• Keine Ergebnisse gefunden

All experiments discussed in this thesis were performed in the group of Prof. Klaus Kern at the Max Planck Institute for Solid State Research in Stuttgart, Germany. Unless oth-erwise stated, the STM and AFM data was exclusively recorded at the same home-built experimental system discussed in Chapter 3. The first microscope was designed and built from 2001 – 2005 by Messina and Wittich [27, 28] and was since then constantly changed

1.3. Location and History of the Experiment

and improved. This is especially true for the microscope head, where four versions are still partly operational in storage. During 2009 a new combined STM/AFM head was designed by the author [14] and put into operation during this thesis in 2011. In 2013 the whole system was moved to a new high precision laboratory, where some significant technical changes were made that will be discussed.

All STM/AFM images are processed either withWSxM[29],MatLaborGWYDDION. For data processing Origin, Excel, Scilab and Matlab were used. Figures were made with the following programs: Power Point, Corel DRAW X6, Corel PHOTO-PAINT X6, Inkscape andAdobe Photoshop CS6. The designing and the technical drawings shown in the thesis were done with Creo ElementsfromPTC.

CHAPTER 2

Theory and Techniques

The STM was invented in 1981 by G. Binnig and H. Rohrer [30, 31] and earned them the Nobel prize in physics already five years later in 1986 (jointly with E. Ruska for the invention of the electron microscope). While a full coverage of the technique and the more than 30 years of evolution is not feasible, only the very basic principle of STM will be discussed. The same holds true for other scanning probe methods that emerged already shortly after, like AFM in 1986 [32]. For an introduction Chen’s “Introduction to Scanning Tunneling Microscopy” [33] and “Noncontact Atomic Force Microscopy” from Morita et al. [34] are recommended.

After looking at the STM in the following section AFM is introduced in Section 2.3 with a special focus on the design where STM and AFM are combined in the tuning fork design (Section 2.4).

2.1. Scanning Tunneling Microscopy

2.1.1. Working Principle

In a simplified picture the working principle and the physics of STM can be summarized as follows: A conducting wire (the tip) is placed with a distance of the order of a few

˚Angstrom (1 ˚A= 10−10 m) to a conducting sample. When a voltage is applied between tip and sample, a small current (in the range of pico- to nanoamperes) can be detected, which flows due to the quantum mechanical nature of the electrons. This effect is called

“tunneling”. By scanning the tip over the sample and recording the current, a map can be recorded that is (in a first approximation) an image of the topography of the sample.

Figure 2.1 illustrates the working principle.

Two modes of operation are commonly deployed: In constant height mode the tip is scanned at a fixed height and an image of the current is recorded. Modern electronics also allow adjustment of the scan plane to align it parallel to the sample surface. This mode requires rather flat surfaces and stable overall performance to preclude crashing of the tip into the surface. In constant current mode the current is used as a feedback signal and the tip-sample distance is adjusted to maintain a constant current and is recorded as the signal. For the feedback a proportional-integral controller has to be set up according to

z-control x-/y-control

Control electronics (Nanonis)

Feedback loop I

VBias PC software

(Nanonis)

Scanner

Tip

z y

x

Figure 2.1.: Working principle of an STM: The metallic tip is brought into close proximity of the sample. The tunneling current flowing due to the applied bias is measured and used as a feedback signal to regulate the distance between tip and sample (zdirection) while the tip is scanned in thexand ydirection to acquire an image.

scan speed, resolution and sample roughness to obtain optimum results.

The quantification of the tunneling process is based on the overlap of the electronic wave functions, which describe the states of the tip and the sample. The probability of tunneling depends on the electron energy (bias) and the height and thickness of the barrier. The barrier height corresponds to the work function that accounts for the energy necessary to remove an electron from the tip and inject it into the sample (or vice versa). The thickness of the barrier is the gap (typically vacuum) between the tip and the surface.

As will be derived in Section 2.1.2, the tip-sample distance dependence of the tunneling current is exponential. This is crucial for STM operation, as the current is extremely sensitive to very small height corrugations (it changes by an order of magnitude for a change in height of only about 1 ˚A) and because only the last few atoms at the tip apex contribute to the current. This makes the exact geometry of the tip usually not so crucial for imaging.

STM images do not directly reflect the topography of the sample, but rather reveal the spatial distribution of the electronic structure, which is the convolution of the local density of states (LDOS) of tip and sample, as will be discussed in the following section. Analogous to the system used in this work, the following discussion will be limited to the case where the bias voltage is applied to the sample, i.e. for positive bias electrons flow from occupied states of the tip to the empty states of the sample.

