence spectra (B) and (E) are dominated by = − sec the large amount of σ-bonds in the long alkyl chains. These states are localized at 12eV E > B > 4eV. A shift towards higher binding energies upon fluorination is observed. (C) and (F) Due to the long alkyl chains, the attenuation of any interface state near EF cannot be clearly identified. Spectra are vertically offset for clarity. between the thiolate linker group (-SH) and the molecular backbone. As mentioned above, a fluorinated monolayer usually increases the work function of a metal substrate. This effect is attributed to the high electronegativity of fluorine. Since fluorine is usually incorporated in the molecular backbone and head group, a high dipolar moment along the alkyl chain is induced. This dipole moment points towards the metal, i.e. the negative partial charge is located at the head group. Thus, a work function increase is observed (see also figure 2.4C). These results have been accurately reproduced for the assembly of C10Thiol and FC10Thiol on Au, Ag and Cu, as displayed in figure 6.1.

Each molecule generates secondary electron cutoffs that seem to only weakly depend on the metallic substrate, figure 6.1A andD for C 10Thiol and FC10Thiol, respectively. For all metals, C10Thiol lowers the work function while FC10Thiol induces an upward shift. Again, a common final work function is more likely observed rather than a common shift – especially the modifi-

115 Chapter 6: Work Function Tuning with Tailor-Made Dithiocarbamates

C10Thiol Au Ag Cu Work function Φ[eV] 4.1 0.1 3.7 0.1 3.7 0.1 ± ± ± Surface coverage Θ[1014 mol./cm2] 3.0 0.2 4.0 0.2 6.8 0.5 ± ± ± FC10Thiol Work function Φ[eV] 5.2 0.1 5.3 0.1 5.3 0.1 ± ± ± Surface coverage Θ[1014 mol./cm2] 2.1 0.3 3.3 0.2 5.2 0.5 ± ± ± Table 6.1: Work function and surface coverage of reference thiols on Au, Ag, Cu. cations of Ag and Cu show very similar (almost identical) work functions of Φ 3.7eV C10Thiol = and Φ 5.3eV. On gold, similar values of Φ 4.1eV and Φ 5.2eV are FC10Thiol = C10Thiol = C10Thiol = observed in good agreement with literature values [159, 164, 198]. A distinct correlation of the work function values with the molecular surface density on Au, Ag, Cu cannot be identified, even though the coverages on gold are the lowest determined by XPS using the sulfur-to-metal intensity ratio. Altogether, the values listed in table 6.1 are in good agreement with those reported in pre- vious studies [32, 149]. For densely packed monolayer of decanethiol on gold, a (p3 p3)R30◦ × structure with an average area of 21.4 Å2 is expected [32], which equals 4.6 1014 molecules/cm2. × On Ag(111), a slightly distorted (p7 p7)R19◦ overlayer structure is observed with effectively the × same thiol-to-thiol spacing. However, disorder and surface roughening in the present samples pos- sibly reduce the real packing density compared to those investigated on a single crystal surface. On Cu(111), a massive surface reconstruction has been reported, which compensates for the smaller nearest-neighbor distance of copper atoms compared to gold and silver. Thus enables a more densely packed monolayer with a coverage of Θ 5.9 1014 mol./cm2 [157, 160]. In comparison ≈ × to these literature values, the surface coverage of C10Thiol on Cu is only slightly overestimated by the methods used in this thesis. The different work function values observed on gold may also be a result of a different molecular orientation. If the molecule is tilted with an angle φ to the surface normal z, then the effective dipole moment along z is reduced to z cos(φ). This will directly · reduce the resulting work function. In fact, a tilt angle of φ 30◦ for long-chain alkanethiols have ≈ been reported on Au(111), while more upright-standing molecules are observed on both Ag(111) and Cu(111) with φ 12◦ [32, 149, 202]. ≈

In the valence band region 4eV E 12eV, the effect of fluorination on the frontier orbitals can < B < be directly observed in figure 6.1B andE. The UPS spectra of both molecules are dominated by the contribution of σ-type bonds in this region. Upon fluorination, these states are shifted by ∆E 3eV towards higher binding energies. This trend is also observed for the C 1s core ≈ level (section B.3), in which the CF2 and CF3 units produce shifts to higher binding energies (E (CF ) 291.0eV,E (CF ) 293.4eV). The onset of metallic d-states is still weakly observable B 2 = B 3 = in figure 6.1C andF. Yet, an unambiguous metal/molecule interface state as observed for C 2DTC in figure 5.10C cannot be observed. Here, the attenuation of these states due to large alkyl chains

116 Section 6.2: Novel Dithiocarbamates

are comparable to the trend observed for increasing number of methyl units in CnDTC layers, as shown in figure 5.12.

6.2 Novel Dithiocarbamates

In the following, a set of new dithiocarbamates will be introduced and their influence on the work function of the coinage metals discussed. As presented in the previous section, fluorinated thiols are able to push the work function to values up to 5.3 eV, which strongly resembles the work function of pure gold. However, a passivated gold surface is less affected by adsorbates (no

Pauli push-back) and, therefore, more stable. This also holds for the CnDTC layers, which have been presented to produce very low work functions. Now, further monolayer materials based on the DTC ligand are presented with tailored molecular backbones in order to enlarge the field of possible surface modifications that will eventually lead to a surface tailoring of electrode materials. All molecules in this section are illustrated in the chapter figure on page 113.

6.2.1 Odd vs. Even Effect: Dimethyl-Dithiocarbamate

For alkanethiol monolayers, the “odd-even effect” denotes the dif- ferences of molecular orientation found for an odd or even number of methyl units in the thiolate alkyl backbone, respectively. Re- sulting from these structural differences, a variation of electronic transfer rates in molecular junction has been identified [202–204]. The question, if “odd-even” configurations affect the work func- Figure 6.2: Molecular scheme tion has not been addressed yet. Here, one representative of N,N- of C1DTC. dialkyl dithiocarbamates with an odd number of methyl units has been investigated: Dimethyl-Dithiocarbamate (C1DTC). Being the smallest possible CnDTC, it depicts an optimal system for comparison to C2DTC. Both molecular species are equally small and likely occupy the same area per molecule on the surface.

Indeed, all characteristics of C1DTC are in line with those observed for C2DTC on Au, Ag, Cu. From XPS measurements, figure B.5, S 2p and N 1s core level can be found in close resemblance to those of C2DTC. Only the carbon C 1s reflects the fact that the amount of C-C and C-H bonds is reduced, which is especially prominent for C1DTC on Au(111). On Ag(111) and Cu(111), a less distinct splitting as discussed for C2DTC is found, which hinders the identification of the different chemical configurations of carbon. Using again the area ratios of core levels measured with XPS, the molecular surface coverage is estimated to values in the regime of Θ 2.5 1014 mol./cm2, = × as listed in table 6.2. In particular on Cu(111), the molecular density seems slightly increased

117 Chapter 6: Work Function Tuning with Tailor-Made Dithiocarbamates

A B C

Au C1DTC Ag C1DTC Cu C1DTC Intensity [arb. u.]

19 18 17 12 10 8 6 4 2 0 4 2 0 Binding energy [eV]

Figure 6.3: UPS spectra of C1DTC Au, Ag, Cu. (A) The secondary electron cutoffs Esec reveal work functions Φ 21.22eV Esec close to the values found for C2DTC. (B) The valence spectra = − clearly display molecular states at higher binding energies 11eV EB 7eV and d-states of the metal substrates. (C) The frontier metal/molecule interface states> (MMIS)> strongly resemble those of C2DTC (section 5.3) and are even more accentuated. The binding energy of the MMIS are also equal to those of C2DTC. Spectra are vertically offset for clarity. accompanied with a further reduction of the work function down to Φ (3.4 0.1)eV. Cu/C1DTC = ± This supports the assumption, that for an equal surface packing and orientation a common work function should also be observable for DTCs with the smallest alkyl chains. For C1DTC on Au and Ag, comparable work functions values to C2DTC are extracted.

The work functions listed in table 6.2 have been extracted from the secondary electron cutoffs plotted in figure 6.3 (A). The valence structure of C1DTC monolayers on Au, Ag, Cu (B) reveal distinct molecular states located at E 8.0eV and E 9.9eV, which correspond to σ-type bonds. B = B = While three distinct states could be observed for C2DTC (figure 5.10), only two of those can be identified for C1DTC. This also reflects the fact, that there are less σ-bonds due to the reduced number of methyl units. Consequently, this reduced amount of molecular states also improves the intensity of metal d-states and the frontier metal/molecule interface state (C). In principle, these states strongly resemble those of C2DTC: On Au and Cu, the position and onset of the MMIS are almost identical while a slightly shifted MMIS towards higher binding energies is present on Ag (∆E 0.4eV). ≈

Concluding: Does the “odd-even effect” influence the work function? For C1DTC in compar- ison to C2DTC, no significant differences can be identified. The valence region displays com- mon features with only minor changes due to the different amount of σ-bonds. In particular, the metal/molecule interface states are identical in shape and energetic position. Minor changes of the work function in the order of ∆Φ 0.1eV are observed and a slightly higher packing density ≈

118 Section 6.2: Novel Dithiocarbamates

C1DTC Au Ag Cu Work function Φ[eV] 3.3 0.1 3.2 0.1 3.4 0.1 ± ± ± Surface coverage Θ[1014 mol./cm2] 2.7 0.2 2.6 0.2 2.4 0.5 ± ± ± C2DTC Work function Φ[eV] 3.2 0.1 3.1 0.1 3.5 0.1 ± ± ± Surface coverage Θ[1014 mol./cm2] 2.6 0.2 2.1 0.3 1.5 0.4 ± ± ±

Table 6.2: Work function and surface coverage of C1DTC on Au, Ag, Cu. Values for C2DTC for comparison from figure 5.11. on Cu(111) might be attributed to the shorter alkyl substitution and can thus be interpreted as a result of the “odd-even” effect. However, these changes also remain within the usual changes from sample to sample of different monolayer quality.

6.2.2 Aromatic Modification: Pyrrolyl-Dithiocarbamate

In section 5.4.1, the insulating nature of long alkyl chains has been identified in the contact resistance of PTCDI-C13 OFETs. In con- trast, aromatic thiols have been reported to improve charge trans- fer and contact resistances in both n- and p-type organic transis- tors [169, 205, 206]. Thus, the realization of an aromatic surface modification based on dithiocarbamates with a low work function is highly desirable. It is noteworthy, that aromatic molecules are Figure 6.4: Molecular scheme more affected by depolarization than molecules with fully satu- of PyDTC. rated bonds [207]. Therefore, depolarization is able to lower an initially high dipole moment when molecules form an oriented layer [55]. Here, a monolayer based on pyrrole is presented. The pyrrolyl-dithiocarbamate (abbreviated with PyDTC) can be identified as the aromatic counterpart to C2DTC, since it is equally small and exhibits a compara- ble stoichiometry. However, the reaction of pyrrole with CS2 does not take place as spontaneously as observed for usual secondary amines. Pyrrole is much less basic in nature, which reduces the reaction velocity. A common secondary amine like diethyl amine is able to donate and accept an additional hydrogen atom at the nitrogen. This, however, is not the case for pyrrole: Accepting an additional hydrogen requires the formation of a new bond at the cost of the free electron pair of the nitrogen. In the case of pyrrole, this process would destroy the aromatic nature of the molecule and is therefore energetically unfavored. Thus, an additional reactant has to be added to the reaction scheme. Reactions of pyrrole with CS2 have been reported previously, particularly in combination with strong basic compounds like KOH [208–211]. Here the molecule 4-dimethylaminopyridine (DMAP) has been employed as catalyst [212], since it is less aggressive and thus more suited to

119 Chapter 6: Work Function Tuning with Tailor-Made Dithiocarbamates

A B C

Au PyDTC Ag PyDTC Cu PyDTC Intensity [arb. u.]

18 17 12 10 8 6 4 2 0 4 2 0 Binding energy [eV]

Figure 6.5: UPS spectra of PyDTC Au, Ag, Cu. (A) The secondary electron cutoffs Esec produce work functions Φ 21.22eV Esec (4.0 0.1)eV independent on the initial substrate. (B) The valence spectra reveal= a broad− peak≈ generated± by molecular states at higher binding energies 11eV EB 6eV.(C) The frontier metal/molecule interface states (MMIS) are hardly noticeable. Spectra> are> vertically offset for clarity. be employed within a glovebox.

The UPS spectra of PyDTC layers on Au, Ag and Cu are illustrated in figure 6.5. Again, as has been often observed in this thesis, a common secondary electron cutoff and hence a common work function of Φ (4.0 0.1)eV is observed. In addition, the valence region reveals a broad PyDTC = ± peak at 11eV E 6eV for PyDTC on all metal substrates. This broad, but featureless peak > B > might result from the aromatic nature of the molecular backbone. A metal/molecule interface state as observed for CnDTC, however, is only barely noticeable (C). This is actually a surprising fact, since a PyDTC monolayer is supposed to be of equal thickness and density as C2DTC. A damping of these frontier states can be caused by additional residues on top of the PyDTC layer. At first sight, the N 1s, C 1s and S 2p core level examined with XPS reveal no peculiar features (see figure B.6). Yet, a minor amount of oxygen is observed which is caused by the relative low chemical purity of the pyrrole reactant being only 90 %. Nonetheless, oxidized sulfur can not be detected. A realistic possibility is the adsorption of additional pyrrole and/or DMAP molecules on top of the SAM. Both molecules would not be traceable with XPS, since they exhibit the same chemical structure with nitrogen in an aromatic coordination. Upon examination of the surface coverage, table 6.3, a large amount of molecules is calculated with relative large errors due to the differences of the analytical methods. Here, a more sophisticated synthesis would be desired that separates the pyrrolyl-dithiocarbamates from all other possible reactants prior to the immersion of the metal surface. Moreover, an imaging technique like STM would be advantageous to gain more insight into the actual molecular arrangement.

120 Section 6.2: Novel Dithiocarbamates

Nonetheless, the prospect to generate aromatic monolayers based on dithiocarbamates can be envi- sioned. Actually, a stable work function of 4.0 eV would be an excellent value to support electron injection into the fullerene C60 [168, 213]. In the following, the affect of electron withdrawing head groups on the work function of DTC modified metals will be discussed.

PyDTC Au Ag Cu Work function Φ[eV] 4.0 0.1 3.9 0.1 3.9 0.1 ± ± ± Surface coverage Θ[1014 mol./cm2] 3.6 1.2 4.0 1.2 8.0 1.5 ± ± ± Table 6.3: Work function and surface coverage of Pyrrolyl-DTC on Au, Ag, Cu.

6.2.3 Cyanoethyl-Dithiocarbamate

A cyano group is defined by the form R-C N and can be utilized ≡ as a substituent of the methyl head group of N,N-dialkyl dithio- carbamates. In this position, the top nitrogen generates a strong dipole in the opposite direction of the unmodified dithiocarbamate [56, 59, 123]. Hence, such a substitution counteracts the initially strong dipole moment of this molecule that has generated low work functions of 3.5 eV and less. In order to investigate the influence of such a head group, the molecule di(2-cyanoethyl)-dithiocarbamate (abbreviated with CNC2DTC) has been synthesized following the Figure 6.6: Molecular scheme common reaction of a secondary amine, here di(2-cyanoethyl)- of CNC2DTC. amine, with CS2 in an solution of ethanol.

Thereafter, work function values have been extracted from UPS measurements, figure 6.7A, which are listed in table 6.4. In close similarity to the pyrrolyl-dithiocarbamate, a common work func- tion of Φ (4.0 0.1)eV is observed. However, strong depolarization of the molecule in CNC2DTC = ± a monolayer can be excluded due to the absent aromatic backbone. Instead, the strong intrinsic dipole moment in CnDTC molecules, which is generated by the nitrogen in the molecular back- bone, is compensated by the same element located in the head group of CNC2DTC. Therefore, a relatively nonpolar molecule is created when a cyano group is attached to the end of the alkyl chain. In the end, CNC2DTC generates work functions very similar to those of the (also nonpolar) C10Thiol (section 6.1). The valence spectra of CNC2DTC appear very similar on Au, Ag and Cu with only shallow variations due to the different position of the d-states figure 6.7B. In addition, the region close the Fermi edge EF is blurred (C), but the onset of states at roughly 1 eV can be noticed. This might also result from an incorporation of oxygen ( 6..8%) in the layers. Again, a ≈ relative low chemical purity of the amine could have caused this contamination.

121 Chapter 6: Work Function Tuning with Tailor-Made Dithiocarbamates

A B C

Au CNC2DTC Ag CNC2DTC Cu CNC2DTC Intensity [arb. u.]

18 17 12 10 8 6 4 2 0 4 2 0 Binding energy [eV]

Figure 6.7: UPS spectra of CNC2DTC Au, Ag, Cu. (A) The secondary electron cutoffs Esec produce work functions Φ 21.22eV E (4.0 0.1)eV independent on the initial substrate. = − sec ≈ ± (B) The valence spectra reveal molecular states at binding energies 10eV EB 6eV while the metallic d-states are still observable. (C) The frontier metal/molecule interface> > states (MMIS), however, are hardly noticeable with an onset at EB 1.0eV. Spectra are vertically offset for clarity. ≈

CNC2DTC Au Ag Cu Work function Φ[eV] 4.1 0.1 3.9 0.1 4.0 0.1 ± ± ± Surface coverage Θ[1014 mol./cm2] 3.3 1.2 3.3 0.5 18 4 ± ± ±

Table 6.4: Work function and surface coverage of CNC2DTC on Au, Ag, Cu.

From XPS measurements of the core level (figure B.7), a comparable surface coverage to those of C2DTC can be found on Au and Ag, table 6.4. On Cu, however, a massive overestimation is observed. This is the result of an extremely small Cu 2p peak detected with XPS. The origin of this low intensity is not fully understood, since the other core level of the molecule exhibit a reasonable intensity. Copper is known to be more reactive than gold and silver. In addition, copper nanoparticles have been reported to catalyze the reaction of aryl- and vinyl-dithiocarbamates [214]. It is possible that Cu atoms or nanoparticles dissolved in ethanol have initiated the growth of a multilayer structure on the Cu(111) surface instead of a single monolayer. In such a case, the position of the core level remain constant due to similar binding coordinations, e.g via an azo-type link (-N=N-) [215]. Consequently, the intensity of the Cu 2p level is significantly lowered due to the exponential decrease in equation (3.2): I exp( t/λ). Anyhow, this assumption remains to ∝ − be validated. However, if such a scenario is possible with dithiocarbamates, then this could lead to a new route to extreme work functions: By stacking dipolar layers of dithiocarbamates on top of each other work functions far below Φ 3eV might be generated. =

122 Section 6.2: Novel Dithiocarbamates

6.2.4 Fluorinated Dithiocarbamates

At last, a further step is taken towards higher work functions by using a fluorinated dithiocarbamate: Di(1H,1H-heptafluorobutyl)- dithiocarbamate, abbreviated with FC4DTC. This molecule has also been synthesized using the common routine in ethanol with a mixture of Di(1H,1H-heptafluorobutyl)amine with CS2. To be- gin with, this dithiocarbamate has not accurately formed a mono- layer on Au(111), which can be observed after careful examination from the XPS core levels, figure B.8. Even though all expected el- Figure 6.8: Molecular scheme ements are present on the Au(111) surface, the composition of the of FC4DTC.

C 1s core level reveals a striking difference to those of FC4DTC on Ag and Cu as well as compared to FC10Thiol on Au, Ag and Cu: While both CF2 and CF3 configurations are observed at E (CF ) 291.5eV and E (CF ) 293.7eV, respectively, the main B 2 = B 3 = contributions are centered in the region of 284eV E 289eV, which is characteristic for bonds < B < to hydrogen, carbon, nitrogen and sulfur. This implies that much less fluorine terminated carbon atoms are incorporated in the monolayer on gold as expected. This is also supported by a relative low atomic concentration of fluorine (θF,exp31.5% and θF,exp53.8%). Furthermore, a low coverage of DTC molecules estimated by using only the core levels of nitrogen, table 6.5, suggest that a uniform monolayer of fluorinated DTC molecules has not formed on gold. Instead a mixture of full FC4DTC molecules and additionally reacted CS2 should be considered. This has only been observed for the formation on gold and the reaction scheme might be modified to ensure that CS2 has fully reacted with the fluorinated amines before adding the metallic surface. On silver and copper, however, the more reactive metal species may have stimulated this reaction [214].

From this point on, only FC4DTC on Ag and Cu are discussed, since the adsorption on Au clearly needs further improvement. The formation of a closed-packed monolayer on Ag and Cu can be identified from the XPS core level with a reasonable molecular coverage of Θ ≈ 2.4 1014 mol./cm2. It is possible that the higher reactivity of silver and copper also induce a × catalyzing effect on the reaction scheme for this molecule. Relative similar work functions of Φ 4.5eV and Φ 4.6eV are found from UPS measurements illustrated in Ag/CF4DTC = Ag/CF4DTC = figure 6.9 (A). The valence spectra (B) display a DOS, which is in good agreement with the fluori- nated thiol, figure 6.1. Molecular states are shifted by several eV towards higher binding energies compared to the hydrogenated C4DTC counterpart in figure 5.12. In addition to this shift, the metallic d-states and the metal/molecule interface state are clearly visible. In particular, the inter- face state is located at their characteristic position for dithiocarbamate monolayers.

It is interesting to point out the fact, that this modification has effectively not changed the work function of Ag(111). However, with this passivating layer consisting of a fluorinated surface, a

123 Chapter 6: Work Function Tuning with Tailor-Made Dithiocarbamates

A B C

Ag FC4DTC Cu FC4DTC Intensity [arb. u.]

