Sei sulla pagina 1di 83

TEL AVIV UNIVERSITY

The Iby and Aladar Fleischman Faculty of Engineering


The Zandman-Slaner School of Graduate Studies

Reconstruction of Kelvin Probe Force Microscopy images

A thesis submitted toward the degree of


Master of Science in Electrical and Electronic Engineering

by

Gilad Cohen

November 2012

TEL AVIV UNIVERSITY


The Iby and Aladar Fleischman Faculty of Engineering
The Zandman-Slaner School of Graduate Studies

Reconstruction of Kelvin Probe Force Microscopy images

A thesis submitted toward the degree of


Master of Science in Electrical and Electronic Engineering

by

Gilad Cohen

This research was carried out in the Department of Electrical Engineering Physical Electronics, under the supervision of Prof. Yossi Rosenwaks

November 2012

I would like to express my gratitude for the advices and directions provided by Prof.
Yossi Rosenwaks, my supervisor. The discussions and suggestions from him contributed
substantially to this thesis. I also would like to thank Prof. Amir Boag for his guidance
and assistance along the way.

I want to express my appreciations to my colleagues in Tel Aviv University for


supporting me with everything I needed. Special thanks go to: George Elias, Ezer
Halperin, Iddo Amit and Gil Shalev for lots of constructive suggestions and helpful
conversations.

Table of contents
Abstract

ii

List of Publications

iii

List of Figures

iv

List of symbols and abbreviations

ix

1
2
5
8

Introduction
1.1
Kelvin Probe Force Microscopy (KPFM)
1.2
Frequency Modulation KPFM
1.3
Comparison between Amplitude and Frequency Modulations
Literature Review

Electrostatic Model
3.1
Probe-Sample Electrostatic System
3.2
Minimum Force/Force-gradient Condition
3.3
Point Spread Function (PSF) Integration
3.4
Model Validation

14
14
18
20
22

The Effect of Probe Geometry on PSF


4.1
Force/Force-gradient Distributions
4.2
Properties of AM-PSF and FM-PSF
4.3
The Cantilever Effect

25
25
28
31

5
6

Tip-Sample Distance Effect

33

Image Reconstruction
6.1
Outline of the Deconvolution Method
6.2
Reconstruction Process
6.2.1
PSF Expansion
6.2.2
Calibration Sample Measurements
6.3
Results and Comparison with Measurements
6.3.1
Reconstruction of CdS-PbS Images
6.3.2
Reconstruction of Graphene Images

37
37
39
39
42
49
49
54

Summary and Conclusions

58

References

59

Appendix

62
[i]

Abstract
The main goal of this thesis is to develop an algorithm for reconstructing the surface
potential from its Kelvin probe force microscopy (KPFM) measurements. The KPFM
measures the surface potential, however due to the long range electrostatic forces the
measured potential is a weighted average and not the real potential under the tip apex. In
addition, we develop a method to calculate the point spread function (PSF) of frequency
modulation KPFM (FM-KPFM) and compare it with amplitude modulation KPFM (AMKPFM). In FM-KPFM the probe detects the force-gradient rather than the force and thus
demonstrates better spatial resolution than in AM-KPFM.

In order to reconstruct the real surface potential we estimate the system noise statistics
and calculate the exact PSF for the KPFM measurement. The image reconstruction is
then performed by applying the Wiener filter. The reconstruction algorithm is validated
by measuring a calibration sample under a known bias. We exhibit reconstruction of
surface potential on CdS-PbS nanorods measured with AM-KPFM in argon atmosphere
and surface potential on Graphene layers measured with FM-KPFM in ultra-high
vacuum. We show that in AM-KPFM measurements the averaging effect is very
dominant whereas FM-KPFM measurements demonstrate no averaging effect.

Furthermore, we analyze the effect of the tip-sample distance on the spatial resolution
and on the attenuation factor of the measured potential. By inspecting the full width at
half maximum (FWHM) of the PSFs we show that FM-KPFM demonstrate a superior
spatial resolution than AM-KPFM. We also show that for conventional tip-sample
distances (1nm-50nm) there is almost no attenuation for FM-KPFM, however in AMKPFM measurements conducted above 10nm the measured potential is derived mainly
from the substrate and not from the feature beneath the tip apex.

[ii]

List of Publications

Nanayakkara, S., Cohen, G., Jiang. C., Romero, M., Maturova, K., AlJassim, M., Lagemaat, J., Rosenwaks, Y., Luther, J. Built-in potential
and charge distribution within single heterostructured nanorods
measured by scanning Kelvin probe microscopy. Under review
(2012).

[iii]

List of Figures
FIG.1: SCHEMATICS OF A TYPICAL KPFM TIP ABOVE A P-N JUNCTION SHOWING THAT THE TIP
DIMENSIONS ARE MUCH BIGGER THAN THE JUNCTION REGION UNDERNEATH IT.

FIG. 2: DEFINITION AND BASIC MEASUREMENT SETUP OF CONTACT POTENTIAL DIFFERENCE


(CPD). - WORK FUNCTION OF THE TIP, - WORK FUNCTION OF THE SAMPLE, - FERMI

LEVEL, LVL - LOCAL VACUUM ENERGY LEVEL.

FIG. 3: PRINCIPLE OF FM-KPFM, THE FREQUENCY SHIFT VS. TIP SAMPLE BIAS. THE KELVIN
CONTROLLER MINIMIZES THE AMPLITUDE OF THE FREQUENCY SHIFT, WHICH IS SMALLEST
WHEN VDC = VCPD .

FIG. 4: SCHEMATIC FREQUENCY SPECTRUM OF THE TIP OSCILLATION. THE PEAKS AT fmod

AND 2fmod ORIGINATE FROM THE ELECTROSTATIC FORCE, WHEREAS THE PEAKS AT

f0 fmod AND f0 2fmod SHOW THE FREQUENCY MODULATION AT f0 .

