Scanning tunneling spectroscopy on graphene nanoislands, iron nanoislands and phase change materials

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

vorgelegt von

Dipl.-Phys. Dinesh Subramaniam

aus Mundel (Sri Lanka)

Berichter: Univ.-Prof. Dr. Markus Morgenstern apl. Prof. Dr. Matthias Wuttig

Tag der mündlichen Prüfung: 20.04.2012

Diese Dissertation ist auf den Internetseiten der Hochschulbibliothek online verfügbar.

Contents

1 Introduction 1

2 Theory of Scanning tunneling microscopy 5 2.1 Tersoff–Hamann model ...... 5 2.1.1 Extension to non-spherical symmetric tip ...... 7 2.2 Scanning tunneling spectroscopy (STS) ...... 8 2.3 Spin-polarized scanning tunneling microscopy (SP-STM) ...... 11 2.3.1 Spin-polarized tunneling and constant current modus ...... 12 2.3.2 Spin-polarized scanning tunneling spectroscopy (SP-STS) . . . . 15 2.4 Energy resolution ...... 16

3 Phase change materials 17 3.1 Background phase change alloy Ge1Sb2Te4 ...... 17 3.2 Scanning tunneling microscopy and spectroscopy of the phase change alloy Ge1Sb2Te4 ...... 22 3.2.1 Preparation of Ge1Sb2Te4 ...... 22 3.2.2 Scanning tunneling spectroscopy of Ge1Sb2Te4-sample ...... 24 3.2.3 Structural analysis of Ge1Sb2Te4 ...... 29

4 Spin polarized scanning tunneling spectroscopy of Fe nanoislands on W(110) 37 4.1 Fe nanoislands on W(110) ...... 37 4.2 Preparation of tips for spin polarized scanning tunneling microscopy . . 40 4.3 Preparation of Fe nanoislands on W(110) ...... 42 4.4 Vortex core movement in in-plane magnetic fields investigated by spin polarized scanning tunneling spectroscopy ...... 43 4.5 Hysteresis properties of Fe nanoisland study by spin polarized scanning tunneling spectroscoopy ...... 47

5 Graphene 51 5.1 Background graphene ...... 51 5.1.1 Band structure of graphene ...... 52 5.1.2 Pseudo relativistic Dirac Fermions ...... 53 5.1.3 Graphene quantum dots ...... 54

i Contents

5.1.4 Edge state ...... 57 5.2 Background graphene on Ir(111) ...... 63 5.2.1 Preparation of graphene on Ir(111) ...... 63 5.2.2 Structure of graphene on Ir(111) ...... 63 5.2.3 Band structure of Ir(111) and graphene on Ir(111) ...... 66 5.3 Wave function mapping in graphene quantum dots ...... 69 5.3.1 Scanning tunneling spectroscopy measurement of eigenstates on graphene quantum dots (QDs) ...... 70 5.3.2 Theoretical model ...... 72 5.3.3 Comparison between theory and experiment ...... 76 5.3.4 Ir surface state S0 and moiré structure ...... 79 5.3.5 Lifetime of quantum dot states ...... 85 5.4 Investigation of zigzag graphene edge on Ir(111) ...... 87 5.4.1 Density functional theory calculation of graphene edge on Ir(111) 87 5.4.2 Termination of the graphene edge ...... 91 5.4.3 Scanning tunneling spectroscopy on graphene edge ...... 94 5.4.4 Spin polarized scanning tunneling spectroscopy on graphene edge 98 5.5 Mechanical instability of graphene naonislands on Ir(111) ...... 103

6 Summary 105

Bibliography 107

Publications 121

Curriculum vitae 123

Acknowledgments 125

ii 1 Introduction

In recent years, in the area of technological development increased efforts were made in miniaturizing electronic components. In the case of new transistors and data mem- ory devices, the development has already reached the nanometer range. To obtain such nanometer range materials, either the known materials are scaled-down or exchanged completely by novel materials. An interesting technological application is the use of magnetic polarization to store information. The aim from the beginning of ferrite-core memory, magnetic tape and hard disk technology till to modern non-volatile M-RAM (Magnetic Random Access Memory) is to miniaturize the storage media, to store as much information as possible in a small area [1]. Since the properties of materials change drastically at the nanoscale region, the study of ferromagnetic materials regard- ing the magnetic properties is of enormous interest. Spin polarized scanning tunnel- ing microscopy and spectroscopy is a very useful tool for investigating such nanoscale magnetic systems. In this work, the behavior of the model system Fe nanoisland on W(110) with and without external magnetic field was studied. The investigation was focused on the movement of vortices in an in-plane magnetic-field and hysteretic be- havior of a single-domain island has been observed in all three spatial dimension. Another interesting material which has great potential for future generations of stor- age media are the so-called phase change materials based on Ge-Te-Sb alloys. Such materials are already used in optical re-writable discs, known as CD-RW (compact disc rewritable) or DVD-RW (digital versatile disc rewritable). Still in development is the non-volatile electrical PCRAM (Phase Change Random Access Memory) based on Ge-Te-Sb alloys. Here the large electrical contrast between the amorphous and the crystalline phase is used to store information. The switching between the amorphous and the crystalline phase is done by heating up or cooling down the material. In this work, the scanning tunneling microscopy and spectroscopy have been employed to reveal the evolution of the band gap and the Fermi level as a function of the annealing temperature for Ge1Sb2Te4. Moreover, large scaled and atomic resolution topography images were recorded to analyze the morphology and the complex atomic structure of this phase change material. In 2004, a new material was discovered, which has both technological relevance as well as very interesting aspects for the fundamental physics. The material graphene was first discovered by Geim and Novoselov [2] and was honored with the Nobel Prize in the year 2010. From the perspective of fundamental physics the most interesting as- pects of this new two dimensional electronic system (2DES) is that the charge carriers

1 1 Introduction

are massless Dirac fermions. To describe this relativistic particles, it is necessary to use the Dirac equation instead of the famous Schrödinger equation. This leads to physical effects which have so far been observed only at very high energies and is described by the theory of quantum electrodynamics (QED). With graphene it is possible to observe QED effects on the table of the lab, with low energies and relatively simple preparation techniques and setups. The relativistic description for graphene is valid with the mod- ification that the speed of the particles is 300 times smaller than the speed of light and the carrier energies must be low (E ≤ 1 eV) [3]. One of the exciting relativistic effects in graphene is the so called Klein tunneling [4]. The effect is that Dirac fermions can be transmitted with probability one through a classically forbidden region [3]. To some extent this QED phenomena has been already shown in an experiment on graphene [5]. By switching on the magnetic field the Dirac Fermions show also unconventional behavior, leading to new physical phenomena such as the anomalous integer quantum Hall effect at room temperature [6, 7]. Besides of these beautiful fundamental physics properties, graphene has raised hope for a variety of applications. Graphene is a strong candidate to be a potential solu- tion to the limits of the silicon-based semiconductors. Compared to silicon, graphene has very high electrical mobility up to 200000 cm2/Vs [8, 9], due to the reduced back scattering of the Dirac Fermions and of the high crystal quality. In recent years, big steps were made regarding the application of graphene. IBM had already built a 100 GHz graphene-transistor on SiC substrate, that can be almost produced in industrial processes [10]. Further possible applications for graphene are supercapacitors [11] or touch screens [12]. Another exiting plan is to use graphene quantum dots as spin qubits [13, 14]. The ba- sic requirement to built spin qubits is very long spin coherence time [15, 16], which is likely to exist in graphene [17] due to the absence of hyperfine coupling in iso- topically pure material and the small spin-orbit coupling [18, 19]. In the last years first graphene quantum dots have been produced and probed by transport measure- ments [20, 21, 22, 23]. However, since graphene provides no natural gap, it is difficult to control the electron number in the quantum dot [24]. Moreover, the 2D sublattice symmetry makes the quantum dot properties very susceptible to the atomic edge con- figuration [13, 14], which is different from conventional quantum dots. As a result, chaotic Dirac billiards have been predicted [25] and were even claimed to be realized [20, 26, 27], i.e. the wave functions are assumed to be rather disordered. To get more control on graphene quantum dots, the quantum dot edges must be well defined and a more fundamental understanding of the quantum dot properties is mandatory. A direct insight into quantum dot properties is provided here by scanning tunneling spectroscopy (STS) which maps the squared wave functions of quantum dots [28, 29] and, at the same time, determines the shape of the quantum dot atom by atom. Using STS, in this work the graphene quantum dots with well defined zigzag edges sup-

2 ported on an Ir(111) surface were probed [30]. These quantum dots maintain graphene properties as the filled part of the graphene Dirac cone lies in the projected band gap of Ir [31]. By comparing the measured wave functions with model calculations, the relationship between geometry and electronic properties are determined and general trends are extracted. A further interesting aspect of this system is the pure zigzag edges, which may have a spin-polarized edge state. Comparison of spin polarized scanning tunneling mi- croscopy and spectroscopy measurements with density functional theory calculations are used to determine the termination of the graphene edges and its spin polarization.

3

2 Theory of Scanning tunneling microscopy

2.1 Tersoff–Hamann model

The simplest theoretical description of vacuum tunneling between tip and sample based on a time-dependent, first order perturbation theoretical approach from Bardeen [32]. This approach considers the three-dimensional geometry of the sample and the tip. Instead of solving the Hamilton operator of the entire system from tip elec- trode, vacuum barrier and sample electrode Bardeen calculated the tunneling current through the overlap of the wave function of independent systems by using Fermi’s Golden Rule. The tunneling current can be written as:

2πe 2 I = · ∑{ f (Eµ)[1 − f (Eν + eU)] − f (Eν + eU)[1 − f (Eµ)]} · |Mµν| δ(Eν − Eµ) h¯ µ,ν (2.1) with the Fermi-function f (E)1, the applied bias voltage U and the tunnel matrix el- 2 ement Mµν between the non-perturbed wave function of the tip Ψµ and the sample Ψν. Both electrodes are described by discrete electronic states Eµ and Eν which are summed up3. Due to the tunnel voltage U, a potential gradient eU between tip and sample is generated. The resulting current is the difference between the current from tip to sample and that from the sample to the tip (terms in the curly brackets). The term is multiplied by the tunnel matrix element Mµν. The delta function indicates, that only elastic tunneling electrons are considered. The main problem now is to estimate the tunnel matrix element Mµν. The square of this term describes the transmission prob- ability of the tunneling electron through the potential barrier. Bardeen expressed this by: −h¯ 2 Z M = · dS · (Ψ ∗∇Ψ − Ψ ∇Ψ ∗). (2.2) µν 2m µ ν ν µ 1The eigenvalue of the sample (tip) will be marked with the index ν( µ). 2Non-perturbed wave function means, the wave functions of the tip and sample do not influence each other. This condition is valid for large tip to sample distances. For small tunnel resistance R = Ugap/I, i.e. for small tip to sample distances, this approximation can’t be used. 3Instead of discrete states, there are continuous bands in metallic solid. Therefore, later the sum will be replaced by an integral.

5 2 Theory of Scanning tunneling microscopy

The integral is used on the wavefunction Ψ of the tip and sample and on any surface S in the tunnel barrier between the tip and sample. An exact specification of Mµν in equation 2.2 is not possible, therefore the real geometry of the tip is unknown. And therefore the electronic structure is also unknown. Bardeen proposed the equation 2.1 for a metal-oxide-metal tunneling process in 1961. Based on this model Tersoff and Hamann [33, 34] developed a theory for the metal-vacuum-metal tunneling process in STM. They simplified in the case of small temperatures, i.e. the widening of the Fermi distribution can be neglected4, and for small tunneling voltages the equation 2.1 is used: 2 2πe 2 I = U ∑ |Mµν| δ(Eν − EF) · δ(Eµ − EF) (2.3) h¯ µ,ν where EF is the Fermi energy. Tersoff and Hamann took for the unknown tip structure a simple model of the tip with spherical symmetry (see Figure 2.1). The spherical symmetry of the tip corresponds to an s-like electronic structure (angular quantum number l = 0). With this assumption, the tunnel matrix element can be calculated and

R r z 0 d x,y

Figure 2.1: Sketch illustrating the Tersoff-Hamann model of a STM tip: The forefront of the tip is approximated with a spherical model. R is the curvature radius from center~r0 and the distance from the tip apex to the flat surface is d (after [33]).

the equation 2.3 simplified to:

2κR 2 2κR I ∝ U · ρµ(EF) · e · ∑ |Ψν(r0)| δ(Eν − EF) = U · e · ρµ(EF) · ρν(~r0, EF). (2.4) ν

The sum describes the local density of states (LDOS) of the sample ρν(~r0, EF) at the Fermi energy, while ρµ(EF) represents the constant density of states of the tip at the Fermi energy. The property “local” is defined by the position of the tip. The decay

4The Fermi distribution will be replaced by a Θ step function in the limit T → 0K

6 2.1 Tersoff–Hamann model

√ constant κ is given by κ = 2meΦ/¯h, where Φ is the effective work function of tip and sample. The effective work function can be approximated in the first order with Φ = (Φµ + Φν)/2. The wave functions of the electrons, which have a component in the z-direction, decay exponentially into the vacuum:

−κz Ψν(r) ∝ e , (2.5)

so that the probability to find an electron at the position z = d + R (see Figure 2.1) is:

2 [−2κ(d+R)] |Ψν(r0)| ∝ e . (2.6)

The result will be applied now to equation 2.4 and we obtain an exponential depen- dence of the distance between the tip and the sample:

I ∝ e−2κd. (2.7)

With the perturbation theory model of Tersoff and Hamann and with the realistic nu- merical values of the tip-sample distance and the tip diameter, it is possible to calculate the corrugation of a surface moving the tip across. The calculated values fit well with the measured corrugation5 ∆z of a reconstructed metal surface (a good example is the reconstructed Au(110)-surface [33]. ∆z can be written as:  r  2 π2 2 −2( κ + 2 −κ)z ∆z ≈ e a , z = d + R, (2.8) κ with a the period in real space. The conclusion from 2.8 is that the corrugation increases with smaller distance to the sample, i.e. with larger tunnel current, and decreases with smaller distance between adjacent atoms. This indeed is observed very often in experiments.

2.1.1 Extension to non-spherical symmetric tip Experimental studies on closely-packed metal surfaces (for example Al(111) [35], Au(111) [36]) have shown that the model of Tersoff and Hamann describes not al- ways correctly the z-corrugations. This can be explained partly by an extension of the model of Tersoff and Hamann. In STM experiments, mostly tungsten-, platinum- or iridium-tips are used. The tunneling process in these materials depends not only on the simple s-wave function, but also on the additionally localized tip orbital with dz2 - symmetry [37, 38, 39]. The resolution can drastically increase if the tunneling process goes through the d-orbital of a tip atom, because the d-orbitals extend far into the vac- uum and have a large direction-dependence [1]. Chen extended Bardeen’s approach

5 The corrugation ∆z is the difference between maximum and minimum z value: ∆z = zmax − zmin.

7 2 Theory of Scanning tunneling microscopy for tip-states with different angle-dependence or more specifically with angular mo- mentum numbers l and magnetic quantum numbers m. The main results from the l (~ ) (~r ) d Ψν r0 equation 2.4 is that Ψν 0 will be replaced by dzl , so that the tip-orbitals with an- gular momentum l and orbitals with preferential direction to the sample contribute to the tunneling current. Chen’s model partially explains the observed high corruga- tions on closely-packed surfaces, as well as the occasional change of the height contrast in atomically resolved STM images. Chen describes these phenomena as a switching from a s-orbital to a d-orbital tunneling process. A summary of this model can be found in the book of Wiesendanger [40].

2.2 Scanning tunneling spectroscopy (STS)

The tunneling current is not only dependent on the tip-sample distance, it is also sen- sitive to the change of the applied voltage U. Therefore the voltage of the tunneling between the two metal electrodes was set to the millivolt range. In this case the tun- neling current is linear to the applied voltage (see equation 2.4). For higher voltages the tunneling current response is partially non-linear [41]. From detailed studies of this voltage dependent, spectroscopic data can be obtained. While conventional techniques such as photoemission spectroscopy (PES) and inverse photoemission spectroscopy (IPES) average over large areas of the sample, the spectroscopic modes of the scanning tunneling microscopy (STS) opens new opportunities for the local-resolved study of the electronic structure of a surface down to atomic scale [42]. However, the interpre- tation of the obtained data is far from trivial. Selloni et al. [43] simplifies the tunneling current, which was derived in section 2.1, assuming a structure with low density of states of the tip and ε = E − EF to:

Z eU I ∝ ρν(~r, EF + ε) · T(ε, U)dε. (2.9) 0 In the equation, the sum was replaced by an integration, which is allowed for quasi- continuous electronic states as present in solids and ρµ(E) was assumed to be constant. The main problem is that the applied voltage not only occurs in the density of states of the sample ρν(~r, EF + ε), but also in the unknown transmission coefficients T(ε, U). The transmission coefficient is

T = T(d, U, ε,~k), (2.10) a function of the distance d between the electrodes, the applied voltage U and it de- pends on energy ε and wave vector~k of the tunneling electrons, i.e. on band structure E(~k). In addition, the transmission coefficient depends on the shape of the tip. The

8 2.2 Scanning tunneling spectroscopy (STS)

assumption of free electrons (neglecting band-structure effects) provides an approxi- mation within the semi-classical WKB-theory6 for the transmission coefficient in planar tunnel junctions: v u    u h¯ 2k2 (−2κ(ε,U,)s) u2m eU k T ≈ e , κ(ε, U) = t Φ + − ε − , (2.11) h¯ 2 2 2m

with the effective tunneling distance s = d + R and the mean work function Φ = (Φµ + Φν)/2. The inverse decay length κ(ε, U, kk) is smallest at a certain energy ε or E for a state with vanishing wave vector parallel to the surface (kk = 0). Due to T, the states at Γ-point in the surface Brioullin-zone will be weighted strongly, so that the states at Γ are detected preferably with scanning tunneling spectroscopy [44]. This property of the transmission coefficient is usually not explicitly included. One only considers the effect of U and E under the restriction kk = 0. Figure 2.2 shows this

Figure 2.2: Energy diagram for the sample and tip at T = 0 K. (a) Sample and tip in equilibrium, separated by a small vacuum barrier. (b) Negative sample voltage: electrons tunnel from the sample into the tip. (c) Positive sample voltage: electron tunneling from the tip into the sample (Figure from [45])

dependence of the transmission coefficient for the one-dimensional case. Without ap- plied voltage, the Fermi levels of tip and sample are equal (2.2(a)). Now, if a voltage U is applied to the sample, the energy levels either move upward or downward by the amount of |eU|, depending on whether a negative (2.2(b)) or a positive (2.2(c)) po- larity is applied. In the case of a positive sample voltage, the electrons tunnel from the occupied states of the tip into the empty states of the sample (2.2(c)). The reverse

6Wentzel-Kramers-Brioullin

9 2 Theory of Scanning tunneling microscopy

process takes place at the negative sample voltage. The electrons tunnel from the oc- cupied states of the sample to the empty states of the tip (2.2(b)). By selecting the voltage polarity one can study either the occupied or the empty states of the sample. It is important to notice here, that the tunneling probability of electrons is dependent on their energy relative to the Fermi-level. Electrons near to the Fermi-level "see" a lower barrier than electrons, which are far below the Fermi-level. Therefore, the first case contributes more to the tunnel current than the last case. This is symbolized by the thickness of the arrows in Figure 2.2. It was shown experimentally that the contribution of the density of states in the dif- ferential conductivity dI/dU is not clearly visible for large U, due to the exponential divergence of the tunneling probability [46]. To solve the problem, Feenstra et. al. [46] proposed to normalize the differential conductivity dI/dU by the division of the total conductivity I/U: (dI/dU) dlnI = . (2.12) (I/U) dlnU Differentiation of 2.9 gives7:

dI Z E1 d ∝ eρν(E1) · T(E1, U) + e ρν(E) · [T(E, U)]dE . (2.13) dU EF d(eU) That implies:

divergent background LDOS z }| { = d [ ( )] z }| { Z E1 eU d(eU) T E, U ρν(E1) + ρν(E) · dE (dI/dU) E = T(E , U) = F 0 1 . (2.14) (I/U) 1 Z E1 T(E, U) ρν(E) · dE eU EF T(E1, U) | {z } normalization The first term of the addend gives the local density of states (LDOS) of the sample and the second term gives the divergent background, which exponentially increases with the voltage. For the transmission coefficient or the tunneling probability we would apply T(E, U) = exp(−2κD). The two terms are normalized by the denominator. The derivative appearing in the numerator dT/d(eU) can be set to T by neglecting the inner derivation of T. Due to the appearance of the factors T(E, U) and T(E1, U) as ratios, both in the numerator and in the denominator, the dependence on the applied voltage can be canceled. Stroscio et al. [47] could show for Si and for Ni that this normalization method can derive the LDOS information out of the spectra. A disadvantage of this

7 R y1 The integral of the function h(x, y) will be written in primitive h(x, y) = 0 f (ey, y)dx = y1 [F(x, y)]|0 = F(x1, y) − F(0, y) and then partially derived using x = e · y

10 2.3 Spin-polarized scanning tunneling microscopy (SP-STM)

normalization is the loss of the spectroscopic information in the region of the Fermi energy due to dI/dU ≈ I/U. Furthermore, it follows from equation 2.12 that the prob- lem with the division by zero could occur twice. Both the voltage U and the tunneling current I disappear at one point during the voltage ramp, the result is a bad signal to noise ratio near the Fermi level [42]. Finally, the model assumes that the density of states of the tip ρµ(E) is featureless. This assumption is not always valid, what could be already shown in the field emission measurements on tungsten tips [48]. A randomly attained featureless ρµ(E) is therefore a precondition for a reliable determination of the LDOS ρν(E1). Another method was proposed by Ukraintsev [49]. Here a function F(U, αi) was used, to approach the transmission coefficient with the free fit parameters αi. The function F(U) will be fitted to the experimental dI(U)/dU data of a known density of states and dI(U) afterwards the spectra will be normalized by F: U /F(U). In a simulation the au- thor showed that the relative intensities of maxima for the density of states of a surface could be resolved better by the normalized spectra with F(U) as with normalization with the total conductivity I(U)/U. The method, however, requires an explicit knowl- edge of F(U), which is usually extremely complicated to determine. The method of Ukraintsev can also not reflect the influence of the density of states of the tip ρµ on the differential conductivity. A possible solution of this problem provides the charac- terization of the density of states of the tip by a comparative measurement on a test sample with known featureless density of states. Such a system is Au with dominant s-bands at the Fermi-level, which could be used to eliminate tips with sharp features in the density of states [44]. It can be summarized, that a quantitative evaluation of an individual spectra regarding ρν(E) can be very complex. A careful analysis must consider a huge number of possi- ble influences. There is no generally applicable procedures. It is therefore necessary to develop a method that works for the individual experiment. Particularly helpful are comparisons with the theoretical calculations of the density of states as well as com- parisons with the results of other spectroscopy methods [44].

2.3 Spin-polarized scanning tunneling microscopy (SP-STM)

In the previously described concepts of the STM and STS the spin of the tunneling electrons was neglected. In the spin-polarized scanning tunneling microscopy (SP- STM) and -spectroscopy (SP-STS) a probe will be used, which has spin-polarized states. So that SP-STM is not only sensitive on the spin-averaged sample density of states, but also on the local spin density of states [44]. A very successful method is using tungsten tips, where a thin magnetic coating is applied [44, 42]. In the recent years also bulk anti-

11 2 Theory of Scanning tunneling microscopy

ferromagnetic tips were successfully used as spin sensor [50, 51]. The lateral resolution in SP-STM images is similar to the resolution of the conventional topographic STM images, i.e. it is in the atomic range.