2.1. Scanning Tunneling Microscopy

2.1.2. Theoretical Description

d Re V

Tip

Sample

E

Energy

Figure 2.2.: Schematic of the one-dimensional tunneling barrier, as discussed in the text.

Because the simple one-dimensional model depicted in Figure 2.2 with a square barrier of height V, thickness d and an electron with energy E leads to an exponentially declining tunneling probability, it will be briefly derived in the following.

Starting point is the Schr¨odinger equation:

Eψ(x) = (− ̵h2

2me2+V(x))ψ(x), (2.1)

with me the electron mass and V(x) the potential barrier. This equation is solved by plane waves outside the barrier and by an exponential decay inside the barrier, with the boundary conditions determining the amplitudes. With κ∶=√

2m(V−E)

h̵2 the transmission coefficient that describes the probability of an electron to penetrate through the barrier, can be calculated to be:

T = 4E(V −E)

4E(V −E) +V2sinh2(κd)

≈ 16E

V (1−E

V) ⋅e−2κd forκd≫1. (2.2)

This shows the important exponentially declining probability of an electron to penetrate through the barrier as a function of barrier thickness d.

In a more realistic system, such as that depicted in Figure 2.3, (V −E) is replaced by the average work function of sample and tip Φ= 12ST)and all possible conductance channels are summed up. Therefore, the local density of states (LDOS) of the sample and the tipρS andρT, respectively, are introduced to describe the states available at a certain energy. A net current is obtained when a bias V is applied. For a positive (negative) voltage electrons flow from the filled states of the tip (sample) to the empty states of the sample (tip).

To quantify the current, one starts from Fermi’s golden rule:

I(V) ∝

−∞

ρS(E)ρT(E−eV) ∣M(E, V, z)∣2(fT(E−eV, T) −fS(E, T)) dE , (2.3)

eV

Sample E

ρT(E) ρS(E) ΦT

ΦS Tip

Evac

Evac

EF EF

Figure 2.3.: Simplified energy diagram of the tunneling junction with the density of states of tip and sample depicted in gray.

where fT and fS are the Fermi-Dirac distributions of tip and sample, respectively, T is the temperature, and M is the transmission matrix for the tunneling barrier. According to Bardeen’s theory for tunneling between metals [35], the matrix element is given by the overlap of their wave functions:

M(E, V, z) = h̵2

2m∫ (ψS∇ψT −ψT∇ψS) dA , (2.4) with integration over a surfaceAseparating sample and tip. The calculation of this matrix element is in general not feasible, because the wave functions are not exactly known. In 1983 Tersoff and Hamann introduced a model [36, 37] that makes it possible to calculate M with the fundamental assumptions of a single (spherical) s-orbital for the tip. The sample wave function is described by plane waves. For the matrix element it follows:

M2 ∝ exp(−2z

√m

2ST −2E+eV)), (2.5)

where the effect of a finite temperature is neglected. With all the approximations made, one gets forT = 0, which corresponds to a step function instead of the Fermi distribution:

I(z, V) ∝ρT

EF+eV

EF

ρS(E)exp(−2z

√m

2ST −2E+eV)) dE , (2.6) which shows the same exponential distance dependence as the simple one dimensional model (Equation (2.2)). If one further assumes the bias voltage to be small compared to Φ and thus the matrix element being independent of E, the only voltage dependence which remains is:

I(V) ∝

EF+eV

EF

ρS(E)dE . (2.7)

2.2. Scanning Tunneling Spectroscopy

Furthermore, if the bias is assumed to be small, the only proportionality which remains is:

I ∝ ρS(EF). (2.8)

This emphasizes that the recorded current in STM can be interpreted – with the mentioned limitations – as an image of the LDOS around EF of the substrate. The exponential dis-tance dependence of the current remains valid and contributes the topographic information to the recorded image.

2.2. Scanning Tunneling Spectroscopy

While STM is mainly deployed to learn about the topography of a sample, scanning tun-neling spectroscopy (STS) is a powerful technique to investigate the electronic properties of the system of interest at the atomic scale. A number of fundamental questions were suc-cessfully addressed by this technique. Already before the rise of scanning probe techniques tunneling spectroscopy had shown to be a powerful technique, as nicely summarized in ref. [38], were also spin-flip and Kondo scattering is discussed.

In the following, an introduction to inelastic dI/dV-spectroscopy is given, as a significant part of the results in this thesis was obtained by this technique. The measurement setup deploys a lock-in amplifier and the limitations relevant for the energy resolution are briefly discussed.

(a) (b)

E0

Current

Energy 0

E0

dI/dV

Energy 0

Figure 2.4.: Example of an inelastic tunneling process. (a) Tunneling junction with a molecule inside. A vibrational mode is excited if the electron energyeV is larger thanE0. (b) I(V)-curve showing the additional conductance channel opening for V >E0. In the dI(V)/dV-curve below this is visible as a step (and in the second derivative as a peak).