17 16 12 10 8 6 4 2 0 4 2 0 Binding energy [eV]

Figure 6.9: UPS spectra of FC4DTC Ag and Cu. (A) The secondary electron cutoffs Esec produce work functions Φ 21.22eV E (4.6 0.2)eV on Ag and Cu with only a minor difference = − sec ≈ ± of 0.2eV.(B) The valence spectra resemble those of the fluorinated thiol with molecular states ≈ shifted towards higher binding energies of 13eV EB 9eV. The metallic d-states are clearly observable. (C) The frontier metal/molecule interface> > states (MMIS) are relatively clearly ob- servable in good comparison to those of the hydrogenated C4DTC. Spectra are vertically offset for clarity. strong dewetting could be directly observed when the sample has been rinsed with fresh ethanol after the monolayer has been formed. Such a strong dewetting effectively keeps the surface free of unwanted adsorbates, while conserving the work function of 4.5 eV. Weakly physisorbed material can be easily removed by standard cleaning procedures. Hence, this layer acts as a protection layer.

FC4DTC Au Ag Cu Work function Φ[eV] 4.5 0.1 4.6 0.1 − ± ± Surface coverage Θ[1014 mol./cm2] 1.1 0.5 2.2 0.6 2.6 0.6 ± ± ±

Table 6.5: Work function and surface coverage of CNC2DTC on Ag and Cu.

6.3 A New Window of Work Functions for Au, Ag, Cu

In summary, the combination of common thiols with the class of dithiocarbamates enables to achieve a broad range of work functions for the coinage metals Au, Ag, Cu. While thiols, es- pecially aromatic and (partly) fluorinated ones, are useful to induce high work function values [163, 198], dithiocarbamates can be employed to tune the work function below 4 eV. This regime is actually difficult to reach with thiols. Also, modifications based on other linker groups such as

124 Section 6.3: A New Window of Work Functions for Au, Ag, Cu

A B Ag Cu Au Ag Cu Au 5.5 C1DTC 1.0 5.0 C2DTC CnDTC 4.5 0.0 PyDTC FC4DTC

4.0 [eV] [eV] CNC DTC

Φ 2 ∆Φ -1.0 3.5 C10Thiol FC10Thiol 3.0 -2.0 metals 2.5 4.4 4.6 4.8 5.0 5.2 4.4 4.6 4.8 5.0 5.2 Φ Φ metal [eV] metal [eV]

Figure 6.10: A new window of work functions for Au, Ag, Cu based on thiols and dithiocarba- mates. (A) An overview of all measured work functions Φ of this work as a function of the initial metal work function. CnDTC represents the N,N-dialkyl dithiocarbamates with n 4. Addition- ally, the values for the plain, unmodified metals are added as gray circles as reference.≥ A range of 3.2 eV up to 5.4 eV is now available, in which the initial metal species appears to have only minor influence on the resulting work function. Consequently, different work function shifts ∆Φ are observed for each molecular compound on the different metals (B). cyanide (R-NC) have not been reported to remain stable at room temperature [158]. Those reports, that discuss stable monolayers based on DFT calculations only, should be taken with caution. DFT is a ground-state technique, which implies that all calculations presume a temperature at absolute zero Kelvin. Hence, weak bonds, which are stable at in DFT calculation or in experiments at cryogenic temperature, might dissipate at elevated temperatures.

Here, a set of stable dithiocarbamate molecules has been introduced that generates reproducible work functions in a range of 3.2 eV to 4.5 eV. The lowest work function values are produced by the smallest compounds, i.e. C1DTC and C2DTC. Figure 6.10A highlights the results of all work functions that have been achieved in this study. By a successive alteration of the molecular back- bone and head group, the work function could be constantly raised beginning from the lowest value of 3.2 eV (3.4 eV in the case of Cu). This enables the optimization of the charge injec- tion barrier ∆e in electronic devices for different organic semiconductors. Moreover, the use of aromatic dithiocarbamates might even reduce the losses due to the charge transfer through these molecular adsorbents. The fully aromatic pyrrolyl-dithiocarbamate is one representative for such modifications. However, when dealing with aromatic molecules depolarization has always to be considered. This might also be one of the reasons that PyDTC generates “only” a work function of 4.0 eV. The importance of the dipole moment on the work function shift has been demonstrated in the case of CNC2DTC, in which the attachment of a cyano head group effectively compensates the initially large dipole moment of the dithiocarbamate ligand. Altogether, the trend to common

125 Chapter 6: Work Function Tuning with Tailor-Made Dithiocarbamates

work functions seems to persist also for those DTC molecules with substituted head group, which is in line with N,N-dialkyl dithiocarbamate monolayers and the observations made for thiols.

At this point it should be mentioned that the large work function reductions as plotted in fig- ure 6.10B are not a unique property of dithiocarbamate monolayers. E.g., the molecule MV0 has also been reported to generate a work function of 3.3 eV when up to one layer is deposited on gold [138], while the molecule α-sexithiophene has been reported to generate a work function of 3.85 eV on Ag(111) [45]. Yet, such adsorbents remain more reactive to ambient conditions than strongly bound molecules, which form a closed packed monolayer with saturated -CH3 or -CF3 groups at the surface. As has been demonstrated with alkanethiols, such monolayers are able to prevent copper from oxidation [153].

Therefore, dithiocarbamates can be envisioned to not only improve metallic electrodes in hybrid organic/inorganic devices. They may also serve as model system to generate new, effective surface modifications for the class of conductive, transparent metal-oxides. This prospect will be briefly discussed with preliminary results in the next part.

126 Part III

Beyond Metal Surfaces: Modifications of Metal Oxides

Chapter 7

Amino-based SAMs for Metal Oxides

In present light emitting as well as energy conversion devices, the active semiconducting material is sandwiched between two electrodes [216, 217]. The necessity of at least one electrode to be transparent requires conductive metal oxides [218]. Among the large variety of metal oxides, only the compound indium tin oxide (ITO) has been established in industry as transparent electrode, since it excels with the two most crucial physical properties: Electrical conductivity accompanied with a high transparency in the visible light [219]. Hence, minimizing the interface barrier between this electrode and the organic semiconducting layer is one of the key steps towards energy-efficient devices [22]. In line with the discussions in this thesis, control over the work function of ITO is a promising route to achieve low contact resistances. It has been demonstrated for hole transfer using fluorinated derivatives of the phosphonic acid [166, 220–223]. ITO is commonly employed for hole injection and extraction due to its intrinsically high work function of Φ 5eV. In ITO ≈ normal structures of organic solar cells (OSC) and light emitting diodes (OLED), ITO then serves as both the anode and as substrate for the subsequent layers [224]. Consequently, a low function electrode has to be placed on top of the stack for efficient electron transfer into the LUMO of

129 Chapter 7: Amino-based SAMs for Metal Oxides

the semiconducting material. With the use of natural low work function materials such as lithium (Φ 2.2eV), a very reactive surface is created that has to be carefully encapsulated. This defi- Li ≈ ciency defines the need for stable, low work function modifications for ITO or other transparent conductive oxides (TCO). Then, the device stack could be inverted with the low function electrode buried under the full stack and a stable, high work function electrode on top. Previously, only the polymer PEIE (polyethylenimine ethoxylated) has been identified to produce “universal low work function electrodes” [139]. Yet, OLED devices incorporating PEIE have not established so far. Indeed, experiments with PEIE in OLEDs at the institute of GaN Device Technology (RWTH Aachen University) showed an insufficient device stability [225].

In the previous chapters, the versatility of dithiocarbamates to tailor metal work functions has been elucidated. Therein, very low and stable work functions in the range from 3.2 eV to 4.5 eV have been observed. Based on the knowledge that is gained from the interfacial dipole formation of dithiocarbamate monolayers, the next step to efficient electron transfer at interfaces of organic semiconductors to ITO is aimed at in this chapter. A set of potential molecular adsorbates is pre- sented to enable low work functions for ITO. These first results mark the path to a new generation of surface modifiers for metal oxides. The following results have been obtained in a joint study with Julia Rittich and Sebastian Mäder at our research group at the I. Institute of Physics (IA).

7.1 Chemical Structure of ITO

At first, the material ITO is briefly introduced. ITO can be interpreted as tin-doped indium oxide, in which tin occupies cationic sites [226, 227]. This mixture of indium(III) oxide and tin(IV) oxide creates oxygen vacancies in this structure, which distort the original lattice. A typical composition of ITO is denoted as (In2O3)0.9 (SnO2)0.1, which displays the high doping concentration of tin. · ITO coated glass substrates have been ordered from Kintec Company, Hong Kong and cleaned with standard techniques as recommended by the supplier. A measured stoichiometry of such a

film is listed in table 7.1 and compared with the nominal composition of (In2O3)0.9 (SnO2)0.1. · The match between these values seems reasonable, since Kintec does not disclose the true doping concentration. Yet, a higher tin concentration is observed compared to the literature value, which apparently is accompanied by a higher concentration of oxygen vacancies.

O In Sn C measured 49.9 % 42.1 % 3.9 % 4.1 % (In2O3)0.9 (SnO2)0.1 60.4 % 37.5 % 2.1 % 0 % · Table 7.1: Stoichiometry of ITO after UV/ozone treatment.

130 Section 7.1: Chemical Structure of ITO

A

Survey spectrum of ITO In3d C1s

288 284 280 In O1s InMNN 3p

Intensity [arb. u.] Sn Sn Sn OKLL Sn In3s 3p 3d Sn4s Sn4p MNN 3s In In C1s 4s 4p

1000 800 600 400 200 0 Binding Energy in eV B C D

XPS O1s XPS Sn3d XPS In3d 530.3 531.5 532.8 Intensity [arb. u.] Intensity [arb. u.] Intensity [arb. u.]

536 534 532 530 528 526 500 495 490 485 455 450 445 440 Binding energy [eV] Binding Energy [eV] Binding energy [eV]

Figure 7.1: XPS spectra of bare ITO after UV/ozone treatment. (A) The survey spectrum reveals the complex structure of ITO with a set of core levels for each element. Residues of carbon ( 4%) can be found in addition to all expected elements. The annotation MNN and KLL refer to≈ Auger peaks of In, Sn and O. (B-D) High resolution scans of O 1s, Sn 3d and In 3d. (B) The asymmetric O 1s level is dominated by bulk oxygen in a In2O3 coordination at EB 530.2eV (blue line). Oxygen adjacent to a vacancy and surface configurations like In-OH generate= a second peak (green line) at E 531.4eV while interactions to contaminants correspond to E 532.8eV (red B = B = line). (C) Tin 3d doublet splitting. For Sn 3d5/2, a main component is assigned to bulk SnO2 at EB 486.5eV and a smaller contribution of surface Sn-OH at EB 487.3eV.(D) For In 3d5/2, = = the same configurations are observed with bulk In2O3 at EB 444.5eV and surface In-OH at = E 445.0eV. B =

Figure 7.1 illustrates the chemical composition of the ITO surface as measured after chemical cleaning with an UV/ozone treatment as the final step. The survey spectrum of ITO (A) displays a more complex structure compare to those of the noble metals, since a set of core levels is observed for each of the elements. The most prominent core levels that are further investigated are O 1s, Sn 3d and In 3d, figure 7.1(B-D). Peak positions and shapes fit well with respect to previously reported data [228].

Oxygen (B) exhibits an asymmetric O 1s core level at 531 eV. This peak reflects the different binding configurations that are accompanied by the complex crystal structure. The main peak, centered at EB 530.2eV, corresponds to the coordination of In2O3 like oxygen as well as SnO2 =

131 Chapter 7: Amino-based SAMs for Metal Oxides

Low intensity XPS cutoff

Figure 7.2: Work function measurements of ITO at differ- ent sample positions. The low intensity XPS (P 50W) XPS = cutoff is employed to deter- Intensity [arb.u.] mine the work function. Since the intensity is plotted against the kinetic energy EK, the work function directly correlates to 4.0 4.5 5.0 5.5 6.0 6.5 7.0 the onset of photoionization Kinetic Energy [eV] Φ E on 5.1eV. = K ≈ and surface In-O-In configurations [220, 229]. Second, a large contribution is observed at E B = 531.4eV. This peak is assigned to be a convolution of oxygen atoms in the vicinity of a vacancy and hydroxide states at the surface (e.g. In-OH). Oxygen located next to a vacancy is supposed to donate an electron to adjacent indium, which is not fully coordinated due to this unoccupied lattice site. This donation then leaves a positive charge at the oxygen, which increases the binding energy of the remaining electrons. The third peak at E 532.8eV is supposed to correspond B = to bonds to contaminants. As listed in table 7.1, a remaining 4 % of carbon is still detected. It is expected that theses residues are not incorporated within the ITO film, but are more loosely bound via physisorption to the surface and thus may generate some extra oxygen surface states at 532.8 eV.

In the case of the metal species, the 3d orbitals of tin (C) and indium (D) are examined. For both species, the peak reveals only a minor asymmetry with additional contributions towards higher binding energies. Sn 3d5/2 exhibits a main component at EB 486.5eV, which originates from = bulk SnO2, while surface Sn-OH are shifted to higher binding energies at EB 487.3eV. The = same configurations are expected for indium, with In 3d5/2 centered at EB 444.5eV for bulk = In2O3 and a smaller contribution at EB 445.0eV for surface In-OH. [220, 229] =

Within the scope of this thesis, the work function of ITO before and after modification with polar molecules is essential. Yet, the examination of the work function of ITO is a demanding step, as has also been discussed by Schlaf et al. [26]. This metal oxide, similar to the surface reactivity of TiO2, is sensitive to the illumination with UV light, which leads to light induced work function shifts. These shifts in the order of 0.5 eV instantaneously arise by illumination with UV light and further shift to towards lower work function values with prolonged exposure. Electrons are effectively transferred to the surface of ITO, which then reduces their binding energy. Thus, a

132 Section 7.2: Surface Modifications of ITO

modified procedure to extract the work function has been applied with the use of low intensity X-ray irradiation of the XPS source (Al Kα at 50 W). The onset of photoionization can be equally validated with XPS compared to UPS. However, no further information of the valence structure is gained, see figure 7.2. Thereafter, a work function Φ (5.1 0.1)eV is extracted, which ITO = ± matches the literature value of the supplier (Kintec: Φ 4.9eV)[230]. Here, it is noteworthy ITO = that a broad range of ITO work functions have been reported depending on the cleaning procedure [25].

7.2 Surface Modifications of ITO

Now, the final question will be addressed: How effective are molecular adsorbates on ITO to lower the work function? Similar to the weak bond of thiols on ZnO [231], a test with C2DTC on ITO has also shown that the dithiolate (CS2) linker group is ineffective on metal oxides. As mentioned in the beginning of this chapter, surface modifications based on phosphonic acids and esters are well known to strongly chemisorb on metal oxide surfaces [232]. Still, there are other linker groups that are equally attractive to be employed for surface functionalization such as derivates of the carboxylic acid (-COOH), sulfonyl chloride (-SO2Cl) and dichlorophosphonates (-POCl2) [165, 168]. Still, modifications that effectively reduce the work function of ITO remain seldom. While pyridine has been reported to significantly lower the work function of ZnO on the basis of DFT [233], experiments have also revealed the instability of this monolayer at room temperature [234]. It is noteworthy, that theoretical studies are also rare, since the complex structure of the oxide with Sn dopants and O vacancies necessitates a large slab to reproduce the surface with sufficient accuracy [229, 235]. Lately, the focus of theoretical research on the basis of DFT has turned to the binding of molecules on undoped ZnO surfaces [233, 236, 237]. Hence, the following results are mainly based on experimental data.

The key question is, how to transfer the strong intrinsic dipole moment of the DTC moiety to a linker that is suitable for metal oxide surfaces? The first potential candidates are illustrated in figure 7.3, in which the secondary amine is directly bound to the linker group:

(A) N,N-Dimethylphosphoramic dichloride (CH3)2NPOCl2, and (B) N,N-Dimethylsulfamoyl chloride (CH3)2NSO2Cl. Both molecules strongly resemble C1DTC with the difference that the CS2 linker is replaced by sulfonly chloride (A) or phosphonic dichloride (B). Yet, based on the core level observed with XPS, it is assumed that the N-S and N-P bonds of the free molecules are relatively weak and easily break during the process of chemisorption. In particular, the characteristic nitrogen core level is not found after deposition. In any case, an improper stoichiometry is found, in which at least one element is strongly underrepresented or even missing at all. Therefore, it is concluded

133 Chapter 7: Amino-based SAMs for Metal Oxides

Figure 7.3: Molecular structures of first guess ITO modifier, (A) N,N-Dimethylphosphoramic dichloride ((CH3)2NPOCl2) and (B) N,N-Dimethylsulfamoyl chloride ((CH3)2NSO2Cl). (C) Molecular structure of 4-dimethylamino benzoic acid (DMABA, C9H11NO2). In contrast to the molecules (A) and (B), the polar dimethylamino (-N(CH3)2) group is spatially separated from the linker group (-COOH) by a benzyl core. that no stable monolayer has formed, but only fragments of the molecules have adsorbed on the ITO surface. Thus, an additional molecular group is needed to effectively link the amine with the anchoring group. This is the case for the two following molecules: 4-dimethylamino benzoic acid and lysine.

7.2.1 4-Dimethylamino Benczoic Acid

The molecule 4-dimethylamino benzoic acid (C9H11NO2, abbreviated with DMABA) is the first stable compound, that incorporates all of the desired units to achieve a considerable work function reduction of ITO. As illustrated in figure 7.3C, DMABA is based on the carboxylic acid as a linker group (-COOH) and comprises an aromatic core. This combination is generally denoted as “benzoic acid”. In addition, the polar N,N-dimethylamine is placed as the head group on top of the molecular backbone. This head group substitution generates a much stronger dipole moment1 as compared to the plain benzoic acid: p 4.9D and p 1.6D. Therefore, this molecule DMABA = BA = strongly resembles the dithiocarbamate ligand.

In fact, work function measurements performed with the low intensity XPS cutoff highlight a valuable reduction of ∆Φ 1.1eV. Figure 7.4 shows a set of measurements at different positions ≈ of one sample. As can be observed, there are little differences leading to work function ranging from 3.9 eV to 4.1 eV. These differences may result from a local variation of the monolayer quality (molecular density). Nonetheless, a general trend towards a work function of (4.0 0.1)eV can be ± 1Values for the molecular dipole moments p are calculated for the single, neutral molecules using the Gaussian package with the B3LYP method [30] and projected onto the long molecular axis, which is supposed to align parallel to the surface normal.

134 Section 7.2: Surface Modifications of ITO

Low intensity XPS cutoff Figure 7.4: Work function measurements of DMABA on ITO at different sample positions. The low intensity XPS (PXPS 50W) cutoff is employed= to determine

Intensity [arb.u.] the work function, which directly correlates to the onset of photoionization Φ E on (4.0 0.1)eV. Thus, = K ≈ ± 3.0 3.5 4.0 4.5 5.0 5.5 a reduction of ∆Φ 1.1eV is achieved with DMABA= on Kinetic Energy [eV] ITO. observed and confirms the effect of the N,N-dimethylamine head group. A work function of an unsubstituted benzoic acid has been reported to achieve Φ 4.5eV [168]. BA =

Apparently, the molecular coverage on the ITO surface is less than observed for CnDTC on noble metals (table 7.2). This is not entirely surprising since the surface is not as perfect as the highly textured metal (111) planes. Molecular surface densities for derivatives of the phosphonic acids have been reported to vary over a broad range of one order of magnitude from 5 1013 mol./cm2 × to 5 1014 mol./cm2 [220, 222, 223, 232, 235]. In this work, values have been extracted using × different intensity ratios of XPS core level. Since oxygen is incorporated in both the molecular adsorbent and the oxide substrate, this element is unsuitable for this examination. Thus only carbon and nitrogen are employed. Otherwise, both indium and tin can be used as reference levels. This is indeed a useful test to validate this examination. The values for DMABA on ITO are listed in table 7.2 and present reasonable molecular densities in the order of 1 1014 mol./cm2. × The difference by a factor of two for the values extracted from carbon to those from nitrogen most probably result from contaminants on the surface. As listed in table 7.1, a non-vanishing amount of carbon is already present on the surface prior to the adsorption of DMABA. In addition, it cannot be excluded that the molecules may have formed an imperfect monolayer due to the sample preparation. In literature, a variety of different procedures are reported [165, 168, 220, 232]. Thus, further experiments may improve the monolayer quality of DMABA. In comparison with benzylphosphonic acids, a surface coverage of 3 1014 mol./cm2 seems to be a realistic estimation × for a closed-packed monolayer [222, 223].

It is interesting to point out that a common work function of 4.0 eV is observed for similar molecu- lar compounds on entirely different substrates: DMABA on ITO resembles the aromatic pyrrolyl- dithiocarbamate on Au, Ag, Cu (section 6.2.2) as well as diphenyl-dithiocarbamate on Au reported

135 Chapter 7: Amino-based SAMs for Metal Oxides

by Philip Schulz [136]. All these molecules inhibit a strong dipole moment originating from an amine group and all of these compounds are aromatic, which may result in stronger depolarization as compared to fully saturated modifier (also discussed in section 6.2.2 and reference [207]). Here, a detailed investigation including a dipole analysis with DFT techniques is required to generate further insight into the symmetries of these different systems.