2.3.1 Spin-polarized tunneling and constant current modus To understand the basic principle of the method, a tunnel junction between two ferro- magnetic electrodes is considered (see figure 2.3). Due to the exchange interaction in

Figure 2.3: Schematic of spin-polarized tunneling between two ferromagnetic elec- trodes in a parallel orientation (a) and an anti-parallel orientation (b) of the magnetization. In the case of elastic and spin-preservative tunneling, the tunnel current in (a) is larger than in (b). Figure from [44].

the Stoner model of the band ferromagnetism, the individual sub-bands with opposite spin orientations are shifted against each other. The energetically lowered majority band (↑) and the lifted minority band (↓) are filled with a different number of electrons up to the Fermi energy [44]. The unequal occupation of the states leads to a net magne- tization. To describe the tunneling probability between the ferromagnetic electrodes, an energy-dependent spin-polarization P(E) is defined [44]:

ρ↑(E) − ρ↓(E) P(E) ≡ . (2.15) ρ↑(E) + ρ↓(E)

In Figure 2.3, the relevant tunneling processes at the Fermi-energy EF are sketched. Since spin-flip process can be neglected (P ≤ 1 %), the spin is conserved during the tunneling. For parallel magnetization (a) only the majority electrons from the tip tunnel into the empty majority states of the sample and the minority electrons into the empty minority states. In the case of antiparallel magnetization 2.3 (b), combina- tions of occupied or empty majority- and minority states appear [52]. In the real sys- tem, the preferred directions are not always parallel, the angle is arbitrarily oriented. Slonczewski [54] extended the method in section 2.1 for the case of ferromagnetic elec-

12 2.3 Spin-polarized scanning tunneling microscopy (SP-STM)

Figure 2.4: Tunneling conductance of a Fe/Al2O3/Fe-system as a function of the an- ~ ~ gle θ(M1, M2) between the magnetization of the ferromagnetic electrodes −1 (From [44, 53] with σ0 = 96, 2 Ω , P1P2 = 9 %, at T = 300 K). The experi- mental data confirm the cosine dependence. trodes. Within the model for free electrons in the limit of small voltage he got the following expression for conductivity:

σ = σ0[1 + cos θ P1(EF)P2(EF)]. (2.16)

Here, θ is the angle between the two magnetization directions, and σ0 for the con- ductivity at a vertical orientation of the magnetization [44, 52]. The equation 2.16 shows that the tunneling current or the conductivity is increased in the parallel case (cos θ = 1) compared to the anti-parallel case (cos θ = −1). This applies for spin po- larization Pi with the same algebraic sign. The effect is high for large products [P1 · P2] [44]. Experimentally the cosine dependence of the tunneling current was shown on planar Fe/Al2O3/Fe tunnel junctions by Miyazaki and Tezuka [53] (see Figure 2.4). The theoretical extension of the Tersoff-Hamann-model for the case of magnetic elec- trodes was done by Heinze et. al. [55]. In this model the tip has one constant energy t t density with respect to spin orientation ρ↑, ρ↓. The absolute values are different, so that a non-vanishing spin polarization is given by [44]: ρt − ρt t t ↑ ↓ P (E) = P = t t . (2.17) ρ↑ + ρ↓

The assumption is a similar decay of the s-like tip states into the vacuum (κ↑ = κ↓). Heinze et. al. [55] got for the tunneling current I for small tunneling voltages and neglected spin-flip processes the following expression: t s t s I = I0 + Ip I(~r0, U, θ(~r0)) ∝ ρ · ρ˜ (~r0, U) + m · m˜ (~r0, U) · cos θ(~r0), (2.18)

13 2 Theory of Scanning tunneling microscopy

t = t + t s(~ ) = R eU s ( + ~ ) + here ρ ρ↑ ρ↓ means the total tip density of states, ρ˜ r0, U 0 ρ↑ EF ε, r0 s ρ↓(...)de is the energy-integrated density of states of the sample at the local position of ~ t = t − t s = R eU s ( + ~ ) − s ( ) the tip r0, m ρ↑ ρ↓ is the spin density of the tip, m˜ 0 ρ↑ EF ε, r0 ρ↓ ... dε ~ t ~ s is the energy-integrated spin density of the sample and θ(~r0) = θ(M , M (~r0)) is the angle between the local sample magnetization M~ s(x, y, z = 0) and the fixed tip ~ t magnetization M . The first term in 2.18 describes the unpolarized part I0 of the tunneling current and the second term gives the polarized part Ip. With the defini-

Figure 2.5: Schematic drawing of constant-current mode of spin-polarized scanning tunneling microscopy. The tip-sample distance is increased for the paral- lel orientation of the magnetization of tip and sample and decreased in the case of antiparallel orientation. Figure from [44].

s tion of the locally resolved energy-integrated polarization of the sample, P˜ (~r0, U) ≡ s s m˜ (~r0, U)/ρ˜ (~r0, U), we get the following equivalent relation to 2.18:

t s t s  I(~r0, U, θ(~r0)) ∝ ρ · ρ˜ (~r0, U) 1 + P P˜ (~r0, U) cos θ(x0, y0) (2.19)

According to 2.19, the tunneling current changes with a variation of cos θ (where PtP˜s 6= 0), and the magnetic structure of a surface can be imaged in the constant- current mode. Here the magnetic contrast becomes apparent as a height variation [44]. In figure 2.5 this is schematically shown. In the case of same decay of the sample states into vacuum, the maximum difference in the tip-sample distance 4z|max = z↑↑ − z↑↓ can be estimated between a part with parallel (cos θ = 1) and a part with antiparallel (cos θ = −1) sample to tip magnetization [44]. √ 1 1 + Pt · P˜s(U) 2mΦ¯ 4 z|max = ln mit κ = . (2.20) 2κ 1 − Pt · P˜s(U) h¯

14 2.3 Spin-polarized scanning tunneling microscopy (SP-STM)

An estimation using realistic values for the effective barrier height Φ¯ = 4 eV and the effective polarization of the tunnel junction P˜ts = Pt · P˜s ≈ 0.2 results in a corrugation of 4z|max ≈ 0.02 nm [44]. Such height differences can be measured effectively with the scanning tunneling microscope. Nevertheless, the topographic mode is not often used, because it has one major disadvantage. The height variations in equation 2.21 are difficult to separate from the topography of the sample. To overcome this problem spin-polarized scanning tunneling spectroscopy is used, which is described in the next section.

2.3.2 Spin-polarized scanning tunneling spectroscopy (SP-STS)

The spectroscopic mode allows to separate the magnetic information from the topog- raphy [56]. The differential conductance at low voltages from the derivation of the equation 2.18 results in [44]:

dI t s t s (~r0, U, θ(~r0)) ∝ ρ ρ (~r0, EF + eU) + m m (~r0, EF + eU) cos θ(~r0), (2.21) dUsp or using equation 2.19:

dI dI t s  (~r0, U, θ(~r0)) ∝ (~r0, U) · 1 + P P (~r0, EF + eU) cos θ(x0, y0) . (2.22) dUsp dUsa

In a spin-polarized measurement dI , both a non-polarized (spin-averaged) part dI dUsp dUsa and a polarized part will be measured. The spectroscopic mode has a much higher sen- sitivity compared to the constant-current mode. Due to that the magnetic contrast is no longer measured by the tunneling current, which is logarithmically-dependent on the tip-sample distance, but directly on the variation of the polarized part of the differ- ential conductivity [44]. So it is possible to select a certain voltage U to get a high value t s of the product P P (~r0, EF + eU). If an electronically homogeneous surface is provided, the spin-average of part dI and also Ps(~r , E + eU) are laterally constant. Then, at a dUsa 0 F fixed voltage and constant tip-sample distance any signal variation in dI/dUsp can di- ~ s ~ t ~ s rectly associate with a change of local sample magnetization M (~r0) by cos θ(M , M ) [44]:

dI ts  ts t s (U, x, y) = C 1 + P cos θ(x, y) with P = P · P (EF + eU). (2.23) dUsp

Simultaneous recording of spectroscopy and topography images could help to verify, whether the tip to sample distance is constant or z-variations are present.

15 2 Theory of Scanning tunneling microscopy

2.4 Energy resolution

So far all the derivations were done for a temperature T = 0 K. The temperature de- pendence can be considered by introducing the Fermi function. T is one of the factors, which limits the energy resolution of the experiment. The other parameter which re- stricts the energy resolution is the lock-in modulation voltage Umod. The energy reso- lution is given by [57]: q 2 2 δE ≈ (3.3 · kBT) + (2.5 · eUmod) . (2.24)

In order to get high-resolution spectroscopy images, the temperature should be as low as possible. Our low temperature measurements were done at T = 6 K and usually with Umod = 10 mV, this lead to an energy resolution of δE = 25 meV.

16 3 Phase change materials

In the past two decades, the demand for faster and larger storage devices has increased exponentially. Phase change media are promising to overcome this problem. Phase change media based on Ge-Sb-Te (GST) alloys are nowadays already preferred in op- tical re-writable discs, known as CD-RW (compact disc rewritable), DVD-RW (digi- tal versatile disc rewritable) and Blue Ray. The interesting feature of these materials lies in the large optical and electrical contrast between the amorphous and the crys- talline phase. Therefore, GST is also a promising candidate for non-volatile electrical data storage, such as PCRAM. The electrical switching between the two phases was al- ready predicted by Ovishinsky [58] in the sixties, but is still not fully understood. The schematic drawing of the optical storage process is shown in Figure 3.1, and is based on the switching process between crystalline and amorphous phase using a laser pulse. The material is initially in the crystalline phase. By local heating to the melting temper- ature Tm and subsequent very rapid cooling (quenching), we obtain sub-micron large amorphous marks. The cooling must be done very quickly with a rate of 106 to 109 K/s, so that the crystallization is suppressed. This is called the writing process. Now in order to get back into the crystalline phase, the “written” area is heated just above the crystallization temperature Tx. Thus, the information will be deleted. To read the information, the high optical contrast between the two phases are used. Therefore, the intensity of the reflected laser pulse, with much lower power than during the writing process, is measured. The following work was made during his diploma thesis to- gether with C. Pauly [60]. Section 3.1 provides a review of the structural and physical properties of Ge1Sb2Te4. The following section 3.2 describes the scanning tunneling microscopy and spectroscopy measurements of the phase change alloy Ge1Sb2Te4.

3.1 Background phase change alloy Ge1Sb2Te4

Ge1Sb2Te4 is an amorphous semiconductor in its initial phase, e.g. after sputter depo- sition. In its crystalline phase, there is a distinction between a metastable and a stable phase. The difference between them is the lattice structure. The transition from one phase to another occurs continuously with increasing temperature, so that the mixed phase is limited to a small temperature window.

17 3 Phase change materials

Tm

temperature

Tg

write read erase time

crystalline amorphous

Figure 3.1: Schematic drawing of optical data storage with phase change materials. Figure from [59].

Figure 3.2 shows DSC8 measurements [61] revealing the activation-energy, which is required for the phase transitions. The transition from amorphous to metastable phase occurs at 390 K and the transition from metastable phase to stable crystalline phase at around 500 K [61]. The stable phase remains up to the melting point. XRD9 measurements show that the metastable phase is a face-centered cubic rock salt structure. The tellurium atoms (Te) occupy the first fcc sublattice. The other two atoms, germanium (Ge) and antimony (Sb) occupy the second fcc sublattice, where 25 % of the lattice sites are vacancies (see Figure 3.3). The stable phase is a complicated 21-cubic closely-packed layer structure. Matsunaga et. al. [61] suggested that the transition from the metastable to the stable phase occurs due to the movement of vacancies. Re-

8Differential Scanning Calorimetry 9X-Ray Diffraction

18 3.1 Background phase change alloy Ge1Sb2Te4

Figure 3.2: DSC measurement of an amorphous Ge1Sb2Te4 powdered specimen. The heating rate is 10 K/min over a range of 300-1000 K. Figure from [61]. cent detailed X-ray absorption spectroscopy (EXAFS10) measurements [62] and density functional theory (DFT) calculations [63] show that the structure of Ge1Sb2Te4 in the metastable phase differs slightly from the ideal rock salt structure. Here mainly Ge and Sb atoms deviate from their ideal lattice position. Welnic et. al. [63] calculated using DFT, that the ideal rock salt structure is energetically unfavorable compared to the dis- torted configuration, where Ge-, Sb-atoms and the vacancies deviate statistically from their ideal positions. According to DFT calculations this deviation is on averages 18 pm for Ge1Sb2Te4 [64]. For the amorphous phase the spinel structure appears to be an appropriate configuration, assuming that the Ge atoms are arranged tetrahedrally and the Te- and Sb-atoms are bonded octahedrally [63]. The high optical contrast between the amorphous and the crystalline phase is an out- standing property of GST. Infrared spectroscopy and spectroscopic ellipsometry for Ge1Sb2Te4 show that the optical dielectric constant is 118 % higher for the metastable crystalline phase than for the amorphous phase at an excitation energy of 0.05 eV [65]. This large difference is explained by the difference in binding configuration between

10Extended X-ray Absorption Fine Structure

19 3 Phase change materials

Figure 3.3: Model of the metastable rock-salt structure of Ge1Sb2Te4. The first sublat- tice is occupied by Te atoms. Ge atoms, Sb atoms and vacancies randomly occupy the second sublattice. The nearest-neighbor spacing is close to 3 Å. Figure from [? ]. the two phases. In the crystalline phase the bonds consist of p-orbitals. Since the p- level of the Sb atoms is only half filled, but Sb has six nearest neighbors, this leads to so-called resonance bonding. This can be explained as a superposition of two covalent bonding configurations. Due to this effect, the electrons are delocalized, which leads to an increased polarizability [65]. The main reason for the high optical contrast between the amorphous and crystalline phase is because there is no resonance bonding in the amorphous phase. Due to the distortion of the rock salt structure, which leads to peri- odic short and long bonds (Peierls distortion), the resonance bonding effect is reduced. The interplay between distortion and resonance bonding seems to be an intrinsic prop- erty of phase-change materials. With its help it is possible to classify different materials [66]. The mechanisms of the fast changing from crystalline to amorphous phase have still to be identified. However, one interesting proposal according to Kolobov et. al. [62] is that GST has an octahedral arrangement of p-like bonds in the crystalline phase. By changing the energy, for example using a laser pulse for the writing process, the structure is easily transformed from an octahedral arrangement into a tetrahedral con- figuration (amorphous phase). In this process the Ge atoms play an important role. In Figure 3.4 this transformation is shown. Here, the Ge atom is shown in the fcc

20 3.1 Background phase change alloy Ge1Sb2Te4

Figure 3.4: Fragments of the local structure of GST around Ge atoms in the crystalline (left) and amorphous (right) states. Stronger covalent bonds are shown as thick lines whereas weak interblock bonds are shown as thin lines [62].

lattice of Te atoms and occupies the symmetric octahedral and tetrahedral positions in the crystalline and amorphous phase, respectively. As shown on the left in Figure 3.4, the Ge atom has three strong and three weak covalent bonds. By transfering en- ergy to the weak bonds, they will break and the Ge atom jumps into the tetrahedral position (on the right in Figure 3.4) [62]. In addition to the optical contrast between amorphous and crystalline phase, Ge1Sb2Te4 has also a high difference in electrical conductivity between the two phases. This advantage can be used for non-volatile electronic data storage [67]. However, the origin of this large difference in conductivity for phase change materials is still not clearly understood. What is known from mea- surements and calculations is that the electrical resistance between the amorphous and crystalline phases varies by several orders of magnitude, depending on stoichiometry [58]. Switching from amorphous to crystalline phase can be realized with short current pulses. If the material is in the amorphous phase, a short current pulse is sufficient to achieve the transition temperature Tx (see Figure 3.2). The increased mobility of the atoms allow then the rearrangement into the crystalline structure. To characterize the electronic properties of Ge1Sb2Te4 scanning tunneling measurements provide the perfect tool. With the help of topographic and spectroscopic mode both structural and electronic aspects are accessible. Especially with the scanning tunneling spectroscopy, the position of the occupied and empty states can be determined with respect to the Fermi energy EF and with high spatial resolution.

21 3 Phase change materials

3.2 Scanning tunneling microscopy and spectroscopy of the phase change alloy Ge1Sb2Te4

In this chapter, scanning tunneling microscopy measurements of the phase change ma- terial Ge1Sb2Te4 are presented. The measurements were done in a modified Omicron STM at room temperature. The first section describes, how to prepare Ge1Sb2Te4 for surface sensitive STM and STS measurements. The other two sections 3.2.2 and 3.2.3 describe the analysis of the structure, band gap and Fermi level behavior in Ge1Sb2Te4.

3.2.1 Preparation of Ge1Sb2Te4

For the studies, a 100 nm thick amorphous film Ge1Sb2Te4 was used, which was de- posited by dc-magnetron sputtering on a silicon (100) substrate. The sample was pre- pared by the group of Prof. Dr. Wuttig. The dc-sputtered sample was mounted with tantalum metal sheets on a molybdenum sample holder and then transferred into the ultra high vacuum chamber. Since the sample was exposed to ambient conditions, the surface was cleaned by ion sputtering. Therefore, the sample was sputtered with a beam of 600 eV Ar+ ions at a chamber pressure of 1 × 10−5 mbar for 70 min with a sputtering rate of R = 0.0024 nm/s.11 The sputtering rate describes the average num- ber of surface atoms, which are emitted per incident ion, and can be calculated with the following equation: M R = · Y(Ei) · ji, (3.1) e · NA · ρ

with the molar mass M and the sample density ρ, the sputter yield Y(Ei) with beam energy Ei and the ion current density ji. The sputtering yield was determined using 16 2 Monte Carlo simulation [68] to Y(Ei) = 1.2. An ion flux of 3.2 · 10 ions/cm is mea- sured for a surface area of A = 0.35 cm2. The selected parameter leads to a thickness reduction of about 10 nm, so that after the cleaning, the thickness of Ge1Sb2Te4 is only about 90 nm. To check the contamination and stoichiometry of Ge1Sb2Te4, Auger electron spec- troscopy (AES) was done before and after each sputtering cycle. Figure 3.5 shows the measured Auger spectra of Ge1Sb2Te4 before and after the sputtering. The interesting part here is the energy range where the peaks for antimony (Sb) (at 454 eV), tellurium (Te) (at 483 eV) and oxygen (O) (at 503 eV) are located. The atomic concentration Xi of the elements i can be calculated with the following equation [69]:

Yi/Sidi Xi = , (3.2) ∑α Yα/Sαdα

11 sputter setting: acceleration voltage UB = 600 V, filament current If il = 4 A, emission current Iemi = 1.5 mA and ion current Ii = 0.5 µA

22 3.2 Scanning tunneling microscopy and spectroscopy of the phase change alloy Ge1Sb2Te4

a) Oxide (a.u) 100 nm Ge1Sb2Te4

XSb = 24 % X = 47 % Si(100) Te XO = 29 %

b) XSb = 33 % XTe = 62 % Ge Sb Te 90 nm XO ≤ 5 % 1 2 4 (a.u)

Si(100)

Figure 3.5: Auger spectrum of Ge1Sb2Te4 sample (a) before sputtering and (b) after sputtering. During the sputtering process 10 nm of the sample surface was removed. The rate of the atomic concentration of the elements are given in percent. The main adsorbates on the sample surface were oxides. The germanium-peak, which is detected at higher energies, is shown in spec- trum (b) inset.

23 3 Phase change materials

where Yi is the peak to peak intensity, Si the relative sensitivity of the elements i and di = Li · ip is a scale factor. Here Li is equivalent to the gain factor of the lock-in am- plifier and ip is the current flowing through the sample. The values for the relative sensitivity Si of the elements were taken from [69]. The spectrum of the non-sputtered sample (see Fig. 3.5 (a)) shows peaks for antimony, tellurium and oxygen. Here the largest contamination of the sample surface is oxygen (29 %). The stoichiometric ratio of 1 : 2 between antimony and tellurium is confirmed by the measurement with 24 % to 47 %. Figure 3.5 (b) shows the AES measurement of the sputtered sample. Note, that this measurement was done after STM and STS investigation to avoid contamina- tion (mainly carbon) due to AES measurement. The spectrum again shows antimony and tellurium peaks with a ratio of 33 % to 62 %. However, the oxygen peak is almost vanished and is determined to be less than 5 %. To check the overall stoichiometry of the sample, the germanium peak has to be detected at an energy of 1147 eV. Since the AES optics provide only a maximum electron energy of 1000 eV, an external high- voltage source was connected to the sample to shift the spectrum by the applied volt- age. The measured germanium peak is shown in Figure 3.5 (b) inset. After evaluation of the germanium peak in relation to the proportions of antimony and tellurium the stoichiometry of Ge1Sb2Te4 could be confirmed within an accuracy of 15 %, which is limited by the lower resolution at high voltages. This shows that the sputtering pro- cess does not affect the stoichiometry of the sample severely. Figure 3.6 shows two STM images before and after sputtering. In both images, the Ge1Sb2Te4 film is in the amorphous phase. Before sputtering, the sample is covered with 10-20 nm large grains (see Fig. 3.6 (a)). These have an average height of 2-3 nm. Since AES measurement shows that the sample surface contains 29 % oxygen, most of these clusters are prob- ably oxide compounds with the elements of the phase change material. After seventy minutes of sputtering of the sample, the majority of the oxide is removed. The surface shows strong irregularities with just about 3 nm large clusters (see Fig. 3.6 (b)). The ditches between each region are 15 nm deep. It is clear that the corrugation of the sam- ple has increased, due to the argon bombardment and has lead to a disordered surface structure. By comparing the roughness of both images, it is shown that the sputtered sample is approximately four times rougher. In summary, the Ge1Sb2Te4 sample is adsorbates-free, due to the sputtering process. The stoichiometry of the material still remains. Now, the sample is well prepared for further microscopic and spectroscopic measurements.

3.2.2 Scanning tunneling spectroscopy of Ge1Sb2Te4-sample

In this section, the scanning tunneling spectroscopy measurements on the Ge1Sb2Te4 sample are described. The aim was to determine the band gap in all phases of this material and to analyze the changes in conductivity. To obtain the various phases the

24 3.2 Scanning tunneling microscopy and spectroscopy of the phase change alloy Ge1Sb2Te4

a) RMS roughness: 0.92 nm

15 nm

T = 300 K

b) RMS roughness: 4.11 nm

3 nm

small cluster

T = 300 K

2 Figure 3.6: 200 × 200 nm STM images of Ge1Sb2Te4 with line profiles: (a) Image of a non-sputtered Ge1Sb2Te4 film in the amorphous phase (U = −2 V, I = 0.7 nA), (b) Ge1Sb2Te4 film after 70 min sputtering (U = −1.5 V, I = 0.4 nA). Morphological differences are clearly visible. cleaned sample was heated for several minutes at a certain constant temperature. The transition from amorphous to metastable phase occurs at about 390 K, from metastable to stable phase at about 500 K [61]. The heating of the sample was done using a re- sistive heater. The pressure in the chamber increased at maximum to 2 · 10−9 mbar during the last heating step at T ≈ 503 K. The heating temperature was measured with a thermocouple, which was located approximately 1 cm from the sample, so that the real temperature of the sample is approximately 10 % higher than the measured value. After heating, the sample was cooled down and the spectroscopic measure- ments were performed at room temperature. The size of the band gap can be directly measured from the dI/dU-spectra. To get the dI/dU-spectra, Lock-In technique with a modulation voltage of 20 mV and 30 mV were used. For each temperature, ten to twenty spectroscopy curves were measured and the average was calculated. In the spectroscopic measurements, a Pt-Ir-tip was used. The stabilization parameters U0

25 3 Phase change materials

and I0 of the average spectra were different. For the determination of the band gap, this has no relevance, since this only influences the absolute values of dI/dU curve, but not the size of the band gaps. Since the effects of scanner creep or thermal drift has an influence on the stability of the measurement, the time for recording a spec- trum was set to 5 seconds. We defined the band gap size by considering the apparent band gap, which is given by the area where the differential conductivity is zero. After- wards we added the energy resolution of 100 mV on both sides of the apparent band gap, to account the limiting energy resolution due to room temperature and modu- lation voltage (see 2.24). All values of band gaps in this chapter are effective band gaps, which is added on both sides with the energy resolution. Figure 3.7 shows the average dI/dU spectra of Ge1Sb2Te4 in different phases. Spectra (a) and (b) represent the band gap in the amorphous phase. The spectra (d) to (g) were measured at the metastable phase, and spectrum (h) shows the differential conductance in the stable crystalline phase. In addition, spectrum (c) is measured at about 293 K in the so-called mixed phase. In this phase, parts of the sample are still in the amorphous phase and the other parts are already transferred to the metastable crystalline phase. In addition to the band gap size the dI/dU spectra directly reveal also the position of the Fermi level EF with respect to the valence band edge EV and conduction band edge EC. As displayed in Figure 3.7, Ge1Sb2Te4 has a constant band gap of 0.65 eV in the amor- phous phase. The Fermi energy lies near the band gap center and at room temperature or after heating to 353 K the distance from the valence band edge is 0.26 eV. In the metastable crystalline phase, the band gap decreases from 0.49 eV (T = 408 K) to 0.40 eV (T = 473 K). In addition, the Fermi level moves towards the valence band edge. At the beginning of the metastable phase, the distance between EF and EV is 0.15 eV and towards the end of this phase it lies very close to the valence band edge. In the stable phase the tendency of reduction of the band gap size is continued. This is shown in the last dI/dU-spectra (Fig. 3.7 spectrum (h)), where the band gap disappears within the energy resolution, i.e. the gap is definitely smaller than 200 meV. Previous measure- ments on Ge2Sb2Te5 showed slightly different and partly contradictory results. Kato et. al. [70] derived from transport measurements a similar movement of the Fermi level from the mid gap towards the valence band including both phase transitions, thereby ignoring the possibility of closure of the band gap. In contrast, photoemission spectroscopy measurements on the same material deduced a larger Fermi level shift by 0.14 eV during the amorphous/metastable phase transition, but no shift during the metastable/stable phase transition [71]. Since Ge1Sb2Te4 and Ge2Sb2Te5 lie along the GeTe-Sb2Te3 pseudo-binary line, they have similar optical and electronic proper- ties according to DFT calculations and can be compared. For clarity Figure 3.8 shows the measured band gap and the Fermi energy with respect to the valence band as a function of temperature. Again the reduction of the band gap from the amorphous (0.65 eV) via the metastable phase (0.4 eV) to the stable phase (0 eV) (Fig. 3.8 (a)) can

26 3.2 Scanning tunneling microscopy and spectroscopy of the phase change alloy Ge1Sb2Te4

a) bandgap: 0.65 V b) bandgap: 0.65 V c) bandgap: 0.56 V

EV EF EL EV EF EL EV EF EL dI/dU (a.u.) dI/dU (a.u.) dI/dU (a.u.)

T = 293 K T = 353 K T = 393 K

sample voltage (V) sample voltage (V) sample voltage (V)

d) bandgap: 0.49 V e) bandgap: 0.44 V f) bandgap: 0.40 V

EV EF EL EV EF EL EV EF EL dI/dU (a.u.) dI/dU (a.u.) dI/dU (a.u.)

T = 408 K T = 423 K T = 453 K

sample voltage (V) sample voltage (V) sample voltage (V) g) bandgap: 0.40 V h) no bandgap

EV EF EL EF dI/dU (a.u.) dI/dU (a.u.)

T = 473 K T = 503 K

sample voltage (V) sample voltage (V)

Figure 3.7: Average dI/dU spectra of the different phases of Ge1Sb2Te4 sample. The sample was heated for several minutes to the temperatures T marked and spectroscopic investigation were performed after cooling back to room tem- perature. The dashed lines show the valance band edge EV and conduc- tion band edge EC. The Fermi level EF is at zero sample voltage. (a) T = 293 K, stabilization parameters (stab.): I0 = 0.5 nA, and U0 = −0.8 V. (b) T = 353 K, stab.: I0 = 1 nA, and U0 = −0.8 V. (c) T = 393 K, stab.: I0 = 0.4 nA, and U0 = −1 V. (d) T = 408 K, stab.: I0 = 0.4 nA, and U0 = −1 V. (e) T = 423 K, stab.: I0 = 0.5 nA, and U0 = −0.7 V. (f) T = 453 K, stab.: I0 = 0.45 nA, and U0 = −0.7 V. (g) T = 473 K, stab.: I0 = 0.5 nA, and U0 = −0.7 V. (h) T = 503 K, stab.: I0 = 0.5 nA, and U0 = −0.7 V.