FIG. 5: A MODEL IN WHICH THE AFM TIP IS REPRESENTED BY A SERIES OF PARALLEL PLATE
CAPACITORS Ci AT DISTANCES Zi FROM THE SAMPLE SURFACE. ONE PLATE OF A CAPACITOR
IS LOCATED ON THE TIP AND THE OTHER ON THE SAMPLE SURFACE.

FIG. 6: SCHEMATIC DESCRIPTION OF THE DIHEDRAL CAPACITORS. THE TWO CAPACITOR


PLATES ARE INDICATED TOGETHER WITH LENGTH OF THE ARC, THAT CONNECTS THEM.

11

FIG. 7: SCHEMATIC DESCRIPTION OF THE KPFM SETUP PROPOSED BY JACOBS ET AL. THE

SAMPLE IS DIVIDED INTO A SYSTEM OF IDEAL CONDUCTORS WITH ELECTROSTATIC


INTERACTIONS REPRESENTED BY MUTUAL CAPACITANCES .

12

HYPERBOLA TO THE MEASURED TIP FOR EVERY ANGLE .

13

FIG. 8: CALCULATING THE EQUIPOTENTIAL LINES AND FIELD LINES BY FITTING A


FIG. 9: BOUNDARY CONDITIONS FOR THE PROBE-SAMPLE ELECTROSTATIC SYSTEM.

14

FIG. 10: TWO CONTRIBUTION TO THE PROBE VOLTAGE: (A)FROM THE HOMOGENOUS SYSTEM
AND (B)FROM THE CPD OF THE SAMPLE.

16

FIG. 11: TRANSFORMING THE BOUNDARY CONDITION ON THE SAMPLE SURFACE, VCPD , TO

DIPOLES IN FREE SPACE. (A) VCPD IS REPLACED WITH SURFACE CHARGE DENSITY s . (B)THE

[iv]

SURFACE CHARGE DENSITY IS DESCRIBED WITH VARYING DIPOLE DENSITY ABOVE A

GROUNDED PLANE. (C)THE GROUNDED PLANE IS REPLACED WITH IMAGE CHARGES TO


FORM AN ADDITIONAL DIPOLE LAYER. (D)THE ORIGINAL DIPOLE LAYER COINCIDES WITH
THE IMAGE DIPOLE LAYER, FORMING A DIPOLE LAYER WITH DOUBLED DENSIT.

17

FIG. 12: TIP-INFINITE FLAT SURFACE SYSTEM AND CHARACTERISTIC DIMENSIONS.

23

FIG. 13: COMPARISON BETWEEN THE FORCE AND FORCE-GRADIENT CALCULATED IN OUR
MODEL TO ANALYTICAL EXPRESSIONS, AS A FUNCTION OF THE TIP-SAMPLE DISTANCE. THE
SOLID LINES (RED) INDICATE ANALYTICAL VALUES AND THE MARKED LINES (GREEN)
INDICATE CALCULATED VALUES FROM OUR MODEL. (A)FORCE VS. TIP-SAMPLE DISTANCE.
(B)GRADIENT-FORCE VS. TIP-SAMPLE DISTANCE. THE INSETS IN (A) AND (B) DESCRIBE THE
PERCENTAGE OF THE ABSOLUTE ERROR FROM THE FORCE AND FORCE-GRADIENT,
RESPECTIVELY. (C)LOGLOG PLOT OF THE FORCE VS. TIP-SAMPLE DISTANCE. (D)LOGLOG
PLOT OF THE FORCE-GRADIENT VS. TIP-SAMPLE DISTANCE. FOR BOTH THE CALCULATION
AND SIMULATIONS WE CONSIDERED: R = 30nm, H = 3.88m AND 0 = 10.

24

FIG. 14: (A)CROSS-SECTION OF A TIP WITH A SPHERICAL APEX RADIUS , CONE LENGTH ,
HALF APERTURE ANGLE 0 , CONNECTED TO A CANTILEVER WITH LENGTH, WIDTH AND

THICKNESS OF L, W AND t, RESPECTIVELY. (B)CROSS SECTION OF A PROBE-SAMPLE SYSTEM

COMPOSED FROM A CANTILEVER TILTED IN DEGREES TOWARDS THE SAMPLE AND A TIP-

SAMPLE DISTANCE OF d. ALL THE ANALYSES WERE CARRIED OUT USING THE PARAMETER
VALUES: L = 225m , W = 40m, t = 7m, l = 13m, R = 30nm AND 0 = 17.5.

26

FIG. 15: DISTRIBUTION ALONG THE PROBE OF THE HOMOGENEOUS AND INHOMOGENEOUS
PARTS OF THE (A)FORCE AND (B)FORCE-GRADIENT. LEFT AXES - RELATIVE CONTRIBUTION

OF THE HOMOGENEOUS FORCE/FORCE-GRADIENT ON DIFFERENT SECTION OF THE PROBE.


RIGHT AXES - RELATIVE AREA OF EACH SEGMENT OF THE PROBE OUT OF THE TOTAL AREA
OF THE PROBE. THE INSETS IN (A) AND (B) OUTLINE THE DISTRIBUTIONS OF THE
INHOMOGENEOUS FORCE AND INHOMOGENEOUS FORCE-GRADIENT, RESPECTIVELY, ALONG
THE PROBE. THE PROBE IS DIVIDED INTO ELEVEN DISTINCT SEGMENTS DEFINED AS
FOLLOWS (FROM LEFT TO RIGHT): THE TIP APEX, THE TIP SPHERE, THE LOWER PART OF THE
CONE (VERTICAL LENGTH OF 5m), THE UPPER PART OF THE CONE (VERTICAL LENGTH OF

5m) AND SEVEN CANTILEVER SEGMENTS EACH WITH A LATERAL LENGTH OF 26.7m. THE
SIMULATIONS WERE CALCULATED FOR = 20 AND PROBE-SAMPLE DISTANCE OF

d = 5nm.