A constant DOS is desirable (and in the following assumed) for the tip in the bias range of interest. In the experiments the tip termination is usually unknown, therefore reference spectra on the substrate have to be recorded allowing for unambiguous identification of

features that are only present on the structure of interest.

Starting from Equation (2.7) and differentiating it with respect to V one finds the pro-portionality:

dI

dV(V0) ∝ρS(eV0). (2.9)

By recording I(V)−curves and calculating the derivative one has hence direct access to the LDOS of the sample as a function of bias. In practice a more elegant method de-ploying lock-in detection is used, which allows for direct measurement of the differential conductance. Therefore a small AC voltage is added to the tunneling bias by the lock-in amplifier:

V =VBias+VLIsin(2πf t), (2.10)

with VLI being the lock-in modulation voltage and f the frequency of the modulation.

The detection signal, which is proportional to the LDOS atVBias averaged over the range of±VLI, is then recovered from the tunneling current by the lock-in amplifier electronics.

The lock-in excitation voltageVLI has to be chosen according to the experimental require-ments: For the experiments performed in this work, where usually features with widths of a few meV are investigated at T = 1.5 K, it should be in the range of 0.1 – 1 mV. It can be shown that at 1 mV the obtained resolution is mainly limited by the modulation voltage to 2.5 meV, while at 0.1 mV excitation the temperature broadening of the features is more crucial, allowing an increase in resolution down to 0.3 meV at 1.5 K [39]. The frequency of the excitation voltage has to be chosen below the bandwidth of the current amplifier and above the bandwidth of the scan feedback loop to avoid interference, and is typically in the range of 600 – 800 Hz. System inherent parasitic frequencies should be also avoided to obtain a clear signal.

A broad range of phenomena can be investigated by inelastic electron tunneling spec-troscopy (IETS) where the tunneling electrons excite processes in the junction by losing some of their energy. The first application of this method dates back to 1966 where Jaklevic and Lambe used it to investigate molecular vibrations (O-H and C-H bending and stretching modes) in metal-oxide-metal junctions [40]. Only the invention of STM made it possible to address those (and further) properties on the individual molecule level and with unprecedented knowledge of the environment on the supporting substrate. Here, the first measurements of inelastic vibrational spectra on single molecules by STM in the group of W. Ho in 1998 are worth highlighting [41].

Inelastic tunneling occurs if the sample provides two discrete energy states and if the applied bias provides enough energy for the tunneling electron to excite the state by over-coming the energy difference between the two states, i.e. if eV ≥ E0. This additional channel results in an increase of the observed tunneling current above the threshold volt-age E0, as depicted in Figure 2.4. Note that this kink in the I(V)−curve will show up as step and peak in the dI/dV and second derivative (d2I/dV2−signal), respectively. As the excitation does not depend on the direction of the current the observed features are symmetric around zero bias.

Further aspects of IETS, especially spin-flip spectroscopy, where the pioneering work by Heinrichet al. [42] on the observation of the excitation of single electron spins has to be mentioned, will be discussed in the context of the experiments in Chapter 5.

2.3. Atomic Force Microscopy

2.3. Atomic Force Microscopy

(a) (b)

z y x

I

Figure 2.5.: Illustration of the working principle of an AFM: (a) In the most common setup util-ising an optical detection mechanism a laser beam is reflected from the cantilever and the bending motion due to the interaction with the sample is measured with a detector array. (b) In the setup implemented in this system an oscillating quartz tuning fork with a tip attached to it is used as a force sensor that is scanned over the surface. Due to forces acting between tip and sample the frequency and amplitude of the oscillat-ing prong change from its unperturbed values. The oscillation amplitude is directly accessible as a small current, which is induced due to the piezoelectric properties of the quartz.

2.3.1. Working Principle

In AFM a sharp tip is attached to a cantilever and scanned over the sample. Due to the forces that act between tip and surface (see Figure 2.6 (a)) the cantilever bends and thereby the interaction force can be indirectly measured. The cantilever deflection can be measured in different ways: While in the first AFM an STM was used [32], also capacitive, piezoelectric, or optical schemes are deployed. The deflection of a laser beam as sketched in Figure 2.5 (a) is the most common implementation. The most important advantage of AFM over STM is its ability to also investigate insulating samples. This makes AFM also popular in the life sciences.

The different characteristics of AFMs are as broad as the range in which they are used.

To understand the huge advantages of the chosen sensor setup deployed in this work, it is worth quickly summarizing the most popular modes of operation. In ref. [44] an extensive review on the AFM modes is given by Garcia and Perez.