27 3 Phase change materials

(a) amorphous metastable stable 0.7 0.6 0.5 0.4

0.3 0.2

band gap (eV) gap band 0.1 0.0 300 350 400 450 500 annealing temperature (K)

(b) amorphous metastable stable 0.30 0.25 0.20 (eV)

0.15 F 0.10 E-E 0.05 0.00 300 350 400 450 500 annealing temperature (K)

Figure 3.8: (a) Temperature dependence of band gap. The band gap decreases from 0.65 eV in the amorphous phase to 0 eV in the stable phase. (b) Temperature dependence of the Fermi level with respect to the valence band edge Ev. Within the crystalline phase EF moves continuously towards the valence band edge.

28 3.2 Scanning tunneling microscopy and spectroscopy of the phase change alloy Ge1Sb2Te4

be observed and EF moves from mid gap in the amorphous phase continuously to EV within the metastable phase (Fig. 3.8 (b)). However, this observation of EF move- ment with the annealing temperature is contradictory to the measurement of Siegrist et. al. [72]. They proposed that the conductivity variation with the temperature in phase change material is due to Anderson localization. A possible solution of this con- tradictory measurements could be that our measurement is surface sensitive and the conductivity measurement of Siegrist et. al. is bulk sensitive. In summary, with STS we could observe the variation of the band gap in the different phases and the position of the Fermi energy with respect to the valence band.

3.2.3 Structural analysis of Ge1Sb2Te4

In this section, the STM images of Ge1Sb2Te4 after the sputtering process are displayed. In Figure 3.9, two images of the amorphous phase are illustrated. Figure 3.9 (a) shows

a) RMS: 4.11 nm b) RMS: 0.51 nm 3 nm 8 nm

T = 300 K T = 388 K

2 Figure 3.9: 200 × 200 nm STM images of the clean Ge1Sb2Te4-surface in the amorphous phase. (a) The sample is in the initial state. The irregularities are from the sputtering process (compare with Figure 3.6 (a)). The cluster size is around 3 nm (see line profile) (U = −1.5 V, I = 0.4 nA). (b) The sample was heated to 388 K, which corresponds to a state just below the mixed phase. The size of the clusters is increased (about 8 nm) (U = −1 V, I = 0.5 nA).

an overview image of the unheated Ge1Sb2Te4 sample. The surface irregularities are caused by the argon impacts during the sputter process (see Figure 3.6 (a) for com- parison), which extend over the entire surface. The surface consists of many small clusters, which have a size of about 3 nm. The planes are separated by a few nanome- ter deep ditches (see line profile). The second image (Fig. 3.9 (b)) shows the amorphous phase close to the mixed phase (heated to 388 K). The grain size has increased slightly

29 3 Phase change materials

and the ditches have become less deep (see line profile). By comparing the RMS12 roughness, the structural change in the surface can be seen even more clearly. The roughness is reduced by almost a factor of ten to 0.5 nm compared to the unheated sample. The fact that the surface structure is changed, is already obvious in the amor- phous phase. Whereas the decrease in roughness between the two images is primarily through the annealing of the surface after sputtering. Figure 3.10 shows STM images of the metastable crystalline phase. The STM image in Figure 3.10 (a) was measured after annealing the sample to 408 K. The line profile shows that the grain size is al- most unchanged compared to the amorphous phase (Fig. 3.9 (b)). The clusters are slightly converged together leads to a flat film. The analysis of the roughness of the film confirmed this statement. Here, the RMS value has fallen again slightly from 0.5 nm to 0.38 nm. The next Fig. 3.10 (b) shows an STM image of Ge1Sb2Te4 heated to 473 K. The sample is at the upper limit of the metastable phase. The image shows an almost fully converged film with relatively large clusters. The roughness of the sample has increased slightly to 0.69 nm. This means, that the clusters have not only become larger in area, but have also increased in height (see line profile in Figure 3.10 (b)). The areas formed here are more homogeneous. The magnified image shows again the different clusters. All of them are greater than 10 nm. By increasing the cluster area within the metastable phase, it is possible to rearrange the vacancies in the material. This structural change could also affect the electronic properties. The shift of the Fermi level towards the valence band (see Figure 3.8 (b)) within the metastable phase could be explained by the continuous change of the structure (cluster size). To support this analysis, it requires more STM investigations at low temperature. Figure 3.11 shows the surface of the sample, which is heated to T = 393 K for 30 min. At this temperature, the transition between amorphous and metastable crystalline phases occurs. The STM image shows two different heights, suggesting that there are two phases involved. The height of the film differs by about 4.5 nm as demonstrated in the histogram shown in Figure 3.11 (b). If we consider that the initial Ge1Sb2Te4 film was 100 nm thick and after sputtering only 90 nm, this corresponds to a reduc- tion of the film thickness of 5 %. This is consistent with the previous studies [73], where they observed also a reduction of 5 % in the transition from amorphous to metastable crystalline phase. The reason for the height reduction is the different structures of both phases, thereby the mass density is increased in the metastable phase. From these studies, we conclude that in Figure 3.11 (a), the amor- phous areas appear bright and the crystalline regions dark. Ge1Sb2Te4 in its metastable phase has a rock salt structure, with the Ge and Sb atoms slightly displaced from their ideal lattice position [62, 63]. DFT calculations [74] show a dominant Te density of states for occupied states and a dominant Ge and Sb den- sity of states for unoccupied states. Therefore, it should be possible to image the Te

12Root Mean Square

30 3.2 Scanning tunneling microscopy and spectroscopy of the phase change alloy Ge1Sb2Te4

RMS:

a)

0.38 nm 8 nm z [Å] z

x [nm]

T = 408 K 8 nm

RMS: b)

0.69 nm 10 nm z [nm] z

x [nm]

T = 473 K

2 Figure 3.10: 200 × 200 nm STM images of the clean Ge1Sb2Te4-surface in the metastable crystalline phase. (a) The sample was heated to 408 K. Com- pared to the amorphous phase (T = 388 K) the cluster size remained al- most the same. However, they have moved closer together and have be- come significantly flat (U = −1.5 V, I = 0.4 nA). The Figure on the right shows an enlarged image (50 × 50 nm2, U = −0.65 V, I = 0.4 nA). (b) Image of the sample heated to 473 K. The size of the clusters is further in- creased (≥ 10 nm, U = −0.94 V, I = 0.75 nA). The right image shows an enlarged area (50 × 50 nm2, U = −0.94 V, I = 0.75 nA).

31 3 Phase change materials

a) b) crystalline

crystalline amorphous counts

amorphous

T = 393 K height (nm)

Figure 3.11: (a) STM image after subsequent annealing at 390 K for 30 min. Accord- ing to [61] the transition between amorphous and metastable crystalline phase occurs at this temperature. Both phases are shown in the image. The crystalline regions are characterized by 4.5 nm reduction in thickness (U = −1.3 V, I = 0.5 nA). (b) The histogram shows the mean height differ- ence between crystalline and amorphous phase of 4.5 nm.

atoms with a negative sample bias and the Ge atoms and Sb atoms with a positive sample bias. Figure 3.12 shows a large scale atomic resolution image of Ge1Sb2Te4 in the metastable crystalline phase. The image is differentiated along the x-axis in order to enhance the atomic contrast. Crystallites with different atomic orientation indicated by blue lines are visible. It is made sure that these lines have the same direction in images without contrast enhancement. Only the Te lattice structure is imaged, due to the positive sample bias. A stripe-like structure for different crystallites in Figure 3.12 suggests a preferential bonding along one direction. This is probably an intrinsic prop- erty of Ge1Sb2Te4. Since the stripes for different crystallites have different directions, it can not be caused by tip effects. X-ray diffraction (XRD) data reveals that 95 % of the crystallites exhibit a (111) surface orientation. Figure 3.13 shows the intensities of the reflected beam as a function of the detection angle 2Θ. The sample is in the stable crystalline phase during the XRD mea- surement. Figure 3.13 shows two different curves, recorded in different modes. The blue curve was measured with the same angle Θ with respect to the surface normal for incident and reflected beam. Thereby only the planes will be detected, which sat- isfy the Bragg condition and lie parallel to the sample surface. The black curve was measured with a constant 2◦ incident beam and moving the detector by an angle of 2Θ. This is similar to the Debye-Scherrer method, where all crystal planes satisfy the Bragg condition at certain angles. All reflections, which occur for this structure at dif- ferent angles, were calculated with the Bragg condition. For this calculation we take the structural parameter proposed for the 21-cubic close-packed layers at stable phase

32 3.2 Scanning tunneling microscopy and spectroscopy of the phase change alloy Ge1Sb2Te4

4 nm

Figure 3.12: Atomic resolution image of Ge1Sb2Te4 crystallites obtained after annealing the sample to 410 K (metastable phase). The image is differentiated along the x-axis to enhance the atomic contrast. by Matsunaga et. al. [61]. The software Diamond was used to determine the intensities of the reflected beam. The calculated reflections (black dots in Figure 3.13) are in good agreement with the measured peaks, besides a constant deviation for all calculated val- ues. The possible reflections for the metastable phase were also calculated (red circle in Figure 3.13). For the interpretation of the atomically resolved STM images, the deter- mination of the crystal planes at the surface (blue curve) is important. The blue curve shows a distinct peak at 26◦, its intensity is dominating compared to the other peaks. In the stable phase this peak belongs to the (0012) reflection, which corresponds to the (111) plane in the metastable phase. Another indication for the (111) plane is that there are reflexes in the black curve (marked with arrows), that occur only in the case of a (111) surface. After evaluation of the peak heights, it follows that the (111) reflection obtains 95 % of the total intensity and consequently the surface orientation exhibits 95 % (111) plane. Figure 3.14 (a) shows a higher resolution image of the Te atoms. The hexagonal pat-

33 3 Phase change materials

measured 2° incidence

measured Θin = Θout stable phase metastable phase Intensity (a.u.) Intensity

2Θ (deg)

Figure 3.13: XRD measurement of Ge1Sb2Te4 sample in the stable crystalline phase to determine the surface plane [75]. The blue curve shows a Θin = Θout scan, which detects all lattice planes parallel to the sample surface. The black curve shows all plane reflections, independently from their orientation to- wards the surface. The incident angle of the X-ray beam was constant at 2◦ during the scan. The calculated reflections, which should occur theo- retically, are shown together with their intensity for the stable (black dots) and metastable phases (red circles). tern of (111) plane is visible (compared with Figure 3.14 (d)) and is marked with gray dots. The deviation from an ideal hexagonal structure is probably due to thermal drift. The hexagonal Te pattern is not threefold symmetric, but seems to prefer one of the three orientations, i.e. a stripe-like appearance of the atomic structure implying that a preferential bonding along this direction is visible. The fact that rather different pref- erential orientations are observed in Figure 3.12 and that we always observe stripe structures using several different microtips, strongly suggests that the atomic surface arrangement of Te is not threefold symmetric. The mean row distance of 3.4 ± 0.2 Å, which is evaluated from several images, differs from the distance of the XRD mea- surement [61] (see Figure 3.14 (d)) by about 9 %. This deviation is probably due to the thermal drift, which leads to the distortion of the hexagonal structure. At posi- tive sample bias (see Figure 3.14 (b)) the second fcc sublattice is visible with Ge, Sb atoms or vacancies. In contrast to the image of the Te atoms, here a complex stripe structure is visible. The atoms are not clearly separable. All the positive sample bias

34 3.2 Scanning tunneling microscopy and spectroscopy of the phase change alloy Ge1Sb2Te4

a) b) 3.4 Å

2.4 Å 2.4 Å c) d) 3.7 Å

Ge/Sb Te

Figure 3.14: Atomically resolved STM images of Ge1Sb2Te4 in the metastable crys- talline phase at (a) negative sample bias (U = −0.4 V, I = 1.9 nA) and (b) positive sample bias (U = 0.4 V, I = 1.8 nA). Gray points mark the Te atoms in (a). The deviation in (a) from an ideal hexagonal structure is most likely caused by thermal drift. (c) Bulk atomic structure model of the metastable fcc Ge1Sb2Te4 [61]. The two planes indicate two adjacent (111) planes. (d) Surface atomic model of the (111) surface with lattice parame- ter according to XRD measurements [61].

35 3 Phase change materials

images show also a different preferred directions with distorted stripes, so that they are not caused by tip artifacts. Similar distortions especially for Ge and Sb atoms have been found by DFT calculations of Ge1Sb2Te4 and traced back to two different prefer- ential bond lengths [63]. Together with the vacancies, the distortions appear to be an intrinsic feature of phase change materials [64], which, in addition, employ "resonance bonding" [66], [65]. Thus, they are prone to local lattice distortions and electronic insta- bilities, frequently identified as Peierls distortions. It is obvious that these instabilities will affect the atomic arrangement. Hence the distortions in the atomic arrangement observed in Figure 3.14 (b) could be a direct manifestation of the electronic instabilities which characterizes phase change materials.

36 4 Spin polarized scanning tunneling spectroscopy of Fe nanoislands on W(110)

Spin polarized scanning tunneling microscopy and spectroscopy are used to study the behavior Fe nanoisland on W(110) in an external magnetic field. The first section 4.1 summarizes the previous spin-polarized scanning tunneling spectroscopy (SP-STS) work on Fe nanoislands on W(110) and discusses the different domain configurations of this system. Section 4.2 and 4.3 describe the preparation of magnetic tips and the nanoisland sample, respectively. The vortex core movement in in-plane magnetic fields measured by SP-STS is discussed in 4.4. In the last section 4.5, hysteresis curves in all three spatial direction were studied by SP-STS.

4.1 Fe nanoislands on W(110)

Using SP-STS Bode et. al. [76] studied the different domain configuration of Fe islands on W(110). The total energy of the ferromagnetic sample contains various energy con- tributions [77]: Etot = Eex + EK + Ezeeman + Ed + E0, (4.1) with the exchange energy Eex, the magneto-crystalline anisotropy energy EK, the magneto-static energy in external field or Zeeman energy Ezeeman and the stray field en- ergy Ed. The term E0 contains the magneto-elastic coupling energies as well as surface anisotropy energies. For a detailed explanation of the different energy contributions, I suggest the Ph.D.-work of A. Wachowiak [44]. Following the thermodynamic princi- ples, the ferromagnetic nanoisland aims the configuration with the minimum energy. The SP-STS measurements on Fe on W(110) system [76] show that the domain configu- ration strongly depends on the thickness of the island (see Figure 4.1). This is obvious, since for thicker islands the energy contribution of the W(110) substrate plays a smaller role than in thin islands. The mono-crystalline Fe islands with 3.5 nm thickness have a magnetization parallel to the surface plane, i.e. an in-plane magnetization. Figure 4.1 (a) shows a difference in the spin-polarized dI/dU signal between different islands. These are single-domain states with magnetization in two different directions. The

37 4 Spin polarized scanning tunneling spectroscopy of Fe nanoislands on W(110)

Figure 4.1: Topographic images (1st column), spin-resolved dI/dU maps (2nd col- umn), and topographic line sections (3rd column) of Fe islands on W(110) with different mean island heights h: (a) h = 3.5 nm, (b) h = 4.5 nm, (c) h = 7.5 nm, and (d) h = 4.5 nm. The measurement was done at T = 14 ± 1 K. In the 4th column the resulting island domain configuration are schematically represented. I = 0.5 nA, U = -400 mV (1st and 2nd row), U = 100 mV (3rd row), U = -350 mV (4th row). Figure from [76].

38 4.1 Fe nanoislands on W(110)

dark (bright) Fe island is always surrounded by a dark (bright) Fe monolayer. Since the easy axis of the Fe monolayer on W(110) is along [110¯ ] direction, the magnetic is- lands show in the same direction. For larger volume of the Fe islands, a multi-domain state becomes energetically favorable, due to decreasing of the anisotropy energy. Fig- ure 4.1 (b) shows an island with a two domain state and a height of d ≈ 4.5 nm. At even larger thicknesses (d ≈ 7 − 10 nm) more complicated structure are visible, such as the vortex state (Figure 4.1 (c)) or the diamond state (Figure4.1 (d)). The magnetization shows in [110¯ ]- and [001]-direction in these thick islands, i.e. in the direction of the easy axis of bulk iron [76]. A magnetic configuration with a ring closure, as in the case of the vortex state, is generally favored by a small magnetic anisotropy of the system [44]. Figure 4.2 shows such a vortex state of a thin disk. The magnetization rotates continu-

Figure 4.2: Schematic view of the vortex state in a thin disk. a) Continuous rotation of the magnetization in the plane at the center of the disk. At the edge the magnetization is parallel to the bounding surface. b) Detailed view of the vortex core: In a small area around the center the magnetization rotates synchronously out of the plane and directly in the center the magnetization is perpendicular to the surface plane (bright area). c) The four energetically degenerate vortex states: The magnetization in the center is aligned either parallel or antiparallel to the surface normal and the rotation in the plane has two options. Figure from [44].

ously in the plane around the disc core. At the edge the magnetization is parallel to the surface and it rotates out of plane in the core area. Directly at the vortex core the mag- netization is oriented perpendicular to the surface plane. By applying a small external perpendicular magnetic field, it is possible to widen or to narrow the vortex core [44].

39 4 Spin polarized scanning tunneling spectroscopy of Fe nanoislands on W(110)

If the direction of the external B-field is oriented like the core magnetization (see in Fig. 4.3 (b)) the vortex core at B = +800 mT), the vortex core becomes larger compared to the vortex core in zero field. For the antiparallel case (see in Fig. 4.3 (b)) the vortex core at B = −800 mT), the vortex core gets smaller compared to the zero field. As seen in Figure 4.3 (a) and (b), in this measurement the vortex core is also moved laterally. In an external field of −800 mT the vortex core is at the right side of the island. In the op- posite field of +800 mT, it takes a position at the left side. This is caused by the effect, that the external field was not exactly perpendicular to the sample surface, but had still additional in-plane field component. Our 4-Kelvin-System provides both, out-of-plane and in-plane magnetic fields. Hence it was possible to study vortex core movement in a purely in-plane magnetic field.

4.2 Preparation of tips for spin polarized scanning tunneling microscopy

In recent years, two methods have been established to produce tips for spin polar- ized measurements. The first one is to evaporate magnetic material on a etched tung- sten tip and the other one is to prepare the tip directly from bulk material. Therefore, chromium, which is an antiferromagnet, seems to be useful. Tungsten tips are often covered with thin films of iron, cobalt/gold, or chromium [78, 79, 44]. The preparation of an iron covered tungsten tip was done in the following way. The previously etched tungsten tip was transfered into the ultra high vacuum system and heated to 2000 K for few seconds using electron beam heater. The power was regulated manually till the color of the tip corresponds to the color of a tungsten crystal at the same temperature14. The tip radius is increased after this procedure [44]. Particularly the oxide and other impurities are removed by the heating. The result is a clean surface for the vapor deposition of magnetic material. The iron was deposited to the tungsten tip from a molecular beam evaporator13. The deposited amount of iron was 10 ML. The chamber pressure during evaporation was about 3 · 10−10 mbar. As a final preparation step, the film was annealed to approximately 700 K for 10 min. The orientation of the tip magnetization is comparable to the system Fe/W(110). So the local magnetization of the tip is parallel to the sample surface. A big disadvantage of these ferromagnetic Fe-tips is the dipolar interaction between tip and sample magnetization, which can lead to a switching of the sample or tip magnetization. To solve this problem, bulk tips are made from the anti-ferromagnetic chromium [51]. Therefore, a polycrystalline Cr bar was cutted from a 99.99 % Cr foil using a diamond saw. The tip was then formed

14 Power settings: acceleration voltage UB = 500 V and emission current Iemi = 80 mA. 13 Evaporator setting: acceleration voltage UB = 960 V, filament current If ill = 2.9 A and emission current Iemi = 9 mA.

40 4.2 Preparation of tips for spin polarized scanning tunneling microscopy

Figure 4.3: (a) dI/dU map of Fe island measured with a nearly out-of-plane sensitive chromium covered tungsten tip at zero field. (b) dI/dU map measured in a ext ext perpendicular field B = µ0H = −800 mT. (c) The inset shows the vor- tex core position for Bext = +800 mT. Tunneling parameter: U = −300 mV and I = 0.5 nA. The lateral movement of the vortex core is due to a small in- plane component of the external field. The in-plane component results from a non-perfect alignment of the perpendicular field relative to the surface normal of the island. Below: Schematic drawing of the lateral displacement of the vortex core. Due to the in-plane component of the external field, the outer domain with the magnetization parallel to the in-plane field is ener- getically favored. The domain grows in size and thus the vortex core moves laterally. Figure from [44].

41 4 Spin polarized scanning tunneling spectroscopy of Fe nanoislands on W(110)

by electrochemically etching the bar in 8 % NaOH with a dc voltage of 5 − 8 V. The Cr- tip was bonded to the tip-holder using two-component conducting ultra high vacuum glue. After transfering the tip into the STM, short voltage pulses up to 10 V and 12 hours field emission13 between tip and sample were applied for an in situ preparation. The magnetization of the tip has both in-plane and out-of-plane components.

4.3 Preparation of Fe nanoislands on W(110)

Before Fe can be evaporated on the substrate, the W(110)-substrate must be cleaned. The main impurity of a new tungsten crystal is the carbon on the surface. The W-crystal forms an W(110)/C-R(15 × 3)-tungsten carbide reconstruction. To clean the surface from carbon, a analogous method to that in [42, 44, 1] was used. First the W-crystal is annealed in an oxygen atmosphere of 1 · 10−7 mbar for 10 min at a temperature of 1500 K. During heating, the carbon segregates to the surface and reacts with the oxy- gen to form carbon-monoxide or dioxide. With a short heat-up to 2000 K, the molecules desorb from the surface, which results in a depletion region of carbon defects. Addi- tionally, on the surface tungsten oxide is formed. In contrast to the thermally stable tungsten carbide, tungsten oxide can be removed also by short heating about 2000 K for 5 − 10 seconds. The cycle of 10 minute annealing and flashing to 2000 K will be repeated for approximately one hour. After this procedure, the result is verified using electron diffraction (LEED) technique. Now the W(110) is ready for further prepara- tion. The iron was deposited from a molecular-beam evaporator14. Iron grows on W(110) in bcc structure. The lattice constant of tungsten aW = 0, 3155 nm is larger than that of iron with aFe = 0, 2866 nm. Nevertheless, iron grows pseudomorphic up to second monolayer (ML) on tungsten [80]. The large mismatch of nearly 10 % means, that the pseudomorphic growth is connected with stresses in the layer. The strain is reduced by dislocations in the iron layer. In larger iron layers islands are formed. They can reduce the high strains and the associated elastic energy by a lateral relaxation more effectively than a closed film. These three-dimensional islands are not formed automatically after the vapor deposition. The formation is blocked by heterogenous nucleation, e.g. step edges, impurities or dislocations [44]. By adding thermal energy, the iron particles can move and form islands. The temperature dependence of the size of the Fe-particles has been used to form compact lateral islands. Here, 5 ML iron was deposited on W(110) and annealed to 800 K for 10 minutes. This procedure results in islands of different sizes with heights ranging from 5 − 9 nm.

13Parameter for field emission: U = 80 V and I = 1 − 2 µA. 14 Evaporator setting: acceleration voltage UB = 960 V, filament current If ill = 2.9 A and emission current Iemi = 9 mA.

42 4.4 Vortex core movement in in-plane magnetic fields investigated by spin polarized scanning tunneling spectroscopy 4.4 Vortex core movement in in-plane magnetic fields investigated by spin polarized scanning tunneling spectroscopy

All measurement were done with a home-built scanning tunneling microscope op- erating in ultrahigh vacuum at 6 K inside a bath cryostat. An etched tungsten wire was used as the tunneling tip, covered with 10 ML iron. In order to characterize the tip/sample system, a single-domain magnetic Fe island was searched. The selection was made through the thickness of the island, i.e. searching for a thin island, which has probably single-domain magnetic configuration. Figure 4.4 shows such an island with a size of about 40 × 80 nm2 and an average height of 9 nm. The signal of the differential conductivity dI/dU was recorded as a function of an externally applied magnetic field at the point marked with a blue dot. The field direction was parallel to the image vertical (see blue arrow). The result is a hysteresis curve. At B = 0 T tip- and island-magnetization are aligned parallel. At a field strength of −70 mT and +55 mT an abrupt decrease of the dI/dU-signal occur, which can be explained by the switching of the tip magnetization. Due to the antiparallel alignment of tip and sam- ple magnetization the dI/dU signal is significantly smaller. A further increase of the field of about 10 mT results in the switching of the sample magnetization. This leads to an increase of the dI/dU signal again. We end up with a parallel spin orientation along the external field of tip and sample. This process demonstrates that it is a real spin-polarized signal. Now this tip-/sample-configuration was used to visualize more complex spin structures with spatial resolution. Therefore a larger Fe-island was investigated. Figure 4.5 (a) shows a Fe-island with a length of 350 nm and a width of 170 nm. In Figure 4.5 (b) the simultaneously recorded spin polarized image at 0.9 V is shown. The contrast of the brightness is equal to the angle between the magnetization directions of tip and sample (bright areas correspond to a parallel spin alignment, dark areas of an anti-parallel orientation). The island is investigated without an external magnetic field and shows a vortex structure. The next step was to manipulate the magnetic vortex with the help of an applied magnetic field in the sample plane. Figure 4.6 (a)-(d) shows field-dependent spin-polarized spectroscopic images. The green arrows indicate the direction of spin of the Fe island, the red arrows show the spin orientation of the tip. The yellow dot marks the vortex core. The contrast inverting between image (a) and (b) is caused by a flipping of the tip magnetization due to the stray field of the island. In addition, we see that the vortex core moves, as expected, perpendicular to the applied field (in the image to the right). At a field strength of B = 110 mT (Fig. 4.6 (d), the entire island- and the tip-magnetization is in the direction of the field. Now the tip was positioned next to the vortex core (see the white dot in Figure 4.7 (b))

43 4 Spin polarized scanning tunneling spectroscopy of Fe nanoislands on W(110)

[001] [1-10] B

0,00262 30nm 30nm

(a.u) 0,0026

dI/dU 0,00258

-0,1 -0,05 0 0,05 0,1

B (T)

Figure 4.4: Signal of the differential conductivity dI/dU as a function of applied field, measured on a single domain Fe island (inset: STM topography image of the Fe island). The red curve and the black curve indicate the direction of the magnetic field changes (see red and black arrows). The hysteresis effect is due to the relative magnetic orientation of the island and the tip. The red and green arrows show the orientation of the magnetization of tip and sample.