[v]

27

FIG. 16: TWO DIMENSIONAL PSFS OF (A)AM-PSF AND (B)FM-PSF. THE PSFS WERE
CALCULATED FOR = 20 AND TIP-SAMPLE DISTANCE OF d = 5nm.

28

FIG. 17: COMPARISON BETWEEN ONE DIMENSIONAL FM-PSFS(I)(RED) AND AM-PSFS(II)(BLUE).


(A)ONE DIMENSIONAL PSFS ALONG THE Y-AXIS. THE HORIZONTAL LINES REPRESENT THE
WIDTH AT HALF MAXIMUM, THE SOLID AND DASHED LINES CORRESPOND TO THE FM- AND
AM-PSFS, RESPECTIVELY. (B)LOG OF THE PSFS ALONG THE X-AXIS. (C)LOG OF THE PSFS
ALONG THE Y-AXIS. THE SIMULATIONS WERE PERFORMED WITH = 20 AND TIP-SAMPLE

DISTANCE OF d = 5nm.

30

FIG. 18: ONE DIMENSIONAL AM- AND FM-PSFS SIMULATED FOR TWO DIFFERENT TIP-SAMPLE
DISTANCES d WITH AND WITHOUT THE CANTILEVER, REPRESENTED BY THE SOLID (RED) AND
DASHED (BLUE) LINES, RESPECTIVELY. (A)AM-PSF FOR d=5nm. (B)AM-PSF FOR d=30nm.

(C)FM-PSF FOR d=5nm. (D)FM-PSF FOR d=30nm. THE HORIZONTAL LINES REPRESENT THE
FWHM OF THE PSFS - SOLID LINES FOR THE FULL PROBE (TIP+CANTILEVER) AND DASHED
LINES FOR THE TIP ONLY. THE SIMULATIONS WERE PERFORMED WITH THE PROBE
ILLUSTRATED IN FIG. 14 WITH = 20 FOR THE FULL PROBE.

32

FIG. 19: (A)FULL WIDTH AT HALF MAXIMUM OF BOTH AM-PSF (DASHED LINE) AND FM-PSF
(SOLID LINE) AS A FUNCTION OF THE TIP-SAMPLE DISTANCE. (B)PSF PEAK VALUE (MAXIMAL
PSF VALUE) AS A FUNCTION OF THE TIP SAMPLE DISTANCE FOR AM-PSF (DASHED LINE) AND
FM-PSF (SOLID LINE). ALL SIMULATION WERE PERFORMED WITH THE PROBE IN FIG. 14 FOR

= 15.

33

FIG. 20: HOMOGENOUS FORCE (LEFT AXIS) AND HOMOGENOUS FORCE-GRADIENT (RIGHT
AXIS) VS. THE TIP-SAMPLE DISTANCE. THE FORCE AND FORCE-GRADIENT ARE MARKED WITH
BLUE AND GREEN LINES, RESPECTIVELY.

34

FIG. 21: SUMMATION OVER THE AM-PSF (LEFT AXIS) AND FM-PSF (RIGHT AXIS) VS. THE TIPSAMPLE DISTANCE. ALL PSFS WERE SIMULATED FOR 192nm 192nm SAMPLE AREA.
FIG. 22: INPUT AND OUTPUT SIGNALS IN THE KPFM SYSTEM.

36
37

FIG. 23: (A)LOGLOG PLOT OF THE AM-PSF VS. X AND Y LINES. A LINEAR RELATION IS
OBSERVED FAR FROM THE ORIGIN. (B)EXPANDING THE AM-PSF TO AN INFINITE AREA. FIRST,
THE X AND Y LINES ARE EXTRAPOLATED (ILLUSTRATED BY BLUE ARROWS). NEXT,
EXTRAPOLATION IS PERFORMED ON THE DIAGONAL DIRECTION (ILLUSTRATED BY RED
ARROWS). IN THE FINAL STEP, THE REMAINING PIXELS ON THE PLANE (BLACK DOMAINS) ARE
INTERPOLATED USING 2 PIXELS WHICH WERE PREVIOUSLY EXTRAPOLATED. AN EXAMPLE

[vi]

FOR THIS INTERPOLATION IS MARK WITH A GREEN AREA.

40

FIG. 24: (A)CONTACT MODE TOPOGRAPHY OF THE SAMPLE: TWO Ni CONTACTS ON A LAYER
OF SiO2 . (B)HEIGHT PROFILE (AS INDICATED IN (A)). (C)CPD IMAGE OF THE SAMPLE. (D)CPD
PROFILE (AS INDICATED IN (C)). (E)CPD IMAGE OF THE SAMPLE WHILE THE LEFT Ni

CONTACT WAS BIASED WITH 0.5V AND THE RIGHT Ni CONTACT WAS BIASED WITH -0.5V,
RELATED TO THE SiO2 . (F)CPD PROFILE (AS INDICATED IN (E)). FROM THIS PROFILE ONE CAN
OBSERVE AN ALMOST LINEAR SLOPE IN THE SURFACE POTENTIAL. A SMALL DEVIATION

UPWARDS IS VISIBLE DUE TO THE HIGHER CPD VALUE ON THE SiO2 COMPARED TO THE Ni.
ALL KPFM MEASUREMENTS WERE CONDUCTED WITH AM USING LIFT-MODE WHERE THE LIFT
HEIGHT WAS 5nm.