Figure 7.5 illustrates the XPS spectra of DMABA on ITO, which have been employed to estimate the surface coverage as listed in table 7.2. Within the survey spectrum (A), all traceable elements are labeled and indicate a molecular passivation of the ITO substrate. Unwanted adsorbates are not directly identified, but cannot be excluded either, since the most prominent impurities would be oxygen and carbon. Thus, high resolution scans of the O 1s, N 1s and C 1s levels have been recorded (B-D). The O 1s level of oxygen has conserved its asymmetric peak shape from the plain

ITO, but small differences are identified. While bulk oxygen in In2O3 and SnO2 coordination still dominates the peak at lower binding energies of E 530.2eV, the contribution of surface InOH B = and new CO-In as well as COH OIn(ITO) generate a peak at E 530.9eV. Compared to O 1s ··· B = in unmodified ITO, this peak is slightly broadened (FWHM 1.9 and FWHM 1.5) and DMABA ≈ ITO ≈ shifted by 0.5eV. These two effects are attributed to the additional states of oxygen incorporated in the linker group of DMABA, which cannot be separated from those of oxygen inside of the ITO based on the current energy resolution. In addition, the fitting routine may also affect the position and width of the peaks in such complex peak shapes, in which many configurations are located closely to each other (roughly within the FWHM of the X-ray source). Finally, a minor peak can be found in analogy to the unmodified surface at E 532.5eV, which most probably contains B = bonds to impurities.

In figure 7.5C, the nitrogen N 1s core level is located at E 400.2eV. Compared to the DTC B = ligand, this peak is slightly shifted by 0.2 eV towards higher binding energies. This shift therefore reflects the different position of nitrogen. In DMABA, the nitrogen atom is not incorporated in an aromatic coordination, but located on top of the aromatic core. Thus a partially lower electron density and, hence, a higher binding energy for the N 1s orbital is observed.

At last, carbon comprises three major configurations (D). At first, the well-known C-H and C-C bonds (partly in an aromatic coordination) generate the peak at E 285.1eV. Second, the amino B = moiety as head group shifts the electron binding energy at carbon in a C-N bond towards higher

4-dimethylamino benzoic acid C/In C/Sn N/In N/Sn Surface coverage Θ[1014 mol./cm2] 2.2 0.2 2.4 0.2 1.0 0.1 1.1 0.1 ± ± ± ± Table 7.2: Surface coverage of DMABA on ITO using the intensity ratio of different molecule to substrate core levels. Since oxygen is incorporated in both ITO and DMABA at different atomic concentration, this species cannot be employed to estimate the molecular surface density.

136 Section 7.2: Surface Modifications of ITO

A

Survey spectrum of DMABA on ITO In 3d C 1s In4s

N1s

O1s 400 300 200 100 InMNN In3p Sn

Intensity [arb. u.] Sn 3d SnMNNOKLL Sn In3s 3p Sn4s Sn4p 3s In In N1s C1s 4s 4p

1000 800 600 400 200 0 Binding Energy in eV B C D

XPS O1s XPS N 1s XPS C 1s Intensity [arb. u.] Intensity [arb. u.] Intensity [arb. u.]

536 534 532 530 528 526 404 402 400 398 396 292 290 288 286 284 282 Binding energy [eV] Binding energy [eV] Binding Energy [eV]

Figure 7.5: XPS spectra of DMABA on ITO. (A) The survey spectrum confirms the presence of all elements of ITO and DMABA. Unwanted residues may be incorporated in the core levels of carbon and oxygen and cannot be excluded at first glance. (B-D) High resolution scans of O 1s, N 1s and C 1s. (B) The asymmetry of the O 1s level of the unmodified ITO is conserved and comprises bulk oxygen in a In2O3 coordination at EB 530.2eV (blue line). Oxygen adjacent to a vacancy and surface configurations like In-OH and= COH OIn(ITO) generate a second peak ··· (green line) at E 530.9eV. This peak is broadened compared to the plain ITO due to the B = additional molecular bonds like CO, COOH. Interactions to contaminants contribute to E B = 532.5eV.(C) The N 1s core level is exclusively incorporated in DMABA and centered at N 1s E 400.2eV.(D) C 1s comprises aromatic C-C bonds at E 285.1eV as well as C-N moieties B = B = at higher binding energies of E 285.9eV. The COOH group is assigned to E 288.8eV. B = B =

137 Chapter 7: Amino-based SAMs for Metal Oxides

binding energies to EB 285.9eV. This trend is in good agreement with CnDTC with short alkyl = chain lengths (figure 5.3 and B.5). The broad peak area and width may indicate additional carbon species, presumably those of impurities. A third, small peak can be identified at E 288.8eV and B = is assigned to the carbon atom within the COOH linker group.

A useful stoichiometry cannot be evaluated due to the impossible separation of oxygen within ITO and DMABA. Yet, the surface densities in table 7.2 indicated that a significant amount of unwanted species are present on the surface due to the different values from carbon and nitrogen. Again, similar to DTC derivatives on copper, nitrogen can be best employed to validate the molecular density. Apparently, the sample preparation and synthesis remain to be optimized. The influence of a more densely packed layer of DMABA on the final work function will be an interesting point for the comparison to the aromatic DTCs on metals.

7.2.2 Lysine

Figure 7.6: Molecular structure of lysine. (A) Neutral configura- tion and (B) twitter ionic repre- sentation, in which the hydrogen of the carboxyl group is moved to the top amino group. The result- ing molecular dipole is desired in both magnitude and direction.

Finally, the last compound is considered, which can be envisioned to considerably lower the work function of ITO: Lysine. In its natural form, denoted as L-lysine and depicted in figure 7.6, this molecule is an essential amino acid [238]. Proteins such as lysine are indispensable for living organisms (including humans) and considered essential when such nutrients cannot be synthe- sized by the organisms themselves [239, 240]. Within the scope of this thesis, lysine can also be envisioned to become an essential building block for efficient energy-conversion electronics. As can be inferred from its full name, 2,6-diaminohexanoic acid comprises of both an amine as head group as well as a carboxyl linker. This configuration already results in a dipole moment of 6.3 D. In contrast to the neutral configuration, figure 7.6A, the twitter ionic representation is more often observed, in which the free proton of the COOH linker group is transferred to one of the amine groups (B). This charge transfer massively increases the dipole moment to 24.1 D.

Indeed, a first publication has reported an enormous work function reduction of ITO using lysine by Deng et al. [241]. In this work, a value of 2.5 eV is reported after the modification with lysine. Yet, the results of this publication seem questionable, since technical mistakes may have been

138 Section 7.2: Surface Modifications of ITO

made that considerably affect the work function measurements. At first, the authors determine the work function with both UPS and Kelvin Probe measurements. The effect of UV illumination has been referred to at the beginning of this chapter (page 132). The authors do not comment on shifts of the secondary electron cutoff due to the illumination with a high intensity UV source. In addition, work function measurements with a Kelvin Probe are also strongly affected by the illumination with UV light, if the sample is sensitive to such a treatment. This is especially present for TiO2 [242]. Therefore, if the Kelvin Probe measurements of ITO in reference [241] have been performed after or during the illumination of UV light, the induced work function shift is also measured with this technique. Here, the authors also do not comment on the exact procedure of their measurements. Second, the authors seem to misinterpret their data by assuming that “the non-washed ITO surface shows a further reduced work function value from 2.5 eV to 1.8 eV (...) as a result of the Lys treatment” (p. 3 of reference [241]). This statement is questionable: If the ITO samples have not been washed after the treatment of lysine, then additional, physisorbed residues of lysine and water (as solvent) are present on the surface and may even form a multilayer structure by van-der-Waals forces due to the high polar moment of lysine. Yet, a layer of non- conductive molecules is very likely to experience a charging due to the intensive emission current of photoelectrons. Such a charging always shifts all electronic states including the secondary electron cutoff towards higher binding energies (i.e. lower kinetic energies). Thus, a lower work function is extracted, but as a result of charging. It is essential to validate, that no charging is present. A common way is to check the position of the Fermi level, since this would also shift as a result of charging. However, the Fermi level of a band gap material like ITO cannot be observed. Thus, such a shift is more difficult to observe2 and care should be taken not to misinterpret such large shifts. In this context, a lysine modification of ITO has been reproduced for this thesis.

Figure 7.7 illustrates the XPS core level of an ITO sample after treatment with lysine. Survey spectrum (A) and core levels of oxygen, nitrogen and carbon (B-D) strongly resemble those of DMABA, figure 7.5, and indicate an adsorption of lysine on ITO as expected. Indeed, only minor peak shifts are observed compared to DMABA. However, employing the area ratios of the C 1s and N 1s core levels with respect to In 3d or Sn 3d reveals a significant difference, table 7.3. Using the carbon to metal area ratio results in a surface coverage of Θ 3 1014 mol./cm2, which is an ≈ × acceptable number for a closed packed monolayer. Yet, with the use of the nitrogen core level, a coverage of only Θ 0.7 1014 mol./cm2 is extracted. Again, impurities are accounted for this ≈ × difference, but they seem to be more significant for lysine compared to DMABA. Most common impurities comprise of hydrocarbons, so that the surface coverage of lysine is more affected by those species when taking the C 1s core level for this calculation.

Based on the dependency of the work function with the molecular surface density, a coverage

2A possible method to validate an uncharged ITO surface can be considered by a careful validation of the core level positions carbon or indium.

139 Chapter 7: Amino-based SAMs for Metal Oxides

A

Survey spectrum of Lysine on ITO In 3d C1s In4s

N1s

In O 400 300 200 100 InMNN 3p 1s

Intensity [arb. u.] Sn Sn OKLL Sn In3s 3p Sn3d Sn4s Sn4p MNN 3s In In N1s C1s 4s 4p

1000 800 600 400 200 0 Binding Energy in eV B C D

XPS O1s XPS N 1s XPS C 1s Intensity [arb. u.] Intensity [arb. u.] Intensity [arb. u.]

536 534 532 530 528 526 404 402 400 398 396 292 290 288 286 284 282 Binding energy [eV] Binding energy [eV] Binding Energy [eV]

Figure 7.7: XPS spectra of Lysine ITO. (A) The survey spectrum confirms the presence of all elements of ITO and DMABA. Unwanted residues may be incorporated in the core levels of carbon and oxygen and cannot be excluded at first glance. (B-D) High resolution scans of O 1s, N 1s and C 1s. (B) The asymmetry of the O 1s level of the unmodified ITO is conserved and comprises bulk oxygen in a In2O3 coordination at EB 530.2eV (blue line). Oxygen adjacent to a vacancy and surface configurations like In-OH, CO= and COOH generate a second peak (green line) at E 530.8eV. Interactions to contaminants and COH OIn(ITO) may contribute to B = ··· E 531.9eV.(C) The N 1s core level is exclusively incorporated in lysine and centered at N 1s B = E 400.3eV.(D) C 1s comprises C-C bonds at E 285.1eV as well as C-N moieties at higher B = B = binding energies of E 285.9eV. The COOH group is assigned to E 288.5eV. B = B =

L-Lysine C/In C/Sn N/In N/Sn Surface coverage Θ[1014 mol./cm2] 2.6 0.3 2.9 0.3 0.6 0.1 0.7 0.1 ± ± ± ± Table 7.3: Surface coverage of Lysine on ITO using the intensity ratio of different molecule to substrate core levels. Since oxygen is incorporated in both ITO and Lysine at different atomic concentration, this species cannot be employed to estimate the molecular surface density.

140 Section 7.2: Surface Modifications of ITO

of less than 1 1014 mol./cm2 seems to be a realistic value, since a large work function shift × as reported by Deng et al. is not observed (even though the authors do not report any surface coverage) [241]. Instead, a (reproducible) work function of (4.1 0.1)eV is found for lysine on ± ITO. While this still implies a considerable work function shift of 1 eV, it can be envisioned that values similar to those of CnDTC on noble metals can be achieved with a more sophisticated sample preparation. Thus, a work function in the range of (3..4) eV might be realistic with a closed-packed monolayer of lysine. Nonetheless, a significant reduction of the work function of ITO can be achieved even with a moderate packing density of lysine. As mentioned in the beginning of this chapter, the lowest work functions reported 3 so far remained limited to values of 4.4 eV [213, 220].

In summary, taking advantage of the same polar unit (amine group) in surface modifications for ITO, the original approach of using dithiocarbamates could successfully be transferred to the class of metal oxides. This finding should mark the way to achieve a tailored work function of metal oxides. As elucidated in chapter 6, the amine group can be employed as polar backbone to reduce the work function and additional electron withdrawing substitutions (cyano, fluorine) may generate small steps back towards higher values. However, one of most important steps will be taking new ways to synthesize these molecules. In particular, stronger linker groups (such as the phosphonic acid) should be linked with the amine group. In addition, different ways to chemically modify an oxide surface have been reported. The optimal procedure for a combination of molecule and oxide has to be tested to ensure a reproducible, densely packed monolayer. At last, the effect of such surface modifier on a broader range of TCO materials will be particularly interesting. Taking these next step will be crucial towards the next generation of energy-efficient optoelectronic devices based on molecular materials.

Low intensity XPS cutoff

Figure 7.8: Work function measurements of lysine on ITO at different sample po-

Intensity [arb.u.] sitions. The low intensity XPS (PXPS 50W) cutoff is employed= to determine the work function with Φ E on (4.1 0.1)eV. Thus, 3.0 3.5 4.0 4.5 5.0 5.5 = K ≈ ± a reduction of ∆Φ 1.0eV is Kinetic Energy [eV] achieved with lysine= on ITO.

3For this discussion, the reported but questionable results from Deng et al. are not considered [241].

141 Chapter 7: Amino-based SAMs for Metal Oxides

142 Chapter 8

Conclusion and Outlook

Optimizing the interface between electrode and organic semiconductor remains one of the crucial steps towards energy efficient electronics for light-emission and energy-conversion based on or- ganic materials. This last chapter presents a brief summary and outlook of the highlights of the interface physics discussed in this thesis. The current work aimed at the characterization of (new) surface modifiers based on the dithiocarbamate (DTC) ligand to effectively tune the work function of electrode materials such as noble metals as well as the conductive metal oxide ITO.

Prior to this thesis, dithiocarbamates have been reported by our research group to produce ex- tremely low work functions of gold down to 3.2 eV [27]. Based on this result, the assembly of N,N-dialkyl dithiocarbamates on the three noble metals Au, Ag and Cu has been analyzed in detail. In this context, the comparison of dithiocarbamates on different metallic substrates enables new insights into the binding chemistry as well as the alignment of energy levels at this interface. In chapter 5, densely packed monolayers have been identified with XPS for dithiocarba- mates on gold and silver. On copper, a mismatch between the nearest-neighbor distance of copper atoms of the Cu(111) plane with the sulfur-sulfur distance within one DTC molecule is observed. DFT simulations reveal that a symmetric binding position, e.g. a top-top binding on gold or a symmetric bridge-bridge position on silver, cannot be established. These simulated results sup- port the experimental data, which revealed an asymmetric S 2p core level. This asymmetric peak shape indicates two different binding configurations for the two sulfur atoms of one molecule. In combination of experimental and theoretical data, DTCs are expected to arrange in an asymmet- ric coordination with one sulfur located on top of a copper atom and the other sulfur atom in a more weakly bound hcp-like binding position. Due to this asymmetry, the incorporation of CS2 molecules within the DTC monolayer is expected on the basis of the XPS measurements. Here, a more sophisticated synthesis is desired for future experiments, in which the incorporation of CS2

143 Chapter 8: Conclusion and Outlook

residues can be avoided. The effect of these residues on the packing density of dithiocarbamates will also be an interesting point to investigate.

The most striking result is the observation of a common work function of (3.5 0.1)eV for almost ± all dithiocarbamates, irrespective of the metal species. Only DTC with the smallest number of methyl units in the alkyl chain (i.e. C2DTC and also supported with the results of C1DTC in section 6.2.1) generate even lower work function values of (3.2 0.1)eV on gold and silver, while ± it remains almost constant at (3.5 0.1)eV for copper. The origin of these work functions and ± the nature of the common work function is an important finding to understand the energy level alignment featuring chemisorption. There are several factors that influence the work function of a modified metal surface:

1. Interface dipole formation: It could be demonstrated using DFT calculations, that the

charge rearrangements upon chemisorption of C2DTC as model material strongly vary with the metal species. This has been obtained from the bond dipole analysis. In addition, dithio- carbamates exhibit a strong intrinsic dipole moment, which generates a large work function shift in combination with the bond dipole. Based on these calculations, a common work function for dithiocarbamates on all three metals can be envisioned in a scheme that the charge rearrangements (i.e. the bond dipole) compensates the different work functions of the different metals. However, further boundaries conditions have to be met as will be dis- cussed in the following.

2. Molecular arrangement: Since molecules in an arranged monolayer represent ordered dipoles, the packing motif of these molecules play an important role for the generation of the final work function [54, 90]. This has also been identified for dithiocarbamates. In the

case of C2DTC, a difference of the packing density is extracted with XPS measurements. For longer alkyl chains, this difference is less distinct. In addition to the molecular cov- erage, the orientation and geometry of molecules have to be taken into account. A depen- dence of orientation and work function can be envisioned for thiols (in section 6.1), but also

the small differences between C1DTC and C2DTC may originate from differently oriented molecules due to the odd-even effect in alkyl-SAMs. Furthermore, depolarization of dipole moments often accompanies monolayer formation, when polar molecules arrange closely to each other [90, 207].

3. Energy level alignment at the interface: The identification of the frontier metal/molecule interface state (MMIS) is a key objective to understand the energy level alignment upon chemisorption. In this work, the combination of experimental with theoretical results gener-

ates insights regarding the fundamental processes of chemisorption. For C2DTC on all three noble metals, the MMIS is identified as a hybridization of molecular S 3p states with those of the frontier metal d-states. Hence, the original HOMO of the DTC molecule (lone pair of

144 Section

the sulfur in the CS2 linker group) forms the MMIS with the noble metal, but located at a fixed distance to the Fermi level. Moreover, the position of this state is independent on the initial work function of the unmodified metal. This is an important finding, since it clearly excludes any kind of vacuum level alignment to control the interface energetics. Likewise, Fermi level pinning seems unrealistic, since the energetic position of the MMIS is too far

away from the Fermi level for all systems. In the case of Fermi level pinning, EF is located closely, i.e. within a few hundred meV to the frontier molecular orbitals [41, 53]. How-

ever, the MMIS in DTCs on Au, Ag, Cu are located roughly 1.5 eV beneath EF. Instead, a new kind of pinning might be considered such as a “bond-induced pinning”, in which the

nature of the bond between the molecular linker group (CS2 in the case of DTCs) and the metal dominates the position of the pinning position. This mechanism also fits well to the

observed bond dipoles of C2DTC: To achieve this pinning position, a different amount of charges have to be rearranged. Indeed, this is observed by comparing the low work function metal silver (lowest bond dipole) to the other species.

Indeed, a systematic comparison of molecular surface modifier on various (noble) metals has generated the most intriguing insight and should be further investigated in the future. In particular, the interplay of Fermi level pinning with a “bond-induced pinning” as promoted in this thesis, can be expected to produce a deeper knowledge of the electronic alignment in the limit of strongly chemisorbing molecules. Consequently, a set of further experiments and simulations have been proposed in section 5.3. Therein, polar molecules that experience Fermi level pinning on gold (as reported in reference [59]) can likewise be investigated on other noble metals. Additionally, pyridine is a suitable candidate for further analysis on noble metals based on the reports of Heimel et al. [53–55].

The versatility of dithiocarbamates to support electron injection in organic semiconductors has been successfully verified with organic thin film transistors based on the n-type material PTCDI-

C13 in section 5.4.1. For an unmodified noble metal, the charge injection barrier is too large for efficient OTFT performance: The energetic difference between Fermi level of the metal EF to the LUMO of PTCDI-C13 is too large to inject electron, as depicted in figure 8.1. This is correlated to the large work function of noble metals, in this case gold. With the use of dithiocarbamate surface modifications, control over the energy level alignment between metal and semiconductor is gained. A large work function shift ∆Φ realigns the vacuum levels of the metal with respect to that of PTCDI-C13. In this case, the charge injection barrier for electrons is significantly reduced, which could be directly validated in the contact resistance of the thin film transistors.

Moreover, the electrode functionalization with dithiocarbamates provides the opportunity to ob- serve the evolution of charge transport properties of PTCDI derivatives in an in-situ setup, which has been created in collaboration with Ingolf Segger in our research group at the I. Institute of

145 Chapter 8: Conclusion and Outlook

ΔΦ

Φ Φ

Figure 8.1: Simplified energy level diagram of metal/molecule interface upon interface modifica- tion with dithiocarbamates. In reference to the PTCDI-C13 OTFTs with DTC modified electrodes in section 5.4, this scheme illustrates the effect of DTC modification. When the molecular semi- conductor PTCDI-C13 is brought into contact with the bare, unmodified gold surface (or any other noble metal), the injection barrier for electrons is too large for efficient charge transfer. Upon modification with a DTC layer, the work function of gold ΦAu is significantly reduced by ∆Φ. Thereafter, an improved interface is generated for electron transfer from the electrode into the LUMO of PTCDI-C13.

Physics (IA). A brief overview of this powerful setup in section 5.4.2 has revealed how the mo- bility µ µ(Θ,T ) is a function of both the layer thickness Θ and the substrate temperature T .A = temperature-dependent mobility µ(T ) is one of the fundamental properties of organic semicon- ductors [62, 69]. However, it is not known how many layers are actually incorporated in charge transport within an OTFT. The results presented here indicate a strong correlation of the growth mode with the number of charge transport layers, which has recently been investigated in extended studies in our research group [183–186].