44 4.4 Vortex core movement in in-plane magnetic fields investigated by spin polarized scanning tunneling spectroscopy

a) b)

[001] [1-10]

Figure 4.5: (a) 300 × 300 nm2 STM topographic image of a Fe nanoisland on W(110). (b) Spin-polarized spectroscopic images (U = −0.9 V and I = 0.7 nA) of the same island. The image shows a vortex structure (spin direction with green arrows, the vortex core is marked with dotted circle). The red arrow indicates the direction of the tip magnetization. This in-plane tip direction is arbitrary and has been adapted to the green arrows, with the aim that in the case of parallel orientation of tip and sample magnetization the highest contrast is visible (bright signal on the island).

a) b) c) d) B = 0mT B = 40 mT B = 60 mT B = 110mT

Figure 4.6: Spin-polarized scanning tunneling spectroscopy images of a Fe nanoisland with a vortex magnetization structure at different external magnetic fields (the field direction is marked with a blue arrow). The magnetization direc- tion of the island is marked with green arrow and the magnetization of the tip is marked with red arrows.

45 4 Spin polarized scanning tunneling spectroscopy of Fe nanoislands on W(110)

Figure 4.7: (a) Spin-polarized dI/dU signal as a function of the magnetic field, mea- sured at the point marked by the white dot in (b). The arrows indicate the direction of the field change. (b) dI/dU spectroscopic image at the field of B = 20 mT applied in the direction of the blue arrow. The frequent change in contrast along a scan line (dotted lines) is due to change of the magneti- zation direction of the tip (red arrows). The area, where tip magnetization is changed by the stray field of the tip, is marked in a gray box in (a). The size of the box has been determined by analyzing spectroscopy images at different field values, not showing contrast inversion in consecutive scan lines above |B| = 50 mT.

46 4.5 Hysteresis properties of Fe nanoisland study by spin polarized scanning tunneling spectroscoopy

and the spin-polarized dI/dU signal is recorded field dependent. The result is shown in Figure 4.7 (a). Here we find also a hysteresis curve, but showing a very complex behavior. This is partly a result of changing spin orientation of the tip, which is in- fluenced by the stray fields of the magnetic island. This effect could be observed only in field strength between ±50 mT (gray framed area in Figure 4.7 (a)). Figure 4.7 (b) shows a spectroscopic image, where the tip has changed several times the magneti- zation direction (see red arrows). For field strength above 50 mT the dI/dU signal is dominated from the magnetization of the island, i.e. vortex core movement and the associated changes in spin orientation. For more detailed analysis of this movement, further experiments are necessary. These experiments must be performed with tips, which have a very small stray field, such as anti-ferromagnetic chromium-tip, to avoid tip-sample magnetic interaction.

4.5 Hysteresis properties of Fe nanoisland study by spin polarized scanning tunneling spectroscoopy

To avoid tip-sample magnetic interaction, an anti-ferromagnetic chromium tip was used (see section 4.2). Figure 4.8 (a) shows a Fe island with a size of about 80 × 220 nm2 and an average height of 6 nm. From the average height, we suggest that the island consists of a single magnetic domain. The chromium tip is stabilized on the Fe island at the blue dot, which is marked in Figure 4.8 (a). The differential conductivity dI/dU has been recorded as a function of an externally applied magnetic field for x-, y- and z-direction. The result is shown in Figure 4.8 (a). The Fe island shows hysteresis be- havior in all three spatial directions. As discussed in section 4.1, the magnetism of Fe on W(110) has a strong in-plane magnetic surface anisotropy, with an easy axis in [110] direction (x-y-plane). A detailed analyze of the coercive fields for x- and y-direction shows a small difference between those both directions. Using the average of coer- cive fields found in forward and backward scan, in x-direction the coercive field is determined to 45.5 mT and the coercive field in y-direction is 39.5 mT. Surprisingly the difference between those both directions is only visible at the positive magnetic field side. Calculation is necessary to clarify the cause of this phenomenon. The coercive field of the hard-axis (z-direction) is 109 mT. It is also not expected, that the hard-axis contains a hysteric loop. The reason for this hysteric loop remains unsolved here. From the dI/dU-amplitude of the curves in Figure 4.8 (b), it is possible to deduce the tip magnetization direction. Since the amplitude of the in-plane dI/dU-signals is al- most twice the signal of the out-of-plane dI/dU-signal, the magnetization of the tip should lie in the x-y-plane with a small component showing in the z-direction. The

47 4 Spin polarized scanning tunneling spectroscopy of Fe nanoislands on W(110)

(a) 50 nm

x [001] [110]- y

(b) x-Magnetic-field y-Magnetic-field z-Magnetic-field 2.00x10 4

12 mT 1.75x10 4 dI/dU (a.u.) dI/dU

1.50x10 4

-0.15 -0.10 -0.05 0.00 0.05 0.10 0.15 Magnetic Field (T)

Figure 4.8: (a) 80 × 220 nm2 STM topographic image of a Fe nanoisland on W(110). The x- and y-direction of the external in-plane magnetic-field are shown with the [001]- and [110]-axis. (b) Spin-polarized dI/dU signal as a function of the magnetic field for x-, y- and z-direction, measured at the point marked by the blue dot in (a).

48 4.5 Hysteresis properties of Fe nanoisland study by spin polarized scanning tunneling spectroscoopy

magnetic contrast of the dI/dU-signal can be calculated with [44]:

dI/dU − dI/dU A = max min , (4.2) dI/dUmax + dI/dUmin where dI/dUmax (dI/dUmin) is the value for the parallel (antiparallel) spin configura- tion of tip and sample. The value is A = 12 % for the in-plane signal and A = 6 % for the out-of-plane signal from Figure 4.8. The spin polarization of the Fe sample is 40 %. This leads to a 30 % spin polarization of the tip for the in-plane signal and 15 % for the out-of-plane signal.

49

5 Graphene

Since the discovery of 2D graphene by Geim and Novoselov [2] in 2004, this material had taken an extraordinary development. Graphene has already a big impact in the technological development towards possible application such as high frequency tran- sistors [10], supercapacitors [11] or touch screens [12]. A further interesting plan is to use graphene quantum dots (QDs) as spin qubits [13, 14]. Graphene might be an ideal candidate for qubits, due to his very long spin coherence time [15, 16, 17]. The first transport measurements on graphene QDs show some difficulties to control the electron number [24], caused by the fact, that graphene has no natural gap. Further, the 2D sublattice symmetry of graphene leads to a strong dependence of the electron transport on the edge orientation of the QD. As a result, disordered wave functions are predicted for the graphene QDs in line with results from transport spectroscopy [25, 20, 26, 27]. In this work, we investigated graphene QDs with well defined zig- zag edges supported on an Ir(111) surface. The first section 5.1 provides a review of graphene with a detailed analysis of transport through QD and graphene edge state. In section 5.2 the properties of graphene on Ir(111) substrate is discussed. Section 5.3 shows the experimental data of the wave function of the graphene quantum dots and the interaction of the iridium surface state with graphene. In the following section 5.4 the zigzag graphene edge on Ir(111) is investigated using spin-polarized scanning tunneling spectroscopy. The last section 5.5 deals with the mechanical instability of graphene naonislands on Ir(111).

5.1 Background graphene

In this section, the properties of graphene and graphene quantum dots are described. Subsection 5.1.1 gives an overview of the Dirac band structure of graphene. Due to the linear band structure, the electrons in graphene are pseudo-relativistic Dirac fermions. This is discussed in subsection 5.1.2. The subsection 5.1.3 provides a review of trans- port properties of graphene QDs. The final subsection 5.1.4 deals with the spin polar- ized edge state in graphene.

51 5 Graphene

5.1.1 Band structure of graphene

Graphene consists of sp2 bonded carbon atoms arranged in hexagonal structure with bond-length a ≈ 1.42 Å as shown in Figure 5.1. Graphene features two equivalents

b 1 AB k y

δ δ 3 1 K Γ a δ M 1 2 k x K’ a 2

b 2

Figure 5.1: Left: Graphene honeycomb lattice with the two triangular sublattices A and B. a1 and a2 are the lattice vectors. Right: Graphene Brillouin zone in mo- mentum space with the inequivalent corners at K and K’. b1 and b2 are the reciprocal lattice vectors. Figure from [3]. triangular sublattices A and B, which are distinguishable because√ of translational√ in- a = a ( ) a = a ( − ) variance. The lattice vectors can be written as 1 2 3,√ 3 and 2 2 3, √3 2π 2π [3]. The reciprocal lattice vectors are given by b1 = 3a (1, 3) and b2 = 3a (1, − 3) [3]. The interesting physical properties of graphene occur from the pz orbitals, which form the π band. In particular, this band has a linear dispersion at the corners of the first Brillouin-zone (1.BZ) at low energies. The occupied and unoccupied bands touch each other at the six corners of the hexagonal 1.BZ. The contact points are called Dirac points and are distinguished between two non-equivalent points K = ( 2π , √2π ) and 3a 3 3a K’ = ( 2π , − √2π ). The reason for the distinction between K and K’ is due to the trans- 3a 3 3a lation symmetry and leads to an additional degree of freedom and to a degeneration (valley degeneration) of the system. To calculate the band structure of graphene, the tight binding theory can be used. The calculation assumes one electron per carbon atom (the pz electrons) and a certain probability for the next nearest neighbor hop- ping and the second neighbor hopping. The band structure was already calculated by Wallace in 1947. As a result he gets [81]:

q 0 E±(k) = ±t 3 + f (k) − t f (k), (5.1)

52 5.1 Background graphene

Figure 5.2: Graphene band structure and enlargement of the band structure close to the K and K’ points. It shows the linear energy dispersion and the Dirac point at E = 0 eV. Figure from [3].

where t and t0 are the hopping energy between neighbor and second neighbor sites, respectively. The functional f (k) is defined as: √ √ 3 3 f (k) = 2cos( 3k d) + 4cos( k d)cos( k d). (5.2) y 2 y 2 x 0 Figure 5.2 shows the band structure E±(k) with t = 2.7 eV and t = −0.2t [3]. Since the occupied (electrons) and unoccupied band (holes) only touch at the six corners of the 1.BZ, graphene is a semi-metal. Only considering the low energy part of the dispersion (about −1 to 1 eV), the dispersion relation can be described by: q 2 2 E±(q) ≈ ±hv¯ F · qx + qy (5.3)

6 with q = k − K and vF ≈ 1x10 m/s is the Fermi velocity. The most striking difference between this result and the usual case, E(q) ∝ q2/(2m) where m is the electron mass, is that the Fermi velocity in 5.3 does√ not depend on the energy or momentum. In the usual case we have v = hk¯ /m = 2E/m and hence the velocity changes with energy [3].

5.1.2 Pseudo relativistic Dirac Fermions The linear behavior of the band structure near K and K0 gives us a 2D electron system with massless Dirac fermions. To describe this relativistic particles, the Dirac equa- tion is used. In the case of two dimensions, one can write the Dirac equation in the

53 5 Graphene following way [3]: ivFσ · ∇ψ(r) = Eψ(r), (5.4) with the two component electron wavefunction ψ(r) and Pauli matrices σ = (σx, σy). The two components of the wave function come from the distinction of the graphene lattice in two sublattices A and B. Equation 5.4 is closely related to Dirac’s equation for massless particles in the terminology of this relativistic quantum mechanical theory. The two components of the wavefunction are called pseudo-spin and the electrons and holes in graphene take the role of electrons and positrons. With the real spin of charge carriers in graphene, the full wave function consist of eight components: spin up and down for electrons and holes living in the sublattice A and B [3].

5.1.3 Graphene quantum dots This section is concerned with transport in graphene quantum dots. The motivation to use graphene quantum dots instead of other 2D-materials, like GaAs, is the expected long spin-coherence times due to the absence of hyperfine coupling and the small spin- orbit coupling in graphene [18, 19]. The long spin-coherence time is important for the idea to use graphene quantum dots as hosts for qubits based on spins [13, 14]. The weak hyperfine coupling in carbon materials can be explained with the nuclear spin free 12C isotope (99 %). The small spin-orbit interaction is due to the low atomic num- ber in carbon nuclei [82]. Before discussing the properties of graphene quantum dots, the transport in graphene nanostructures will be briefly discussed. Figure 5.3 (a) shows a sketch of contacted graphene Hall-bar device. By changing the carrier density via the back gate voltage Vbg and applying a current between source (s) and drain (d) the longitudinal resistiv- ity ρxx can be measured. The variation of the carrier density lead to a charge transport from hole to electron like transport by passing the so-called charge-neutrality (or Dirac- ) point at minimum conductivity [82] (see Figure 5.3 (b)). The offset of the Dirac-point from Vbg = 0 is due to residual doping of the graphene [82]. Figure 5.3 (b) shows a nearly linear behavior of the conductivity with the carrier density for electrons and holes. With the Drude Formula σ = nseµe the mobility for both carriers can be cal- 2 culated to µe ≈ 5000 cm /Vs [82]. In comparison to the mobility record values up to 200000 cm2/Vs [8, 9], this mobility value seems to be very low. The reason for the low mobility is charged impurities in SiO2 substrate and residues on graphene. Us- ing hexagonal boron-nitride as a substrate and removal of residues by annealing, large values of mobilities can be achieved [82, 83, 84]. In comparison to silicon, the big challenge for graphene devices is the fact that graphene does not exhibit a band gap. Part of the problem can be overcome by cut- ting (etching) graphene into narrow constrictions allowing to control the transmission via gating [82]. Such constrictions can then be used as tunable tunneling-barriers for

54 5.1 Background graphene

Figure 5.3: (a) Schematic of a contacted graphene Hall-bar device separated from the conductive silicon back gate by an insulating SiO2 layer. Longitudinal resis- tivity ρxx are measured in a four-probe geometry by driving a current from source (s) to drain (d). (b) Field effect i.e. conductivity as a function of back gate voltage Vbg at a fixed source-drain current Isd = 20 nA (T = 2 K). Text and Figure from [82].

quantum dots [82]. The band gap size can be approximated by 4Econ = 2πvFh¯ /3W [85]. For a nanoribbon with the constrictions width of W = 45 nm, the expected band gap size is 4Econ ≈ 30 meV. The band gap opening in graphene is strongly depending on the edge orientation of the ribbon and is discussed in the next section 5.1.4. Figure 5.4 (b) shows a conductance curve as a function of the back-gate voltage mea- sured in a graphene constriction. The width is W = 45 nm (see scanning force mi- croscopy image in Figure 5.4 (a) and provides a tunable tunneling barrier for quantum dots [82]. In contrast to the back gate characteristics of 2D-graphene shown in Figure 5.3 (b), the measurement shows many reproducible conductance fluctuations [82]. Fur- ther, transport between hole and electron conductance is suppressed within a voltage regime 4Vbg, denoted as a transport gap [82]. This gap is related to a gap in Fermi q energy 4EF ≈ hv¯ F 2πCg 4 Vbg/|e|, where Cg is the back gate capacitance per area [86]. In comparison to the band gap 4Econ ≈ 30 meV, the transport gap is significantly larger 4EF ≈ 110 − 340 meV. Figure 5.4 (d) shows the bias dependence of the conduc- tance suppression in the transport gap (white). In the magnification of the transport gap region (Figure 5.4 (e)) individual diamonds are visible [82]. This diamond shaped regions of suppressed conductance can be seen as a signatures of Coulomb blocked transport with a constant number of electrons on the dot. From the size of the dia-

55 5 Graphene

Figure 5.4: a) Scanning force microscope image of graphene constrictions with the width of Wmin = 45 nm. (b) Conductance as a function of back-gate volt- age measured at low source-drain bias (Vb = 300 µV). (c) Magnification of a sharp resonance recorded in the transport gap (see arrow in (b)). (d) Bias dependence of the conductance suppression in the transport gap (white). A close-up is shown in panel (e) where individual Coulomb diamonds are visible. Text and Figure from [82].

56 5.1 Background graphene

Figure 5.5: (a) Coulomb blockade diamonds (i.e differential conductance as a function of bias and gate voltage) for a small (W = 70) nm quantum dot. (b) Zoom around the diamond corner indicated by the arrow in (a). Quantization states are indicated by white arrows. Text and Figure from [82].

monds the charging energy Ec of the quantum dot can be extracted. In addition to these charge quantization effects, for small quantum dots devices d ≤ 200 nm, the energy quantization due to confinement becomes also observable in trans- port measurements at low temperatures [82]. To see the quantum confinement effect in graphene, one has to look closer to the Coulomb diamonds. The energy quantization manifests itself in additional lines running parallel to the edges of the diamond and faint lines at constant bias within the blocked region (see arrows in Figure 5.5) [82]. The quantization energy for large devices (W = 200 nm) is approximately 1 meV and increases for smaller devices to 3 meV (W = 50 nm). However, not only the quantization energy, but also the charging energy and the con- finement band gap scales inversely with the device size [82]. Therefore, the full un- derstanding of the energy spectrum in graphene quantum dots is not straight forward. Further, many questions remain open in this complex system, i.e., how the edge disor- der affects the transport properties, how to reduce bulk disorder effects or alternative ways to create a band gap [82]. In the next section, the edge configuration on graphene will be discussed in detail.

5.1.4 Edge state

As mentioned in the previous section 5.1.3 the transport properties of graphene quan- tum dots strongly depend on the edge configuration. Here, the different edge orien- tation and their properties are discussed. First calculations on graphene edge states were already performed in 1996 by Fujita et. al. [87]. To calculate the band structure,

57 5 Graphene

they used the one-orbital mean-field . This model considers only the 2 π-symmetric electronic states, which are formed by the unhybridized pz orbitals of sp carbon atoms [88]. For the next neighbor tight-binding Hamiltonian they used [88, 87]:

† H0 = −t ∑ [ciσcjσ + h.c.], (5.5)

† in which the operators ciσ and ciσ annihilate and create an electron with spin σ at site i, respectively. h.c. is the Hermitian conjugate counterpart and t ≈ 2.7 eV is the hopping integral, which defines the energy scale of the Hamiltonian. However, to describe magnetic properties theoretically, the electron-electron interaction must be considered. Within the Hubbard model this interaction is introduced through the repulsive on-site Coulomb interaction [88, 87]: 0 H = U ∑ ni↑ni↓, (5.6) i † where niσ = ciσciσ is the spin-resolved electron density at site i. The parameter U > 0 defines the magnitude of the on-site Coulomb repulsion. This model considers only the short-range Coulomb repulsion, two electrons interact only if they occupy the pz atomic orbital of the same atom [88]. Despite its apparent simplicity, this term is no longer trivial from the computational point of view [88]. The mean-field approxima- tion [88, 87] : 0 Hm f = U ∑(ni↑hni↓i + hni↑ini↓ − hni↑ihni↓i) (5.7) i allows to overcome this difficulty. Before we discuss the band structure calculations for large graphene flakes, we will introduce a very simple method to analyze magnetism in small graphene fragments. If we consider the Stoner-model of magnetism, it is necessary that at the Fermi level many states occur to get magnetism in the system. The number of states at the Fermi level for small nano fragments is given by the following equation [89, 88]:

η = 2α − N (5.8) where N is the total number of sites and α is the maximum possible number of non- adjacent sites, i.e. the sites which are not the nearest neighbors to each other [88]. Figure 5.6 (a) shows an example, where the non-adjacent sites are marked by circles. Although the theory from Fajtlowicz et. al. is able to predict the occurrence of zero- energy states, it is not clear how the electron spins align in these states. The comple- mentary knowledge is supplied by Lieb’s theorem [90] which determines the total spin of a system described by the Hubbard model [88]:

1 S = |N − N |, (5.9) 2 A B

58 5.1 Background graphene

a) N = 24 d) α = 12

b) N = 22 e) α = 12

c) N = 38 α = 20

Figure 5.6: Atomic structures and tight-binding energy spectra of graphene fragments. (a) , (b) triangulane and (c) bowtie-shaped fragment. Non-adjacent sites are labeled by circles. Empty and filled circles correspond to sublattice A and sublattice B, respectively. Tight binding energies are plotted as a function of band filling. The dashed line belongs to the energy spectrum of ideal graphene. (d) Mean-field Hubbard model calculations of the tri- angulane with local magnetic moments and spin-resolved energy levels. (c) Mean-field Hubbard model calculations of the bowtie-shaped graphene fragment. Area of each circle is proportional to the magnitude of the lo- cal magnetic moment at each atom. Filled and empty circles correspond to spin-up and spin-down densities. Figure taken from [88].

59 5 Graphene

where NA and NB are the numbers of sites in sublattices A and B, respectively. Three simple examples of nanometer sized graphene fragments are shown in Figure 5.6. The graphene fragment in Figure 5.6 (a) is called coronene molecule. The number of sub- lattices A and B is the same (NA = NB = 12) and the number of non-adjacent sites is therefore also α = 12. There are no states at the Fermi level, hence the total spin S is zero. The tight binding model predicts a wide band gap of 1.08t ≈ 3.0 eV for this graphene molecule. As expected, the mean-field Hubbard-model solution for this fragment does not reveal any magnetism [88]. The second graphene fragment (Figure 5.6 (b)) is called triangulane molecule, due to its shape. The sublattice A consists of NA = 12 sites and the sublattice B consists of NB = 10 sites. This means, that there are two states at the Fermi level and the total spin is S = 1. Hence the spin configuration must be a triplet state and has a ferromagnetic order. This could be confirmed by tight binding calculations (see Figure 5.6 (d)), where most of the spin-up electron density is localized around the zigzag edges of the molecule. One can see that spin polarization lifts the degeneracy of the zero-energy electronic states and opens an energy gap of 4S = 0.30t ≈ 0.8 eV [88]. Figure 5.6 (c) shows a graphene fragment, where two tri- angulane molecules are connected (called bowtie-shaped fragment). Here the number of sublattice sites A and B is NA = NB = 19. There are two states at the Fermi level and the total spin is S = 0. The spin configuration must be a singlet state and has an anti-ferromagnetic order. The tight binding calculation (see Figure 5.6 (e)) also shows that the two opposite spin states are spatially segregated in the two triangular parts of the molecule [91, 88]. Chemical derivatives of this triangular graphene fragments are already synthesized successfully [92, 93]. The spin-triplet ground state of these chemical compounds was verified by the electron spin resonance measurements [88]. In principle, this example can be considered as an indirect proof of edge magnetism in graphene systems, at least in finite fragments produced by means of the chemical bottom up approach [88]. Larger magnetic triangular molecules have not been synthe- sized so far [88]. Moving towards larger graphene fragments or infinite systems, the application of counting rules becomes impractical [88]. An alternative approach considers the ef- fects of edges of graphene nanostructures which can be conveniently modeled using one-dimensional periodic stripes of graphene [88]. Figure 5.7 shows the tight binding calculations for the two high-symmetry crystallographic directions in graphene (arm- chair and zigzag). For the calculation, they assumed that the dangling bonds at the edge are all terminated by atoms. Only the zigzag graphene nano-ribbons feature a flat band extending over one-third of the one-dimensional Brillouin zone. Figure 5.8 (a) shows the magnetic configuration of this state, if the number of sublat- tice sites are equal (NA = NB). For the same number of sublattices Lieb’s theorem predict a total spin of S = 0 and an anti-ferromagnetic mutual orientation of the edge spins. In the case of an anti-ferromagnetic inter-edge, the band-structure calculation

60 5.1 Background graphene

Figure 5.7: Atomic structures and tight-binding bandstructures of (a) armchair and (b) zigzag graphene nanoribbons. Figure from [88].

61 5 Graphene

Figure 5.8: (a) Mean-field Hubbard model calculations for local magnetic moments in a zigzag graphene nanoribbon. Area of each circle is proportional to the magnitude of the local magnetic moment at each atom. Filled and empty circles correspond to spin-up and spin-down densities, respectively. (b) Mean-field Hubbard model band structure (solid lines) compared with the tight-binding band structure (dashed lines) for the solution in (a). The band structure for spin-up and spin-down electrons are equivalent. (c) Mean- field Hubbard model band structure for the same graphene nanoribbon with the ferromagnetic mutual orientation of the edge spins. The band structure for the majority-spin electrons and the minority-spin electrons are distinguished by colors. Text and Figure from [88].

62 5.2 Background graphene on Ir(111) shows a band gap (Fig. 5.8 (b)), due to the electron-electron interactions [88]. Compare to this semiconductor behavior, one gets for the ferromagnetic inter-edge ordering a semi-metallic ribbon. Figure 5.8 (c) shows that in this case the majority-spin and the minority-spin bands cross the Fermi level at k = 2π/3a [88]. So far there is no direct experimental evidence of magnetism in graphene nano-ribbons. However, the experiment by Tao et al. [94] seems to be very promising: They measured a spin-split state at the edge of chiral graphene nanoribbons on Au(111). The measured spin-splitting of 23 meV is very close to the theoretical value of 29 meV. A very inter- esting future experiment would be a magnetic-field dependent measurements of this system.