43

FIG. 25: EVALUATION OF THE STATISTICS OF THE NOISE. (A)SECTIONS CONTAINING ONLY
CLEAN SiO2 WERE FRAMED FOR DISTINGUISHING THE NOISE FROM THE CPD SIGNAL. (B)THE
CPD OF THE SiO2 SUBSTRATE TAKEN FROM (A). (C)HISTOGRAM OUTLINES THE DISTRIBUTION
OF THE SUBSTRATE CPD (BARS) AND A NORMAL FIT (DASHED-LINE) WITH MEAN AND
VARIANCE OF 0.5132V AND 3.23 105 V 2 RESPECTIVELY. (D) AUTOCORRELATION OF THE
NOISE EXTRACTED FROM THE MARKED AREA IN (A).

44

FIG. 26: (A)BAND DIAGRAM OF THE UNBIASED SAMPLE. (B)BAND DIAGRAM OF THE SAMPLE
WHERE THE LEFT AND RIGHT Ni ELECTRODES ARE BIASED WITH -0.5V AND 0.5V,
RESPECTIVELY, RELATED TO THE TIP. Ef(Si),Ef(Ni),Ef(tip) REFERS TO THE FERMI LEVELS OF
THE SILICON, NICKEL AND TIP, RESPECTIVELY. LVL(tip) AND LVL(sample) ARE THE LVLS OF
THE TIP AND SAMPLE, RESPECTIVELY, AND = 50meV. THE CPD IS MARKED IN RED. UNDER

BIAS, THE CPD INCREASES FROM -0.035V (LEFT ELECTRODE) TO 0.965V (RIGHT
ELECTRODE).

46

FIG. 27: (A)2D IMAGE OF THE THEORETICAL CPD ON THE BIASED SAMPLE. (B)MEASURED CPD
(RED), THEORETICAL CPD (BLUE) AND CONVOLUTION OF THE THEORETICAL CPD WITH PSFS
GENERATED FOR TIP-SAMPLE DISTANCE OF 6nm (BROWN), 8nm (BLACK) AND 10nm(GREEN).
ALL CPD LINESCANS ARE PLOTTED ALONG THE WHITE LINE IN (A).

47

FIG. 28: MEASURED CPD (RED), THEORETICAL CPD (BLACK) AND DECONVOLVED CPD (BLUE).
ALL CPD PROFILES ARE RELATED TO THE LINESCAN INDICATED IN FIG. 24(E).

48

FIG. 29: SCHEMATIC OF KPFM EXPERIMENTAL APPARATUS. THE AFM TIP USED FOR THIS
SETUP HAS A TIP APEX OF 30nm. THE MEASURED NRS ARE APPROXIMATELY 80nm IN LENGTH
AND 4nm IN DIAMETER.

49

[vii]

FIG. 30: (A)REPRESENTATION OF THE PROBE IN THE Y-Z AXES. IS THE ANGLE BETWEEN THE
CANTILEVER AND THE HORIZONTAL AXIS. DISTANCE d IS THE AVERAGE TIP-SAMPLE

DISTANCE. (B)REPRESENTATION OF THE PROBE IN THE Y'-Z' AXES. A IS THE OSCILLATION


AMPLITUDE.

50

FIG. 31: (A)AM-KPFM MEASUREMENT OF CDS-PBS NRS ON HOPG SUBSTRATE. TWO NRS (ROD1
AND ROD2) ARE SELECTED FOR FURTHER ANALYSIS. (B)THE CPD OF THE HOPG SUBSTRATE
TAKEN FROM (A). (C)HISTOGRAM PRESENTS THE DISTRIBUTION OF THE SUBSTRATE CPD
(BARS) AND A NORMAL FIT (DASHED-LINE) WITH A MEAN AND VARIANCE OF 0.3862V AND

5.4 105 V 2 , RESPECTIVELY. (D)AUTOCORRELATION OF THE NOISE EXTRACTED FROM THE


MARKED AREA (BLUE) IN (A).

52

FIG. 32: (A)MEASURED CPD ON CDS-PBS NRS. (B)ACTUAL CPD ON CDS-PBS NRS OBTAINED BY
DECONVOLUTION WITH THE WIENER FILTER.

53

FIG. 33: LINESCANS OF THE MEASURED (DASHED LINE) AND DECONVOLVED (SOLID LINE)
CPD ACROSS THE LONGITUDINAL AXIS OF TWO SYMMETRIC PBS-CDS-PBS NRS (INSET KPFM
IMAGE) USING THE WIENER FILTER, EFFECTIVE PSF AND NOISE STATISTICS. (A)CPD ALONG
ROD1. (B)CPD ALONG ROD2. BOTH NRS ARE HIGHLIGHTED IN FIG. 31(A).

54

FIG. 34: OBTAINING SYSTEM NOISE STATISTICS FROM CPD MEASUREMENTS. (A) AND (B) SHOW
RAW CPD MEASUREMENTS OF SINGLE LAYERS AND DOUBLE LAYERS OF GRAPHENE. (C) AND
(D) PRESENT ONLY PURE AREAS OF SINGLE LAYERS OR DOUBLE LAYERS OF GRAPHENE
WITHOUT THE INTERFACIAL AREA BETWEEN THEM. THE SINGLE LAYERS ARE MARKED WITH
II,IV WHEREAS THE DOUBLE LAYERS ARE MARKED WITH I,III,V. (E) AND (F) SHOW THE CPD
DISTRIBUTION (BLUE BARS) ON THE AREAS PRESENTED IN (C) AND (D), RESPECTIVELY,
ALONG WITH FITTED PDFS OF BIMODAL GAUSSIAN FITS CORRESPONDING TO THEM (DASHED
LINE).

56

FIG. 35: (A) AND (B) SHOW LINESCANS OF THE MEASURED (BLUE LINE) AND DECONVOLVED
(RED LINE) CPD ALONG THE PROFILES ILLUSTRATED IN FIG. 34(A) AND FIG. 34(B),
RESPECTIVELY. THE DECONVOLVED CPD VALUES WERE CALCULATED USING THE WIENER
FILTER, EFFECTIVE PSF AND NOISE STATISTICS OBTAINED BEFOREHAND.