A first reasonable criterion to distinguish operation modes is based on whether or not a feedback is controlling the scan, and if so, which kind. In constant height mode, where the cantilever is scanned at a fixed height above the surface, no feedback is active and only the deflection is measured. While in normal setups this mode is rarely used, as it requires extremely flat samples to avoid crashing, it has to be deployed in a slightly modified way

0.4 0.6 0.8 1.0 -0.1

0.0 0.1

Force (nN)

Distance (nm)

(a) (b)

attractive repulsive

Surface

Cantilever

approx. 10µm: Fluid damping

0.1-1 µm: Electrostatic forces (attractive or repulsive) 10-200 nm: Capillary forces (attractive) few Angstroms: Van der Waals forces (attractive) sub-Angstrom: Pauli repulsion

Figure 2.6.: (a) Forces acting between a tip and a surface, based on [43]. (b) The force-distance curve based on a Lennard-Jones potential (with parameters as discussed in Chapter 4).

The attractive part of the potential is assumedr−6 (based on van der Waals inter-action, green curve) and the repulsive part r−12 (Pauli repulsion, red curve). Note that the forces are plotted and not the potential (F = −∇E).

in order to correctly quantify the tip-sample interaction, as will be discussed below.

On the other hand, several signals like deflection, amplitude, frequency shift, and phase can be used to establish a feedback loop regulating the tip-sample distance. In contrast to STM with its monotonic exponential dependance of the current on the tip-sample distance, no such simple functional dependence between the force and the distance exists. This is due to the interplay of several different types of forces present. Figure 2.6 (a) gives an overview of the forces that might be involved, including typical distances. As the work performed here is carried out in UHV fluid damping and capillary forces can be disregarded. If only van der Waals and repulsive forces due to the Pauli repulsion are considered, a good approximation is given by the force-distance curve shown in Figure 2.6 (b), which is based on a Lennard-Jones potential [45].

Chronologically, the first AFMs operated in contact mode, where the tip, attached to a very soft cantilever, is “scratched” over the surface. This quickly allowed fairly high resolution as reported by Martiet al. in 1987 [46]. A serious advancement was the introduction of tapping mode AFM. This is a dynamic operation mode, where the cantilever is externally excited to oscillate, while the amplitude is used as a feedback signal. Amplitudes in the range of 10 – 100 nm and fast scanning speeds make it popular for application in biology.

The impact from the tip to the sample in contact mode is highly reduced in tapping mode AFM, as is nicely shown by Zhonget al. where the two methods are compared [47].

A further advancement, which is also relevant for this work, is the invention of the true non-contact mode, also a dynamic mode, with amplitudes typically below 10 nm. This mode can be achieved using the oscillation amplitude as feedback (“AM nc-AFM”) [48], or the frequency (“FM nc-AFM”) as introduced by Albrecht et al. in 1991 [49]. The latter requires rather stable oscillation, hence relatively stiff cantilevers are used. This technique provided true atomic resolution to resolve the reactive Si(111) surface and its

2.3. Atomic Force Microscopy

7×7 reconstruction [50].

Similar to STM, where the acquired image is a convolution of topography and electronic structure, the interpretation of dynamic mode AFM images is nontrivial. In the following section the method of choice to calculate the forces quantitatively from the AFM signal will be presented.

2.3.2. Calculation of the Force

Mount / Piezo

Sample

m Tip

Cantilever

Interaction k

ki

Figure 2.7.: A simple model illustrating the frequency shift of the free cantilever oscillation (with spring constantkand effective massm) due to the interactionki between the tip and the sample.

The calculation of the interaction forces between tip and sample from the frequency shift is not straightforward. Because the cantilever oscillation is harmonic it can be characterized by a spring constant k (f ∝ √

k/m). If it is exposed to a force gradient, which can be described by a spring constant (dF/dz =ki), the oscillation frequency will change [49].

This interplay is depicted in Figure 2.7. The two springs are considered to be in series resulting in the following frequency of the oscillation:

f = 1 2π

√k+ki

m , (2.11)

which is only valid if Hooke’s law holds, i.e. if ki ≠ ki(z) (i.e. ki is constant over the oscillation cycle). For highly nonlinear force-distance laws found at the atomic scale (Figure 2.6 (b) shows an example of a Lennard-Jones type force) this requirement is not met. To circumvent this limitation, the averaged value ki(z) is introduced. If the inter-action is small compared to the stiffness of the used cantilever (ki(z) ≪k), the frequency shift can be expressed as:

∆f = f−f0 (2.12)

≈ f0

ki(z)

2k . (2.13)

Giessibl calculated ki(z) with first order perturbation theory using the Hamilton-Jacobi