5.2 Background graphene on Ir(111)

This section give an overview of the system graphene on Ir(111). The preparation of graphene on Ir(111) is discribed in subsection 5.2.1. In the following subsections 5.2.2 and 5.2.3 the structure of graphene on Ir(111) and its band structure is discussed, respectiveley.

5.2.1 Preparation of graphene on Ir(111) All the following preparations were done in ultra high vacuum at the basic pressure of 2 x 10−10 mbar . The Ir(111) substrate was cleaned by repeated cycles of annealing to 1500 K and Argon sputtered with a beam energy of 2.5 keV at 300 K. Fig. 5.9 shows an STM topography image of the clean Ir(111) substrate and the atomic configuration of the fcc surface. Graphene was prepared using chemical vapor deposition [95]. There- fore, the substrate was exposed to > 99.95% pure ethylene gas (C2H4) for 4 min at a chamber pressure of 1 x 10−7 mbar (see Figure 5.10). After annealing at 1320 K for 30 s the graphene covers around 22 % of the substrate surface. Figure 5.11 (a) shows an STM images of a hexagonal graphene nano-islands with a typical lateral dimension of 10-40 nm. The atomically resolved images (Figure 5.11 (b) and (c)) reveal the carbon atoms arranged in a hexagonal lattice and the complete enclosure of the quantum-dot by zigzag edges. A nice summary of different preparation methods of graphene on Ir(111) can be found in the Ph.D.-thesis of N’Diaye [96].

5.2.2 Structure of graphene on Ir(111) The most striking feature of graphene on Ir(111) is the presence of a superstructure (see Figure 5.12). This moiré-like structure is due to different lattice constants of graphene and Ir. The graphene and Ir lattice constant is 2.45 Å and 2.72 Å [96], respectively, leading to a lattice mismatch of about 10 %. The vector of the superstructure unit cell

63 5 Graphene

1 nm

U = 1 V, I = 0.3 nA

Figure 5.9: (100 x 100) nm2 STM image of the clean Iridium substrate. The inset shows the atomic configuration of the Ir(111) surface (U = 0.9 V, I = 0.5 nA). has a length of 25.2 Å [96]. One closer look at the structure of graphene on Iridium identifies three specific regions (see Figure 5.13). The regions fcc, hcp and atop differ by the arrangement of the carbon atoms with respect to the underlying surface iridium sites [96], named for whether an fcc-, an hcp-hollow, or an Ir atom shows through the local carbon hexagons. In his Ph.D.-thesis, N’Diaye identifies the dark regions in the moiré-structure as the atop region. In Figure 5.13 this atop positions are marked as red triangular. Further density functional theory calculation [97] shows that in the atop positions the carbon to Ir distance is largest (3.62 Å). The lowest height was found for the hcp region with 3.27 Å, slightly lower than the one of 3.29 Å in the fcc region. The structural inhomogenity of the moiré structure also leads to different chemical binding energies. The hcp and fcc regions behave similar: both can harbour iridium clusters, as shown in post decoration experiments [96]. The local binding energy minimum for evaporated iridium clusters on graphene/Ir(111) is deepest in the hcp regions [96].

64 5.2 Background graphene on Ir(111)

Preparation

1) C H 2 4 4 min

10-7 mbar

4 4

4

4

4

4

4 4

H H

H

H

H

H

H H

2 2

2

2

2

2

2 2

C C

C

C

C

C

C C Ir(111) 300 K 2) H H 2 H2 2 H

2 H2

C C

C

C

C

C C C Graphene Ir(111) 1320 K

Figure 5.10: Illustration of the Graphene preparation: 1) Ethylene gas (C2H4) is ex- posed to the Ir(111) substrate at room-temperature for 4 min. 2) After an- nealing the ethylene/Ir(111) system to 1320 K for 30 sec the hydrogen (H2) is vaporized and a monolayer graphene remains on the substrate.

However, at the atop regions they never found iridium clusters. The nucleation of iridium clusters on this regions seems to be energetically unfavorable. This behavior could be explained with small charge transfer from the fcc and hcp carbon atoms to the iridium atom sitting under the atop position. This lead to a weak covalent bond formation between atop carbon atom and the atop Ir atom [97]. Thus the charge deficit of the neighbors (fcc- and hcp-carbon atoms) explains their tendency to bind addition- ally deposited metal atoms [97]. Core-level photoelectron spectroscopy in combination with ab initio calculation show that the shape of the graphene quantum-dots is dome like [30]. The dome like shape leads to a strong coupling of the edge carbon atoms to the iridium and to a quasi-free- standing graphene layer between the edges. The carbon-iridium distance at the edge is only 1.62 Å. In the center of the island the distance is 3.13 Å. This large graphene- Ir(111) separation also explains the defect free growth of the graphene flakes over step

65 5 Graphene

a) c)

Ir(111)

graphene b)

Figure 5.11: (a) (100 x 100) nm2 STM image of Ir(111) surface covered by monolayer graphene islands (U = −0.3 V, I = 0.3 nA). (b) atomically resolved (12 x 12) nm2 image of graphene island. (c) Magnified view of zigzag edge with a ball model of the graphene lattice (U = 0.7 V, I = 20 nA). edges [98]. And compared to other epitaxially grown systems like graphene on Ni(111) [99, 100] and Ru(0001) [101, 102, 103], the large separation in this system corresponds to a weakly interacting graphene.

5.2.3 Band structure of Ir(111) and graphene on Ir(111) The most interesting aspect of graphene on iridium is the electronic band structure. Angle resolved photoemission spesctroscopy (ARPES) measurements show that the electronic bands near the Fermi level of graphene on Ir(111) largely have the form of the Dirac cone of pristine graphene, only slightly shifted by 0.1 eV to lower binding energies due to a marginal p doping by the substrate and having a small gap of about 0.2 eV [31, 104]. The presence of the Dirac cone can be ascribed by the existence of a band gap around the K point of the surface BZ of Ir(111). In the ARPES spectrum of clean Ir(111) (see Figure 5.14 (a)) the energy gap is visible in the K-M area. The energy gap is enclosed by three surface states S1, S2 and S3. DFT calculation has shown that a large fraction of the S1 surface band is above the Fermi level. Further the calculation show that the S2 and S3 bands are spin orbit split surface states with a spin-orbit split- ting smaller than 0.3 eV [105]. The ARPES spectrum (Figure 5.14 (b)) of Ir(111) covered by graphene shows the presence of the Dirac cone. Besides the primary Dirac cone in Figure 5.14 (b), the data displays an additional band, a faint replica labeled R [31]. The existence of the replica cones can be explained by the moiré structure. The horizontal arrows in Figure 5.14 (b) denote the position, where the replica cones intersects the dirac cone and open an energy mini-gap. The measured gap width is in the range of

66 5.2 Background graphene on Ir(111)

50Å

Figure 5.12: Moiré structure of graphene (right side) on Ir(111) with the superstructure unit cell (marked as a white rhombus). U = −0.17 V, I = 21 nA. Figure from [96]

67 5 Graphene

Iridium

fcc hcp atop 25Å Graphene

Figure 5.13: Enlargement of the STM image in Figure 5.12 of a graphene flake (right side) attached to an iridium terrace. Here the Ir-grid (deduced from the Ir atom position on the left side) is extrapolated over the whole image. The different stacking types are marked by green circles (fcc), blue triangles (hcp) and red triangles (on-top). Sketch on the right illustrates the fcc, hcp

and on-top stacking. Figure from [96].

)

eV

(

F

E

- E

Γ kII (Å) M

Figure 5.14: (a) ARPES spectrum of clean Ir(111). The position of K points of iridium and graphene are marked as KIr and Kg, respectively. S1-S3 are surface states. (b) ARPES spectrum of Ir(111) covered by graphene along the same polar angle as in (a). The mini-gaps are denoted with horizontal arrows. A visible replica band is labeled as R. Text and Figure from [31].

68 5.3 Wave function mapping in graphene quantum dots

Figure 5.15: Dispersion of the spin-orbit split surface state around Γ-point on (a) bare Ir(111) and (b) graphene grown on Ir(111). Figure from [106].

0.1 − 0.2 eV. Apart from the interesting area around K − M in the Brillouin zone, there is another surface state of Ir(111) at the Γ-point (see Figure 5.15 (a). This state with sp-symmetry is spin-polarized and is shifted by 150 meV towards the Fermi energy for graphene covered Ir(111) (see Figure 5.15 (b)).

5.3 Wave function mapping in graphene quantum dots

In this section, the wave function of scanning tunneling spectroscopy measurements of the graphene quantum dots and the interaction of the iridium surface state with graphene are discussed. Tight binding and density functional theory (DFT) calcula- tions are used to describe the local density of states of graphene dots and the intrusion of the iridium S0 surface state into graphene. The tight binding calculations were done by Raphael Reiter and Florian Libisch from TU Vienna. The DFT calculations were per- formed by Yan Li and Riccardo Mazzarello from RWTH Aachen University. The first subsection 5.3.1 describes the scanning tunneling spectroscopy data of the eigenstates of the graphene quantum dots. The tight binding and DFT calculations are discussed in subsection 5.3.2. The next subsection 5.3.3 compares the scanning tunneling spec- troscopy data with the theoretical model. Subsection 5.3.4 describes the Ir surface state S0. The final subsection 5.3.5 deals with the lifetime of quantum dot states.

69 5 Graphene

5.3.1 Scanning tunneling spectroscopy measurement of eigenstates on graphene quantum dots (QDs) All measurements were done with a home-built scanning tunneling microscope in ul- trahigh vacuum at 6 K inside a liquid helium bath cryostat. The local density of states (LDOS) of 15 islands is mapped by STS with a modulation amplitude of Umod = 10 mV. This results in an energy resolution of 25 meV [57]. Figure 5.16 shows an STM im-

U= -190 mV, I= 200 pA U= -190 mV, I= 300 pA U= -146 mV, I= 300 pA

7.5nm a)8.8nm b)2.2nm c) U= -107 mV, I= 300 pA U= -197 mV, I= 500 pA U= -360 mV, I= 500 pA

6.5nm d)10nm e)2.5nm f) U= -200 mV, I= 200 pA U= -280 mV, I= 300 pA U= -200 mV, I= 200 pA

6.1nm g)1.8nm h)3.0nm i) U= -800 mV, I= 200 pA U= -206 mV, I= 200 pA U= -235 mV, I= 300 pA

1.4nm j)6.0nm k)4.5nm l)

U= -250 mV, I= 300 pA U= 57 mV, I= 100 pA U= -320 mV, I= 1000 pA

3.8nm 8.0nm 2.7nm m) n) o)

Figure 5.16: Constant current STM images of investigated graphene quantum dots (QDs). Tunneling parameters are marked in the images.

age gallery of all investigated graphene QDs on the Ir(111) substrate. For dI/dU(U)

70 5.3 Wave function mapping in graphene quantum dots

curves, we stabilize the tip at sample voltage Ustab and current Istab. Figure 5.17 (a)

a) 150

100 8 nm

(arb. units) exp. 50 dI/dU theory 0 1 1 2 1 1 2 -0.7 -0.6 -0.5 -0.4 -0.3 -0.2 -0.1 E (eV)

b) exp. c) d)

-0.26 eV -0.42 eV -0.63 eV

e) theory f) g)

-0.32 eV -0.44 eV -0.66 eV

Figure 5.17: (a) Black line: dI/dU(U) curve spatially averaged over the graphene QD shown at the right; Ustab = 0.5 V, Istab = 0.5 nA, Umod = 10 mV; grey line: DOS(E) of the same island as obtained by tight binding calculation (see text); vertical bars mark the calculated eigenstate energies with degen- eracies indicated as numbers. (b)-(d) dI/dU images recorded at energies E = U · e as marked; I = 0.2 nA; Umod = 10 mV. (e)-(g) LDOS maps calcu- lated with soft edge potential at energies indicated. shows a dI/dU(U) curve (black line) laterally averaged over the hexagonal QD shown on the right. The dI/dU(U) curve displays three maxima with the most pronounced at −0.25 eV, below the Fermi energy EF. Figure 5.17 (b)-(d) show dI/dU maps recorded at the energies of the maxima. For the first peak (U = −0.26 V), one maximum of the LDOS in the center of the island is found. At U = −0.42 V, a ring shaped structure appears. The state at U = −0.63 V shows a maximum-minimum-maximum sequence from the center towards the rim and an additional star-shaped angular dependence. It was checked that no other LDOS shapes are present in the voltage range of −1.4 V to

71 5 Graphene

0 V. The peaks are below the Dirac point ED [31], and, thus, they belong to hole states. The order of observed shapes suggests that the peaks represent confined states of the QD. To support this conjecture, tight binding and density functional theory (DFT) cal- culation are employed. A more detailed description of the theoretical model follows in the next section.

5.3.2 Theoretical model The following Hamiltonian is used for the third-nearest neighbor tight binding approx- imation [107, 108, 109]:

H = ∑ |φi,si Vi hφi,s| + ∑ γ(i,j) |φi,si φj,s + h.c.. (5.10) i,s (i,j),s

The γ(i,j) are hopping amplitudes between sites i and j being γ(i,j) = (3.14, 0.042, 0.35) eV for the coupling to the (first, second, third) nearest-neighbors [107]. The Vi represent local on-site potentials, into which is included the moiré po- potential 200 meV

0 meV 0 1 2 3 4 5 6 7 lateral position (nm)

Figure 5.18: Map of the moiré potential used in the tight binding calculation of Figure 5.17.

tential displayed in Figure 5.18. Using a spatially constant Vi within the islands, i.e. hard-wall-confinement, regular and very irregular wave functions result as shown in Figure 5.19 (e)-(h). The irregular wave functions often display a large intensity at the rim of the QDs and illustrate the sensitivity of graphene QDs to details of the edge

72 5.3 Wave function mapping in graphene quantum dots

a) b) c) d)

soft confinement

-0.44 eV -0.46 eV -0.57 eV -0.67 eV

e) f) g) h)

hard confinement

-0.40 eV -0.50 eV -0.56 eV -0.66 eV

Figure 5.19: (a)-(h) Calculated LDOS = |Ψ|2 for individual confined states with en- ergies marked: (a)-(d) with soft edge potential; (e)-(h) without soft edge potential. configuration [13, 14, 25]. These irregular shapes, however, were never found in the present STS experiments featuring about 50 different states. To solve this puzzle, a spatially inhomogeneous Vi is deduced from ab initio DFT cal- culations. The DFT simulations were carried out using the plane-wave PWSCF code included in the QUANTUM-ESPRESSO package [110]. The band structure for differ- ent graphene-Ir surface distances D was determined, starting from the known distance between extended graphene and Ir(111), D = 3.4 Å [111], towards the shortest dis- tance found at the edge of a graphene island, D = 1.6 Å [30]. A proper description of Ir(111) surface states requires thick slabs which makes it unfeasible to use the large 10 × 10 supercell necessary to account for the graphene-Ir lattice mismatch. Instead, a slightly compressed Ir(111) surface matching the cell size of graphene is considered allowing to use a slab of 24 Ir layers with graphene on both sides and a vacuum space of 20 Å between the slabs. A gradient-corrected exchange correlation functionals [112] and fully-relativistic ultrasoft pseudopotentials including spin-orbit interactions is in- cluded [113]. The wave functions are expanded in plane waves with a kinetic energy cutoff of 30 Ry and a charge-density cutoff of 300 Ry. 20 × 20 × 1 Monkhorst-Pack meshes [114] of k-points are used for the integration over the Brillouin Zone. The electronic band structure of the system was computed along the T¯, T¯ 0 and Σ¯ lines of the surface Brillouin zone, corresponding to the path Γ¯ − K¯ − M¯ − Γ¯ in the recip- rocal space. The graphene states were identified by projecting the wave functions of the slab on the atomic wave functions centered on the C atoms with a threshold of 50 %. With this assumption it is possible to calculate the gap 4E between the lower and upper graphene cone as a function of the graphene-Ir distance. The insets in Figure

73 5 Graphene

5.20 exhibit the resulting band structures for two different D. The energy gap at ED is marked. Its size 4E(D) is plotted in the main image increasing with decreasing D. To model this band gap within the tight binding calculation, a Berry-Mondragon like potential Vi is used [25]: Vi = 4E(D(ri))/2 · σz, (5.11) where the Pauli matrix σz acts on the sublattice degree of freedom. A homoge- neous Vi would open a gap of size 4E at ED. The functional form of 4E(D) =

1.2 (eV) k F E-E 0.8

(eV) E (eV) ∆ k 0.4 F ∆E E-E

0.0 0.24 0.28 0.32 Graphene-Iridium distance D (nm)

Figure 5.20: Energy gap 4E versus graphene-Ir distance D as deduced from DFT cal- culations. Insets: band structure around ED for two different D as marked by arrows with 4E indicated. Grey area: projected bulk bands of Ir. Thick black lines: graphene states.

(0.7 · (3.6 − D[A˚])2 + 0.23) eV is taken from the fit to the DFT calculations (see Figure 5.20). There are two contributions to the spatially varying Vi: (i) the entire graphene flake features a moiré type corrugation leading to mini gaps of about 200 meV (see section 5.2.3), which can be reproduced by an amplitude of Vi of 400 meV, and (ii) the graphene flake bends downward from D = 3.4 Å in the center of a QD to D = 1.6 Å at its rim [30, 111]. The discrepancies between photoemission and DFT bandgap are prob- ably caused by the compression of the Ir(111) within the DFT calculations, but does not affect the general trend of increasing the band gap with shortening the graphene-Ir dis- tance. For the sake of simplicity, D(r) is interpolated linearly from the rim towards 10 Å inside the island as suggested by the DFT calculations of [30]. To assert that this choice does not influence the conclusions, calculations for several different functional forms of the distance between the quantum dot and the Iridium surface, keeping the outmost distance of 1.6 Å fixed, are performed. Excluding unphysical, vertical kinks

74 5.3 Wave function mapping in graphene quantum dots

a) b) c)

-284 meV -406 meV -533 meV d) e) f)

-282 meV -400 meV -525 meV g) h) i)

-282 meV -400 meV -525 meV

Figure 5.21: The three lowest-lying eigenstates with their eigenenergies marked for different parametrizations of the distance development between the graphene quantum dot and the iridium substrate at the edge of the quan- tum dot: (a)-(c) linear interpolation ∝ x over 15 Å, i.e. D(r⊥) = 3.4 − 0.12 · (15 − r⊥) · θ(15 − r⊥) with r⊥, D in Å and θ(x) being the step function. 3/2 3/2 (d)-(f) x over 10 Å, i.e. D(r⊥) = 3.4 − 0.057 · (10 − r⊥) θ · (10 − r⊥). 2 (g)-(i) quadratic interpolation (∝ x ) over 10 Å, i.e. D(r⊥) = 3.4 − 0.018 · 2 (10 − r⊥) ··(10 − r⊥). In Figure 5.17 a linear interpolation over 10 Å was used.

75 5 Graphene

in the shape of the graphene flake, no noticeable changes in the wavefunction patterns were found (see Figure 5.21). The variations of the calculated resonance energies for different types of edge potentials are below 15 meV, which is smaller than the experi- mental energy resolution. Finally, the peak width Γ of the eigenstates is adapted to the experiment leading to Γ(E) = 0.33 · |E|.

5.3.3 Comparison between theory and experiment

The calculated DOS(E) for the hexagonal graphene dot is added as a grey line in Figure 5.17 (a). The calculated LDOS maps at the energies corresponding to the experiment are shown in Figure 5.17 (e)-(g). Both, LDOS curve and LDOS maps exhibit excellent agreement with the experimental data. Notice that the calculated energies are shifted to account for the graphene doping on Ir(111). In Figure 5.17 (b) the two states belong- ing to K and K0 and looking nearly identical are superposed. In Figure 5.17 (c) also two states are superposed. In Figure 5.17 (d) already 4 states, which have a very similar energy, had to be superposed. The multiplicity of the calculated states is greater than the number of observed peaks for two reasons: (i) for every observed state, two eigen- states localized at the K and K’ cone were found, and (ii) due to the linearly increasing DOS in graphene, experimental peaks at high energies are, in our theoretical calcula- tions, created by superposition of several eigenstates of different wave functions. In- deed, the good agreement between theoretical wave function shapes and experimental LDOS measurements at higher energy could only be achieved by superposing up to 18 individual eigenstates in an energy window of 50 meV. The theoretical K-K’ splitting between these states is of the order of 1 meV, due to the symmetry-conserving bound- ary potential and thus smaller than the experimental energy resolution. This suggests that K-K’ scattering is sufficiently suppressed in favor of scattering into the iridium substrate in experiment to keep the splitting below the experimental energy resolution of 1.4 meV. More importantly, the modified calculation only reveals states reflecting the hexagonal symmetry of the QD shape in agreement with experiment, but none of the irregular states found without smooth confinement. This increased geometrical sym- metry is illustrated in Figure 5.19, comparing single state wave functions of the same quantum dot with soft (hard) confinement leading to symmetric (irregular) states. To show this fact more accurately, Figure 5.22 shows STM image of another island with the corresponding dI/dU maps. Here the calculation was done neglecting the soft edge potential, i.e. Vi = ∞ outside the island. Inside the island, the moiré potential is still maintained. Superpositions of squared wave functions at the energies corresponding to the dI/dU images are shown in Figure 5.22 (a)-(d). Only the two states belonging to K and K0 are superposed in Figure 5.22 (a)-(c). In Figure 5.22 (d) 6 states had to be superposed. The shapes of the wave functions and the LDOS correspond to the mea- sured dI/dU maps and the energies reasonably fit, but the calculated shapes are less

76 5.3 Wave function mapping in graphene quantum dots

a) b) c) d)

-0.25 eV -0.35 eV -0.41 eV -0.52 eV

e) f) g) h)

-0.25 eV -0.35 eV -0.40 eV -0.50 eV

i) j) k)

3.8nm -0.44 eV -0.44 eV

Figure 5.22: (a)-(c) Calculated squared wave functions of the island shown in (i) at ener- gies marked. (d) Calculated LDOS of the same island (consisting of 6 wave functions) at the energy marked. (e)-(h) dI/dU maps of the island shown in (i) at different energies E = U · e as marked; I = 0.3 nA, Umod = 10 mV. (i) STM image of the graphene QD; U = −250 mV, I = 0.3 nA. (j)-(k) Cal- culated squared wave functions of the QD shown in (i) not found in exper- iment. All calculations are with abrupt edge.

77 5 Graphene

regular and more extended towards the edge. Moreover, the calculated wave functions displayed in Figure 5.22 (j)-(k) do not resemble the experimental LDOS maps. Like in the previous hexagonal dot, these wave functions feature a strong weight at the edge of the QD. They are suppressed in experiment due to the (partial) hybridization of the graphene pz orbitals with the iridium substrate. Thus, softly opening a band gap at the QD edge leads to strongly improved control on the states residing in its interior. To illustrate this crucial finding, it is shown that the state energies in the QDs can be correctly estimated by a simplified circular flake geometry. Thus resulting in 6 En = hv¯ Dkn with Dirac velocity vD = 10 m/s and kn being deduced from the ze- ros of the zeroth and first Bessel function:

Jn(kn · r) = 0, n = 0, 1. (5.12)

The first two peak energies E0 and E1 of the experiment are determined with respect to the Dirac point ED for different islands. Since the shapes of larger islands become more

6 5 4

3 kr 2 1 0 1 2 3 4 5 6 7 8 Quantum dot radius r (nm)

Figure 5.23: Experimental kn · r = En/(hv¯ D) · r for the two peaks closest to ED at dif- ferent average island radius r. Circles: n = 0, squares: n = 1. Dotted lines: zeros of the first two Bessel functions. irregular (see Figure 5.16), only the nine smallest islands are√ considered. The average island radius r is deduced from the island area A by r = A/π. The resulting kn · r is plotted as a function of r for the two peak energies closest to ED (see Figure 5.23). For a few of the islands, the energy of the 2nd resonance in the dI/dU curve is not well defined due to broadening of lineshapes by finite state lifetime (see 5.3.5), and thus not considered in the present analysis. For the smallest island, only the first resonance en- ergy lies within the Ir projected band gap. The k · r values corresponding to the zeros of

78 5.3 Wave function mapping in graphene quantum dots

the first two Bessel functions are plotted as dotted lines. Obviously, they are in reason- able agreement with the experiment. The estimate fits the experimental peak energies to within 20 % accuracy for the lowest energy state of all islands up to an area of 150 nm2. Moreover, all detectable 2nd lowest energy states of these islands fit within 30 %. The discrepancy increases with radius which is attributed to the more non-circular shape of the larger islands. This verifies the linear dispersion relation (k ∝ E) within the graphene QDs on Ir(111), and the validity of the simplified Dirac equation for the lowest QD energies. Obviously, neither the sensitive sublattice symmetry of graphene [13, 14] or other graphene symmetries, nor the influence of the iridium substrate enter Eq. (5.12) showing the simplicity of softly confined graphene QDs. In conclusion, the LDOS of graphene QDs supported on Ir(111) is mapped. For small island, properties of an isolated graphene QD with soft edge potential reproduce the measured wave functions. Importantly, the soft edge induced by the substrate is re- quired for the experimentally observed large symmetry of the wave functions.