[viii]

57

List of symbols and abbreviations


0

Vacuum permittivity

Work function

Elementary charge

Voltage potential difference

Angular frequency

Frequency

Capacitance

Energy

Force

Amplitude

Mass

Spring constant

Potential

Standard deviation

Surface dipole density

Surface charge density

Expected value (mean value)

Tip apex radius

Tip cone length

Cantilever length

Cantilever width

[ix]

Tip-sample distance

Cone half aperture angle

Noise

V CPD

Actual potential distribution on the sample

Vsub

Substrate voltage

Probe self-interaction matrix (Greens matrix)

Ch

Probe-surface homogenous capacitance

Cantilever tilt angle

2
VDC

Variance
Measured potential, output of Kelvin controller

Vprobe

Applied bias voltage on the probe

VAC

Alternating voltage

Dipole-probe interaction matrix

Cinh

Probe-surface inhomogeneous capacitance

DC

Direct current

AC

Alternate current

KPFM

Kelvin probe force microscopy

AFM

Atomic force microscopy

EFM

Electrostatic force microscopy

AM-KPFM

Amplitude modulation KPFM

FM-KPFM

Frequency modulation KPFM

CPD

Contact potential difference

LVL

Local vacuum level

[x]

PLL

Phase lock loop

SNR

Signal-to-noise ratio

LIA

Lock-in amplifier

AM-PSF

Amplitude modulation PSF

FM-PSF

Frequency modulation PSF

BEM

Boundary element method

FWHM

Full width at half maximum

LPF

Low pass filter

AWGN

Additive white Gaussian noise

NR

Nanorod

HOPG

Highly-oriented pyrolytic Graphite

UHV

Ultra high vacuum

PDF

Probability density function

[xi]

Chapter 1 - Introduction

1 Introduction
Kelvin Probe Force Microscopy (KPFM) has already been demonstrated as a powerful
tool for measuring electrostatic potential distribution with nanometer spatial resolution.
However, due to the long range electrostatic forces the measured potential is a weighted
average of the real surface potential distribution. This effect is demonstrated in Fig. 1,
showing to scale an atomic force microscopy (AFM) tip above a p-n junction. The figure
clearly demonstrates that the size of the tip apex alone (right) is huge compared to the
typical junction dimensions, which emphasizes further the significance of the averaging
effect in KPFM.

Fig.1: Schematics of a typical KPFM tip above a p-n junction showing that the tip
dimensions are much bigger than the junction region underneath it1.

The actual surface potential can be recovered from the measured surface potential by
using deconvolution. The main goal of this thesis is to develop a reconstruction algorithm
for KPFM measurements.

Chapter 1 introduces the KPFM method and compare between two different scanning
modes: amplitude modulation KPFM (AM-KPFM) and frequency modulation KPFM
(FM-KPFM). In Chapter 2 we review some of the previous works, where a special
attention is paid to those who correlated the measured potential with the real potential on
[1]

Chapter 1 - Introduction

the sample. Chapter 3 presents the computational model that was used which is based on
the work of Elias et al.2,3 and Strassburg et al.4,5. In addition, this chapter improves the
model by analyzing the influence of the averaging effect on FM-KPFM and validates it
with comparison to the work of Hudlet et al.6 who obtained an expression for the
electrostatic force acting on a probe above a homogeneous plane. In Chapter 4 and 5 we
examine the dependence of the probe geometry (especially the cantilever) and tip-sample
distance on the averaging effect in AM- and FM-KPFM. Chapter 6 presents a
reconstruction algorithm for the surface potential from a KPFM signal and displays some
results. The thesis is summarized in Chapter 7.

1.1 Kelvin Probe Force Microscopy


Kelvin Probe Force Microscopy (KPFM, sometimes KFM or KPM) is a further
development of Electrostatic Force Microscopy (EFM)7, which measures the contact
potential difference (defined below) between tip and sample. It was first developed by
Nonnenmacher et al.8 in 1991. If the work function of the tip is known, the surface
potential of the material under study can be observed.

The work function of a material in vacuum is defined as the minimum energy required
for emitting an electron in the Fermi level to the vacuum level outside the material. The
Contact Potential Difference (CPD) is the difference between the work function of two
materials, defined as:
VCPD =

1 2
q

(1.1)

Where 1 and 2 are the work function of the first and second materials, respectively,
and q is the elementary charge.

The term Kelvin Probe Force Microscopy relates to macroscopic Kelvin probe
techniques, which were invented by William Thomson, known as Lord Kelvin, in 1898
for the measurement of surface potentials using a vibrating parallel plate capacitor
arrangement: The sample constitutes one plate of a parallel plate capacitor, with a known
[2]

Chapter 1 - Introduction

metal forming the other plate, which is vibrated at frequency . A DC-voltage applied to
one of the plates is used to minimize the induced current by the vibration9. This voltage
corresponds to the CPD of the two materials.
The KPFM employs the same principle, applying a DC-voltage in order to compensate
the CPD between the Atomic Force Microscope (AFM) probe and the sample10.
However, in a KPFM setup the electric force is used as the controlling parameter instead
of the current.
Fig. 2 illustrates the working principle of Kelvin probe force microscopy. When the
probe and sample are not electrically connected their local vacuum levels (LVL) are
aligned, as shown in Fig. 2(a). When they are electrically wired together, current will
flow from the material with the higher work function to the one with the lower work
function until the Fermi levels are aligned, inducing opposite charges between the probe
and the sample which form a binding electrostatic force, as shown in Fig. 2(b). This force
is nullified by a Kelvin feedback that applies bias between the probe and the

sample11. The magnitude of this bias is the CPD between the probe and the sample, as
shown in Fig. 2(c).