To achieve an optimal injection barrier for any kind of organic semiconductor, a new set of dithiocarbamates have been synthesized and discussed in section 6. Using different polar head groups attached to the dithiocarbamate ligand, the work function of noble metals could be succes- sively varied from 3.5 eV to 4.5 eV. In addition, conventional thiols can be employed to generate work function up to 5.4 eV. This range is expected to be sufficient to generate optimal charge injection barriers for both electrons and holes in organic electronic devices.

At last, the knowledge gained form the analysis of dithiocarbamate monolayers offers the oppor- tunity to generate new molecules for the surface functionalization of transparent conductive oxides such as ITO, which are commonly incorporated in organic electronic devices as transpar- ent electrode. While the results generated in chapter 7 remained limited to work function values

146 Section

of 4.0 eV, these molecular modifiers mark a first step towards low work function electrodes of ITO (and possibly other TCOs). These results should highlight the “road ahead” for future in- vestigations of metal oxide modifications. In particular, the procedure for sample preparation and synthesis should be improved to achieve densely-packed monolayers on these surfaces. Further molecular modifications should also be considered, to increase the molecular dipole moment and allocate an optimal binding to the oxide surface.

If these next steps can be accomplished, then a new generation of organic electronic devices can be envisioned, in which the layer stack will be reversed. Thereafter, electron injection will be allocated at the transparent TCO electrode and the volatile LiF layer to lower the work function of the reflective (Ag or Al) electrode can be omitted. Instead, the natural, stable work function of noble metals can be employed for hole transfer and encapsulation. Thus, even though Samsung has declared to reduce its investment into new generations of OLED-TVs, a bright future for OLED technology is still expected.

147

Part IV

Appendix

Appendix A

Supporting Information

A.1 Experimental Details

Here, information of sample preparation of substrates, monolayer synthesis and thin film growth of PTCDI-C13 OFETs are listed.

A.1.1 Metal Substrates on Mica

All metal layers for monolayer formation have been produced on freshly cleaved mika substrates. Mika substrates were ordered from Plano GmbH, Germany. Highly purified gold, silver and cop- per have been purchased at MaTeck GmbH, Germany. The preparation procedure for the metal deposition is based on a publication by R. Waser and coworkers [140], which has been adapted and modified in the course of the Master thesis of T. Schäfer [126]. All three noble metals follow the same steps with variations only regarding the substrate temperature and deposition rate. In general, elevated temperatures at all times of the procedure are essential to achieve a highly tex- tured metal layer with a flat surface. Thus, a post annealing process has been performed directly after the deposition. A detailed summary of each step is given in table A.1. A time consuming annealing of the plain mika substrates, as proposed by Waser and Schäfer turned out to be not essential for the quality of the metal films. Hence, this step of initially 6 hours has been reduced to roughly 1 hour. Two factors have been checked prior to the deposition of the metals: First, the substrate has to reach the nominal temperature and second, the pressure should be in the order of 9 1 10− mbar due to degassing of samples and sample holder. ·

I Appendix A: Supporting Information

substrate temperature deposition post annealing

Au(111) 400 ◦C 200 nm at 1 Å/s 4 h at 400 ◦C 150 nm at 30-50 Å/s Ag(111) 350 C 4 h at 350 C ◦ 50 nm at 0.5 Å/s ◦ 150 nm at 30-50 Å/s Cu(111) 400 C 4 h at 400 C ◦ 50 nm at 0.5 Å/s ◦

Table A.1: Deposition procedures of Au, Ag, Cu on mika.

A.1.2 Recipes for SAM Synthesis

In this thesis, a set of thiolate and dithiocarbamate monolayers has been synthesized, some of which have not been reported before. All monolayers on metals have been prepared in a glovebox with a minimum amount of oxygen and water (p 1ppm). Ethanol (V 10ml) has always served ≤ = as the solvent for both thiols and dithiocarbamate layers. For the synthesis of thiolate layers, a 5 mM solution has been used, while the synthesis of dithiocarbamate layers comprises a 2:1 molar ratio of the amine (50 mM) to carbon disulfide (25mM, equivalent to V 15µl). = Mass [mg] Volume [µl]

C10Thiol decanethiol C10H22S 8.7 10.6 FC10Thiol 1H,1H,2H,2H-perfluorodecanethiol C10H5F17S 2.4 14.3

C2DTC diethylamine C4H11N 36.6 52.2 C4DTC dibutylamine C8H19N 64.6 85.0 C6DTC dihexylamine C12H27N 92.7 116.6 C8DTC dioctylamine C16H36N 120.7 151.1 C10DTC didecylamine C20H45N 148.8

C1DTC dimethylamine C2H7N 22.5 33.6 pyrrole C H N 16.7 17.4 PyDTC 4 5 4-(dimethylamino)pyrridine C7H10N2 30.5 CNC2DTC Bis(2-cyanoethyl)amine C6H9N3 61.6 60.4 FC4DTC Bis(1H,1H-heptafluorobutyl)amine C8H5F14N 190.6 118.1

Table A.2: Experimental details of monolayer synthesis for noble metals.

Modifications of ITO with DMABA and lysine have been carried out with different procedures. The results presented in section 7.2 are based the following synthesis steps: The DMABA SAM has been prepared in a 10mM concentration in a solution of ethanol. After one hour at 80 ◦C, the sample has been left in this solution for additional 24 hours. Lysine has been prepared by an immersion in distilled water at 80 ◦C (10 mM) for two hours.

II Section A.2: Electronic Structure: Projected DOS of C2DTC on metals

A.2 Electronic Structure: Projected DOS of C2DTC on metals

A S 3p S 3pz S 3px S 3py Au 5dz2 [states/eV] [states/eV] Au 5dzy sulfur metal PDOS PDOS

7 6 5 4 3 2 1 0 -1 Binding Energy EB [eV]

B S 3p S 3pz S 3px S 3py Ag 4dz2 [states/eV] [states/eV] Ag 4dzy sulfur metal PDOS PDOS

7 6 5 4 3 2 1 0 -1 Binding Energy EB [eV]

C S 3p S 3pz S 3px S 3py Cu 4dz2 [states/eV] [states/eV] Cu 4dzy sulfur metal PDOS PDOS

7 6 5 4 3 2 1 0 -1 Binding Energy EB [eV]

Figure A.1: Projected DOS of C2DTC on (A) Au, (B) Ag, (C) Cu. For all graphs, red and orange colors represent sulfur 3 p states, while black lines correspond to the metal d-states. Only those components are drawn that shows any hybridization with those of sulfur.

III Appendix A: Supporting Information

A.3 Binding sites of Dithiocarbamates on Noble Metals

Figure A.2: Binding sites for DTC molecules on Au, Ag, Cu. (A) On Au(111), the most stable and reproducible coordination is observed for top-top configurations, in which both sulfur atoms occupy a location directly atop of surface gold atoms. (B) On Ag(111), C2DTC generally moves to a symmetric bridge-bridge position with both sulfur atoms located in between two surface silver atoms. (C) On Cu(111), an asymmetric top-hcp binding coordination is observed with DFT and supported with XPS results of the S 2p core level.

A.4P TCDI-C13 OFETs with Modified Copper Electrodes

A B x104 10-7 30 ] m] -1 C8DTC Ω Ω -8 25 10 C10DTC W [

ON 20 10-9 15 10-10 10 -11 10 C8DTC 5 C10DTC Channel conductance [ 10-12 Total resistance R 0 -10 0 10 20 30 40 50 0 50 100 150 200

Gate Voltage VG [V] Channel length L [µm]

Figure A.3: PTCDI-C13 OTFTs with modified Cu electrodes. (A) Transfer curves of OTFTs with C8DTC and C10DTC modified copper electrodes. Devices with CnDTC modifications with shorter alkyl chains (n 8) the device preparation did not work with sufficient reliability: During immersion the copper layers< cleaved from the substrate surface, making it impossible to extract a reliable contact resistance. Here, a modified device architecture might be considered with an adhesion layer for the copper electrodes and a top gate contact. (B) TLM plots for C8DTC and C10DTC modified PTCDI-C13 OTFTs. Values of Rc 2.5MΩcm are observed, similar to those of modified silver electrodes. ≈

IV Section A.5: Molecular Orbitals of Pyridine-based Phenylene Molecules

A.5 Molecular Orbitals of Pyridine-based Phenylene Molecules

π π

π

Figure A.4: Molecular structures and orbitals of pyridine-based phenyl molecules from DFT cal- culations with the GAUSSIAN software package [30]. Isosurfaces ( Ψ 2 0.025) of the HOMO | | (center row) and HOMO-1 (bottom row) help to identify the nature of= these orbitals. The an- notation “N lp” corresponds to the nitrogen lone pair, which is essential for the formation of a chemical bond to a metal surface. “π” labels a delocalized π-type orbital.

V

Appendix B

Supporting XPS Measurements

B.1 Nitrogen on Ag(111)

Ag3d satellite N1s C2DTC Intensity [arb.u.]

404 402 400 398 396 Binding Energy [eV]

Figure B.1: N 1s core level on Ag 3d satellite peak. The filled blue dots represent the N 1s core level of a C2DTC layer on Ag(111) as investigated in this thesis. The same region of clean, unmodified silver sample is depicted with open circles. A broad feature in the range of 400 eV to 398 eV can be observed and represents a satellite peak of the Ag 3d core level. This satellite peak generates a very complex background for the examination of the N 1s core level. Thus, this measurement of the satellite peak has been employed as background subtraction for all DTC layers on Ag(111).

VII Appendix B: Supporting XPS Measurements

B.2 Carbon Disulfide on Cu(111)

A C1s B S2p C O1s Intensity [arb. u.] Intensity [arb. u.] Intensity [arb. u.]

292 288 284 280 170 168 166 164 162 160 540 536 532 528 524 Binding energy [eV] Binding energy [eV] Binding energy [eV]

Figure B.2: XPS core level of CS2 on Cu(111). (A) C 1s level of CS2 revealing different elec- tronic configurations of carbon. A homogeneous layer of CS2 cannot be expected from this peak shape. Rather a mixing with impurities can be expected including C-O bonds (289.2 eV). (B) The sulfur core levels strongly resemble those of CnDTC layers on Cu(111). Thus, the proposed binding scenario of figure 5.7 with an asymmetric configuration of the two sulfur atoms (top/hcp) on Cu(111) is supported. Apparently, CS2 bind in an equal coordination with S 2p3/2 centered at 161.5 eV and 162.3 eV. (C) O 1s core level revealing the large amount of oxygen on the sur- face, which is not observed for CnDTC layers on Cu(111). Surface oxidation as well as ethanol molecules can account for this large peak area.

VIII Section B.3: Reference Systems: Thiols

B.3 Reference Systems: Thiols

B.3.1 Decanethiol on Au, Ag, Cu

Au Ag Cu

A C1s B C1s C C1s Intensity [arb. u.] Intensity [arb. u.] Intensity [arb. u.]

300 296 292 288 284 280 300 296 292 288 284 280 300 296 292 288 284 280 Binding energy [eV] Binding energy [eV] Binding energy [eV]

D S2p E S2p F S2p Intensity [arb. u.] Intensity [arb. u.] Intensity [arb. u.]

170 168 166 164 162 160 170 168 166 164 162 160 170 168 166 164 162 160 Binding energy [eV] Binding energy [eV] Binding energy [eV]

Figure B.3: XPS measurements of C10Thiol on Au, Ag, Cu. (A-C) Carbon C 1s level of C10Thiol on noble metals with one peak centered at 285.2 eV. (D-F) Sulfur S 2p core level in good compar- ison to those observed for DTC layers.

IX Appendix B: Supporting XPS Measurements

B.3.2 1H,1H,2H,2H-Perfluorodecanethiol on Au, Ag, Cu

Au Ag Cu

A C1s B C1s C C1s Intensity [arb. u.] Intensity [arb. u.] Intensity [arb. u.]

300 296 292 288 284 280 300 296 292 288 284 280 300 296 292 288 284 280 Binding energy [eV] Binding energy [eV] Binding energy [eV]

D S2p E S2p F S2p Intensity [arb. u.] Intensity [arb. u.] Intensity [arb. u.]

170 168 166 164 162 160 170 168 166 164 162 160 170 168 166 164 162 160 Binding energy [eV] Binding energy [eV] Binding energy [eV]

G F1s H F1s I F1s Intensity [arb. u.] Intensity [arb. u.] Intensity [arb. u.]

694 692 690 688 686 684 694 692 690 688 686 684 694 692 690 688 686 684 Binding energy [eV] Binding energy [eV] Binding energy [eV]

Figure B.4: XPS measurements of FC10Thiol on Au, Ag, Cu. (A-C) Carbon C 1s level of FC10Thiol on noble metals with a distinct splitting into three components due to the substitution of fluorine atoms. The C-C coordination is only slightly shifted towards higher binding energies at 285.0 eV. CF2 units are located at 291.0 eV and CF3 at 293.5 eV (D-F) Sulfur S 2p core level resemble those observed for the C10Thiol layers. (G-I) Fluorine F 1s core level with a peak center at 688.3 eV. A clear distinction between CF2 and CF3 unit cannot be observed.

X Section B.4: Dimethyl-Dithiocarbamate

B.4 Dimethyl-Dithiocarbamate

Au Ag Cu

A N1s B N1s C N1s Intensity [arb. u.] Intensity [arb. u.] Intensity [arb. u.]

404 402 400 398 396 404 402 400 398 396 404 402 400 398 396 Binding energy [eV] Binding energy [eV] Binding energy [eV]

D C1s E C1s F C1s Intensity [arb. u.] Intensity [arb. u.] Intensity [arb. u.]

292 288 284 292 288 284 292 288 284 Binding energy [eV] Binding energy [eV] Binding energy [eV]

G S2p H S2p I S2p Intensity [arb. u.] Intensity [arb. u.] Intensity [arb. u.]

170 168 166 164 162 160 170 168 166 164 162 160 170 168 166 164 162 160 Binding energy [eV] Binding energy [eV] Binding energy [eV]

Figure B.5: XPS measurements of C1DTC on Au, Ag, Cu. (A-C) N 1s core level of C1DTC at its characteristic binding energy of 400.0 eV. (D-F) Carbon C 1s spectra with the common peaks representing C-H (and CS2) bonds at 285.5 eV and a dominating C-N coordination at 286.2 eV. (G-I) Sulfur S 2p3/2 core level in analogy to those of higher order CnDTC.

XI Appendix B: Supporting XPS Measurements

B.5 Aromatic Modification: Pyrrolyl-Dithiocarbamate

Au Ag Cu

A N1s B N1s C N1s Intensity [arb. u.] Intensity [arb. u.] Intensity [arb. u.]

404 402 400 398 396 404 402 400 398 396 404 402 400 398 396 Binding energy [eV] Binding energy [eV] Binding energy [eV]

D C1s E C1s F C1s Intensity [arb. u.] Intensity [arb. u.] Intensity [arb. u.]

292 288 284 292 288 284 292 288 284 Binding energy [eV] Binding energy [eV] Binding energy [eV]

G S2p H S2p I S2p Intensity [arb. u.] Intensity [arb. u.] Intensity [arb. u.]

170 168 166 164 162 160 170 168 166 164 162 160 170 168 166 164 162 160 Binding energy [eV] Binding energy [eV] Binding energy [eV]

Figure B.6: XPS measurements of PyDTC on Au, Ag, Cu. (A-C) N 1s core level of PyDTC at EB 400.0eV.(D-F) Carbon C 1s spectra with three main peaks representing aromatic C-C (and C-H)= bonds at 285.0 eV, a C-N coordination at 286.3 eV and a partial incorporation of bonds to oxygen (287.8 eV). (G-I) Sulfur S 2p3/2 core level are comparable to those of CnDTC.

XII Section B.6: Cyanoethyl-Dithiocarbamate

B.6 Cyanoethyl-Dithiocarbamate

Au Ag Cu

A N1s B N1s C N1s Intensity [arb. u.] Intensity [arb. u.] Intensity [arb. u.]

404 402 400 398 396 404 402 400 398 396 404 402 400 398 396 Binding energy [eV] Binding energy [eV] Binding energy [eV]

D C1s E C1s F C1s Intensity [arb. u.] Intensity [arb. u.] Intensity [arb. u.]

292 288 284 292 288 284 292 288 284 Binding energy [eV] Binding energy [eV] Binding energy [eV]

G S2p H S2p I S2p Intensity [arb. u.] Intensity [arb. u.] Intensity [arb. u.]

170 168 166 164 162 160 170 168 166 164 162 160 170 168 166 164 162 160 Binding energy [eV] Binding energy [eV] Binding energy [eV]

Figure B.7: XPS measurements of CNC2DTC on Au, Ag, Cu. (A-C) N 1s core level of CNC2DTC at EB 400.0eV.(D-F) Carbon C 1s spectra with different contributions of C-C, C-N and C-O contributions.= Since two different carbon-nitrogen groups are present (the amine and the cyano group), the peak splittings are difficult to separate. The C N and C-C bonds are associated with a low binding energy at 285.0 eV. The common C-N coordination≡ of the amine contributes to 286.6 eV and a partial incorporation of bonds to oxygen are located again at 288.4 eV. (G-I) Sul- fur S 2p3/2 core level are comparable to those of CnDTC, except on Cu(111), where the distinct splitting is not observed.

XIII Appendix B: Supporting XPS Measurements

B.7 Fluorinated Dithiocarbamate

Au Ag Cu

A N1s B N1s C N1s Intensity [arb. u.] Intensity [arb. u.] Intensity [arb. u.]

404 402 400 398 396 404 402 400 398 396 404 402 400 398 396 Binding energy [eV] Binding energy [eV] Binding energy [eV]

D C1s E C1s F C1s Intensity [arb. u.] Intensity [arb. u.] Intensity [arb. u.]

300 296 292 288 284 280 300 296 292 288 284 280 300 296 292 288 284 280 Binding energy [eV] Binding energy [eV] Binding energy [eV]

G S2p H S2p I S2p Intensity [arb. u.] Intensity [arb. u.] Intensity [arb. u.]

170 168 166 164 162 160 170 168 166 164 162 160 170 168 166 164 162 160 Binding energy [eV] Binding energy [eV] Binding energy [eV]

Figure B.8: XPS measurements of FC4DTC on Au, Ag, Cu. (A-C) N 1s core level of FC4DTC at E 400.0eV.(D-F) Carbon C 1s spectra the distinct splitting due to the substitution with B = fluorine, as observed for FC10Thiol, with CF2 at 291.8 eV and CF3 at 294.0 eV. The remaining C-C bonds assigned to the peak at 286.5 eV. On gold and copper, an additional coordination at 284.3.3 eV is observed and assigned to impurities and configurations including non-reacted CS2. (G-I) Sulfur S 2p3/2 core level are comparable to those of CnDTC. Figure continues on the next page.

XIV Section B.7: Fluorinated Dithiocarbamate

Au Ag Cu

J F1s K N1s L F1s Intensity [arb. u.] Intensity [arb. u.] Intensity [arb. u.]

694 692 690 688 686 684 694 692 690 688 686 684 694 692 690 688 686 684 Binding energy [eV] Binding energy [eV] Binding energy [eV]

Figure B.9: XPS measurements of FC4DTC (continued from figure B.8). (J-L) Core level spec- tra of fluorine F 1s with a peak center at 689.0 eV comparable to those observed for FC10Thiol (figure B.4). Again, a distinct peak splitting of CF2 and CF3 cannot be identified.

XV

Bibliography

[1] A. Bernanose, “Electroluminescence in organic compounds,” British Journal of Applied Physics, vol. 6, pp. 54–56, 1955. Cited on page1.

[2] D. Kearns and M. Calvin, “Photovoltaic Effect and Photoconductivity in Laminated Organic Systems,” Journal of Chemical Physics, vol. 4, pp. 950–951, 1958. Cited on page1.

[3] “The Nobel Prize in Chemistry 1961,” 1961. [Online]. Available: http://nobelprize.org/ nobel_prizes/chemistry/laureates/1961/ Cited on page1.

[4] “The Nobel Prize in Physics 1956.” [Online]. Available: www.nobelprize.org/nobelprizes/ chemistry/laureates/1956/ Cited on page1.

[5] “Global Semiconductor Sales Increase in October; Substantial Growth Projected for 2014,” 02. Dec. 2014, The Semiconductor Industry Association (SIA). [Online]. Avail- able: http://www.semiconductors.org/news/2014/12/02/global_sales_report_2014/global_ semiconductor_sales_increase_in_october_substantial_growth_projected_for_2014/ Cited on page1.

[6] C. W. Tang, “Two layer organic photovoltaic cell,” Applied Physics Letters, vol. 48, no. 2, pp. 183–195, 1986. Cited on page1.

[7] “Best Research-Cell Efficiencies,” 12. Aug. 2014, National Institute of Renewable Energy Laboratories (NREL). [Online]. Available: http://www.nrel.gov/ncpv/images/efficiency_ chart.jpg Cited on page1.

[8] “ORGANEXT,” 2014, INTERREG project Nr. EMR. INT4-1.2.-2009-04/054. [Online]. Available: www.organext.org Cited on page1.

[9] “the District of Tomorrow,” 2014, The University of Heerlen. [Online]. Available: www.sustainablebuilding.info/districtoftomorrow Cited on page1.

XVII Bibliography

[10] C. W. Tang and S. A. van Slyke, “Organic electroluminescent diodes,” Applied Physics Letters, vol. 51, no. 12, pp. 913–915, 1987. Cited on page2.

[11] H. Koezuka, A. Tsumura, and T. Ando, “Field-effect transistor with polythiophene thin film,” Synthetic Metals, vol. 18, p. 699, 1987. Cited on page2.