5.3.4 Ir surface state S0 and moiré structure

For energies closer to ED, the moiré potential is less important as the wavelength of the envelope function cannot resolve its structure. But at energies E < −0.6 eV and, in par- ticular, in large islands, the influence of the moiré potential on wave function patterns becomes clearly visible in both, theory and experiment. Figure 5.24 (a) shows an STM

a) b) c)

topography STS: E = -0.65 eV theory: E = -0.65 eV

Figure 5.24: (a) STM image and (b) dI/dU map of a large graphene QD. 30 × 30 nm2, U = −0.65 V, I = 0.5 nA, Umod = 10 mV. (c) Calculated LDOS of the same QD at E = −0.65 eV. image of a large QD exhibiting a regular moiré pattern. The dI/dU map in Figure 5.24 (b) and the calculated LDOS in Figure 5.24 (c) reproduce the moiré topography albeit with inverted amplitude. The same result is found for all larger islands studied (see Figure 5.25). It was checked that normalizing the dI/dU images in order to account

79 5 Graphene

25×25 nm²

-0.74 eV -0.77 eV 33×33 nm²

-0.67 eV -0.7 eV 38×38 nm²

-0.67 eV -0.7 eV

Figure 5.25: dI/dU maps (left images) of different graphene QDs imaged at energies below -0.6 eV (I = 0.4 nA, Umod = 10 mV). The images are dominated by the moiré potential and are largely reproduced by the calculation (right images).

80 5.3 Wave function mapping in graphene quantum dots

for a spatially varying tip-surface distance [115] still leads to moiré patterns within the LDOS. Thus, the presence of the moiré potential within the QDs is also directly proven by experiment. For small QDs as shown in Figure 5.17 the LDOS can nicely be reproduced. The LDOS of larger QDs, instead, is dominated by a standing wave pattern which could not be explained by the graphene states only. This feature is already visible by comparing Figure 5.24 (b) and (c), where a bright rim of the islands in the dI/dU image is not re- produced by describing the interaction with the substrate in terms of a spatially vary- ing Vi. This bright rim is found for all islands, but was never reproduced by the tight binding calculations. Closer to ED, it develops into a standing wave pattern that finds its counterpart outside the island with slightly larger wave length λ (Figure 5.26 (a)- (c)). Figure 5.27 shows dI/dU maps of another three different quantum dots where the contrast is tuned in order to see this standing wave outside the QD. It was determined quantitatively by line scans as shown in Figure 5.27 (m) (for the dI/dU map in Figure 5.27 (f)) in order to deduce the E(π/λout) dispersion shown in Figure 5.26 (d). Since the measured standing wave pattern exhibits half the wave length of the impinging Bloch wave, 4k = π/λout is used such that the dispersion E(4k) can directly be compared with data from angular resolved photoelectron spectroscopy (ARPES) [106]. Such standing waves are also observed at step edges of the Iridium(111) surface not covered by graphene. Figure 5.28 (a) shows an STM image of the uncovered Ir(111) surface with two step edges. The two white dots are larger adsorbates on the surface. The dI/dU images exhibit standing waves at the step edges also decreasing in wave length with decreasing energy. In addition, remaining oxygen adsorbates on the sur- face are visible as black dots which induce an additional complicated scattering pattern on the terraces. The wave lengthsλIr of the standing waves at the step edges are deter- mined by line scans averaging along the step edge and the resulting E(π/λIr) is also plotted in Figure 5.26 (d). The E(4k) curves in Figure 5.26 (d) are linear according to E = hv¯ D 4 k + ED 5 5 with vD ' 4.9 · 10 m/s, ED = −0.3 eV outside the island and vD ' 4.5 · 10 m/s, ED = −0.2 eV inside the island. The absolute values are lowest for the pure Ir(111) sur- face, slightly higher (about 50 meV) for the standing waves around the graphene QDs and 100 meV higher for the standing waves inside the graphene QDs. The dispersion values agree with the Ir surface resonance S0 around Γ found by photoemission [106] (dashed line in Figure 5.26 (d)) including the energy shift of the dispersion by about 150 meV between uncovered Ir(111) and Ir(111) completely covered with graphene. The values disagree with vD for the graphene Dirac cone on Ir(111) by a factor of two and with ED for the Ir S2 surface state by 0.5 eV [31]. Thus, the standing wave patterns within the QD are attributed to an intrusion of S0 into graphene. The amplitude of the standing wave in the islands AG is found to be close to the amplitude outside the island AIr for several islands and energies. This is surprising considering the fact that

81 5 Graphene

-0.36 eV -0.55 eV -0.66 eV

λλλ in λλλ out

a) b) c)

d) 0.0 e) ARPES S 0.32 8 R = I /I I 0 -0.2 114 meV DFT C IR Ir 0.28 6 E = -0.4eV 4

-0.4 0.24 LDOS graphene 2 I 0.20 C -0.6 0

0.16 3.5 4.0 4.5 5.0 5.5 vertical displacement (nm) (eV) -0.8 R E 0.12 Iridium S 0.08 0 -1.0 v = 4.9 ·10 5 m/s D, Ir 0.04 v = 4.5 ·10 5 m/s -1.2 D, G S 0.00 2 0.0 0.5 1.0 1.5 2.0 2.5 3.0 -1.0 -0.9 -0.8 -0.7 -0.6 -0.5 -0.4 ∆k (1/nm) U (V)

Figure 5.26: (a)-(c) dI/dU maps of a graphene QD recorded at the energies marked is- 2 land size 27 × 30 nm , I = 0.5 nA, Umod = 10 mV. Deduced wave lengths λout (λin) outside (inside) the QD are marked in (b). (d) resulting disper- sion relations E(4k = π/λin/out) inside (stars) and outside (triangles) of the QD as well as from standing waves scattered at Ir(111) step edges (cir- cles). Full lines are linear fits with resulting vD indicated. Energy offset is marked. Dashed line is deduced from photoemission on clean Ir(111) [106]. (e) Relative intensity R of S0 and S2 in graphene as deduced from STS data (squares) and from DFT calculations (S0: circles, S2: triangles). Inset: calculated LDOS of S0 at E = −0.4 eV along the direction perpen- dicular to the surface. IIr and IC as used for determination of R are marked.

82 5.3 Wave function mapping in graphene quantum dots

a) e) i)

2.5nm 3.0nm 6.1nm -0.36 eV -0.42 eV -0.41 eV b) f) j)

-0.51 eV -0.58 eV -0.55 eV c) g) k)

-0.67 eV -0.63 eV -0.70 eV d) h) l)

-0.74 eV -0.72 eV -0.85 eV

20

m) 19

18 λλλ in 17

(arb. units) (arb. 16

15 dI/dU λλλ out 14 0 2 4 6 8 10 12 14 lateral displacement (nm)

Figure 5.27: (a)-(l) dI/dU maps of three different graphene QDs recorded at the ener- gies E = U · e marked; Umod = 10 mV; (a)-(d),(i)-(l) I = 0.2 nA, (e)-(h) I = 0.5 nA. All images exhibit a standing wave around the island. Wave length decreases with decreasing energy. Blue line in (j) marks the line for the line profile shown in (m). (m) Line profile along the line shown in (j). The deduced wave lengths of the standing waves inside (λin) and outside (λout) the QD are marked.

83 5 Graphene

-0.65eV

4.0nm a) b) -0.75 eV -0.91 eV

c) d)

Figure 5.28: (a) STM image of the Ir(111) surface with two step edges. U = −0.2 V, I = 0.3 nA. (b)-(d) dI/dU maps of the same area recorded at the energies E = U · e marked (I = 0.3 nA, Umod = 10 mV). Standing waves are visible at the upper and the lower side of the step edges.

84 5.3 Wave function mapping in graphene quantum dots

the tip is 0.23 nm further away from the Ir surface, when positioned above graphene, normaly leading to a reduction in dI/dU intensity by a factor of 100 [116]. However, DFT calculations reveal that S0, exhibiting sp-symmetry, penetrates into graphene. The ratio between the LDOS in the graphene layer IC and the LDOS in the Ir surface layer IIr is RDFT = IC/IIr ' 8 − 12 % (inset of Figure 5.26 (e)). For comparison, S2 shows only RDFT ' 0.02 %. Figure 5.26 (e) favorably compares RDFT of S0 with the data from αδ STS RSTS where the apparent AG/AIr is rescaled according to RSTS = AG/AIr · e [116] with α = 1.1 − 1.2 /Å deduced from I(z) curves and δ = 1.1 Å being the differ- ence between real height (3.4 Å [111]) and apparent STM height (2.3 Å) of the graphene above the Ir(111). Thus, we can quantitatively reproduce the strength of S0 intrusion into graphene. A simple explanation for the strong S0 intrusion is not obvious, but 2 note that, according to DFT, also the dz-like surface state S1, located at EF and exhibit- ing no dispersion [31], penetrates into graphene by R ' 10 − 40 % and the π-electrons of graphene penetrate back into Ir by R ' 1 − 4 %. In conclusion, large islands show an additional standing wave pattern caused by an intruding Ir surface resonance and signatures of the moiré potential.

5.3.5 Lifetime of quantum dot states

The width ∆E of the dI/dU-curve maxima can be used to estimate the lifetime τ = h¯ /∆E of a corresponding states [117]. Here the spectral broadening ∆E of the states of the graphene QD is determined. The dI/dU curves were fitted by Lorent- zian lineshapes after subtracting a linear background (Figure 5.29 (a), inset). The energy dependence of the width shown in Figure 5.29 (a) is approximately given by ∆E(E) = 0.33 · |E| (line in Figure 5.29 (a)). A log-log plot of ∆E(E) reveals an exponent of 1.03 ± 0.01. Several contributions to the effective linewidth have to be considered. The number of overlapping states within the spectral resolution increases linearly with |E| due to the linearly increasing density of states. Indeed, an increasing number of states contributing to spectral peaks was found (see, e.g., bars in Figure 5.17 (a), but the resulting a := ∆E/E = 0.1 is too small to account for the observed width. Thus, the electron lifetime τ must dominate ∆E. A linear ∆E(E) can be caused by electron- phonon coupling within graphene at energies above the energy of the in-plane optical phonon (200 meV) [118, 119, 120]. Also electron-electron interaction leads to a linear ∆E(E) in intrinsic graphene due to a marginal overlap of possible excitations and de- cay processes [121]. In doped graphene, the parabolic ∆E(E)-dependence is restored at energies close to EF, but again one gets a mostly linear slope, if |E| > |ED − EF| [122, 123]. Thus, determination of the exponent is not sufficient to deduce the under- lying hot electron decay process. Instead, quantitative considerations are necessary to uncover the mechanism. DFT reveals that electron couplings to in-plane phonons in freely suspended graphene

85 5 Graphene

a) 4 0.6 3 0.5

2 (arb. units) (arb. 0.4 1 dI/dU

fit 0 0.3 -0.8 -0.6 -0.4 -0.2 (eV) E (eV) E ∆ 0.2 0.1 0.0 -1.2 -1.0 -0.8 -0.6 -0.4 -0.2 0.0 0.2 E (meV) b) 30

25

20

15

10

island size (nm) 5

0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0 mean free path (nm)

Figure 5.29: a) Black dots: peak widths 4E as deduced from Lorentzian fits as shown in the inset; Line: linear fit. b) Mean free path vs. graphene island size. No island size dependence is found.

86 5.4 Investigation of zigzag graphene edge on Ir(111)

give rise to peak widths below 1 meV up to about E − ED = 1 eV. This rules out this process, even though slight changes of the coupling might be induced by the sub- strate. It is rather unlikely that the coupling is strongly increased by the presence of the substrate. Moreover, a strong electron-phonon coupling parameter λ ' a leads to a 5 slowing down of vD by 1/(1 + λ) [3, 119, 120] which would result in vD ' 7 · 10 m/s taking even the small effect of 2.2 % compression of graphene on Ir(111) into account [124, 95]. Such a small vD is neither found in photoemission of extended graphene lay- ers on Ir(111) [31] nor it does give a good correspondence between the peak energies 6 of tight binding calculations and experiments. Instead, only the usual vD ' 1 · 10 m/s is compatible with both results. The role of out-of plane phonons is less clear [125, 126, 127], in particular, since they are associated with a dipole layer oscillation between graphene and Ir with Dipole moment of about 0.4 Debeye/unit cell [111]. However, it is unlikely that the result- ing decay rate is increased by eight orders of magnitude relative to DFT results of suspended graphene [127]. Note that the phonon spectrum of a graphene on Pt(111), which is bonded similarly to the one on Ir(111), has been measured experimentally and is barely changed with respect to graphite [128]. The broadening induced by electron- electron interaction depends on doping and dielectric screening [129]. But even at a four times larger doping than in our experiment and on a SiO2 substrate, that screens less effectively, one finds only a ' 0.1 [122]. Thus, an additional interaction is required to explain the short lifetimes found in ex- periment. However, which particular interaction is responsible for the broadening re- mains here unsolved.

5.4 Investigation of zigzag graphene edge on Ir(111)

In this section, the zigzag edge of the graphene on Ir(111) is further investigated. Spin polarized scanning tunneling spectroscopy measurements and density functional the- ory calculation are used to clarify the question, if this system has a spin-polarized edge state. The first section 5.4.1 describes the density functional theory (DFT) calcu- lation of graphene edge on Ir(111). This calculation was done by Yan Li and Riccardo Mazzarello. In section 5.4.2 the termination of the graphene edge is discussed. In the following sections scanning tunneling spectroscopy curves on graphene edge with a tungsten-tip (5.4.3) and with a spin polarized chromium tip (5.4.4) are discussed.

5.4.1 Density functional theory calculation of graphene edge on Ir(111) In the literature up to now all spin-polarized edge state prediction refers to an isolated single-hydrogen terminated graphene ribbon. Figure 5.30 shows a simulated STM im-

87 5 Graphene

(a) (b) H H

H H

occupied states empty states U = -0.5 V U = +0.5 V

Figure 5.30: Simulated STM images of isolated graphene ribbon for occupied (a) and empty (b) states. Both configuration feature a spin polarized edge state.

age of an isolated graphene ribbon for the occupied and the unoccupied electronic states. The ab-initio DFT simulation was carried out using the plane-wave code [110]. The wave function were expanded in plane waves with a kinetic energy cutoff of 30 Ry and a charge-density cutoff of 300 Ry. The graphene edges are simulated within a supercell geometry using a vacuum layer of 8.1 Å between two edges and of 10 Å between the two graphene planes. The atomic positions are allowed to fully relax. Electronic integrations are done using Monkhorst Pack meshes of 64 k points along the direction parallel to the zigzag graphene nanoribbon. The simulated STM images were obtained using the Tersoff-Hamann approximation [33, 34, 130]. The most dominant feature in the simulated STM images for occupied and empty states is the signature of the edge state that decays toward the interior of the ribbons. In addition to that, in the interior of the ribbon, where the edge state has decayed, the negative bias STM im- age highlights horizontal structures in the zigzag direction along the ribbon [130]. In contrast to that, the empty electronic states at positive bias shows a regular triangular features. In this work, the edge states of graphene nanoribbon have been studied including the Ir(111) substrate. The adsorption of graphene has been modeled by overlaying the graphene nanoribbon over a 9 × 9 Ir(111) supercell and a slab of 4 Ir layers, using the Γ-point only. The graphene edges are simulated within a supercell geometry using a vacuum layer of 9.6 Å between two edges and of 4.8 Å between two graphene planes.

88 5.4 Investigation of zigzag graphene edge on Ir(111)

The size of the nanoribbon corresponds to 10 × 8 graphene unit cells. Figure 5.31 shows

atop bridge

Figure 5.31: Model of the relaxed atomic structure of graphene on a slab of 4 Ir layers. Yellow (grey) color represents graphene (Ir) atoms. Blue box: atop position of the carbon edge atom. Red box: bridge position of the carbon edge atom. The height difference between atop and bridge position is 0.15 Å. a model of the atomic structure of graphene on Ir(111). The edge has two areas with different configurations. The position where the carbon atoms occupy the atop site with respect to the underlying Ir atoms is marked with a blue box and the position where the carbon atoms occupy the bridge site is marked with a red box. The bridge site position is approximately 0.15 Å higher than the atop-position. The hybridization of all edge carbon atoms is sp3-like. Two bonds of the edge carbon atom are bonded to the next two carbon atoms and the other two bonds are bonded to the Ir substrate. Figure 5.32 (a) and (b) show the simulated STM images of graphene on Ir(111) for oc- cupied and empty states, respectively. The negative and positive bias STM images of

89 5 Graphene

(a) U = -0.5 V (e)

(b) U = +0.5 V (f)

(c) U = -0.5 V with hydrogen (g)

(d) U = +0.5 V with hydrogen (h)

Figure 5.32: Simulated STM images of graphene nanoribbon on Ir(111) for (a) occupied state (U = −0.5 V) and (b) empty state (U = +0.5 V). Single-hydrogen sat- urated graphene nanoriboon on Ir(111) for (c) occupied state (U = −0.5 V) and (d) empty state (U = +0.5 V). Yellow box: (a)-(b) bridge site, (c)-(d) hollow site. (e)-(f) Enlarged image of the graphene edge marked with the red box. The intensity is high (low) for bright (dark) area and is arbitrary for every image.

90 5.4 Investigation of zigzag graphene edge on Ir(111)

the interior graphene region show a regular triangular feature. At the edge, we can dis- tinguish between the atop sites and the bridge sites, which are marked with a yellow box in Figure 5.32 (a) and (b). The atop and the bridge regions on the edge show also a triangular configuration with bonds, which are slightly extended perpendicular to the edge. Additionally the bonds at the bridge sites show a relatively higher density of occupied and empty states. The distribution of the charges in the occupied and empty states seems to be very similar for the interior and the edge region on the graphene. The next step, hydrogen was offered to the graphene on Ir(111) system. The parame- ters for the DFT calculations are the same as in the case without hydrogen. The only difference here is that only one Ir layer slab instead of four Ir layers is used. In the case of supported graphene without hydrogen, we have checked that the reduction of four Ir layers to one Ir layer has no influence on the bonds. Figure 5.33 shows the relaxed atomic structure of single-hydrogen terminated graphene on Ir(111). Again two areas with different configurations can be seen. The part where the edge carbon atoms occu- pied nearly the atop positions is sp3-like hybridized (marked with blue box). In Figure 5.33 (b), it can be seen that the two bonds of the carbon atoms at the edge are bonded to the next carbon atoms, one bond to Ir surface atom and one bond to hydrogen atom. In the other part the edge carbon atoms occupied the hollow sites (marked with red box). Here the hybridization is different, it is sp2-like. Two bonds are formed to the next carbon atoms and one bond to the hydrogen (see Figure 5.33 (c)). The distance between a carbon atom and the iridium atom in the sp3 hybridized area is 2 Å. The sp2 hybridized area reveals a larger height difference, it is 3.5 Å. Figure 5.32 (c) and (d) show simulated STM images of single-hydrogen terminated graphene on Ir(111) for the occupied and empty states, respectively. A clear distinction between the occupied and empty states can be seen. In the case of occupied states, the carbon bonds built hexagons in the interior region and at the edge of the graphene. In some regions the bonds of the hexagons perpendicular to the edge are weak. Thus appearing as hori- zontal lines parallel to the edge. The sp2 hybridized area (marked in the yellow box) shows a relatively high density of occupied hexagon states. The distribution of the empty states is triangular in the interior and at the edge of the graphene (see Figure 5.32 (d)). The sp2 hybridized area (marked in the yellow box) shows also a relatively high density of empty triangular states. For both cases, with and without hydrogen on graphene/Ir(111), the DFT calculations show no spin polarized edge state.

5.4.2 Termination of the graphene edge

Figure 5.34 (a) and (d) show STM images with atomic resolution of the edge of an hexagonal graphene dot for the occupied and the empty states, respectively. The graphene edge shows a clear distinction between the occupied and empty states. In

91 5 Graphene

a)

atop hollow

b) c)

Figure 5.33: (a) Topview image of a relaxed atomic structure of graphene on a slab of 1 Ir layer with single-hydrogen terminated edge. Yellow (grey) color repre- sents graphene (Ir) atoms and blue color represents the hydrogen atoms. Blue box: atop position of the carbon edge atom. Red box: hollow position of the carbon edge atom. The height difference between atop and hollow position is 1.5 Å. (b) Enlarged and sideview image of the atop positions: hybridization of the edge carbon atom is sp3-like. (c) Enlarged and side- view image of the hollow positions: hybridization of the edge carbon atom is sp2-like.

92 5.4 Investigation of zigzag graphene edge on Ir(111)

(a)

Theory: U = -0.5 V (b)

Experiment: U = -0.5 V

(c)

Theory: U = +0.5 V (d)

Experiment: U = +0.5 V

Figure 5.34: STM images of zigzag graphene edge on Ir(111) for (b) occupied state and (d) empty state. The corresponding simulated STM images are shown in (a) occupied state and (c) empty state. The green box highlighted the edge region, where the carbon atoms occupy the hollow sites (see Text). The hight difference between the atop site and hollow site is 40 ± 10 pm. Con- trast and brightness are enhanced for the middle part of (b) and (d).

93 5 Graphene

Figure 5.34 (a) the bonds at negative bias built hexagons at the edge region. In contrast to that the feature at positive bias shows a triangular configuration (see Figure 5.34 (d)). Let us compare the STM images with the calculated images for hydrogen terminated edges and edges without hydrogen termination. The case without hydrogen can be excluded, because the simulated STM images show similar configuration for the occu- pied and the empty states (see Figure 5.32 (a) and (b)). The simulated STM images for single-hydrogen terminated graphene edge shows similarity to the experimental im- ages. In the simulated image the bonds built hexagons in the occupied state (see Figure 5.34 (a)). This fits very well with the hexagons in the experimental STM image in Fig- ure 5.34 (b). For the empty states, the simulation image shows triangular features (see Figure 5.34(c)). This also fits well with the STM image take at negative bias voltage. Beside the similarities in the atop region, Figure 5.34 shows also bright areas (marked with red boxes), where the carbon atoms probably occupy the hollow sites. And like in theory, this sp2-like regions have a relatively high density of occupied hexagonal and empty triangular states. To conclude, the comparison of the STM experiments and the STM image simulations strongly suggests that the graphene edges are singly hydrogen terminated.

5.4.3 Scanning tunneling spectroscopy on graphene edge

The local density of states (LDOS) of the graphene edge is measured by STS with a tungsten tip. The modulation amplitude was set to Umod = 10 mV resulting in an en- ergy resolution of δE = 25 meV. To measure the dI/dU curves, the tip was stabilized at a sample voltage Ustab and current Istab. Figure 5.35 shows the dI/dU curves of the graphene edge at the bright areas, which were already identified as hollow site po- sitions (see previous section 5.4.2). The LDOS varies depending on the edge of the hexagonal graphene dot. For three hollow site positions (red, green and orange), we can identified a peak at positive bias between 170 meV and 260 meV, which shifts in the energy depending on the edge side. At negative bias, we found a peak, which shifts from −80 meV to −120 meV. The dI/dU curves at the atop site positions (see Figure 5.36) show similar behavior. The dI/dU curves vary also depending on the edge. The peak at positive bias vary from 160 meV to 240 meV. At negative bias only the yellow curve shows a shift to −120 meV. The other peak is at −90 meV and it appears at the same energy position for the other curves. The origin of these peaks is not clear. The peak at the negative bias might be the first resonance peak of the confined state in graphene (see section 5.3). If there is an edge state in graphene on Ir(111), the density of states at the Fermi energy EF should be increased and we should measure a pronounced peak at the Fermi level. Figure 5.37 shows dI/dU curves at the atop- and hollow-site edge positions measured on different graphene quantum dots with different micro-tips. Different micro-tips

94 5.4 Investigation of zigzag graphene edge on Ir(111)

4.5

4.0

3.5

3.0

2.5

2.0

1.5 dI/dU (arb. units) 1.0

0.5

0.0 -300 -200 -100 0 100 200 300 U (meV)

Figure 5.35: dI/dU-curves taken at the graphene edge at hollow site positions. Stabi- lization parameter: Ustab = 0.3 V, Istab = 0.3 nA. Each curve was averaged from 5 single dI/dU-curves.

95 5 Graphene

4.5

4.0

3.5

3.0

2.5

2.0

1.5 dI/dU (arb. units) 1.0

0.5

0.0 -300 -200 -100 0 100 200 300 U (meV)

Figure 5.36: dI/dU-curves on graphene edge at atop site positions. Stabilization pa- rameter: Ustab = 0.3 V, Istab = 0.3 nA. Each curve was averaged from 5 single dI/dU-curves.

96 5.4 Investigation of zigzag graphene edge on Ir(111)

(a) (b) 3.0 atop 3.0 atop quantum dot 1 hollow 2.5 quantum dot 2 hollow 2.5 micro-tip a micro-tip b 2.0 2.0 1.5 1.5 1.0 1.0 dI/dU(arb. units) dI/dU(arb. units)

0.5 0.5

0.0 0.0 -300 -200 -100 0 100 200 300 -500 -400 -300 -200 -100 0 100 200 300 400 500 U (meV) U (meV)

(c) 3.0 quantum dot 3 atop 2.5 micro-tip b hollow

2.0

1.5

1.0 dI/dU (arb. units)

0.5

0.0 -500 -400 -300 -200 -100 0 100 200 300 400 500 U (meV)

Figure 5.37: dI/dU-curves on graphene edge at atop site and hollow site positions. (a) The curves are averaged from the contribution of Figure 5.35 and Figure 5.36. Stabilization parameter: Ustab = 0.3 V, Istab = 0.3 nA. (b)-(c) dI/dU- curves of different quantum dots, but with the same micro-tip. Stabiliza- tion parameter: Ustab = 0.3 V, Istab = 0.3 nA.

97 5 Graphene means tunneling from different tip orbitals. The structure of the micro-tip is changed by short voltage pulses (3 − 10 V) between tip and sample during the STM measure- ment. The changing of the tip apex orbital is arbitrary and the identification whether we tunnel through a s-, p- or d-orbital is not obvious. The dI/dU curves on atop- and hollow-sites look similar. For energies close to EF, the LDOS is low and no peak is mea- sured. However, for approximately 5 % of the used micro-tips a peak at EF is mapped (see Figure 5.38). Moreover, the peak is small in the inner part of the graphene dot and on Ir(111). Only at the edge of the graphene, the LDOS increases by the factor of 3 com- pared to the other part of the quantum dot and Ir(111). The observed result suggests that the peak at the Fermi level might represent an edge state of graphene. To confirm this conjecture, spin polarized scanning tunneling spectroscopy was employed. The results will be discussed in the next section.