Fig. 2: Definition and basic measurement setup of contact potential difference (CPD).
- work function of the tip, - work function of the sample, - Fermi level, LVL -

local vacuum energy level.

[3]

Chapter 1 - Introduction

KPFM measurement is usually conducted with Amplitude Modulation (AM-KPFM). In


AM-KPFM the cantilever is excited electrically by inducing a bias potential of
Vprobe = VDC + VAC sin(t), where VDC is the output of the Kelvin controller and is the
frequency of the Kelvin modulation. Therefore, the potential difference between the
probe and the sample is:
V = VCPD Vprobe = VCPD VDC VAC sin(t)

(1.2)

The electrostatic force can be derived from the energy of a capacitor:


U=

1 2
CV
2

(1.3)

Where C is the local capacitance between the probe and the sample. The vertical force is
then the derivative of the energy with respect to the probe-sample separation z:

Fz =

dU
1 dC 2
=
V
dz
2 dz

(1.4)

By substituting Eq. (1.2) in Eq. (1.4) we get that the electrostatic force has spectral
components at DC and at frequencies and 2:

Fz = FDC + F + F2

FDC =
F =

1 dC
1 2
(VCPD VDC )2 + VAC

2 dz
2
dC
(V
VDC )VAC sin(t)
dz CPD

F2 =

1 dC 2
V cos(2t)
4 dz AC

(1.5a)

(1.5b)

(1.5c)

(1.5d)

The Kelvin controller nullifies the force component at the excitation frequency ( ) by
[4]

Chapter 1 - Introduction

adjusting = . If the work function of the tip is known, we can infer the work

function on the sample.

The KPFM measurement can be conducted simultaneously with the topography with
single-pass technique12, or alternatively by using lift mode technique13.
In single-pass the cantilever is exited in 2 eigenmodes simultaneously. The first
eigenmode is used for distance control and KPFM is performed at the second flexural
eigenmode. To this end, the cantilever is mechanically excited by a dither-piezo in the
first resonance and electrically excited in the second resonance.
In lift-mode technique the measurement of topography and surface potential are
alternated. Each scan line of topography is first recorded in tapping mode; the measured
trajectory is retracted in order to perform AM-KPFM at a constant lift height from the
sample.

1.2 Frequency Modulation KPFM


Frequency Modulation KPFM (FM-KPFM) was first proposed by Kitamura et al.14 in
1998. In contrast to AM-KPFM where the force is detected, in FM-KPFM the control
parameter is the force-gradient.
When applying force F with gradient

dF
dz

on a cantilever, the fundamental mechanical

resonance frequency of the free cantilever f0 =

changes as follows:

dF
1 k dz
f =
2
m

(1.6)

Where k is the spring constant of the cantilever and is the effective mass of the

cantilever. For low force-gradient the frequency shift can be approximated by15:

f =

f0 dF
2 dz

For the spectral components of the frequency shift we find:


[5]

(1.7)

Chapter 1 - Introduction

fDC

dFDC
1 d2 C
1 2
2
(
)

V
+
V
CPD
DC
dz
2 dz 2
2 AC

dF
d2 C
f
= 2 (VCPD VDC )VAC sin(2fmod t)
dz
dz
f2

dF2 1 d2 C 2

=
V cos(2 2fmod t)
dz
4 dz 2 AC

(1.8a)

(1.8b)

(1.8c)

Where fmod is the modulation frequency.

The Kelvin controller nullifies the force-gradient component at the modulation frequency
(f ) by adjusting VDC = VCPD , as shown in Fig. 316.

Fig. 3: Principle of FM-KPFM, the frequency shift vs. tip sample bias. The Kelvin
controller minimizes the amplitude of the frequency shift, which is smallest when
VDC = VCPD .

[6]

Chapter 1 - Introduction

In contrast to AM-KPFM, in FM-KPFM the cantilever is mechanically excited in its first


resonance and electrically excited in the modulation frequency fmod , simultaneously. The
tip oscillation spectrum in FM-KPFM is shown in Fig. 417.

Fig. 4: Schematic frequency spectrum of the tip oscillation. The peaks at fmod and 2fmod
originate from the electrostatic force, whereas the peaks at f0 fmod and f0 2fmod
show the frequency modulation at f0 .

The resonance frequency is measured using a frequency demodulator18, or a phase lock


loop (PLL)13. Then, the Kelvin controller nullifies the sidebands at 0 .

[7]

Chapter 1 - Introduction

1.3 Comparison between Amplitude and Frequency Modulations


As was shown in the previous two sections, the AM-KPFM detects the electrostatic force,
whereas the FM-KPFM detects the electrostatic force-gradient. Therefore, the properties
of the two measurements are different.
The decay of the electrostatic force is inversely proportional to the square of the distance
1

2 and the decay of the electrostatic force-gradient is steeper and inversely proportional
r

to third power of the distance 3 . Consequently, the detection of electrostatic force is


r

considered long-ranged detection, whereas the detection of the electrostatic forcegradient is short-ranged detection. Thus, in AM-KPFM the probe averages over a larger
area on the sample, which leads to more prominent averaging effect in the
measurement13,19-21. On the other hand, in FM-KPFM the short-ranged electrostatic
interactions exist mainly between the tip apex and the sample, therefore averaging takes
place over a much smaller area below the tip. Hence, FM-KPFM has better spatial
resolution than AM-KPFM22.
However, AM-KPFM is considered to have better energy resolution of the measured
CPD than FM-KPFM. The signal-to-noise ratio (SNR) in AM-KPFM is very large
because the lock-in amplifier (LIA) detects the high resonance peak of the oscillating
cantilever23, whereas in FM-KPFM the LIA detects the relatively low sidebands
f0 fmod , as portrayed in Fig. 4. In addition, more noise is generated at the output of the
frequency demodulator (or PLL) which further degrades the SNR in FM-KPFM.