[12] A. Tsumura, H. Koezuka, and T. Ando, “Polythiophene field-effect transistor: Its charac- teristics and operation mechanism,” Synthetic Metals, vol. 25, p. 11, 1988. Cited on page 2.

[13] “The Nobel Prize in Chemistry 2000.” [Online]. Available: www.nobelprize.org/ nobelprizes/chemistry/laureates/2000/ Cited on page2.

[14] “CES 2015 TVs 4k all the way,” 12. Dez. 2014. [Online]. Available: www.cnet.com/uk/ news/ces-2015-tvs-4k-all-the-way/ Cited on page2.

[15] “LG unveils flexible TV panel that can be rolled up to 3 cm,” 10. Jul. 2014. [Online]. Avail- able: www.cnet.com/news/lg-unveils-flexible-tv-panel-that-can-be-rolled-up-to-3cm/ Cited on page2.

[16] “Samsung warps possibilities with user-bendable TV,” 9. Jan. 2014. [Online]. Available: http://www.cnet.com/products/samsung-85u9b/ Cited on page2.

[17] R. Søndergaard, M. Hösel, D. Angmo, T. T. Larsen-Olsen, and F. C. Krebs, “Roll-to-roll fabrication of polymer solar cells,” Materials Today, vol. 15, no. 1-2, pp. 36–49, 2012. [Online]. Available: http://dx.doi.org/10.1016/S1369-7021(12)70019-6 Cited on page2.

[18] M. Hösel, R. R. Søndergaard, M. Jø rgensen, and F. C. Krebs, “Fast Inline Roll-to-Roll Printing for Indium-Tin-Oxide-Free Polymer Solar Cells Using Automatic Registration,” Energy Technology, vol. 1, pp. 102–107, 2013. Cited on page2.

[19] “LG roll out 77 inch flexible OLED TV,” 6. Jan. 2014. [Online]. Available: www.cnet.com/news/lg-rolls-out-77-inch-flexible-oled-tv/ Cited on page2.

[20] T. R. Andersen, H. F. Dam, M. Hösel, M. Helgesen, J. E. Carlé, T. T. Larsen-Olsen, S. A. Gevorgyan, J. W. Andreasen, J. Adams, N. Li, F. Machui, G. D. Spyropoulos, T. Ameri, N. Lemaître, M. Legros, A. Scheel, D. Gaiser, K. Kreul, S. Berny, O. R. Lozman, S. Nord- man, M. Välimäki, M. Vilkman, R. R. Søndergaard, M. Jørgensen, C. J. Brabec, and F. C. Krebs, “Scalable, ambient atmosphere roll-to-roll manufacture of encapsulated large area, flexible organic tandem solar cell modules,” Energy & Environmental Science, vol. 7, no. 9, pp. 2925–2933, 2014. Cited on page2.

XVIII [21] “LG G Flex 2 Release Date 2015: Device May Debut At CES In January With Real ,” 28. Nov. 2014. [Online]. Available: http://www.ibtimes.com/ lg-g-flex-2-release-date-2015-device-may-debut-ces-january-real-flexible-display-1730668 Cited on page2.

[22] W. J. Potscavage, A. Sharma, and B. Kippelen, “Critical Interfaces in Organic Solar Cells and Their Influence on the Open-Circuit Voltage,” Accounts of chemical research, vol. 42, no. 11, pp. 1758–1767, 2009. [Online]. Available: http: //www.ncbi.nlm.nih.gov/pubmed/19708653 Cited on pages2 and 129.

[23] “Samsung says no OLED TVs next year,” 02. Nov. 2014. [Online]. Available: www.cnet.com/news/samsung-says-no-oled-tvs-next-year Cited on page2.

[24] “LG forms new didicated OLED division,” 2014. [Online]. Available: www.cnet.com/ news/lg-forms-new-dedicated-oled-division Cited on page2.

[25] K. Sugiyama, H. Ishii, Y. Ouchi, and K. Seki, “Dependence of indium-tin-oxide work function on surface cleaning method as studied by ultraviolet and x-ray photoemission spectroscopies,” Journal of Applied Physics, vol. 87, no. 1, p. 295, 2000. [Online]. Available: http://scitation.aip.org/content/aip/journal/jap/87/1/10.1063/1.371859 Cited on pages3 and 133.

[26] R. Schlaf, H. Murata, and Z. Kafafi, “Work function measurements on indium tin oxide films,” Journal of Electron Spectroscopy and Related Phenomena, vol. 120, no. 1-3, pp. 149–154, 2001. [Online]. Available: http://linkinghub.elsevier.com/retrieve/pii/ S0368204801003103 Cited on pages3 and 132.

[27] P. Schulz, T. Schäfer, C. D. Zangmeister, C. Effertz, D. Meyer, D. Mokros, R. D. van Zee, R. Mazzarello, and M. Wuttig, “A New Route to Low Resistance Contacts for Performance- Enhanced Organic Electronic Devices,” Advanced Materials Interfaces, vol. 1, 2014. Cited on pages3, 43, 44, 58, 60, 69, 70, 71, 72, 73, 76, 83, 86, 90, 103, and 143.

[28] S. Tatemichi, M. Ichikawa, T. Koyama, and Y. Taniguchi, “High Mobility n-Type Thin-Film Transistors based on N,N’-Ditridecyl Perylene Diimide with Thermal Treatments,” Applied Physics Letters, vol. 89, no. 11, p. 112108, 2006. [Online]. Available: http://scitation.aip.org/content/aip/journal/apl/89/11/10.1063/1.2349290 Cited on pages7, 103, and 111.

[29] R. J. Chesterfield, J. C. Mckeen, C. R. Newman, P. C. Ewbank, D. da Silva Filho, J.-L. Brédas, L. L. Miller, K. R. Mann, and C. D. Frisbie, “Organic Thin Film Transistors Based on N-Alkyl Perylene Diimides: Charge Transport Kinetics as a Function of Gate Voltage

XIX Bibliography

and Temperature,” Journal of Physical Chemistry B, vol. 108, no. 50, pp. 19 281–19 292, 2004. Cited on pages8, 104, and 105.

[30] M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G. A. Petersson, H. Nakatsuji, M. Caricato, X. Li, H. P. Hratchian, A. F. Izmaylov, J. Bloino, G. Zheng, J. L. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, J. A. Montgomery, Jr., J. E. Peralta, F. Ogliaro, M. Bearpark, J. J. Heyd, E. Broth- ers, K. N. Kudin, V. N. Staroverov, R. Kobayashi, J. Normand, K. Raghavachari, A. Ren- , J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, N. Rega, J. M. Millam, M. Klene, J. E. Knox, J. B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, R. L. Martin, K. Mo- rokuma, V. G. Zakrzewski, G. A. Voth, P. Salvador, J. J. Dannenberg, S. Dapprich, A. D. Daniels, O. Farkas, J. B. Foresman, J. V. Ortiz, J. Cioslowski, and D. J. Fox, “Gaussian 09 Revision A.02,” gaussian Inc. Wallingford CT 2009. Cited on pages8, 60, 134, andV.

[31] Y. Zhao, W. Pérez-Segarra, Q. Shi, and A. Wei, “Dithiocarbamate Assembly on Gold,” Journal of the American Chemical Society, vol. 127, no. 20, pp. 7328–7329, 2005. [Online]. Available: http://www.ncbi.nlm.nih.gov/pubmed/20356283 Cited on pages8, 69, and 114.

[32] A. Ulman, “Formation and Structure of Self-Assembled Monolayers.” Chemical Reviews, vol. 96, no. 4, pp. 1533–1554, 1996. Cited on pages8,9, 23, 43, 45, 46, 76, 78, 79, 114, and 116.

[33] S. Vollmer, G. Witte, and C. Wöll, “Structural Analysis of Saturated Alkanethiolate Mono- layers on Cu(100): Coexistence of Thiolate and Sulfide Species,” Langmuir, vol. 17, no. 24, pp. 7560–7565, 2001. Cited on pages9 and 81.

[34] N. Koch, N. Ueno, and A. T. S. Wee, Eds., The Molecule-Metal Interface. 69469 Wein- heim, Germany: WILEY-VCH Verlag GmbH & Co. KGaA, 2013. Cited on pages 10, 16, 17, 19, 20, 27, and 28.

[35] S. Braun, W. R. Salaneck, and M. Fahlman, “Energy-Level Alignment at Organic/Metal and Organic/Organic Interfaces,” Advanced Materials, vol. 21, pp. 1450–1472, 2009. [Online]. Available: http://doi.wiley.com/10.1002/adma.200802893 Cited on pages 10, 15, 17, 19, and 22.

[36] N. D. Lang and W. Kohn, “Theory of Metal Surfaces: Induced Surface Charge and Image Potential,” Physical Review B, vol. 7, no. 6, pp. 3541–3550, 1973. Cited on page 10.

XX [37] H. Ishii, K. Sugiyama, E. Ito, and K. Seki, “Energy Level Alignment and Interfacial Elec- tronic Structures at Organic/Metal and Organic/Organic Interfaces,” Advanced Materials, vol. 11, no. 8, pp. 605–625, 1999. Cited on pages 11, 13, 16, and 38.

[38] Y. C. Chen, J. E. Cunningham, and C. P. Flynn, “Dependence of rare-gas-adsorbate dipole moment on substrate work function,” Physical Review B, vol. 30, no. 12, pp. 7317–7319, 1984. Cited on page 11.

[39] W. Pauli jr., “Über den Zusammenhang des Abschlusses der Elektronengruppen im Atom mit der Komplexstruktur der Spektren,” Zeitschrift für Physik, vol. 31, no. 1, pp. 765–783, 1925. Cited on pages 11, 56, and 57.

[40] G. Witte, S. Lukas, P. S. Bagus, and C. Wöll, “Vacuum level alignment at organic/metal junctions: "Cushion" effect and the interface dipole,” Applied Physics Letters, vol. 87, no. 26, p. 263502, 2005. [Online]. Available: http://scitation.aip.org/content/aip/journal/ apl/87/26/10.1063/1.2151253 Cited on page 13.

[41] M. T. Greiner, M. G. Helander, W.-M. Tang, Z.-B. Wang, J. Qiu, and Z.-H. Lu, “Universal energy-level alignment of molecules on metal oxides,” Nature Materials, vol. 11, no. 1, pp. 76–81, 2012. Cited on pages 17, 20, 98, and 145.

[42] C. Tengstedt, W. Osikowicz, W. R. Salaneck, I. D. Parker, C.-H. Hsu, and M. Fahlman, “Fermi-level pinning at conjugated polymer interfaces,” Applied Physics Letters, vol. 88, no. 5, p. 053502, 2006. [Online]. Available: http://scitation.aip.org/content/aip/journal/apl/ 88/5/10.1063/1.2168515 Cited on page 17.

[43] S. Braun, W. Osikowicz, Y. Wang, and W. R. Salaneck, “Energy level alignment regimes at hybrid organic-organic and inorganic-organic interfaces,” Organic Electronics, vol. 8, no. 1, pp. 14–20, 2007. Cited on page 17.

[44] N. Koch, “Interface and bulk doping of organic semiconductors,” in 10th International Con- ference on Electroluminescence and Organic Optoelectronics, Cologne, 2014. Cited on pages 17 and 20.

[45] P. Amsalem, J. Niederhausen, A. Wilke, G. Heimel, R. Schlesinger, S. Winkler, A. Vollmer, J. Rabe, and N. Koch, “Role of charge transfer, dipole-dipole interactions, and electrostatics in Fermi-level pinning at a molecular heterojunction on a metal surface,” Physical Review B, vol. 87, no. 3, p. 035440, 2013. Cited on pages 18, 20, 96, and 126.

[46] W. L. Kalb and B. Batlogg, “Calculating the trap density of states in organic field-effect transistors from experiment: A comparison of different methods,” Physical Review B, vol. 81, no. 3, p. 035327, 2010. [Online]. Available: http://link.aps.org/doi/10.1103/ PhysRevB.81.035327 Cited on pages 20 and 104.

XXI Bibliography

[47] W. L. Kalb, S. Haas, C. Krellner, T. Mathis, and B. Batlogg, “Trap density of states in small-molecule organic semiconductors: A quantitative comparison of thin-film transistors with single crystals,” Physical Review B, vol. 81, no. 15, p. 155315, 2010. [Online]. Available: http://link.aps.org/doi/10.1103/PhysRevB.81.155315 Cited on pages 20, 27, 28, 104, and 107.

[48] H. Vázquez, F. Flores, and A. Kahn, “Induced Density of States model for weakly- interacting organic semiconductor interfaces,” Organic Electronics, vol. 8, no. 2-3, pp. 241–248, 2007. Cited on pages 21 and 22.

[49] H. Vázquez, Y. J. Dappe, J. Ortega, and F. Flores, “Energy level alignment at metal/organic semiconductor interfaces: "Pillow" effect, induced density of interface states, and charge neutrality level,” The Journal of Chemical Physics, vol. 126, no. 14, p. 144703, 2007. Cited on pages 21 and 22.

[50] F. Schreiber, “Structure and growth of self-assembling monolayers,” Progress in Surface Science, vol. 65, pp. 151–526, 2000. Cited on pages 23, 43, and 114.

[51] G. Heimel, L. Romaner, J.-L. Brédas, and E. Zojer, “Interface Energetics and Level Alignment at Covalent Metal-Molecule Junctions: π-Conjugated Thiols on Gold,” Physical Review Letters, vol. 96, no. 19, p. 196806, 2006. [Online]. Available: http://link.aps.org/doi/10.1103/PhysRevLett.96.196806 Cited on pages 23, 58, 60, 61, and 62.

[52] G. Heimel, L. Romaner, J.-L. Brédas and E. Zojer, “Organic/metal interfaces in self- assembled monolayers of conjugated thiols: A first-principles benchmark study,” Surface Science, vol. 600, no. 19, pp. 4548–4562, 2006. Cited on page 23.

[53] G. Heimel, L. Romaner, E. Zojer, and J.-L. Brédas, “Toward Control of the Metal-Organic Interfacial Electronic Structure in Molecular Electronics: A First-Principles Study on Self- Assembled Monolayers of π-Conjugated Molecules on Noble Metals,” Nano Letters, vol. 7, no. 4, pp. 932–940, 2007. Cited on pages 23, 24, 25, 58, 60, 62, 63, 73, 89, 100, 101, 102, and 145.

[54] G. Heimel, F. Rissner, and E. Zojer, “Modeling the Electronic Properties of π-Conjugated Self-Assembled Monolayers,” Advanced Materials, vol. 22, pp. 2494–2513, 2010. [Online]. Available: http://www.ncbi.nlm.nih.gov/pubmed/20414885 Cited on pages 23, 25, 60, 61, 62, 73, and 144.

[55] F. Rissner, G. M. Rangger, O. T. Hofmann, A. M. Track, G. Heimel, and E. Zojer, “Under- standing the Electronic Structure of Metal/SAM/Organic-Semiconductor Heterojunctions,” ACS Nano, vol. 3, no. 11, pp. 3513–3520, 2009. Cited on pages 25, 119, and 145.

XXII [56] Z. Ma, F. Rissner, L. Wang, G. Heimel, Q. Li, Z. Shuai, and E. Zojer, “Electronic structure of pyridine-based SAMs on flat Au(111) surfaces: extended charge rearrangements and Fermi level pinning,” Physical Chemistry Chemical Physics, vol. 13, no. 20, pp. 9747–60, 2011. [Online]. Available: http://www.ncbi.nlm.nih.gov/pubmed/21503307 Cited on pages 24, 25, 98, and 121.

[57] G. Heimel, E. Zojer, L. Romaner, J.-L. Brédas, and F. Stellacci, “Doping Molecular Wires,” Nano Letters, vol. 9, no. 7, pp. 2559–2564, 2009. Cited on page 25.

[58] D. A. Egger, F. Rissner, G. M. Rangger, O. T. Hofmann, L. Wittwer, G. Heimel, and E. Zo- jer, “Self-assembled monolayers of polar molecules on Au(111) surfaces: distributing the dipoles,” Physical Chemistry Chemical Physics, vol. 12, pp. 4291–4294, 2010. Cited on page 25.

[59] O. T. Hofmann, D. A. Egger, and E. Zojer, “Work-Function Modification beyond Pinning: When Do Molecular Dipoles Count?” Nano Letters, vol. 10, pp. 4369–4374, 2010. [Online]. Available: http://www.ncbi.nlm.nih.gov/pubmed/20939589 Cited on pages 25, 100, 121, and 145.

[60] Z. Bao and J. Locklin, Eds., Organic Field-Effect Transistors. Taylor & Francis Group, LLC, 2007. Cited on pages 26 and 27.

[61] J. C. Scott, “Metal-organic interface and charge injection in organic electronic devices,” Journal of Vacuum Science & Technology A, vol. 21, no. 3, p. 521, 2003. [Online]. Available: http://scitation.aip.org/content/avs/journal/jvsta/21/3/10.1116/1.1559919 Cited on pages 26, 27, and 28.

[62] P. Stallinga, “Electronic Transport in Organic Materials: Comparison of Band Theory with Percolation/(Variable Range) Hopping Theory,” Advanced Materials, vol. 23, no. 30, pp. 3356–62, 2011. [Online]. Available: http://www.ncbi.nlm.nih.gov/pubmed/21671450 Cited on pages 26, 109, and 146.

[63] C. R. Crowell and S. M. Sze, “Current Transport in Metal-Semiconductor Barriers,” Solid- State Electronics, vol. 9, pp. 1035–1048, 1966. Cited on page 27.

[64] V. Coropceanu, J. Cornil, D. da Silva Filho, Y. Olivier, R. Silbey, and J.-L. Brédas, “Charge Transport in Organic Semiconductors,” Chemical Reviews, vol. 107, pp. 926–952, 2007. Cited on pages 29 and 110.

[65] N. Karl, “Charge carrier transport in organic semiconductors,” Synthetic Metals, vol. 134- 134, pp. 649–657, 2003. Cited on page 29.

XXIII Bibliography

[66] C. R. Newman, C. D. Frisbie, A. Demetrio, S. Filho, and J.-L. Brédas, “Introduction to Or- ganic Thin Film Transistors and Design of n-Channel Organic Semiconductors,” Chemistry of Materials, vol. 16, pp. 4436–4451, 2004. Cited on page 29.

[67] G. Horowitz, “Organic Field-Effect Transistors,” Advanced Materials, vol. 10, no. 5, pp. 365–377, 1998. Cited on pages 29 and 30.

[68] G. Horowitz, R. Hajlaoui, H. Bouchriha, R. Bourguiga, and M. Hajlaoui, “The Concept of "Threshold Voltage" in Organic Field-Effect Transistors,” Advanced Materials, vol. 10, no. 12, pp. 923–927, 1998. Cited on page 104.

[69] G. Horowitz, R. Hajlaoui, and P. Delannoy, “Temperature Dependence of the Field-Effect Mobility of Sexithiophene: Determination of the Density of Traps,” Journal de Physique III France, vol. 4, pp. 355–371, 1995. Cited on page 146.

[70] G. Horowitz, P. Lang, M. Mottaghi, and H. Aubin, “Extracting Parameters from the Current-Voltage Characteristics of Organic Field-Effect Transistors,” Advanced Functional Materials, vol. 14, no. 11, pp. 1069–1074, 2004. [Online]. Available: http://doi.wiley.com/10.1002/adfm.200305122 Cited on page 30.

[71] C. H. Kim, Y. Bonnassieux, and G. Horowitz, “Fundamental Benefits of the Staggered Geometry for Organic Field-Effect Transistors,” IEEE Electron Device Letters, vol. 32, no. 9, pp. 1302–1304, 2011. Cited on page 30.

[72] M. Gruber, E. Zojer, F. Schürrer, and K. Zojer, “Impact of Materials versus Geometric Parameters on the Contact Resistance in Organic Thin-Film Transistors,” Advanced Functional Materials, vol. 23, no. 23, pp. 2941–2952, 2013. [Online]. Available: http://doi.wiley.com/10.1002/adfm.201203250 Cited on page 30.

[73] C. H. Kim, Y. Bonnassieux, and G. Horowitz, “Charge Distribution and Contact Resistance Model for Coplanar Organic Field-Effect Transistors,” IEEE Transactions on Electron De- vices, vol. 60, no. 1, pp. 280–287, 2013. Cited on page 30.

[74] A. Einstein, “Über einen die Erzeugung und Verwandlung des Lichtes betreffenden heuris- tischen Gesichtspunkt,” Annalen der Physik, vol. 332, no. 6, pp. 132–148, 1905. Cited on page 33.

[75] M. Cardona and L. Ley, Eds., Photoemission in Solids I: General Principles. Berlin: Springer Verlag, 1978. Cited on pages 33, 36, and 42.

[76] S. Hüfner, Photoelectron Spectroscopy, 2nd ed. Berlin: Springer Verlag, 1995. Cited on pages 33, 34, 36, 37, 42, 93, and 96.

XXIV [77] H. Hertz, “Über einen Einfluss des ultravioletten Lichtes auf die elektrische Entladung,” Annalen der Physik, vol. 267, no. 8, pp. 983–1000, 1887. Cited on page 33.

[78] M. P. Seah and W. A. Dench, “Quantitative Electron Spectroscopy of Surfaces: A Standard Data Base for Electron Inelastic Mean Free Paths in Solids,” Surface and Interface Analysis, vol. 1, no. 1, pp. 2–11, 1979. Cited on pages 34 and 44.

[79] C. N. Berglund and W. E. Spicer, “Photoemission Studies of Copper and Silver: Theory,” Physical Review, vol. 136, no. 4, pp. A1030–A1044, 1964. Cited on pages 34, 35, and 36.