Graphene

Edge Ir Substrate

Figure 5.38: dI/dU-curve shows an edge peak (green) close to Fermi energy. In the inner part of graphene (red) and on Ir-substrate (orange) the peak at the Fermi energy is 3 times smaller. The energy width of the peak is Ewidth = 25 mV and the energy position is Epos = −0.4 mV. The position of the peak Ir on Ir((111) is Epos = +6 mV. For clarity an offset is included between the curves. Stabilization parameter: Ustab = 0.2 V, Istab = 0.2 nA.

5.4.4 Spin polarized scanning tunneling spectroscopy on graphene edge Spin polarized scanning tunneling spectroscopy was used to check, if the pronounced peak at EF at the graphene edge is magnetic. As a spin polarized probe a bulk

98 5.4 Investigation of zigzag graphene edge on Ir(111)

chromium tip was used. The chromium tip was characterized on an Fe-island and shows magnetic signal in all 3 spatial direction (see chapter 4.2 and 4.5). Figure 5.39

1.0

Tip with edge peak a, Graphene dot 1 Tip with edge peak b, Graphene dot 2 Tip with edge peak c, Graphene dot 3

0.5

dI/dU(a.u.)

0.0 0.0 0.5 1.0 1.5 2.0 edge distance (nm) graphene

Figure 5.39: Decay of the edge peak amplitude from the edge atom (= 0 nm) to the inner part of the graphene. The direction is marked with black arrow in the STM image. Three different micro-tips, which show the edge peak and three different graphene quantum dots are investigated. The amplitude of the edge peak decays over three graphene unit cells. The energy width and the energy position of the peak for tip a are: Ewidth = 16 mV, Epos = 10 mV; tip b: Ewidth = 18 mV, Epos = −3 mV; tip c: Ewidth = 16 mV, Epos = −5 mV. Ir Ir The position of the peak on Ir((111) for tip a is: Epos = +7 mV; tip b: Epos = Ir +7 mV; tip c: Epos = −11 mV.

shows the decay of the peak amplitude from the edge atom to the inner part of the graphene. The measurement was done on different graphene dots and with three dif- ferent micro-tips, which show a peak near to EF. The amplitude of the edge peak decays very similar for different micro-tips and different dots, it decays over three graphene unit cells. It should be noticed, that every time if we measured a peak at the graphene edge, we also measured a peak near to EF with relatively low dI/dU- amplitude on the iridium substrate. The energy positions of the peak on iridium is slightly shifted in comparison to the peak on graphene edge. Now an in-plane mag- netic field was applied and the amplitude of the edge peak (see Figure 5.40) was measured. Figure 5.41 shows the variation of the amplitude of the edge peak be- tween +800 mT and −800 mT. The expected theoretical magnetic curve is fitted for a macrospin of S = 1.598. The macrospin value is a sum of the magnetic moments from [131] considering 36 atoms at the edge. The theoretical magnetic curve was calculated

99 5 Graphene

1.5

1.4

1.3

1.2

1.1 Ipeak 1.0 dI/dU (a.u.) 0.9

0.8 Ibg

0.7 -0.2 -0.1 0.0 0.1 0.2 Sample Voltage U (eV)

Figure 5.40: dI/dU-curve shows an edge peak in the near of the Fermi-level. The ex- act energy width and energy position of the peak are Ewidth = 22 mV and Epeak = 6.5 ± 3.5 mV. The ratio between the peak height Ipeak and the back- ground Ibg is Ipeak/Ibg = 0.12. In Figure 5.41 the behavior of this edge peak amplitude in an magnetic field was measured.

100 5.4 Investigation of zigzag graphene edge on Ir(111)

with the following equation:

dI/dUtheo = Mtip · Ipeak/Ibg · F(S, y), (5.13) with Mtip the spin polarization of the tip and Ipeak/Ibg the ratio between the peak height Ipeak and the background Ibg (see Figure 5.40). F(S, y) is the Brillouin function [132]:

F(S, y) = (1 + 1/2S)coth[(1 + 1/2S)y] − 1/2Scoth(y/2J) (5.14) where y = SB0(gµB/kT), (5.15) with macrospin S and magnetic field B0. We consider an in-plane tip spin polarization of 30 % (see section 4.5) and Ipeak/Ibg-ratio of 12 % (see Figure 5.40). The resulting the- oretical curve varies about 1 % between ±800 mT (see Figure 5.41). The experimental curve varies about 4 %. From this experimental data, it is not possible to make a clear statement about the magnetization of this system. The variation of the amplitude of approximately 4 % is in the order of the lateral variation of approximately 3 pm, what we already saw in Figure 5.39. This is probably caused by the error of the tip posi- tion during the different field measurements. In future experiment the magnetic field should be set higher, so that the lateral variation plays a minor role.

101 5 Graphene

1.09 Experiment Theory S=1.598 1.08

1.07

1.06 dI/dU(a.u.)

1.05

-800 -600 -400 -200 0 200 400 600 800 Magnetic Field (mT)

Figure 5.41: Graph shows the behavior of the edge peak amplitude in an in-plane mag- netic field. The theoretical magnetic curve is for a macrospin of S = 1.598, with a tip spin polarization of 30 % and Ipeak/Ibg-ratio of 12 %.

102 5.5 Mechanical instability of graphene naonislands on Ir(111)

5.5 Mechanical instability of graphene naonislands on Ir(111)

A further interesting feature of graphene on Ir(111) is the movable areas within the moiré structure. Figure 5.42 (a) shows the moiré structure of graphene on Ir(111). By

(a)U = 0.2 V, I = 3.4 nA (b) U = 0.2 V, I = 4.8 nA

(c)

Figure 5.42: (a) Moiré structure of graphene on Ir(111) at I = 3.4 nA. (b) Bistable areas appear at a current of I = 4.8 nA. (c) The same image like in (a) with grey ellipses added to mark the bistability areas within the moiré structure. increasing the tunneling current from I = 3.4 nA to I = 4.8 nA, arrays of bistability appear (shown in Figure 5.42 (b)). In Figure 5.42 (c) the bistability regions are marked

103 5 Graphene with grey ellipses. The bistability appears in transition between the dark region and the bright region of the moiré structure. From chapter 5.2.2, it is clear that the fcc and hcp regions are weakly bonded to the substrate and appear as bright regions in the moiré structure. Due to this weak bonding to the substrate, these areas could be easily lifted with the tungsten tip via electrostatic and van-der-Waals forces. The ar- eas appear slightly shifted to the left probably due to a lateral tunnel path, caused by a non-symmetrical tungsten-tip. Mashoff et. al. [133] show already such bista- bility nanomembranes on an exfoliated graphene flake on SiO2. They deduced the ∼ resonance frequencies of the nanomembrane up to ν0 = 400 GHz correspponding to a vibronic energy of h · ν0 = 1.7 meV. In our case, the maximal radius of the bistabil- ity area is ∼ 0.8 nm and this leads to a resonance frequency of 1.8 THz corresponding to a vibronic energy of 7.5 meV. Such a large value might provide easy cooling to the ground state leading to a novel access to quantum-mechanical manipulation of vibrons [134, 135] or in combination with the low membrane mass to ultrasensitive mass de- tection, which likely goes down to a single hydrogen atom [133]. Compared to the exfoliated graphene system, the size of the nanomembranes on graphene on Ir(111) are almost equal and therefore it can be stimulated parallel.

104 6 Summary

In this work several sample systems were examined using the scanning tunneling mi- croscope (STM) at room and low temperature in terms of morphology, electrical and magnetic properties. An amorphous Ge1Sb2Te4 sample was prepared for ultra-high- vacuum room temperature scanning tunneling measurements by dc magnetron sput- tering and in situ ion bombardment. Upon crystallization, a volume reduction of 5 % was observed. First scanning tunneling spectroscopy measurements revealed a de- crease of the surface band gap from 0.65 eV in the amorphous phase via 0.3 eV in the metastable phase to zero gap in the stable phase. Moreover, the Fermi level moves from midgap in the amorphous to the valence band edge in the metastable crystalline phase. Thus, the origin of the large conductivity difference between amorphous and crystalline phase, which makes this material so interesting for future non-volatile elec- tronic storage system, could be shown here directly. The surface atomic structure of the (111)-crystallites appears different at positive and negative bias. The negative bias exhibiting a hexagonal structure, corresponding to the Te-atoms. This Te pattern is not threefold symmetric, but seems to prefer a stripelike structure. At positive bias a complex stripe pattern for the Ge/Sb/vacancy lattice appears. Such distortions have already shown by DFT calculations of Ge1Sb2Te4 and traced back to two different pref- erential bond lengths. Spin polarized scanning tunneling spectroscopy method was used to analyze the thin ferromagnetic iron-islands on W(110) substrate in external magnetic field at 6 K. The vortex core of a 9 nm thick Fe-island was mapped with a 10 monolayer iron coated tungsten tip. By applying an in-plane external magnetic field the movement of the vor- tex core was studied. As expected, the vortex core shift perpendicular to the applied field and disappears from the island for field strength higher than 110 mT. To inves- tigate the magnetic hysteresis properties of Fe nanoisland in x-, y- and z-direction, a bulk chromium tip is used. Compare to the ferromagnetic Fe/W-tips, chromium tips are more stable to tip-sample magnetic interaction, due to its anti-ferromagnetic order. The strong in-plane magnetic surface anisotropy, with the easy axis in [110] direction for the thin 6 nm Fe-island on W(110) could be confirmed. The bulk chromium tip was characterized to have both magnetization in the x-y-plane and a small component showing in the out-of-plane direction. Using low-temperature scanning tunneling spectroscopy, the local density of states (LDOS) of graphene quantum dots with 2 − 40 nm diameters supported on Ir(111) was

105 6 Summary mapped. Due to a band gap in the projected Ir band structure around the graphene K point, the electronic properties of the quantum dots are dominantly graphene-like. By comparing the measured wave functions with the tight binding calculations on the honeycomb lattice based on parameters derived from density functional theory, the relationship between geometry and electronic properties are determined. Most no- tably, the interaction with the substrate towards the edge of the island softly opens a gap in the Dirac cone, which implies soft-wall confinement. Interestingly, this con- finement is required for highly symmetric wave functions. This is intimately related to the additional sublattice symmetry (pseudospin) which makes graphene so special [136, 3, 71]. The resonances of the confined states could be correctly estimated by a sim- plified circular flake geometry and with the first Bessel function using the Dirac equa- tion En = hv¯ Dkn. The estimate fits the experimental peak energies to within ∼ 20 % for the two lowest energy states. Neither the sensitive sublattice symmetry of graphene or other graphene symmetries, nor the influence of the iridium substrate enter the first Bessel function showing the simplicity of softly confined graphene quantum dots. Larger islands show an additional standing wave pattern caused by an intruding sur- face resonance S0 and signatures of the moiré potential. The experiment data and DFT calculation show a penetration of S0 into graphene by R ' 8 − 12 %, where R = IC/IIr with IC and IIr the LDOS of graphene and Ir, respectively. This is surprising consider- ing the fact that the tip is 0.23 nm further away from the Ir surface, when positioned above graphene, normaly leading to a reduction in dI/dU intensity by a factor of 100 [116]. By comparing STM topographic images with simulated STM images, the termination of the zigzag graphene edge was checked. It could be shown, that the graphene edges are single-hydrogen terminated. This is not surprising, due to the fact that during the preparation of graphene on Ir(111) C2H4 is offered. For single-hydrogen terminated zigzag graphene edge tight binding calculation predicts a spin polarized edge state [87]. Here spin polarized scanning tunneling spectroscopy (SP-STS) are performed to search for this magnetic edge state. An important evidence for an edge state would be a peak at the Fermi level in the STS experiment. Indeed, approximately 5 % of the used micro-tips show a pronounce peak at the Fermi-level. The spin polarized scan- ning tunneling spectroscopy measurement of this peak shows a variation of 4 % in a magnetic field of ±800 mT. Because the theoretical variation is 1 %, a clear statement about the magnetization of the edge of graphene is not possible with this experimental data. The high 4 % variation is probably caused by the error of the tip position during the different field measurements. However, although the graphene on Ir(111) system exhibit single-hydrogen terminated zigzag edges, the DFT calculation shows no spin polarized edge state. In order to also have an experimental proof, future experiments with the spin-polarized scanning tunneling spectroscopy need to be measured at much higher magnetic fields.

106 Bibliography

[1] M. Pratzer. Untersuchung magnetischer Nanostrukturen mittels Rastertunnelspek- troskopie und Kerr-Magnetometrie am Beispiel von Fe, Co, Co-Fe und Fe-Mn Nanos- trukturen. PhD thesis, Universität Mainz, 2004. [2] K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang, Y. Zhang, S. V. Dubonos, I. V. Grigorieva, and A. A. Firsov. Electric field effect in atomically thin carbon films. Science, 306(5696):666–669, 2004. doi: 10.1126/science.1102896. [3] A. H. Castro Neto, F. Guinea, N. M. R. Peres, K. S. Novoselov, and A. K. Geim. The electronic properties of graphene. Rev. Mod. Phys., 81(1):109–162, Jan 2009. doi: 10.1103/RevModPhys.81.109. [4] M. I. Katsnelson, K. S. Novoselov, and A. K. Geim. Chiral tunnelling and the klein paradox in graphene. Nat Phys, 2:620–625, 2006. doi: 10.1038/nphys384. [5] A. F. Young and P. Kim. Quantum interference and klein tunnelling in graphene heterojunctions. Nat Phys, 5:222–226, 2009. doi: 10.1038/nphys1198. [6] K. S. Novoselov, Geim, A. K., S. V. Morozov, D. Jiang, M. I. Katsnelson, I. V. Grig- orieva, S. V. Dubonos, and A. A Firsov. Two-dimensional gas of massless dirac fermions in graphene. Nat Phys, 438:197–200, 2005. doi: 10.1038/nature04233. [7] Y. Z. Zhang, Y.-W. Tan, H. L. Stormer, and P. Kim. Experimental observation of the quantum hall effect and berry’s phase in graphene. Nat Phys, 438:201–204, 2005. doi: 10.1038/nature04235. [8] K.I. Bolotin, K.J. Sikes, Z. Jiang, M. Klima, G. Fudenberg, J. Hone, P. Kim, and H.L. Stormer. Ultrahigh electron mobility in suspended graphene. Solid State Commun., 146:351, 2008. doi: 10.1016/j.ssc.2008.02.024. [9] X. Du, I. Scachko, A. Barker, and E. Y Andrei. Approaching ballistic transport in suspended graphene. Nature Nanotech., 3:491–495, 2008. doi: 10.1038/nnano.2008.199. [10] Y.-M. Lin, C. Dimitrakopoulos, K. A. Jenkins, D. B. Farmer, H.-Y. Chiu, A. Grill, and Ph. Avouris. 100-ghz transistors from wafer-scale epitaxial graphene. Sci- ence, 327(5966):662, 2010. doi: 10.1126/science.1184289.

107 Bibliography

[11] Y. Zhu, S. Murali, M. D. Stoller, K. J. Ganesh, W. Cai, P. J. Ferreira, A. Pirkle, R. M. Wallace, K. A. Cychosz, M. Thommes, D. Su, E. A. Stach, and R. S. Ruoff. Carbon-based supercapacitors produced by activation of graphene. Science, 332 (6037):1537–1541, 2011. doi: 10.1126/science.1200770. [12] K. S. Kim, Y. Zhao, H. Jang, S. Y. Lee, J. M. Kim, K. S. Kim, J.-H. Ahn, P. Kim, J.-Y. Choi, and B. H. Hong. Large-scale pattern growth of graphene films for stretch- able transparent electrodes. Nature, 457:706–710, 2011. doi: 10.1038/nature07719. [13] B. Trauzettel, D. V. Bulaev, D. Loss, and G. Burkard. Spin qubits in graphene quantum dots. Nat Phys, 3:192–196, 2007. doi: 10.1038/nphys544. [14] S. Das Sarma, S. Adam, E. H. Hwang, and E. Rossi. Electronic transport in two-dimensional graphene. Rev. of Mod. Phys., 83:407–470, 2011. doi: 10.1103/RevModPhys.83.407. [15] D. Loss and D. P. DiVincenzo. Quantum computation with quantum dots. Phys. Rev. A, 57:120–126, 1998. doi: 10.1103/PhysRevA.57.120. [16] C. H. Bennett and D. P. DiVincenzo. Quantum information and computation. Nature, 404:247–255, 2000. doi: 10.1038/35005001. [17] P. R. Struck and Guido Burkard. Effective time-reversal symmetry breaking in the spin relaxation in a graphene quantum dot. Phys. Rev. B, 82:125401, 2010. doi: 10.1103/PhysRevB.82.125401. [18] M. Gmitra, S. Konschuh, C. Ertler, C. Ambrosch-Draxl, and J. Fabian. Band- structure topologies of graphene: Spin-orbit coupling effects from first princi- ples. Phys. Rev. B, 80:235431, 2009. doi: 10.1103/PhysRevB.80.235431. [19] A. H. Castro Neto and F. Guinea. Impurity-induced spin-orbit coupling in graphene. Phys. Rev. Lett., 103:026804, 2009. doi: 10.1103/Phys- RevLett.103.026804. [20] L. A. Ponomarenko, F. Schedin, M. I. Katsnelson, R. Yang, E. W. Hill, K. S. Novoselov, and A. K. Geim. Chaotic dirac billiard in graphene quantum dots. Science, 320(5874):356–358, 2008. doi: 10.1126/science.1154663. [21] C. Stampfer, J. Güttinger, F. Molitor, D. Graf, T. Ihn, and K. Ensslin. Tunable coulomb blockade in nanostructured graphene. Appl Phys. Lett., 92:012102, 2008. doi: 10.1063/1.2827188. [22] F. Molitor, S. Dröscher, J. Güttinger, A. Jacobsen, C. Stampfer, T. Ihn, and K. En- sslin. Transport through graphene double dots. Appl Phys. Lett., 94:222107, 2009. doi: 10.1063/1.3148367.

108 Bibliography

[23] J. Güttinger, C. Stampfer, S. Hellmüller, F. Molitor, T. Ihn, and K. Ensslin. Charge detection in graphene quantum dots. Appl Phys. Lett., 93:212102, 2008. doi: 10.1063/1.3036419. [24] J. Güttinger, C. Stampfer, F. Libisch, T. Frey, J. Burgdörfer, T. Ihn, and K. Ensslin. Electron-hole crossover in graphene quantum dots. Phys. Rev. Lett., 103:046810, 2009. doi: 10.1103/PhysRevLett.103.046810. [25] M. V. Berry and R. J. Mondragon. Neutrino billiards: Time-reversal symmetry-breaking without magnetic fields. Proceedings of the Royal Society of London. A. Mathematical and Physical Sciences, 412(1842):53–74, 1987. doi: 10.1098/rspa.1987.0080. [26] J. Wurm, A. Rycerz, I. Adagideli, M. Wimmer, K. Richter, and H. U. Baranger. Symmetry classes in graphene quantum dots: Universal spectral statistics, weak localization, and conductance fluctuations. Phys. Rev. Lett., 102:056806, 2009. doi: 10.1103/PhysRevLett.102.056806. [27] F. Libisch, C. Stampfer, and J. Burgdörfer. Graphene quantum dots: Beyond a dirac billiard. Phys. Rev. B, 79:115423, 2009. doi: 10.1103/PhysRevB.79.115423. [28] J. Li, W.-D. Schneider, R. Berndt, and S. Crampin. Electron confinement to nanoscale ag islands on ag(111): A quantitative study. Phys. Rev. Lett., 80:3332– 3335, Apr 1998. doi: 10.1103/PhysRevLett.80.3332. [29] T. Maltezopoulos, A. Bolz, C. Meyer, C. Heyn, W. Hansen, M. Morgenstern, and R. Wiesendanger. Wave-function mapping of inas quantum dots by scanning tunneling spectroscopy. Phys. Rev. Lett., 91:196804, Nov 2003. doi: 10.1103/Phys- RevLett.91.196804. [30] Paolo Lacovig, Monica Pozzo, Dario Alfe, Paolo Vilmercati, Alessandro Baraldi, and Silvano Lizzit. Growth of dome-shaped carbon nanoislands on ir(111): The intermediate between carbidic clusters and quasi-free-standing graphene. Phys. Rev. Lett., 103(16):166101, Oct 2009. doi: 10.1103/PhysRevLett.103.166101. [31] I. Pletikosi´c,M. Kralj, P. Pervan, R. Brako, J. Coraux, A. T. N’Diaye, C. Busse, and T. Michely. Dirac cones and minigaps for graphene on ir(111). Phys. Rev. Lett., 102(5):056808, Feb 2009. doi: 10.1103/PhysRevLett.102.056808. [32] J. Bardeen. Tunnelling from a many-particle point of view. Phys. Rev. Lett., 6(2): 57–59, Jan 1961. doi: 10.1103/PhysRevLett.6.57. [33] J. Tersoff and D. R. Hamann. Theory and application for the scanning tunnel- ing microscope. Phys. Rev. Lett., 50(25):1998–2001, Jun 1983. doi: 10.1103/Phys- RevLett.50.1998.

109 Bibliography

[34] J. Tersoff and D. R. Hamann. Theory of the scanning tunneling microscope. Phys. Rev. B, 31(2):805–813, Jan 1985. doi: 10.1103/PhysRevB.31.805.

[35] J. Wintterlin, J. Wiechers, H. Brune, T. Gritsch, H. Höfer, and R. J. Behm. Atomic- resolution imaging of close-packed metal surfaces by scanning tunneling mi- croscopy. Phys. Rev. Lett., 62(1):59–62, Jan 1989. doi: 10.1103/PhysRevLett.62.59.

[36] V. M. Hallmark, S. Chiang, J. F. Rabolt, J. D. Swalen, and R. J. Wilson. Observation of atomic corrugation on au(111) by scanning tunneling microscopy. Phys. Rev. Lett., 59(25):2879–2882, Dec 1987. doi: 10.1103/PhysRevLett.59.2879.

[37] C. Julian Chen. Origin of atomic resolution on metal surfaces in scanning tun- neling microscopy. Phys. Rev. Lett., 65(4):448–451, Jul 1990. doi: 10.1103/Phys- RevLett.65.448.

[38] C. Julian Chen. Tunneling matrix elements in three-dimensional space: The derivative rule and the sum rule. Phys. Rev. B, 42(14):8841–8857, Nov 1990. doi: 10.1103/PhysRevB.42.8841.

[39] C. Julian Chen. Microscopic view of scanning tunneling microscopy. J. Vac. Sci. Technol. A, 9:44, 1991. doi: 10.1116/1.577128.

[40] R. Wiesendanger. Scanning Probe Microscopy and Spectroscopy. Cambridge Uni- versity Press, 2001.

[41] N. D. Lang. Apparent size of an atom in the scanning tunneling microscope as a function of bias. Phys. Rev. Lett., 58(1):45–48, Jan 1987. doi: 10.1103/Phys- RevLett.58.45.

[42] M. Bode. Strukturelle und lokale elektronische Eigenschaften ultradünner Eisenfilme auf W(110). PhD thesis, Universität Hamburg, 1996.

[43] A. Selloni, P. Carnevali, E. Tosatti, and C. D. Chen. Voltage-dependent scanning- tunneling microscopy of a crystal surface: Graphite. Phys. Rev. B, 31(4):2602– 2605, Feb 1985. doi: 10.1103/PhysRevB.31.2602.

[44] A. Wachowiak. Aufbau einer 300mK-Ultrahochvakuum-Rastertunnelmikroskopie- Anlage mit 14 Tesla Magnet und spinpolarisierte Rastertunnelspektroskopie an ferro- magnetischen Fe-Inseln. PhD thesis, Universität Hamburg, 2003.

[45] K. von Bergmann. Iron nanostructures studied by spin-polarised scanning tunneling microscopy. PhD thesis, Universität Hamburg, 2004.

110 Bibliography

[46] R.M. Feenstra, Joseph A. Stroscio, and A.P. Fein. Tunneling spectroscopy of the si(111)2 x 1 surface. Surface Science, 181(1-2):295 – 306, 1987. doi: DOI: 10.1016/0039-6028(87)90170-1. [47] Joseph A. Stroscio, R. M. Feenstra, and A. P. Fein. Electronic structure of the si(111)2 x 1 surface by scanning-tunneling microscopy. Phys. Rev. Lett., 57(20): 2579–2582, Nov 1986. doi: 10.1103/PhysRevLett.57.2579. [48] Vu Thien Binh, S. T. Purcell, N. Garcia, and J. Doglioni. Field-emission electron spectroscopy of single-atom tips. Phys. Rev. Lett., 69(17):2527–2530, Oct 1992. doi: 10.1103/PhysRevLett.69.2527. [49] Vladimir A. Ukraintsev. Data evaluation technique for electron-tunneling spec- troscopy. Phys. Rev. B, 53(16):11176–11185, Apr 1996. doi: 10.1103/Phys- RevB.53.11176. [50] A. Li Bassi, C. S. Casari, D. Cattaneo, F. Donati, S. Foglio, M. Passoni, C. E. Bot- tani, P. Biagioni, A. Brambilla, M. Finazzi, F. Ciccacci, and L. Duo. Bulk cr tips for scanning tunneling microscopy and spin-polarized scanning tunneling mi- croscopy. Applied Physics Letters, 91(17):173120, 2007. doi: 10.1063/1.2800810. [51] A. Schlenhoff, S. Krause, G. Herzog, and R. Wiesendanger. Bulk cr tips with full spatial magnetic sensitivity for spin-polarized scanning tunneling microscopy. Applied Physics Letters, 97(8):083104, 2010. doi: 10.1063/1.3474659. [52] A. Kubetzka. Rastertunnelspektroskopie an magnetischen Nanostrukturen Fe/W(110). PhD thesis, Universität Hamburg, 2002. [53] T. Miyazaki and N. Tezuka. Giant magnetic tunneling effect in fe/al2o3/fe junc- tion. Journal of Magnetism and Magnetic Materials, 139(3):L231 – L234, 1995. doi: DOI: 10.1016/0304-8853(95)90001-2. [54] J. C. Slonczewski. Conductance and exchange coupling of two ferromagnets separated by a tunneling barrier. Phys. Rev. B, 39(10):6995–7002, Apr 1989. doi: 10.1103/PhysRevB.39.6995. [55] S. Heinze. First Principle Theory of Scanning Tunneling Microscopy Applied to Transition-Metal Surfaces. PhD thesis, Universität Hamburg, 2000. [56] M. Bode, M. Getzlaff, and R. Wiesendanger. Spin-polarized vacuum tunneling into the exchange-split surface state of gd(0001). Phys. Rev. Lett., 81(19):4256– 4259, Nov 1998. doi: 10.1103/PhysRevLett.81.4256. [57] M. Morgenstern. Probing the local density of states of dilute electron systems in different dimensions, 2002.