The FM-KPFM energy resolution can be improved by amplifying the modulation


amplitude, VAC . However, high modulation voltage might generate tip induced sample
bend bending in semiconducting surfaces24 or quadratically increase the electrostatic

contribution to the topography signal23, as described in Eq. (1.8a), when scanning in


single-pass technique.

[8]

Chapter 2 Literature Review

2 Literature Review
It is well known that the probe geometry and scan method in KPFM can have a profound
effect on the measured CPD image. Reconstruction of the actual CPD on the sample from
the measured one is therefore of great importance and was carried out by several authors.
In the following we present different models which aim to this target.
Hochwitz et al.25 suggested a simple model where the tip was replaced by a series
(staircase) parallel plate capacitors. The model is presented schematically in Fig. 5 which
shows defragmentation of the tip into small capacitors connected in parallel.

Fig. 5: A model in which the AFM tip is represented by a series of parallel plate
capacitors Ci at distances Zi from the sample surface. One plate of a capacitor is located

on the tip and the other on the sample surface25.

[9]

Chapter 2 Literature Review

After calculating the force acting on each capacitor, the total electrostatic force on the
probe is then nullified by applying a DC voltage of:
VDC =
Where

Ci
z

Ci
z VCPDi
C
i i
z

(2.1)

is the capacitance gradient of capacitor i with respect to its distance from the

surface, Zi , and VCPDi is the CPD between the surface and capacitor i. Several groups26

used this model and showed reasonably good agreement to experimental data. However,
it uses two assumptions which are invalid in KPFM setup, as explained below.

The area below the tip is underestimated because the cones aperture angle is

small (usually between 10 15).

The expression for the parallel plate capacitor (Ci = ) is valid only when the
Zi

capacitor plates area is much larger than the distance between them and the
sample.

A more precise model that takes into consideration the full area of the probe, and
therefore, a larger interacting sample was presented by Hudlet et al6. The tip was
modeled as a cone with a semi-spherical apex and dihedral capacitors replaced the
parallel plate capacitors (in the previous model), as shown in Fig. 6. The electric field on
each infinitesimal tip surface was assumed to be created by the dihedral capacitance
constituted by two infinite planes in the same relative orientation. This approximated
field determines the infinitesimal force, dFz , and the total tip-surface force is then
obtained by summing all these contributions. They formulated analytical expressions for

the electrostatic force of a probe above a homogeneous surface and reported very good
agreement with experimental results.

[10]

Chapter 2 Literature Review

Fig. 6: Schematic description of the dihedral capacitors. The two capacitor plates are
indicated together with length of the arc, that connects them6.

This model was validated by Law et al27 against a large set of experimental data. They
also included the contribution of a tilted rectangular cantilever to the electrostatic force.

Jacobs et al.28 introduced a model which correlates the measured potential with the actual
sample potential distribution. They treat the sample as a surface consisting of ideally

conducting electrodes of constant potential i and a tip of potential t , as shown in

Fig. 7.

[11]

Chapter 2 Literature Review

Fig. 7: Schematic description of the KPFM setup proposed by Jacobs et al28. The sample
is divided into a system of ideal conductors with electrostatic interactions represented by
mutual capacitances .

The vertical force (Fz ) on the probe is calculated by differentiating the expression for the

electrostatic field energy with respect to the tip-sample distance. By setting Fz = 0 they

obtained the KPFM signal:

VDC =

ni=1 Cit i
ni=1 Cit

(2.2)

Where Cit = Cit / z are the derivatives of the mutual capacitances between element

on the sample and the tip. Eq. (2.2) demonstrates that the measured potential is a
weighted average of all potentials i on the surface with Cit being the weighting factors.

Finally, they showed that by dividing an infinitely large surface to infinitesimal small
areas (x y) the KPFM signal can be expressed as:

VDC (xt , yt ) = h(x xt , y yt )(x, y)dxdy


h(x xt , y yt ) = lim
x,y0

C(x xt , y yt )

Ctot
xy

(2.3)
(2.4)

Where (xt , yt ) is the tip location and Ctot


is the derivative of the total tip-surface

capacitance. Eq. (2.3) shows that the measured potential is a two-dimensional


convolution of the actual surface potential (x, y) with the system point spread function

(PSF), h(x, y).

[12]

Chapter 2 Literature Review

Machleidt et al.29 considered the real AFM tip shape in the determination of the PSF.
They used a prolate spheroidal coordinate system to calculate the electric field between
the charged tip and its mirror charge. The tip apex shape was estimated by measuring it
using a calibration sample, such as (NT-MDT) TGT1. The PSF was then derived by
fitting a hyperbola for every angle and calculating the equipotential lines () and the

field lines (), as shown in Fig. 8.

Fig. 8: Calculating the equipotential lines and field lines by fitting a hyperbola to the

measured tip for every angle .

Two assumptions are used by most of the above authors: (a) The bulk material under the
surface does not influence the measurements and (b) the actual CPD on the sample does
not change with the presence or the metallic probe. Shikler et al30,31 validated these
assumptions by simulating a three-dimensional model consisting of both the substrate and
the area above it. These assumptions facilitate the KPFM modeling since we can
disregard the substrate entirely and simulate only a two-dimensional surface.

[13]

Chapter 3 Electrostatic Model

3 Electrostatic Model
In this chapter we use a previously developed algorithm2-5 in order to find a relation
between the real CPD of the sample, VCPD , and the measured CPD of the probe, VDC . The
model calculates the charge distribution on the probe, which yields the electric force and

force-gradient between the probe and the sample. Minimization of the force and the
force-gradient derives the amplitude modulation PSF (AM-PSF) and frequency
modulation PSF (FM-PSF), respectively.

3.1 Probe-Sample Electrostatic System


The model assumes an equipotential probe above a flat sample (Fig. 9). The three inherent
boundary conditions are: (1) a constant probe voltage, Vprobe , (2) zero potential far away

from the probe and (3) the surface potential of the measured sample, VCPD .