[80] C. N. Berglund and W. E. Spicer, “Photoemission Studies of Copper and Silver: Exper- iment,” Physical Review, vol. 136, no. 4, pp. A1044–A1064, 1964. Cited on pages 34 and 35.

[81] J. Braun, “The theory of angle-resolved ultraviolet photoemission and its applications to ordered materials,” Reports on Progess in Physics, vol. 59, no. 10, pp. 1267–1338, 1996. Cited on page 34.

[82] P. M. Dirac, “The Quantum Theory of the Emission and Absorption of Radiation,” Proceedings of the Royal Society A, vol. 114, no. 767, pp. 243–265, 1927. [Online]. Available: http://rspa.royalsocietypublishing.org/cgi/doi/10.1098/rspa.1927.0039 Cited on page 36.

[83] N. V. Smith, “Photoemission properties of metals,” C R C Critical Reviews in Solid State Sciences, vol. 2, no. 1, pp. 45–83, 1971. Cited on page 36.

[84] R. W. Strayer, W. Mackie, and L. W. Swanson, “Work Function Measurements by the Field Emission Retarding Potential Method,” Surface Science, vol. 34, no. 2, pp. 225–248, 1973. Cited on page 38.

[85] S. Berge, P. O. Gartland, and B. J. Slagsvold, “Photoelectric Work Function of a Molyb- denum Single Crystal for the (100), (110), (111), (112), (114), and (332) Faces,” Surface Science, vol. 43, no. 1, pp. 275–292, 1974. Cited on page 38.

[86] D. Cahen and A. Kahn, “Electron Energetics at Surfaces and Interfaces: Concepts and Experiments,” Advanced Materials, vol. 15, no. 4, pp. 271–277, 2003. Cited on pages 39 and 40.

[87] J. H. Scofield, “Hartree-Slater Subshell Photoionization Cross-Sections at 1254 and 1487 eV,” Journal of Electron Spectroscopy and Related Phenomena, vol. 8, pp. 129–137, 1976. Cited on pages 42, 44, and 74.

XXV Bibliography

[88] A. Ulman, J. E. Eilers, and N. Tillman, “Packing and Molecular Orientation of Alkanethiol Monolayers on Gold Surfaces,” Langmuir, vol. 5, no. 5, pp. 1147–1152, 1989. Cited on pages 43, 46, 92, and 114.

[89] S. D. Evans, E. Urankar, A. Ulman, and N. Ferris, “Self-Assembled Monolayers of Alka- nethiols Containing a Polar Aromatic Group: Effects of the Dipole Position on Molecular Packing, Orientation, and Surface Wetting Properties,” Journal of the American Chemical Society, vol. 113, no. 11, pp. 4121–4131, 1991. Cited on page 114.

[90] L. Romaner, G. Heimel, and E. Zojer, “Electronic Structure of Thiol-Bonded Self- Assembled Monolayers: Impact of Coverage,” Physical Review B, vol. 77, no. 4, p. 045113, 2008. [Online]. Available: http://link.aps.org/doi/10.1103/PhysRevB.77.045113 Cited on pages 43, 73, 91, and 144.

[91] A. L. Bramblett, M. S. Broeckl, K. D. Hauch, B. D. Ratner, S. Tomikazu, and J. W. Rogers Jr., “Determination of surface coverage for tetraphenylporphyrin monolayers using ultravio- let visible absorption and x-ray photoelectron spectroscopies,” Surf. Interface Anal., vol. 33, pp. 206–515, 2002. Cited on pages 43, 45, and 83.

[92] F. von Wrochem, D. Gao, F. Scholz, H.-G. Nothofer, G. Nelles, and J. M. Wessels, “Ef- ficient electronic coupling and improved stability with dithiocarbamate-based molecular junctions,” Nature nanotechnology, vol. 5, no. 8, pp. 618–24, 2010. Cited on pages 43, 69, 70, 75, and 76.

[93] P. Morf, N. Ballav, M. Putero, F. von Wrochem, J. M. Wessels, and T. A. Jung, “Supramolecular Structures and Chirality in Dithiocarbamate Self-Assembled Monolayers on Au(111),” The Journal of Physical Chemistry Letters, vol. 1, no. 5, pp. 813–816, 2010. Cited on pages 44, 71, 76, 86, and 92.

[94] M. P. Seah, I. S. Gilmore, and S. J. Spencer, “Quantitative XPS I . Analysis of X-ray pho- toelectron intensities from elemental data in a digital photoelectron database,” Journal of Electron Spectroscopy and Related Phenomena, vol. 120, pp. 93–111, 2001. Cited on page 44.

[95] A. Jablonski, “Database of Correction Parameters for the Elastic Scattering Effects in XPS,” Surface and Interface Analysis, vol. 23, pp. 29–37, 1995. Cited on page 44.

[96] PHOIBOS Hemispherical Energy Analyzer Series, 3rd ed., SPECS GmbH, 2008. Cited on pages 44, 49, and 50.

[97] M. G. Walter, A. B. Rudine, and C. C. Wamser, “Porphyrins and phthalocyanines in solar photovoltaic cells,” Journal of Porphyrins and Phthalocyanines, vol. 14, pp. 759–792, 2010. Cited on page 45.

XXVI [98] A. Yella, H.-W. Lee, H. N. Tsao, C. Yi, A. K. Chandiran, M. K. Nazeeruddin, E. W.-G. Diau, C.-Y. Yeh, S. M. Zakeeruddin, and M. Grätzel, “Porphyrin-sensitized solar cells with cobalt (II/III)-based redox electrolyte exceed 12 percent efficiency.” Science, vol. 334, pp. 629–34, 2011. Cited on page 45.

[99] PHOIBOS CCD Imaging Detector, 2nd ed., SPECS GmbH, 2009. Cited on pages 49 and 50.

[100] XR50 M – X-ray Source for Focus 500, 1st ed., SPECS GmbH, 2006. Cited on page 49.

[101] XRC 1000 M – X-ray Source Control, 2nd ed., SPECS GmbH, 2009. Cited on page 49.

[102] UVS 300 – Very High Intensity UV Lamp and UVS 300-A Power Supply, 1st ed., SPECS GmbH, 2008. Cited on page 50.

[103] W. Kohn, “Walter Kohn - Nobel Lecture: Electronic Structure of Matter – Wave Functions and Density Functionals,” in Nobel Lecture, Chemistry 1996-2000, I. Grenthe, Ed. World Scientific Publishing Co., Singapore, 2003, pp. 213–236. [Online]. Available: http://www.nobelprize.org/nobel_prizes/chemistry/laureates/1998/kohn-lecture.html Cited on page 51.

[104] E. Schrödinger, “Quantisierung als Eigenwertproblem I,” Annalen der Physik, vol. 384, pp. 361–376, 1926. Cited on page 51.

[105] L. H. Thomas, “The calculation of atomic fields,” Mathematical Proceedings of the Cam- bridge Philosophical Society, vol. 23, no. 05, p. 542, 1927. Cited on page 51.

[106] E. Fermi, “Un metodo statistico per la determinazione di alcune priorieta dell’atome,” Ren- dicondi Accademia Nazionale de Lincei, vol. 6, no. 32, pp. 602–607, 1927. Cited on page 51.

[107] P. Hohenberg and W. Kohn, “Inhomogeneous Electron Gas,” Physical Review, vol. 136, pp. 864–871, 1964. Cited on pages 51 and 54.

[108] R. M. Martin, Electronic Structure, 1rst edition ed. Cambridge University Press, 2004. Cited on pages 52, 53, and 54.

[109] R. Dronskowski, Computational Chemistry of Solid State Materials, 1rst edition ed. WILEY-VCG Verlag GmbH & Co., KGaA, Weinheim, 2005. Cited on page 52.

[110] M. Born and R. Oppenheimer, “Zur quantentheorie der molekeln,” Annalen der Physik, vol. 389, pp. 457–484, 1927. Cited on page 52.

[111] E. U. Condon, “Nuclear Motions Associated with Electron Transitions in Diatomic Molecules,” Physical Review, vol. 32, pp. 858–872, 1928. Cited on page 54.

XXVII Bibliography

[112] W. Kohn and L. J. Sham, “Self-Consistent Equations Including Exchange and Correlation Effects,” Physical Review, vol. 140, pp. 1133–1138, 1965. Cited on pages 55 and 58.

[113] P. Giannozzi, S. Baroni, N. Bonini, M. Calandra, R. Car, C. Cavazzoni, D. Ceresoli, G. L. Chiarotti, M. Cococcioni, I. Dabo, A. Dal Corso, S. de Gironcoli, S. Fabris, G. Fratesi, R. Gebauer, U. Gerstmann, C. Gougoussis, A. Kokalj, M. Lazzeri, L. Martin-Samos, N. Marzari, F. Mauri, R. Mazzarello, S. Paolini, A. Pasquarello, L. Paulatto, C. Sbraccia, S. Scandolo, G. Sclauzero, A. P. Seitsonen, A. Smogunov, P. Umari, and R. M. Wentzcov- itch, “QUANTUM ESPRESSO: A Modular and Open-Source Software Project for Quan- tum Simulations of Materials,” Journal of Physics: Condensed Matter, vol. 21, no. 39, p. 395502, 2009. Cited on pages 56 and 60.

[114] R. Feynman, “Forces in Molecules,” Physical Review, vol. 56, no. 4, pp. 340–343, 1939. [Online]. Available: http://link.aps.org/doi/10.1103/PhysRev.56.340 Cited on page 56.

[115] F. Herman, J. P. Van Dyke, and I. B. Ortenburger, “Improved Statistical Exchange Approx- imation for Inhomogeneous Many-Electron Systems,” Physical Review Letters, vol. 22, no. 16, pp. 807–811, 1969. Cited on page 58.

[116] D. C. Langreth and M. J. Mehl, “Beyond the local-density approximation in calculations of ground-state electronic properties,” Physical Review B, vol. 28, no. 4, pp. 1809–1834, 1983. Cited on page 58.

[117] J. P. Perdew, K. A. Jackson, M. R. Pederson, D. J. Singh, and C. Fiolhais, “Atoms, molecules, solids, and surfaces: Applications of the generalized gradient approximation for exchange and correlation,” Physical Review B, vol. 46, no. 11, pp. 6671–6687, 1992. Cited on page 58.

[118] J. P. Perdew, K. Burke, and M. Ernzerhof, “Generalized Gradient Approximation Made Simple,” Physical Review Letters, vol. 77, no. 18, pp. 3865–3868, 1996. [Online]. Available: http://www.ncbi.nlm.nih.gov/pubmed/10062328 Cited on pages 58 and 64.

[119] N. W. Ashcroft and D. N. Mermin, Solid State Physics. München: Oldenbourg Wis- senschaftsverlag GmbH, 2007. Cited on page 58.

[120] C. Cohen-Tannoudji, B. Diu, and F. Laloë, Quantenmechanik. Berlin: Walter de Gruyter GmbH & Co. KG, 1999. Cited on page 58.

[121] F. London, “Zur Theorie und Systematik der Molekularkräfte,” Zeitschrift für Physik, vol. 63, no. 3-4, pp. 245–279, 1930. Cited on pages 59 and 64.

XXVIII [122] S. Grimme, “Semiempirical GGA-type density functional constructed with a long-range dispersion correction,” Journal of Computational Chemistry, vol. 27, no. 15, pp. 1787– 1799, 2006. Cited on pages 59 and 64.

[123] G. Heimel, L. Romaner, E. Zojer, and J.-L. Brédas, “The Interface Energetics of Self-Assembled Monolayers on Metals,” Accounts of Chemical Research, vol. 41, no. 6, pp. 721–9, 2008. [Online]. Available: http://www.ncbi.nlm.nih.gov/pubmed/18507404 Cited on pages 60, 62, and 121.

[124] A. S. Wan, J. P. Long, G. Kushto, and A. J. Mäkinen, “The interfacial chemistry and energy level structure of a liquid crystalline perylene derivative on Au(111) and graphite surfaces,” Chemical Physics Letters, vol. 463, no. 1-3, pp. 72–77, 2008. Cited on page 60.

[125] M. Smeu, R. A. Wolkow, and G. A. DiLabio, “Theoretical investigation of electron transport modulation through benzenedithiol by substituent groups,” The Journal of Chemical Physics, vol. 129, no. 3, p. 034707, 2008. [Online]. Available: http://www.ncbi.nlm.nih.gov/pubmed/18647038 Cited on page 60.

[126] T. Schäfer, “Diethyl-Dithiocarbamatmonolagen als Oberflächenmodifikation für Edelmet- allelektroden,” Master thesis, RWTH Aachen University, 2012. Cited on pages 60, 62, 64, 65, 69, 72, 74, 81, 92, andI.

[127] L. Wang, G. M. Rangger, Z. Ma, Q. Li, Z. Shuai, E. Zojer, and G. Heimel, “Is there a Au-S bond dipole in self-assembled monolayers on gold?” Physical Chemistry Chemical Physics, vol. 12, pp. 4287–4290, 2010. [Online]. Available: http://www.ncbi.nlm.nih.gov/pubmed/20407694 Cited on pages 62 and 63.

[128] D. Meyer, P. Schulz, T. Schäfer, D. Mokros, F. Martin, C. Effertz, R. Mazzarello, and M. Wuttig, “Interface dipole formation of dithiocarbamate monolayers for organic elec- tronics,” in 11th International Symposium on Functional π-Electron Systems, Arcachon (Bordeaux), 2013. Cited on page 64.

[129] D. Meyer, T. Schäfer, P. Schulz, I. Segger, D. Mokros, J. Felter, F. Martin, C. Effertz, R. Mazzarello, and M. Wuttig, “Towards work function tailoring with dithiocarbamate based self-assembled monolayers,” in 10th International Conference on Electrolumines- cence and Organic Optoelectronics, Cologne, 2014. Cited on page 64.

[130] D. Vanderbilt, “Soft self-consistent pseudopotentials in a generalized eigenvalue formal- ism,” Physical Review B, vol. 41, p. 7892, 1990. Cited on page 64.

[131] H. J. Monkhorst and J. D. Pack, “Special points for Brillouin-zone integration,” Physical Review B, vol. 13, no. 12, pp. 5188–5192, 1976. Cited on page 64.

XXIX Bibliography

[132] M. Methfessel and A. T. Paxton, “High-precision sampling for Brillouin-zone integration in metals,” Physical Review B, vol. 40, no. 6, pp. 3616–3621, 1989. Cited on page 64.

[133] H. Zhu, D. M. Coleman, C. J. Dehen, I. M. Geisler, D. Zemlyanov, J. Chmielewski, G. J. Simpson, and A. Wei, “Assembly of Dithiocarbamate-Anchored Monolayers on Gold Surfaces in Aqueous Solutions,” Langmuir, vol. 24, no. 16, pp. 8660–8666, 2008. [Online]. Available: http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid= 2585038&tool=pmcentrez&rendertype=abstract Cited on page 69.

[134] D. Thorn and R. Ludwig, The Dithiocarbamates and Related Compounds. Amsterdam: Elsevier Publishing Company, 1962. Cited on page 69.

[135] H. Debus, “Ueber die Verbindungen der Sulfocarbaminsäure,” Justus Liebigs Annalen der Chemie, vol. 73, no. 1, pp. 26–34, 1850. Cited on page 69.

[136] A. P. Schulz, “Interface Dipole Formation in Organic Surface Functionalizations,” Disser- tation, RWTH Aachen University, 2012. Cited on pages 69, 73, and 136.

[137] P. Schulz, C. D. Zangmeister, Y.-L. Zhao, P. R. Frail, S. R. Saudari, C. A. Gonzalez, C. R. Kagan, M. Wuttig, and R. D. van Zee, “Comparison of the Energy-Level Alignment of Thiolate- and Carbodithiolate-Bound Self-Assembled Monolayers on Gold,” The Journal of Physical Chemistry C, vol. 114, no. 48, pp. 20 843–20 851, 2010. [Online]. Available: http://pubs.acs.org/doi/abs/10.1021/jp107186m Cited on page 70.

[138] B. Bröker, R.-P. Blum, J. Frisch, A. Vollmer, O. T. Hofmann, R. Rieger, K. Müllen, J. P. Rabe, E. Zojer, and N. Koch, “Gold work function reduction by 2.2 eV with an air-stable molecular donor layer,” Applied Physics Letters, vol. 93, no. 24, p. 243303, 2008. [Online]. Available: http://scitation.aip.org/content/aip/journal/apl/93/24/10.1063/1.3049616 Cited on pages 71 and 126.

[139] Y. Zhou, C. Fuentes-Hernandez, J. Shim, J. Meyer, A. J. Giordano, H. Li, P. Winget, T. Papadopoulos, H. Cheun, J. Kim, M. Fenoll, A. Dindar, W. Haske, E. Najafabadi, T. M. Khan, H. Sojoudi, S. Barlow, S. Graham, J.-L. Brédas, S. R. Marder, A. Kahn, and B. Kippelen, “A Universal Method to Produce Low-Work Function Electrodes for Organic Electronics,” Science, vol. 336, no. 6079, pp. 327–332, 2012. [Online]. Available: http://www.ncbi.nlm.nih.gov/pubmed/22517855 Cited on pages 71, 88, and 130.

[140] B. Lüssem, S. Karthäuser, H. Haselier, and R. Waser, “The origin of faceting of ultraflat gold films epitaxially grown on mica,” Applied Surface Science, vol. 249, pp. 197–202, 2005. Cited on pages 74 andI.

XXX [141] D. Marton, K. J. Boyd, A. H. Al-Bayati, S. S. Todorov, and J. W. Rabalais, “Carbon Ni- tride Deposited Using Energetic Species: A Two-Phase System,” Physical Review Letters, vol. 73, no. 1, pp. 118–121, 1994. Cited on page 74.

[142] W. Zheng, K. Xing, N. Hellgren, I. Ivanov, W. S. Salaneck, and J.-E. Sundgren, “N1s Electron Binding Energies of CNx Thin Films Grown by Magnetron Sputtering at Different Temperature,” Journal of Materials Science and Technology, vol. 14, pp. 25–28, 1998. Cited on page 74.

[143] Y. W. Yang and L. J. Fan, “High-Resolution XPS Study of Decanethiol on Au(111): Single Sulfur-Gold Bonding Interaction,” Langmuir, vol. 18, no. 4, pp. 1157–1164, 2002. Cited on pages 74 and 114.

[144] J. F. Moulder, W. F. Strickle, P. E. Sobol, and K. D. Bomben, “Handbook of x-ray photoelec- tron spectroscopy: A reference book of standard specrta for identification and interpretation of xps data,” 1995. Cited on page 74.

[145] P. Morf, F. Raimondi, H.-G. Nothofer, B. Schnyder, A. Yasuda, J. M. Wessels, and T. A. Jung, “Dithiocarbamates: functional and versatile linkers for the formation of self-assembled monolayers,” Langmuir, vol. 22, no. 2, pp. 658–663, 2006. [Online]. Available: http://www.ncbi.nlm.nih.gov/pubmed/16401114 Cited on pages 75, 76, and 81.

[146] D. Gao, F. Scholz, H.-G. Nothofer, W. E. Ford, U. Scherf, J. M. Wessels, A. Yasuda, and F. von Wrochem, “Fabrication of Asymmetric Molecular Junctions by the Oriented Assembly of Dithiocarbamate Rectifiers,” Journal of the American Chemical Society, vol. 133, no. 15, pp. 5921–5930, 2011. [Online]. Available: http://www.ncbi.nlm.nih.gov/pubmed/21443249 Cited on pages 75 and 76.

[147] W. E. Ford, D. Gao, F. Scholz, G. Nelles, and F. von Wrochem, “Conductance modulation in tetraaniline monolayers by HCl-doping and by field-enhanced dissociation of H2O,” ACS Nano, vol. 7, no. 3, pp. 1943–1951, 2013. [Online]. Available: http://www.ncbi.nlm.nih.gov/pubmed/23421952 Cited on pages 76, 78, and 81.

[148] T. M. Willey, A. L. Vance, T. van Buuren, L. J. Terminello, and C. S. Fadley, “Rapid Degra- dation of Alkanethiol-Based Self-Assembled Monolayers on Gold in Ambient Laboratory Conditions,” Surface Science, vol. 576, no. 1-3, pp. 188–196, 2005. Cited on page 76.

[149] P. E. Laibinis, G. M. Whitesides, D. L. Allara, Y.-T. Tao, A. N. Parikh, and R. G. Nuzzo, “Comparison of the Structures and Wetting Properties of Self-Assembled Monolayers of n- Alkanethiols on the Coinage Metal Surfaces, Cu, Ag, Au,” Journal of the American Chemi- cal Society, vol. 113, no. 19, pp. 7152–7167, 1991. Cited on pages 78, 81, 82, 114, and 116.

XXXI Bibliography

[150] L.-A. Soria, G. Zampieri, and M. L. Martiarena, “Sulfur-Induced Reconstruction of Ag (111) Surfaces Studied by DFT,” The Journal of Physical Chemistry C, vol. 115, pp. 9587– 9592, 2011. Cited on page 79.

[151] H. Sellers, A. Ulman, Y. Shnidman, and J. E. Eilers, “Structure and Binding of Alkanethio- lates on Gold and Silver Surfaces: Implications for Self-Assembled Monolayers,” Journal of the American Chemical Society, vol. 115, no. 8, pp. 9389–9401, 1993. Cited on pages 79 and 114.

[152] A. Cossaro, R. Mazzarello, R. Rousseau, L. Casalis, A. Verdini, A. Kohlmeyer, L. Floreano, S. Scandolo, A. Morgante, M. L. Klein, and G. Scoles, “X-ray diffraction and computation yield the structure of alkanethiols on gold(111),” Science, vol. 321, no. 5891, pp. 943–946, 2008. [Online]. Available: http://www.ncbi.nlm.nih.gov/pubmed/18703737 Cited on page 79.