111 Bibliography

[58] S. R. Ovshinsky. Reversible electrical switching phenomena in disordered structures. Phys. Rev. Lett., 21(20):1450–1453, Nov 1968. doi: 10.1103/Phys- RevLett.21.1450.

[59] W. Welnic. Electronic and Optical Properties of Phase Change Alloys studied with ab initio Methods. PhD thesis, Universität Aachen, 2006.

[60] C. Pauly. Rastertunnelmikroskopie-untersuchung von phasenwechselmedien und oxiden, 2009.

[61] T. Matsunaga and N. Yamada. Structural investigation of gesb2te4: A high-speed phase-change material. Phys. Rev. B, 69(10):104111, Mar 2004. doi: 10.1103/Phys- RevB.69.104111.

[62] A. V. Kolobov, P. Fons, A. I. Frenkel, A. L. Ankudinov, J. Tominaga, and T. Uruga. Understanding the phase-change mechanism of rewritable optical media. Nat. Mater., 3:703–708, 2004. doi: 10.1038/nmat1215.

[63] W. Welnic, A. Pamungkas, R. Detemple, C. Steimer, S. Blügel, and M. Wuttig. Unravelling the interplay of local structure and physical properties in phase- change materials. Nat. Mater., 5:56–62, 2006. doi: 10.1038/nmat1539.

[64] M. Wuttig, D. Lüsebrink, D. Wamwangi, W. Welnic, M. Gillesen, and R. Dron- skowski. The role of vacancies and local distortions in the design of new phase- change materials. Nat. Mater., 6:122–128, 2007. doi: 10.1038/nmat1807.

[65] K. Shportko, S. Kremers, M. Woda, D. Lencer, J. Robertson, and M. Wuttig. Reso- nant bonding in crystalline phase-change materials. Nat. Mater., 7:653–658, 2008. doi: 10.1038/nmat2226.

[66] D. Lencer, M. Salinga, B. Grabowski, T. Hickel, J. Neugebauer, and M. Wut- tig. A map for phase-change materials. Nat. Mater., 7:972–977, 2008. doi: 10.1038/nmat2330.

[67] M. H. R. Lankhorst, B. W. S. M. M. Ketelaars, and R. A. M. Wolters. Low-cost and nanoscale non-volatile memory concept for future silicon chips. Nat. Mater., 4:347–352, 2005. doi: 10.1038/nmat1350.

[68] N. Matsunami, Y. Yamamura, Y. Itikawa, N. Itoh, Y. Kazumata, S. Miyagawa, K. Morita, R. Shimizu, and H. Tawara. Energy dependence of the yields of ion- induced sputtering of monatomic solids. IPPJ-AM, 32, 1983.

[69] S. Mroczkowski and D. Lichtman. Calculated auger yields and sensitivity fac- tors for kll–noo transitions with 1–10 kv primary beams. Journal of Vacuum

112 Bibliography

Science & Technology A: Vacuum, Surfaces, and Films, 3(4):1860–1865, 1985. doi: 10.1116/1.572933.

[70] T. Kato and K. Tanaka. Electronic properties of amorphous and crystalline ge2sb2te5 films. Japanese Journal of Applied Physics, 44(10):7340–7344, 2005. doi: 10.1143/JJAP.44.7340.

[71] D. Lee, S. S. Lee, W. Kim, C. Hwang, M. B. Hossain, N. L. Hung, H. Kim, C. G. Kim, H. Lee, H. N. Hwang, C.-C. Hwang, T.-Y. Lee, Y. Kang, C. Kim, D.S. Suh, K. H. P. Kim, and Y. Khang. Valence band structures of the phase change mate- rial ge[sub 2]sb[sub 2]te[sub 5]. Applied Physics Letters, 91(25):251901, 2007. doi: 10.1063/1.2825573.

[72] T. Siegrist, P Jost, H. Volker, M. Woda, P. Merkelbach, C. Schlockermann, and M. Wuttig. Disorder-induced localization in crystalline phase-change materials. Nat. Mater., 10:202–208, 2011. doi: 10.1038/nmat2934.

[73] J. Kalb, F. Spaepen, and M. Wuttig. Atomic force microscopy measurements of crystal nucleation and growth rates in thin films of amorphous te alloys. Applied Physics Letters, 84(25):5240–5242, 2004. doi: 10.1063/1.1764591.

[74] M. Gillessen. Lmto dos daten gst.

[75] P. Merkelbach. Xrd measurement on gst.

[76] M. Bode, Wachowiak A., J. Wiebe, A. Kubetzka, M. Morgenstern, and R. Wiesen- danger. Thickness dependent magnetization states of fe islands on w(110): From single domain to vortex and diamond patterns. Applied Physics Letters, 84(6): 948–950, 2004. doi: 10.1063/1.1644613.

[77] A. Hubert and R. Schäfer. Magnetic Domains. Springer, 2000.

[78] M. Bode, R. Pascal, and R. Wiesendanger. Scanning tunneling spectroscopy of fe/w(110) using iron covered probe tips. 15(3):1285–1290, 1997. ISSN 07342101. doi: DOI:10.1116/1.580577.

[79] J. Prokop, A. Kukunin, and H. J. Elmers. Spin-polarized tunneling spectroscopy and magnetic coupling in au-coated fe-mo(110) nanostructures. Phys. Rev. B, 75: 144423, Apr 2007. doi: 10.1103/PhysRevB.75.144423.

[80] J. Hauschild. Morphologie und Magnetismus ultradünner Eisenfilme auf vizinalem Wolfram(110). PhD thesis, Universität Mainz, 1998.

[81] P. R. Wallace. The band theory of graphite. Phys. Rev., 71(9):622–634, May 1947. doi: 10.1103/PhysRev.71.622.

113 Bibliography

[82] J. Güttinger, F. Molitor, C. Stampfer, S. Schnez, A. Jacobsen, S. Dröscher, T. Ihn, and K. Ensslin. Transport in graphene quantum dots.

[83] J. Moser, A. Barreiro, and A. Bachtold. Current-induced cleaning of graphene. Appl. Phys. Lett., 91:163513, 2007. doi: 10.1063/1.2789673.

[84] C. R. Dean, A. F. Young, I. Meric, C. Lee, L. Wang, S. Sorgenfrei, K. Watan- abe, T. Taniguchi, P. Kim, K. L. Shepard, and J. Hone. Boron nitride sub- strates for high-quality graphene electronics. Nature Nano, 5:722–726, 2010. doi: 10.1038/nnano.2010.172.

[85] C. T. White, J. Li, D. Gunlycke, and J. W. Mintmire. Hidden one-electron inter- actions in carbon nanotubes revealed in graphene nanostrips. Nano Letters, 7(3): 825–830, 2007. doi: 10.1021/nl0627745.

[86] C. Stampfer, J. Güttinger, S. Hellmüller, F. Molitor, K. Ensslin, and T. Ihn. Energy gaps in etched graphene nanoribbons. Phys. Rev. Lett., 102:056403, Feb 2009. doi: 10.1103/PhysRevLett.102.056403.

[87] M. Fujita, K. Wakabayashi, K. Nakada, and K. Kusakabe. Peculiar localized state at zigzag graphite edge. Journal of the Physical Society of Japan, 65(7):1920–1923, 1996. doi: 10.1143/JPSJ.65.1920.

[88] O. V. Yazyev. Emergence of magnetism in graphene materials and nanostruc- tures. Rep. Prog. Phys., 73:056501, 2010. doi: 10.1088/0034-4885/73/5/056501.

[89] S. Fajtlowicz, P. N. John, and H. Sachs. On maximum matchings and eigenvalues of benzenoid graphs. Croatica Chemica Acta, 78:195, 2005. doi: 10.1143/JPSJ.65.1920.

[90] Elliott H. Lieb. Two theorems on the hubbard model. Phys. Rev. Lett., 62(10): 1201–1204, Mar 1989. doi: 10.1103/PhysRevLett.62.1201.

[91] Wei L. Wang, Oleg V. Yazyev, Sheng Meng, and Efthimios Kaxiras. Topological frustration in graphene nanoflakes: Magnetic order and spin logic devices. Phys. Rev. Lett., 102(15):157201, Apr 2009. doi: 10.1103/PhysRevLett.102.157201.

[92] G. Allinson, R. J. Bushby, J.-L. Paillaud, and M. Thornton-Pett. Synthesis of a derivative of triangulene; the first non-kekule polynuclear aromatic. J. Chem. Soc., Perkin Trans. 1, 1:385–390, 1995. doi: 10.1039/P19950000385.

[93] J. Inoue, K. Fukui, T. Kubo, S. Nakazawa, K. Sato, D. Shiomi, Y. Morita, K. Ya- mamoto, T. Takui, and K. Nakasuji. The first detection of a clars hydrocarbon, 2,6,10-tri-tert-butyltriangulene: A ground-state triplet of non-kekule polynuclear

114 Bibliography

benzenoid hydrocarbon. Journal of the American Chemical Society, 123(50):12702– 12703, 2001. doi: 10.1021/ja016751y.

[94] T. Chenggang, J. Liying, O. J. Yazyev, Y.-C. Chen, J. Feng, X. Zhang, R. B. Capaz, J. M. Tour, A. Zettl, S. G. Louie, H. Dai, and M. H. Crommie. Spatially resolving edge states of chiral graphene nanoribbons. Nature Physics, pages 1745–2481, 2011. doi: 10.1038/nphys1991.

[95] A. N. N’Diaye, J. Coraux, T. N. Plasa, C. Busse, and T. Michely. Structure of epitaxial graphene on ir(111). New Journal of Physics, 10(4):043033, 2008.

[96] A. T. N’Diaye. Epitaxial Graphene and Cluster Lattices on Iridium (111). PhD thesis, Universität Aachen, 2010.

[97] C. Busse, P. Lazi´c,R. Djemour, J. Coraux, N. Gerber, T.and Atodiresei, V. Caciuc, R. Brako, A. T. N’Diaye, S. Blügel, J. Zegenhagen, and T. Michely. Graphene on ir(111): Physisorption with chemical modulation. Phys. Rev. Lett., 107:036101, Jul 2011. doi: 10.1103/PhysRevLett.107.036101.

[98] Johann Coraux, Alpha T. N Diaye, Carsten Busse, and Thomas Michely. Struc- tural coherency of graphene on ir(111). Nano Letters, 8(2):565–570, 2008. doi: 10.1021/nl0728874.

[99] Yu. S. Dedkov, M. Fonin, U. Rüdiger, and C. Laubschat. Rashba effect in the graphene/ni(111) system. Phys. Rev. Lett., 100(10):107602, Mar 2008. doi: 10.1103/PhysRevLett.100.107602.

[100] Alexander Grüneis and Denis V. Vyalikh. Tunable hybridization between elec- tronic states of graphene and a metal surface. Phys. Rev. B, 77(19):193401, May 2008. doi: 10.1103/PhysRevB.77.193401.

[101] S. Marchini, S. Günther, and J. Wintterlin. Scanning tunneling microscopy of graphene on ru(0001). Phys. Rev. B, 76(7):075429, Aug 2007. doi: 10.1103/Phys- RevB.76.075429.

[102] A. L. Vázquez de Parga, F. Calleja, B. Borca, M. C. G. Passeggi, J. J. Hinarejos, F. Guinea, and R. Miranda. Periodically rippled graphene: Growth and spa- tially resolved electronic structure. Phys. Rev. Lett., 100(5):056807, Feb 2008. doi: 10.1103/PhysRevLett.100.056807.

[103] P. W. Sutter, J.-I. Flege, and E. A. Sutter. Epitaxial graphene on ruthenium. Nature Materials, 7:406–411, 2008. doi: 10.1038/nmat2166.

115 Bibliography

[104] E. Starodub, A. Bostwick, L. Moreschini, S. Nie, F. E. Gabaly, K. F. McCarty, and E. Rotenberg. In-plane orientation effects on the electronic structure, stability, and raman scattering of monolayer graphene on ir(111). Phys. Rev. B, 83:125428, Mar 2011. doi: 10.1103/PhysRevB.83.125428.

[105] I. Pletikosi´c,M. Kralj, D. Šokˇcevi´c,R. Brako, P. Lazi´c,and P. Pervan. Photoe- mission and density functional theory study of ir(111); energy band gap map- ping. Journal of Physics: Condensed Matter, 22(13):135006 (6 pp.), 7 April 2010. doi: 10.1088/0953-8984/22/13/135006.

[106] A. Varykhalov, D. Marchenko, M. R. Scholz, E. Rienks, T. K. Kim, and O. Rader. Topological surface state under graphene for two-dimensional spintronics in air. ARXIV, 2011. URL http://arxiv.org/abs/1104.3308v1.

[107] S. Reich, J. Maultzsch, C. Thomsen, and P. Ordejón. Tight-binding description of graphene. Phys. Rev. B, 66(3):035412, Jul 2002. doi: 10.1103/PhysRevB.66.035412.

[108] A. Grüneis, C. Attaccalite, L. Wirtz, H. Shiozawa, R. Saito, T. Pichler, and A. Ru- bio. Tight-binding description of the quasiparticle dispersion of graphite and few-layer graphene. Phys. Rev. B, 78(20):205425, Nov 2008. doi: 10.1103/Phys- RevB.78.205425.

[109] F. Libisch, S. Rotter, J. Güttinger, C. Stampfer, and J. Burgdörfer. Transition to landau levels in graphene quantum dots. Phys. Rev. B, 81(24):245411, Jun 2010. doi: 10.1103/PhysRevB.81.245411.

[110] P.Giannozzi, S. Baroni, N. Bonini, M. Calandra, R. Car, C. Cavazzoni, D. Ceresoli, Chiarotti. G. L., M. Cococcioni, I. Dabo, A. Dal Corso, S. de Gironcoli, S. Fabris, Fratesi. G., 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, Scandolo. S., G. Sclauzero, A. P. Seitsonen, A. Smo- gunov, P. Umari, and R. M. Wentzcovitch. Quantum espresso: a modular and open-source software project for quantum simulations of materials. Journal of the American Chemical Society, 21(39):395502, 2009. URL http://stacks.iop.org/ 0953-8984/21/i=39/a=395502.

[111] C. Busse, P. Lazi´c,R. Djemour, J. Coraux, T. Gerber, N. Atodiresei, V. Caciuc, R. Brako, A. T. N’Diaye, S. Blügel, J. Zegenhagen, and T. Michely.

[112] J. P. Perdew, K. Burke, and M. Ernzerhof. Generalized gradient approxima- tion made simple. Phys. Rev. Lett., 77:3865–3868, Oct 1996. doi: 10.1103/Phys- RevLett.77.3865.

116 Bibliography

[113] A. D. Corso and A. M. Conte. Spin-orbit coupling with ultrasoft pseudopo- tentials: Application to au and pt. Phys. Rev. B, 71:115106, Mar 2005. doi: 10.1103/PhysRevB.71.115106. [114] H. J. Monkhorst and J. D. Pack. Special points for brillouin-zone integrations. Phys. Rev. B, 13:5188–5192, Jun 1976. doi: 10.1103/PhysRevB.13.5188. [115] C. Wittneven, R. Dombrowski, M. Morgenstern, and R. Wiesendanger. Scatter- ing states of ionized dopants probed by low temperature scanning tunneling spectroscopy. Phys. Rev. Lett., 81(25):5616–5619, Dec 1998. doi: 10.1103/Phys- RevLett.81.5616. [116] M. Morgenstern. Probing the local density of states of dilute electron systems in different dimensions. Surf. Rev. Lett., 10:933, 2003. [117] P. M. Echenique, R. Berndt, E. V. Chulkov, T. Fauster, A. Goldmann, and U. Höfer. Decay of electronic excitations at metal surfaces. Surface Science Reports, 52(7-8): 219 – 317, 2004. doi: DOI: 10.1016/j.surfrep.2004.02.002. [118] Cheol-Hwan Park, Feliciano Giustino, Marvin L. Cohen, and Steven G. Louie. Velocity renormalization and carrier lifetime in graphene from the electron- phonon interaction. Phys. Rev. Lett., 99(8):086804, Aug 2007. doi: 10.1103/Phys- RevLett.99.086804. [119] M. Calandra and F. Mauri. Electron-phonon coupling and electron self-energy in electron-doped graphene: Calculation of angular-resolved photoemission spec- tra. Phys. Rev. B, 76(20):205411, Nov 2007. doi: 10.1103/PhysRevB.76.205411. [120] J. P. Carbotte, E. J. Nicol, and S. G. Sharapov. Effect of electron-phonon inter- action on spectroscopies in graphene. Phys. Rev. B, 81(4):045419, Jan 2010. doi: 10.1103/PhysRevB.81.045419. [121] J. Gonzalez, F. Guinea, and M. A. H. Vozmediano. Unconventional quasipar- ticle lifetime in graphite. Phys. Rev. Lett., 77(17):3589–3592, Oct 1996. doi: 10.1103/PhysRevLett.77.3589. [122] V. W. Brar, S. Wickenburg, M. Panlasigui, C.-H. Park, T. O. Wehling, Y. Zhang, R. Decker, Ç. Girit, A. V. Balatsky, S. G. Louie, A. Zettl, and M. F. Crommie. Ob- servation of carrier-density-dependent many-body effects in graphene via tun- neling spectroscopy. Phys. Rev. Lett., 104(3):036805, Jan 2010. doi: 10.1103/Phys- RevLett.104.036805. [123] E. H. Hwang, Ben Yu-Kuang Hu, and S. Das Sarma. Inelastic carrier life- time in graphene. Phys. Rev. B, 76(11):115434, Sep 2007. doi: 10.1103/Phys- RevB.76.115434.

117 Bibliography

[124] V. M. Pereira, A. H. Castro Neto, and N. M. R. Peres. Tight-binding ap- proach to uniaxial strain in graphene. Phys. Rev. B, 80(4):045401, Jul 2009. doi: 10.1103/PhysRevB.80.045401. [125] E. Mariani and F. von Oppen. Temperature-dependent resistivity of sus- pended graphene. Phys. Rev. B, 82(19):195403, Nov 2010. doi: 10.1103/Phys- RevB.82.195403. [126] E. Mariani and F. von Oppen. Flexural phonons in free-standing graphene. Phys. Rev. Lett., 100(7):076801, Feb 2008. doi: 10.1103/PhysRevLett.100.076801. [127] K. M. Borysenko, J. T. Mullen, E. A. Barry, S. Paul, Y. G. Semenov, J. M. Zavada, M. B. Nardelli, and K. W. Kim. First-principles analysis of electron-phonon in- teractions in graphene. Phys. Rev. B, 81(12):121412, Mar 2010. doi: 10.1103/Phys- RevB.81.121412. [128] T. Aizawa, Y. Hwang, W. Hayami, R. Souda, S. Otani, and Y.favorable Ishizawa. Phonon dispersion of monolayer graphite on pt(111) and nbc surfaces: bond softening and interface structures. Surface Science, 260(1-3):311 – 318, 1992. doi: DOI: 10.1016/0039-6028(92)90046-9. [129] W. K. Tse, E. H. Hwang, and S. D. Sarma. Ballistic hot electron transport in graphene. Appl. Phys. Lett., 93:023128, 2008. doi: 10.1063/1.2956669. [130] T. Wassmann, A. P. Seitsonen, A. M. Saitta, M. Lazzeri, and F. Mauri. Clars the- ory, pi-electron distribution, and geometry of graphene nanoribbons. Journal of the American Chemical Society, 132(10):3440–3451, 2010. doi: 10.1021/ja909234y. PMID: 20178362. [131] H. Feldner, Z. Y. Meng, A. Honecker, D. Cabra, S. Wessel, and F. F. Assaad. Mag- netism of finite graphene samples: Mean-field theory compared with exact diag- onalization and quantum monte carlo simulations. Phys. Rev. B, 81:115416, Mar 2010. doi: 10.1103/PhysRevB.81.115416. [132] J. Crangle. Solid State Magnetism. Edward Arnold, 1991. [133] T. Mashoff, M. Pratzer, V. Geringer, T. J. Echtermeyer, M. C. Lemme, M. Lieb- mann, and M. Morgenstern. Bistability and oscillatory motion of natural nanomembranes appearing within monolayer graphene on silicon dioxide. Nano Letters, 10(2):461–465, 2010. doi: 10.1021/nl903133w. PMID: 20058873. [134] D Kleckner, I. Pikovski, E. Jeffrey, L. Ament, E. Eliel, J. van den Brink, and D. Bouwmeester. Creating and verifying a quantum superposition in a micro- optomechanical system. New Journal of Physics, 10(9):095020, 2008. doi: 10.1088/1367-2630/10/9/095020.

118 Bibliography

[135] D. V. Anghel and T. Kühn. Quantization of the elastic modes in an isotropic plate. Journal of Physics A: Mathematical and Theoretical, 40(34):10429, 2007. doi: 10.1088/1751-8113/40/34/003.

[136] A. K. Geim and K. S. Novoselov. The rise of graphene. Nat Mater, 6:183–191, 2007. doi: 10.1038/nmat1849.

119

Publications

1. D. Subramaniam, C. Pauly, M. Liebmann, M. Woda, P. Rausch, P. Merkelbach, M. Wuttig and M. Morgenstern. Scanning tunneling microscopy and spectroscopy of the phase change alloy Ge1Sb2Te4. Appl. Phys. Lett., 95 (10):103110, 2009. doi:10.1063/1.3211991.

2. V. Geringer, D. Subramaniam, A. K. Michel, B. Szafranek, D. Schall, A. Georgi, T. Mashoff, D. Neumaier, M. Liebmann, and M. Morgenstern. Electrical trans- port and low-temperature scanning tunneling microscopy of microsoldered graphene. Appl. Phys. Lett., 96 (8):082114, 2010. doi:10.1063/1.3334730.

3. D. Subramaniam, F. Libisch, Y. Li, C. Pauly, V. Geringer, R. Reiter, T. Mashoff, M. Liebmann, J. Burgdoerfer, C. Busse, T. Michely, R. Mazzarello, M. Pratzer, M. Morgenstern. Wave function mapping in graphene quantum dots with soft con- finement. Phys. Rev. Lett., 108:046801, 2012. doi:10.1103/PhysRevLett.108.046801.

121

Curriculum vitae

• Surname: Subramaniam

• Name: Dinesh

• Date of Birth: 25.12.1979

• Place of Birth: Mundel (Sri Lanka)

• Nationality: German

Qualifications:

• 05.06.2000: Abitur

• 2000-2006: Physik Diplom (RWTH Aachen)

• 2006-2012: PhD at II. Physikalischen Institut B der RWTH Aachen

123

Acknowledgments

First and foremost, I sincerely like to thank Prof. Dr. Markus Morgenstern, who gave me the opportunity to perform this thesis in his institute. It is thanks to him that this work was done in an institute, which provide a warm and highly scientific atmosphere. He cared about my studies in every detail and offered a lot of efficient suggestions. His great enthusiasm and deep insights in physics encouraged and helped me a lot whenever I met any difficulty. I would like to acknowledge Prof. Dr. Wuttig, who agreed to referee this thesis. Special thanks to Dr. Marco Pratzer who introduced me to the scanning tunneling microscopy field during my diploma thesis. His technical skills have helped me in many many situations during my phd work. He has always very good ideas and I learnt a lot from him. Warm thanks to Dr. Marcus Liebmann who always had an open ear for my theoreti- cal problems and was very helpful in the evaluation of my data. A lot of thanks to Florian Libisch, Yan Li and Riccardo Mazzarello for the very suc- cessful cooperation. Special thanks to my friend Dr. Viktor Geringer with whom I shared a several end- less nights of experiments in the 4K-lab and also a lot of incredible private time. I look forward to continued the fruitful cooperation with him in our new work. Many thanks to Christian Pauly with whom I have worked the last 2-3 years very closely. We have spent not only hours, days, weekend at the institute 2B, but also made our first small steps toward handling/understanding ARPES-System in Bessy/Berlin. Warm thanks to Mike Pezzotta and Dr. Stefan Becker with whom I shared a great times at the institute. I would like to thanks Dr. Torge Mashoff, Alex Georgi, Nils Freitag, Kilian Floehr, Christian Saunus, Raphael Bindel, Martin Grob, Anne Majerus, Silke Hattendorf, Lisa Felker, Sven Runte and Thomas Marzi for the enjoyable time in the institute, on the soccer field, in pubs... Many thanks to Mr. Kordt, Mr. Retetzki and the workshop for the remarkable and very professional collaboration. I thank Sascha Mohr and Joerg Schirra for the continuous supply of helium. Espe- cially the night and Christmas Helium-support requires a huge gratefulness. I thank the Secretariat for the helpful and friendly work.

125