[153] P. E. Laibinis and G. M. Whitesides, “Self-Assembled Monolayers of n-Alkanethiolates on Copper Are Barrier Films That Protect the Metal against Oxidation by Air,” Journal of the American Chemical Society, vol. 114, no. 23, pp. 9022–9028, 1992. Cited on pages 79, 81, 114, and 126.

[154] G. Henkelman, A. Arnaldsson, and H. Jónsson, “A fast and robust algorithm for Bader decomposition of charge density,” Computational Materials Science, vol. 36, no. 3, pp. 354–360, 2006. [Online]. Available: http://linkinghub.elsevier.com/retrieve/pii/ S0927025605001849 Cited on pages 80 and 82.

[155] T. P. Ang, T. S. A. Wee, and W. S. Chin, “Three-Dimensional Self-Assembled Monolayer (3D SAM) of n-Alkanethiols on Copper Nanoclusters,” Journal of Physical Chemistry B, vol. 108, no. 30, pp. 11 001–11 010, 2004. Cited on page 81.

[156] M. Konôpka, R. Rousseau, I. Štich, and D. Marx, “Detaching Thiolates from Copper and Gold Clusters: Which Bonds to Break?” Journal of the American Chemical Society, vol. 126, no. 38, pp. 12 103–11, 2004. [Online]. Available: http://www.ncbi.nlm.nih.gov/pubmed/15382946 Cited on page 81.

[157] S. M. Driver and D. P. Woodruff, “Adsorption Structures of 1-Octanethiol on Cu(111) Stud- ied by Scanning Tunneling Microscopy,” Langmuir, vol. 16, no. 16, pp. 6693–6700, 2000. Cited on pages 82, 90, and 116.

[158] Y. Sohn and J. White, “Phenyl Isocyanide on Cu(111): Bonding and Interfacial Energy Level Alignment,” Journal of Physical Chemistry C, vol. 111, no. 21, pp. 7816–7825, 2007. [Online]. Available: http://pubs.acs.org/cgi-bin/doilookup/?10.1021/jp070903f Cited on pages 88 and 125.

XXXII [159] C. Schmidt, A. Witt, and G. Witte, “Tailoring the Cu(100) Work Function by Substituted Benzenethiolate Self-Assembled Monolayers,” The Journal of Physical Chemistry A, vol. 115, no. 25, pp. 7234–7241, 2011. Cited on pages 88 and 116.

[160] S. M. Driver and D. P. Woodruff, “Scanning tunnelling microscopy study of the interaction of dimethyl disulphide with Cu(111),” Surface Science, vol. 457, pp. 11–23, 2000. Cited on pages 90 and 116.

[161] Y. Wang, N. S. Hush, and J. R. Reimers, “Understanding the Chemisorption of 2-Methyl-2- propanethiol on Au(111),” Journal of Materials Chemistry C, vol. 111, no. 29, pp. 10 878– 10 885, 2007. Cited on page 93.

[162] R. Paniago, R. Matzdorf, G. Meister, and A. Goldmann, “High-resolution photoemission study of the surface states near Gamma on Cu(111) and Ag(111),” Surface Science, vol. 331-333, pp. 1233–1237, 1995. Cited on pages 93 and 96.

[163] C. D. Zangmeister, L. B. Picraux, R. D. van Zee, Y. Yao, and J. M. Tour, “Energy-Level Alignment and Work Function Shifts for Thiol-bound Monolayers of Conjugated Molecules Self-Assembled on Ag, Cu, Au, and Pt,” Chemical Physics Letters, vol. 442, pp. 390–393, 2007. [Online]. Available: http://linkinghub.elsevier.com/retrieve/pii/S0009261407007427 Cited on pages 99, 100, and 124.

[164] B. de Boer, A. Hadipour, M. M. Mandoc, T. van Woudenbergh, and P. W. M. Blom, “Tuning of Metal Work Functions with Self-Assembled Monolayers,” Advanced Materials, vol. 17, no. 5, pp. 621–625, 2005. [Online]. Available: http://doi.wiley.com/10.1002/adma. 200401216 Cited on pages 103, 114, and 116.

[165] S. Khodabakhsh, B. M. Sanderson, J. Nelson, and T. S. Jones, “Using Self-Assembling Dipole Molecules to Improve Charge Collection in Molecular Solar Cells,” Advanced Func- tional Materials, vol. 16, pp. 95–100, 2006. Cited on pages 133 and 135.

[166] A. Sharma, B. Kippelen, P. J. Hotchkiss, and S. R. Marder, “Stabilization of the work function of indium tin oxide using organic surface modifiers in organic light- emitting diodes,” Applied Physics Letters, vol. 93, no. 16, p. 163308, 2008. [Online]. Available: http://scitation.aip.org/content/aip/journal/apl/93/16/10.1063/1.2998599 Cited on page 129.

[167] F. Li, Y. Zhou, F. Zhang, X. Liu, Y. Zhan, and M. Fahlman, “Tuning Work Function of Noble Metals As Promising Cathodes in Organic Electronic Devices,” Chemistry of Materials, vol. 21, no. 13, pp. 2798–2802, 2009. [Online]. Available: http://pubs.acs.org/doi/abs/10.1021/cm900508t

XXXIII Bibliography

[168] N. Beaumont, I. Hancox, P. Sullivan, R. A. Hatton, and T. S. Jones, “Increased efficiency in small molecule organic photovoltaic cells through electrode modification with self- assembled monolayers,” Energy & Environmental Science, vol. 4, pp. 1708–1711, 2011. Cited on pages 121, 133, and 135.

[169] J. Youn, G. R. Dholakia, H. Huang, J. W. Hennek, A. Facchetti, and T. J. Marks, “Influence of Thiol Self-Assembled Monolayer Processing on Bottom-Contact Thin-Film Transistors Based on n-Type Organic Semiconductors,” Advanced Functional Materials, vol. 22, no. 9, pp. 1856–1869, 2012. [Online]. Available: http://doi.wiley.com/10.1002/adfm.201102312 Cited on pages 103 and 119.

[170] Y. Kanbur, M. Irimia-vladu, E. D. Glowacki, G. Voss, M. Baumgartner, G. Schwabegger, L. Leonat, M. Ullah, H. Sarica, S. Erten-Ela, R. Schwödiauer, H. Sitter, Z. Kücükyavuz, S. Bauer, and N. S. Sariciftici, “Vacuum-processed polyethylene as a dielectric for low operating voltage organic field effect transistors,” Organic Electronics, vol. 13, no. 5, pp. 919–924, 2012. [Online]. Available: http://dx.doi.org/10.1016/j.orgel.2012.02.006 Cited on page 103.

[171] Y. Zhang, B. de Boer, and P. W. M. Blom, “Trap-free electron transport in poly(p-phenylene vinylene) by deactivation of traps with n-type doping,” Physical Review B, vol. 81, no. 8, p. 085201, 2010. [Online]. Available: http://link.aps.org/doi/10.1103/PhysRevB.81.085201 Cited on page 104.

[172] S. Olthof, S. Mehraeen, S. K. Mohapatra, S. Barlow, V. Coropceanu, J.-L. Brédas, S. R. Marder, and A. Kahn, “Ultralow Doping in Organic Semiconductors: Evidence of Trap Filling,” Physical Review Letters, vol. 109, no. 17, p. 176601, 2012. [Online]. Available: http://link.aps.org/doi/10.1103/PhysRevLett.109.176601

[173] S. Olthof, S. Singh, S. K. Mohapatra, S. Barlow, S. R. Marder, B. Kippelen, and A. Kahn, “Passivation of trap states in unpurified and purified C60 and the influence on organic field-effect transistor performance,” Applied Physics Letters, vol. 101, no. 25, p. 253303, 2012. [Online]. Available: http://scitation.aip.org/content/aip/journal/apl/101/25/10.1063/ 1.4772551 Cited on page 104.

[174] J. M. Beebe, V. B. Engelkes, L. L. Miller, and C. D. Frisbie, “Contact Resistance in Metal- Molecule-Metal Junctions Based on Aliphatic SAMs: Effects of Surface Linker and Metal Work Function,” Journal of the American Chemical Society, vol. 124, pp. 11 268–11 269, 2002. Cited on page 104.

[175] L. Venkataraman, J. E. Klare, I. W. Tam, C. Nuckolls, M. S. Hybertsen, and M. L. Steigerwald, “Single-Molecule Circuits with Well-Defined Molecular

XXXIV Conductance,” Nano Letters, vol. 6, no. 3, pp. 458–462, 2006. [Online]. Available: http://www.ncbi.nlm.nih.gov/pubmed/16522042

[176] F. Chen, X. Li, J. Hihath, Z. Huang, and N. Tao, “Effect of Anchoring Groups on Single-Molecule Conductance: Comparative Study of Thiol-, Amine-, and Carboxylic- Acid-Terminated Molecules,” Journal of the American Chemical Society, vol. 128, no. 49, pp. 15 874–81, 2006. [Online]. Available: http://www.ncbi.nlm.nih.gov/pubmed/17147400 Cited on page 104.

[177] C. Effertz, S. Lahme, P. Schulz, I. Segger, M. Wuttig, A. Classen, and C. Bolm, “Design of Novel Dielectric Surface Modifications for Perylene Thin-Film Transistors,” Advanced Functional Materials, vol. 22, no. 2, pp. 415–420, 2012. [Online]. Available: http://doi.wiley.com/10.1002/adfm.201101299 Cited on page 105.

[178] A. Dodabalapur, L. Torsi, and H. E. Katz, “Organic Transistors: Two-Dimensional Trans- port and Improved Electrical Characteristics,” Science, vol. 268, p. 270, 1995. Cited on page 105.

[179] E. L. Granstrom and C. D. Frisbie, “Field Effect Conductance Measurements on Thin Crys- tals of Sexithiophene,” Journal of Physical Chemistry B, vol. 103, pp. 8842–8849, 1999.

[180] F. J. Meyer zu Heringdorf, M. C. Reuter, and R. M. Tromp, “Growth dynamics of pentacene thin films,” Nature, vol. 412, pp. 517–20, Aug. 2001. [Online]. Available: http://www.ncbi.nlm.nih.gov/pubmed/17025827

[181] F. Dinelli, M. Murgia, P. Levy, M. Cavallini, F. Biscarini, and D. M. de Leeuw, “Spatially Correlated Charge Transport in Organic Thin Film Transistors,” Physical Review Letters, vol. 92, no. 11, p. 116802, 2004.

[182] R. Ruiz, A. Papadimitratos, A. C. Mayer, and G. G. Malliaras, “Thickness Dependence of Mobility in Pentacene Thin-Film Transistors,” Advanced Materials, vol. 17, pp. 1795–1798, 2005. Cited on page 105.

[183] A. Hessler, “Real-Time Observation of Charge Transport in OTFTs During Growth,” Bach- elor thesis, RWTH Aachen University, 2014. Cited on pages 105 and 146.

[184] E. Hofmann, “In-situ Characterization of Charge Transport in PTCDI Thin Film Transis- tors,” Bachelor thesis, RWTH Aachen University, 2014. Cited on pages 105 and 111.

[185] A. Leis, “In-situ measurements of charge transport and its dependence on substrate temper-

ature in PTCDI-C13-OTFTs,” Bachelor thesis, RWTH Aachen University, 2014. Cited on pages 105, 107, 108, and 109.

XXXV Bibliography

[186] I. Segger, “Playground Stories,” Dissertation, RWTH Aachen University, 2015. Cited on pages 105, 107, 108, 109, 110, 111, and 146.

[187] A. Shehu, S. D. Quiroga, P. D’Angelo, C. Albonetti, F. Borgatti, M. Murgia, A. Scorzoni, P. Stoliar, and F. Biscarini, “Layered Distribution of Charge Carriers in Organic Thin Film Transistors,” Physical Review Letters, vol. 104, no. 24, p. 246602, Jun. 2010. [Online]. Available: http://link.aps.org/doi/10.1103/PhysRevLett.104.246602 Cited on pages 106 and 108.

[188] F. Liscio, C. Albonetti, K. Broch, A. Shehu, S. D. Quiroga, L. Ferlauto, C. Frank, S. Kowarik, R. Nervo, A. Gerlach, S. Milita, F. Schreiber, and F. Biscarini, “Molecular Reorganization in Organic Field-Effect Transistors and Its Effect on Two-Dimensional Charge Transport Pathways,” ACS nano, vol. 7, no. 2, pp. 1257–64, Feb. 2013. [Online]. Available: http://www.ncbi.nlm.nih.gov/pubmed/23350706 Cited on page 106.

[189] S. Kirkpatrick, “Percolation and Conduction,” Reviews of Modern Physics, vol. 45, no. 4, pp. 574–588, 1973. Cited on page 106.

[190] M. Sahimi, “On the relationship between the critical exponents of percolation conductivity and static exponents of percolation,” Journal of Physics A: Mathematical and General, vol. 17, pp. L601–L607, 1984. Cited on page 106.

[191] J. Felter, “Growth Study of Organic Thin Films,” Bachelor thesis, RWTH Aachen Univer- sity, 2012. Cited on page 108.

[192] J. A. Nichols, D. J. Gundlach, and T. N. Jackson, “Potential imaging of pentacene organic thin-film transistors,” Applied Physics Letters, vol. 83, no. 12, p. 2366, 2003. [Online]. Available: http://scitation.aip.org/content/aip/journal/apl/83/12/10.1063/1.1611278 Cited on page 108.

[193] T. K. Maiti, T. Hayashi, L. Chen, H. Mori, M. J. Kang, K. Takimiya, and H. J. Mattausch, “A Surface Potential Based Organic Thin-Film Transistor Model for Circuit Simulation Veri- fied With DNTT High Performance Test Devices,” IEEE Transactions on Semiconductor Manufacturing, vol. 27, no. 2, pp. 159–168, 2014. Cited on page 108.

[194] J. Lu, E. Delamarche, L. Eng, R. Bennewitz, E. Meyer, and H.-J. Güntherodt, “Kelvin Probe Force Microscopy on Surfaces : Investigation of the Surface Potential of Self-Assembled Monolayers on Gold,” Langmuir, vol. 15, pp. 8184–8188, 1999. Cited on page 108.

[195] K. P. Puntambekar, P. V. Pesavento, and C. D. Frisbie, “Surface potential profiling and contact resistance measurements on operating pentacene thin-film transistors by Kelvin probe force microscopy,” Applied Physics Letters, vol. 83, no. 26, p. 5539, 2003. [Online]. Available: http://scitation.aip.org/content/aip/journal/apl/83/26/10.1063/1.1637443

XXXVI [196] V. Panchal, R. Pearce, R. Yakimova, A. Tzalenchuk, and O. Kazakova, “Standardization of surface potential measurements of graphene domains,” Scientific Reports, vol. 3, no. 2, p. 2597, Jan. 2013. [Online]. Available: http://www.pubmedcentral.nih.gov/articlerender. fcgi?artid=3764438&tool=pmcentrez&rendertype=abstract Cited on page 108.

[197] R. Schmidt, J. H. Oh, Y.-S. Sun, M. Deppisch, A.-M. Krause, K. Radacki, H. Braunschweig, M. Könemann, P. Erk, Z. Bao, and F. Würtner, “High-Performance Air-Stable n-Channel Organic Thin Film Transistors Based on Halogenated Perylene Bisimide Semiconductors,” Journal of the American Chemical Society, vol. 131, pp. 6215–6228, 2009. Cited on page 114.

[198] D. M. Alloway, A. L. Graham, X. Yang, A. Mudalige, R. Colorado, V. H. Wysocki, J. E. Pemberton, T. R. Lee, R. J. Wysocki, and N. R. Armstrong, “Tuning the Effective Work Function of Gold and Silver Using ω-Functionalized Alkanethiols: Varying Surface Com- position through Dilution and Choice of Terminal Groups,” Journal of Physical Chemistry C, vol. 113, no. 47, pp. 20 328–20 334, 2009. Cited on pages 114, 116, and 124.

[199] R. Mazzarello, A. Cossaro, A. Verdini, R. Rousseau, L. Casalis, M. F. Danisman, L. Flore- ano, S. Scandolo, A. Morgante, and G. Scoles, “Structure of a CH3S Monolayer on Au(111) Solved by the Interplay between Molecular Dynamics Calculations and Diffraction Mea- surements,” Physical Review Letters, vol. 98, p. 016102, 2007. Cited on page 114.

[200] H. Häkkinen, “The gold-sulfur interface at the nanoscale,” Nature chemistry, vol. 4, no. 6, pp. 443–55, 2012. [Online]. Available: http://www.ncbi.nlm.nih.gov/pubmed/22614378 Cited on page 114.

[201] Y. Kuzumoto and M. Kitamura, “Work function of gold surfaces modified using substituted benzenethiols: Reaction time dependence and thermal stability,” Applied Physics Express, vol. 7, no. 3, p. 035701, 2014. [Online]. Available: http://stacks.iop.org/1882-0786/7/i=3/ a=035701?key=crossref.b49288c245c8cf5f598ccfcc32741aa9 Cited on page 114.

[202] M. M. Walczak, C. Chung, S. M. Stole, C. A. Widrig, and M. D. Porter, “Structure and Interfacial Properties of Spontaneously Adsorbed n-Alkanethiolate Monolayers on Evap- orated Silver Surfaces,” Journal of the American Chemical Society, vol. 113, no. 7, pp. 2370–2378, 1991. Cited on pages 116 and 117.

[203] J. F. Smalley, S. W. Feldberg, C. E. D. Chidsey, M. R. Linford, M. D. Newton, and L. Yi- ping, “The Kinetics of Electron Transfer through Ferrocene-Terminated Alkanethiol Mono- layers on Gold,” Journal of Physical Chemistry, vol. 99, no. 35, pp. 13 141–13 149, 1995.

XXXVII Bibliography

[204] J. J. Sumner, K. S. Weber, L. A. Hockett, and S. E. Creager, “Long-Range Heteroge- neous Electron Transfer Between Ferrocene and Gold Mediated By n-Alkane and N-Alkyl- Carboxamide Bridges,” Journal of Physical Chemistry B, vol. 104, no. 31, pp. 7449–7454, 2000. Cited on page 117.

[205] P. Marmont, N. Battaglini, P. Lang, G. Horowitz, J. Hwang, A. Kahn, C. Amato, and P. Calas, “Improving charge injection in organic thin-film transistors with thiol-based self- assembled monolayers,” Organic Electronics, vol. 9, pp. 419–424, 2008. Cited on page 119.

[206] C. Bock, D. V. Pham, U. Kunze, D. Käfer, G. Witte, and A. Terfort, “Influence of anthracene-2-thiol treatment on the device parameters of pentacene bottom-contact tran- sistors,” Applied Physics Letters, vol. 91, p. 052110, 2012. Cited on page 119.

[207] D. Cornil, Y. Olivier, V. Geskin, and J. Cornil, “Depolarization Effects in Self-Assembled Monolayers: A Quantum-Chemical Insight,” Advanced Functional Materials, vol. 17, pp. 1143–1148, 2007. Cited on pages 119, 136, and 144.

[208] R. D. Bereman and D. Nalewajek, “Preparation and Characterization of Pyrrole-N- carbodithioate Complexes of Selected Transistion Elements,” Inorganic Chemistry, vol. 16, no. 11, pp. 2687–2691, 1977. Cited on page 119.

[209] R. D. Bereman and D. Nalewajek, “Preparation of Dithiocarbamate Ligands Derived from Indole, Indoline, Carbazole, and Imidazole and Representative Transition-Element Com- plex,” Inorganic Chemistry, vol. 17, no. 4, pp. 1085–1087, 1978.

[210] V. B. Kobychev, N. M. Vitkovskaya, I. L. Zaytseva, E. Y. Larionova, and B. A. Trofimov, “Theoretical analysis of pyrrole anions addition to carbon disulfide and carbon dioxide,” International Journal of Quantum Chemistry, vol. 88, no. 5, pp. 542–548, 2002. [Online]. Available: http://doi.wiley.com/10.1002/qua.10206

[211] B. A. Trofimov, L. N. Sobenina, A. I. Mikhaleva, A. P. Demenev, O. A. Tarasova, I. A. Ushakov, and S. V. Zinchenko, “Reaction of Pyrrole Anions with Carbon Disulfide. Synthesis of Pyrrole-3-carbodithioates,” Tetrahedron, vol. 56, no. 37, pp. 7325–7329, 2000. [Online]. Available: http://linkinghub.elsevier.com/retrieve/pii/S0040402000006414 Cited on page 119.

[212] G. C. Fu, “Asymmetric Catalysis with "Planar-Chiral" Derivatives of 4- (Dimethylamino)pyridine,” Accounts of Chemical Research, vol. 37, no. 8, pp. 542–547, 2004. Cited on page 119.

[213] A. Sharma, A. Haldi, W. J. Potscavage Jr., P. J. Hotchkiss, S. R. Marder, and B. Kippe- len, “Effects of surface modification of indium tin oxide electrodes on the performance

XXXVIII of molecular multilayer organic photovoltaic devices,” Journal of Materials Chemistry, vol. 19, pp. 5298–5302, 2009. Cited on pages 121 and 141.

[214] S. Bhadra, A. Saha, and B. C. Ranu, “One-pot copper nanoparticle-catalyzed synthesis of S-aryl- and S-vinyl dithiocarbamates in water: high diastereoselectivity achieved for vinyl dithiocarbamates,” Green Chemistry, vol. 10, pp. 1224–1230, 2008. [Online]. Available: http://xlink.rsc.org/?DOI=b809200a Cited on pages 122 and 123.

[215] D. Brete, D. Przyrembel, C. Eickhoff, R. Carley, W. Freyer, K. Reuter, C. Gahl, and M. Weinelt, “Mixed self-assembled monolayers of azobenzene photoswitches with trifluo- romethyl and cyano end groups,” Journal of Physics: Condensed Matter, vol. 24, p. 394015, 2012. Cited on page 122.

[216] B. Kippelen and J.-L. Brédas, “Organic photovoltaics,” Energy & Environmental Science, vol. 2, no. 3, pp. 241–261, 2009. Cited on page 129.

[217] T. Ameri, G. Dennler, C. Lungenschmied, and C. J. Brabec, “Organic tandem solar cells: A review,” Energy & Environmental Science, vol. 2, pp. 347–363, 2009. Cited on page 129.

[218] E. Fortunato and R. Martins, “Where science fiction meets reality? With oxide semicon- ductors!” Physica Status Solidi RRL, vol. 9, pp. 336–339, 2011. Cited on page 129.

[219] H. Kim, C. M. Gilmore, A. Piqué, J. S. Horwitz, H. Mattoussi, H. Murata, Z. H. Kafafi, and D. B. Chrisey, “Electrical , optical , and structural properties of indium-tin-oxide thin films for organic light-emitting devices,” Journal of Applied Physics, vol. 86, no. 11, pp. 6451–6461, 1999. Cited on page 129.

[220] S. A. Paniagua, P. J. Hotchkiss, S. C. Jones, S. R. Marder, A. Mudalige, F. S. Marrikar, J. E. Pemberton, and N. R. Armstrong, “Phosphonic Acid Modification of Indium - Tin Oxide Electrodes: Combined XPS/UPS/Contact Angle Studies,” Journal of Physical Chemistry C, vol. 112, pp. 7809–7817, 2008. Cited on pages 129, 132, 135, and 141.

[221] A. Sharma, P. J. Hotchkiss, S. R. Marder, B. Kippelen, A. Sharma, P. J. Hotchkiss, S. R. Marder, and B. Kippelen, “Tailoring the work function of indium tin oxide electrodes in electrophosphorescent organic light-emitting diodes,” Journal of Applied Physics, vol. 105, p. 084507, 2009.

[222] P. J. Hotchkiss, H. Li, P. B. Paramonov, S. A. Paniagua, S. R. Marder, S. C. Jones, N. R. Armstrong, J.-L. Brédas, and S. R. Marder, “Modification of the Surface Properties of Indium Tin Oxide with Benzylphosphonic Acids: A Joint Experimental and Theoretical Study,” Advanced Materials, vol. 21, pp. 4496–4501, 2009. Cited on page 135.

XXXIX Bibliography

[223] P. J. Hotchkiss, S. C. Jones, S. A. Paniagua, A. Sharma, B. Kippelen, N. R. Armstrong, and S. R. Marder, “The Modification of Indium Tin Oxide with Phosphonic Acids: Mechanism of Binding, Tuning of Surface Properties, and Potential for Use in Organic Electronic Ap- plications,” Accounts of Chemical Research, vol. 45, no. 3, pp. 337–346, 2012. Cited on pages 129 and 135.

[224] K.-S. Liao, S. D. Yambem, A. Haldar, N. J. Alley, and S. A. Curran, “Designs and Archi- tectures for the Next Generation of Organic Solar Cells,” Energies, vol. 3, pp. 1212–1250, 2010. Cited on page 129.

[225] H. Kalisch, “Private Communication,” 2014. Cited on page 130.

[226] G. B. González, J. B. Cohen, J.-H. Hwang, T. O. Mason, J. P. Hodges, and J. D. Jorgensen, “Neutron diffraction study on the defect structure of indium-tin-oxide,” Journal of Applied Physics, vol. 85, no. 5, pp. 2550–2555, 2001. Cited on page 130.

[227] G. B. González, T. O. Mason, J. P. Quintana, O. Warschkow, D. E. Ellis, J. Hwang, J. P. Hodges, and J. D. Jorgensen, “Defect structure studies of bulk and nano-indium-tin oxide,” Journal of Applied Physics, vol. 96, no. 7, pp. 3912–3920, 2004. Cited on page 130.

[228] C. Donley, D. Dunphy, D. Paine, C. Carter, K. Nebesny, P. Lee, D. Alloway, N. R. Arm- strong, and D. Dunphy, “Characterization of Indium-Tin Oxide Interfaces Using X-ray Pho- toelectron Spectroscopy and Redox Processes of a Chemisorbed Probe Molecule: Effect of Surface Pretreatment Conditions,” Langmuir, vol. 18, no. 2, pp. 450–457, 2002. Cited on page 131.

[229] P. B. Paramonov, S. A. Paniagua, P. J. Hotchkiss, S. C. Jones, N. R. Armstrong, S. R. Marder, and J.-L. Brédas, “Theoretical Characterization of the Indium Tin Oxide Surface and of Its Binding Sites for Adsorption of Phosphonic Acid Monolayers,” Chemistry of Materials, vol. 20, no. c, pp. 5131–5133, 2008. Cited on pages 132 and 133.

[230] Kintec Co., “www.kintec.hk,” 26.11.2014. [Online]. Available: www.kintec.hk Cited on page 133.

[231] C. L. Perkins, “Molecular Anchors for Self-Assembled Monolayers on ZnO: A Direct Com- parison of the Thiol and Phosphonic Acid Moieties,” Journal of Physical Chemistry C, vol. 113, pp. 18 276–18 286, 2009. Cited on page 133.

[232] C. Queffélec, M. Petit, P. Janvier, D. A. Knight, and B. Bujoli, “Surface Modification Using Phosphonic Acids and Esters,” Chemical Reviews, vol. 112, pp. 377–3807, 2012. Cited on pages 133 and 135.

XL [233] O. T. Hofmann, J.-C. Deinert, Y. Xu, P. Rinke, J. Stähler, M. Wolf, and M. Scheffler, “Large work function reduction by adsorption of a molecule with a negative electron affinity: Pyridine on ZnO(10-10).” The Journal of Chemical Physics, vol. 139, no. 17, p. 174701, 2013. [Online]. Available: http://www.ncbi.nlm.nih.gov/pubmed/24206316 Cited on page 133.

[234] S. Hövel, C. Kolczewski, M. Wühn, J. Albers, K. Weiss, V. Staemmler, and C. Wöll, “Pyridine adsorption on the polar ZnO(0001) surface: Zn termination versus O termination,” The Journal of Chemical Physics, vol. 112, no. 8, p. 3909, 2000. [Online]. Available: http://scitation.aip.org/content/aip/journal/jcp/112/8/10.1063/1.480942 Cited on page 133.

[235] H. Li, P. Paramonov, and J.-L. Brédas, “Theoretical study of the surface modification of indium tin oxide with trifluorophenyl phosphonic acid molecules: impact of coverage density and binding geometry,” Journal of Materials Chemistry, vol. 20, no. 13, pp. 2630–2637, 2010. [Online]. Available: http://xlink.rsc.org/?DOI=b921768a Cited on pages 133 and 135.

[236] H. Li, L. K. Schirra, J. Shim, H. Cheun, B. Kippelen, O. L. A. Monti, and J.-L. Brédas, “Zinc Oxide as a Model Transparent Conducting Oxide: A Theoretical and Experimental Study of the Impact of Hydroxylation, Vacancies, Interstitials, and Extrinsic Doping on the Electronic Properties of the Polar ZnO (0002) Surface,” Chemistry of Materials, vol. 24, no. 15, pp. 3044–3055, 2012. [Online]. Available: http://pubs.acs.org/doi/abs/10.1021/cm301596x Cited on page 133.

[237] C. Wood, H. Li, P. Winget, and J.-L. Brédas, “Binding Modes of Fluorinated Benzylphos- phonic Acids on the Polar ZnO Surface and Impact on Work Function,” The Journal of Physical Chemistry C, vol. 116, pp. 19 125–19 133, 2012. Cited on page 133.

[238] H.-D. Jakubke and H. Jeschkeit, Aminosäuren, Peptide, Proteine. Weinheim, Germany: Verlag Chemie, 1982. Cited on page 138.

[239] P. J. Reeds, “Dispensable and Indespensable Amino Acids for Humans,” The Journal of Nutrition, vol. 130, no. 7, pp. 1835–1840, 2000. Cited on page 138.

[240] S. Spielberg, “Jurassic Park,” 1993, Universal Pictures. Cited on page 138.

[241] X. Deng, R. Nie, A. Li, H. Wei, S. Zheng, W. Huang, Y. Mo, Y. Su, Q. Wang, Y. Li, J. Tang, J. Xu, and K.-y. Wong, “Ultra-Low Work Function Transparent Electrodes Achieved by Naturally Occurring Biomaterials for Organic Optoelectronic Devices,” Advanced Materials Interfaces, 2014. [Online]. Available: http://doi.wiley.com/10.1002/admi.201400215 Cited on pages 138, 139, and 141.

XLI Bibliography

[242] A. Yatim, “Electronic Properties of Titanium Dioxide under UV Illumination,” Diploma thesis, RWTH Aachen University, 2011. Cited on page 139.

XLII List of Figures

1.1 Schematic energy diagram of an OLED stack...... 2 1.2 Schematic work function shifts upon DTC modification...... 4

2.1 Molecular structure of PTCDI-C13 ...... 8 2.2 Formation of dithiocarbamate monolayer...... 9 2.3 Potential energy and charge distribution of metal surface...... 11 2.4 Effects of adsorbates on a metal surface...... 12 2.5 Vacuum level alignment without charge transfer...... 14 2.6 Band bending in an organic semiconductor...... 16 2.7 Fermi level pinning at metal/molecule interface...... 18 2.8 Fermi level pinning: ICT vs gap states...... 19 2.9 Induced density of states model...... 22 2.10 Electrostatic potential of a free molecular layer...... 24 2.11 Fermi level pinning in SAM...... 25 2.12 Charge injection at metal/organic interface...... 27 2.13 Scheme of an OFET with electrode modifications...... 28 2.14 Charge flow in an OFET...... 29

3.1 Secondary electron in UPS measurements of Ag(111)...... 35 3.2 Escape cone for photoexcited electrons...... 37 3.3 Vacuum levels at surfaces...... 38 3.4 Schematic illustration of the energy landscape of vacuum levels of a crystal with different (hkl) facets...... 39 3.5 Energy diagram of a sample and detector during photoexcitation...... 40 3.6 Schematic representation of the energetic processes in a photoelectron measurement 41 3.7 Core level spectra of carbon upon fluorination...... 42 3.8 Illustration of hemispherical energy analyzer...... 50

XLIII List of Figures

4.1 Self-consistency scheme in DFT...... 57 4.2 Surface slab illustration of C2DTC on a gold cluster...... 61 4.3 Electrostatic potential along a metal-molecule interface...... 62 4.4 Complete representation of the electrostatic potential upon SAM formation.... 63 4.5 Initial surface slab geometries for calculations...... 65

5.1 Electronic structure of DTC monolayers on gold...... 71

5.2 Decomposition of a C2DTC on Au(111)...... 72

5.3 XPS spectra of C2DTC on Au(111)...... 75

5.4 XPS spectra of C2DTC on Ag(111)...... 77 5.5 Proposed binding scenario of C2DTC on Ag(111)...... 78

5.6 XPS spectra of C2DTC on Cu(111)...... 80 5.7 Proposed binding configuration of C2DTC on Cu(111)...... 81 5.8 Core level spectra of N,N-dialkyl DTC on Ag(111)...... 84 5.9 Core level spectra of N,N-dialkyl DTC on Cu(111)...... 85 5.10 UPS spectra of C2DTC on Au, Ag, Cu...... 88 5.11 Theoretical work function shift against surface coverage for C2DTC on Au, Ag, Cu 89 5.12 UPS spectra of CnDTC on Ag and Cu...... 91 5.13 Work functions of CnDTC on Au, Ag, Cu...... 92 5.14 Valence structure of C2DTC on Au(111)...... 94 5.15 Valence structure of C2DTC on Cu(111)...... 95 5.16 Valence structure of C2DTC on Ag(111)...... 97 5.17 Proposed bond-induced pinning mechanism...... 99 5.18 Fermi level pinning versus bond induced pinning...... 101 5.19 PDOS of Pyridines on Gold...... 102 5.20 PTCDI-C13 OFET with CnDTC modified S/D electrodes...... 104 5.21 In-situ transfer characteristics of PTCDI-C13...... 106 5.22 In-situ mobility and threshold voltage of PTCDI-C13...... 107 5.23 Charge carrier density of PTCDI-C13...... 108 5.24 Mobility and charge carrier density of PTCDI-C13 vs temperature...... 109 5.25 Evolution of charge transport layers of PTCDI-C13 as a function of temperature.. 110

6.1 UPS spectra of C10Thiol and FC10Thiol on Au, Ag, Cu...... 115 6.2 Molecular scheme of C1DTC...... 117 6.3 UPS spectra of C1DTC Au, Ag, Cu...... 118 6.4 Molecular scheme of PyDTC...... 119 6.5 UPS spectra of PyDTC Au, Ag, Cu...... 120 6.6 Molecular scheme of CNC2DTC...... 121

XLIV 6.7 UPS spectra of CNC2DTC Au, Ag, Cu...... 122 6.8 Molecular scheme of FC4DTC...... 123 6.9 UPS spectra of FC4DTC Ag and Cu...... 124 6.10 A new window of work functions for Au, Ag, Cu...... 125

7.1 XPS spectra of bare ITO after UV/ozone treatment...... 131 7.2 Work function measurements of ITO...... 132 7.3 Molecular structures of possible ITO modifier...... 134 7.4 Work function measurements of DMABA on ITO...... 135 7.5 XPS spectra of DMABA on ITO...... 137 7.6 Molecular structure of lysine...... 138 7.7 XPS spectra of Lysine on ITO...... 140 7.8 Work function measurements of lysine on ITO...... 141

8.1 Simplified energy level diagram of metal/PTCDI-C13 interface upon interface modification...... 146

A.1 Projected DOS of C2DTC on Au, Ag, Cu...... III A.2 Binding sites for DTC molecules on Au, Ag, Cu...... IV A.3 PTCDI-C13 OTFTs with modified Cu electrodes...... IV A.4 Molecular orbitals of pyridine-based phenyl molecules...... V

B.1 N 1s core level on Ag 3d satellite peak...... VII B.2 Core level of CS2 on Cu(111)...... VIII B.3 XPS measurements of C1oThiol...... IX B.4 XPS measurements of FC1oThiol...... X B.5 XPS measurements of C1DTC...... XI B.6 XPS measurements of PyDTC...... XII B.7 XPS measurements of CNC2DTC...... XIII B.8 XPS measurements of FC4DTC...... XIV B.9 XPS measurements of FC4DTC (continued)...... XV

XLV

List of Tables

4.1 Unit cell parameters of C2DTC monolayers on noble metals...... 64

5.1 Stoichiometry of C2DTC monolayer on Au(111)...... 76 5.2 Stoichiometry of C2DTC monolayer on Ag(111)...... 79 5.3 Stoichiometry of C2DTC monolayer on Cu(111)...... 82 5.4 Surface density of N,N-dialkyl dithiocarbamates on Au, Ag, Cu...... 86 5.5 Work functions of C2DTC monolayer on Au, Ag, Cu...... 88

6.1 Work function and surface coverage of reference thiols on Au, Ag, Cu...... 116 6.2 Work function and surface coverage of C1DTC on Au, Ag, Cu...... 119 6.3 Work function and surface coverage of Pyrrolyl-DTC on Au, Ag, Cu...... 121 6.4 Work function and surface coverage of CNC2DTC on Au, Ag, Cu...... 122 6.5 Work function and surface coverage of CNC2DTC on Ag and Cu...... 124

7.1 Stoichiometry of ITO after UV/ozone treatment...... 130 7.2 Surface coverage of DMABA on ITO...... 136 7.3 Surface coverage of Lysine on ITO...... 140

A.1 Deposition procedures of Au, Ag, Cu on mika...... II A.2 Experimental details of monolayer synthesis for noble metals...... II

XLVII

Acknowledgments

Abschließend möchte ich all jene Menschen honorieren, die mir während meiner Zeit am Institut hilfsbereit zur Seite standen. Dabei sind viele gute Freundschaften entstanden, von denen hoffent- lich einige auch in Zukunft so erhalten bleiben.

In Zeiten des Wandels in der deutschen Energielandschaft durfte ich mich als Teil der Arbeits- gruppe Organische Dünnschichten thematisch und experimentell austoben. Dabei konnte ich viele Kenntnisse in den Gebieten Ladungstransport und Grenzflächenphysik im Bereich moderner or- ganischer Elektronik sammeln. Am spannenden Euregio-Projekt ORGANEXT teilzunehmen und viel über die Möglichkeiten organischer Solarzellen zu lernen, war eine tolle Erfahrung. Für diese Möglichkeiten, gepaart mit einer großen kreativen Freiheit zur Gestaltung meiner Forschung, bin ich Prof. Dr. Matthias Wuttig zutiefst dankbar. Ohne ihn wäre diese schöne Zeit nicht möglich gewesen. Herzlichen Dank!

Nicht nur für die Übernahme des Zweitgutachtens meiner Arbeit möchte ich Prof. Dr. Riccardo Mazzarello herzlich danken. Als Experte im Bereich theoretischer Physik und insbesondere DFT- Simulationsrechnungen war er als ständiger Ansprechpartner eine große Hilfe bei der Diskussion und Validierung unserer Ergebnisse. Grazie mille!

Gerade beim Stichwort DFT-Simulationen möchte ich hier auch gerne den Beitrag von meinem Freund und Kollegen Tobias Schäfer hervorheben. Den Grundstein seiner Ergebnisse, die sich auch in meiner Arbeit wiederfinden, legte Tobias bereits in seiner Masterarbeit. Obwohl er danach das Themengebiet der organischen Materialien verlassen hatte, blieb er unseren DTC’s verbun- den und führte weiterhin mit Leidenschaft Simulationsrechnungen durch. Damit leistete er einen wesentlichen Beitrag für die vorliegende Arbeit. Vielen Dank!

Insgesamt 5 Jahre war ich als Teil der Organikgruppe am Institut. Es waren tolle 5 Jahre, in denen wir gemeinsam viele Höhen und Tiefen durchlebt haben, wodurch auch außerhalb der Uni wert- volle Freundschaften entstanden sind. Insbesondere bei meinen langjährigen Bürokollegen Ingolf

XLIX Acknowledgments

und Christian möchte ich mich für die hilfsbereite, sympathische und witzige Atmosphäre bedan- ken. Wir haben gemeinsam mit unserer Phantasialandphysik viel Spaß gehabt und sind dabei auch richtig gute Freunde geworden. Mit Daniel war ich von Anfang an im Geiste verbunden und wir haben Freud und Leid allzu häufig geteilt. Vom Hoffen und Bangen, Fluchen, Reparieren und Ma- rathonmessen am ORPHEUS bis zum Rotweintrinken am Strand von Arcachon inklusive Austern- all-you-can-eat. Serviert von Jean-Luc. Einmalig. Mit Carolin verbinde ich die ehrlichsten und direktesten Diskussionen, die ich je geführt habe, wodurch eine vertrauensvolle Freundschaft ent- standen ist. Auch alle weiteren Mitgliedern der Organikgruppe möchte ich nicht vergessen. Cathy, Philip, Sebastian, Julia, Franziska, Arthur, Emily und Andreas: Es hat Spaß gemacht, mit euch zusammen zu arbeiten!

Ich möchte hier auch meine Freunde erwähnen, die zwar nicht zusammen mit mir in der Organik- gruppe gespielt haben, die aber trotzdem zur schönen Atmosphäre am Institut beigetragen haben. Dazu zählen insbesondere meine langjährigen Kommilitonen, zeitweise Bürokollegen und guten Freunde Felix und Alexandra ebenso wie Rüdiger, Patrick, Matthias, Malte, Hanno, Deepu.

Oliver Lehmann bin ich dankbar für die Hilfsbereitschaft bei zahlreichen IT-Fragen und Pro- blemen. Dass du mich zur wöchentlichen Badminton-Runde eingeladen hast, hat mich besonders gefreut, denn es hat für den sportlichen Ausgleich in stressigen Zeiten gesorgt.

Bei Pascal möchte ich mich für die vertrauensvolle Überlassung seiner grünen Oase bedanken. Dadurch konnte ich meinen tristen Bürotisch in den schönen Dschungel verwandeln, der zuvor schon erfolgreich dein Chaos auf dem Tisch vertuschte. Meine ersten Monate als Doktorand mit dir und Felix in einem Büro haben viel Spaß gemacht.

Für die stets hilfsbereite und lehrreiche Unterstützung im technischen Bereich möchte ich mich sehr herzlich bei Stephan Hermes, Sebastian Moorhenn sowie Gert Kirchoff bedanken.

Ebenso gilt mein Dank Renate Betz, Sarah Schlenter und Kathrin Gerth, die es mit offenen Türen und Ohren schafften, neben der Bearbeitung aller Verwaltungsschritte stets eine herzliche Atmosphäre zu erhalten.

Zum Schluss gilt mein Dank selbstverständlich meinen Eltern Irmgard und Željko sowie meiner großen Schwester Larissa für die Unterstützung in den vergangenen Jahren.

L Hiermit versichere ich, dass ich die vorliegende Arbeit selbständig verfasst und nur die angegebe- nen Quellen und Hilfsmittel benutzt habe.

Aachen, 19. August 2015