Fig. 9: Boundary conditions for the probe-sample electrostatic system.


[14]

Chapter 3 Electrostatic Model

In the absence of space charge, the potential in the electrostatic system follows the
Laplace equation:
2 V() = 0

(3.1)

Where V() is the potential in the point . The potential of interest is the potential of the

probe. The interior potential in a domain bounded by a surface S is given by the integral
equation:
V () =

Sprobe

Where G(, ) is Greens function:

G(, )( )ds + Vs ()

G(, ) =

1
40 | |

(3.2)

(3.3)

() is the surface charge density on the probe, and Vs () represents the potential of

known charges observed in point . Physically, Greens function G(, ) express the
contribution of a unit charge in point to the potential in point .
We define:

h
() =
Vprobe

Sprobe

G(, )( )ds

inh
Vprobe
() = Vs ()

(3.4)
(3.5)

h
() is generated from a homogenous system, consisting the charged probe above a
Vprobe

inh
grounded surface. Vprobe
() is generated exclusively from the CPD of the sample.

Therefore:

h
inh
() + Vprobe
Vprobe () = Vprobe
()

(3.6)

For the homogeneous system, the grounded surface is replaced with equivalent image
charges. In this manner we maintain the boundary condition at z = 0, V(x, y, z = 0) = 0
(Fig. 10(a)).

[15]

Chapter 3 Electrostatic Model

Fig. 10: Two contribution to the probe voltage: (a)From the homogenous system and
(b)from the CPD of the sample.

Both the probe surface charges and the image charges contribute to the homogenous
potential:

h
() =
Vprobe

Sprobe

[G(, ) G(, )]( )ds

(3.7)

Where S G(, )( )ds and S G(, )( )ds are the contribution to the
homogeneous voltage from the probe surface charge and the image charge, respectively,
and where: = (x , y , z ).

For evaluating the inhomogeneous potential we model the sample surface with dipoles.
This transforms the boundary condition calculation to free space. The model replaces the
sample surface potential, VCPD (Fig. 10(b)), with a layer of non-uniformed surface charge

density, s , above a grounded plane in an infinitesimal distance l (Fig. 11(a)). The


[16]

Chapter 3 Electrostatic Model

dipole density is therefore: () = s ()l = VCPD 0 (Fig. 11(b)). To satisfy the


boundary condition for z = 0 the ground plane is replaced with image charges,

generating an additional dipole layer (Fig. 11(c)). Both dipole layers coincide to a dipole
layer with doubled density () = 2 () = 2VCPD 0 (Fig. 11(d)).

Fig. 11: Transforming the boundary condition on the sample surface, VCPD , to dipoles in

free space. (a) VCPD is replaced with surface charge density s . (b)The surface charge

density is described with varying dipole density above a grounded plane. (c)The
grounded plane is replaced with image charges to form an additional dipole layer. (d)The
original dipole layer coincides with the image dipole layer, forming a dipole layer with
doubled density.

The inhomogeneous potential in point on the probe is calculated by integrating all the
dipole contribution from the sample surface, Ssample :
inh ( )
Vprobe
=

( )
( )
1

( )ds
| |3
40 Ssample

( )
( )
1

=
VCPD ( )ds
| |3
2 Ssample

(3.8)

( ) is a unit vector normal to the sample surface. In our system


( ) = z .
Where

In order to solve Eq. (3.7) and Eq. (3.8) numerically the equations are discretized into a
[17]

Chapter 3 Electrostatic Model

linear system of equations using the boundary element method (BEM)2,4. Thus, the total
h
inh ( )
() + Vprobe
voltage on the probe, Vprobe
, can be represented in matrix form :

Vprobe I = G + D VCPD

(3.9)

I is a vector in the length of the number of the probes surface elements whose elements

are all ones, and VCPD is a vector of discrete samples of the surface potential, VCPD , in the
length of the total number of the image pixels. The matrices and vectors sizes and values
are all defined in the Appendix.
From Eq. (3.9) we obtain the surface charge density on the probe:
= G1 Vprobe I D VCPD = Ch Vprobe Cinh VCPD

(3.10)

Where the vector Ch = G1 I describes the capacitance per unit area on the probes

surface elements and the matrix Cinh = G1 D describes the capacitance per unit area
between every probe surface element and sample surface element pair.

3.2 Minimum Force/Force-gradient Condition


The CPD measurement in AM-KPFM is derived by minimizing the harmonic vertical
force applied on the probe, whereas the CPD measurement in FM-KPFM is derived by
minimizing the harmonic vertical force-gradient on the probe17. Using these two
conditions, a relation between VDC and VCPD is extracted for each of the scan methods

under the assumption that the tip-sample distance is maintained constant during the
measurement.
The vertical force on the probe is calculated using the expression2:
Fz = t B

Note that underbars denote vectors and double underbarsmatrices.


[18]

(3.11)

Chapter 3 Electrostatic Model

Where B is related to the geometry of the probe, as defined in the Appendix.

The potential on the probe, positioned above a point on the sample, contains both DC
and AC components:

Vprobe = VDC () + VAC sin(t)

(3.12)

Applying Eq. (3.10) and Eq. (3.12) into (3.11) generates the force harmonic components:
Fz = Fz,DC + Fz, sin(t) + Fz,2 cos(2t)

(3.13)

In AM-KPFM the Kelvin controller minimizes , . The minimum force condition

therefore compels , = 0, extracting the following relation2:


VDC () =

Cht B Cinh
Cht B Ch

VCPD ()

(3.14)

The derivation of the latter expression is followed in the Appendix.


In FM-KPFM the Kelvin controller minimizes the component of the force-gradient,
dFz

dz

. The force-gradient at frequency of the entire probe is:

dFz
= 2VAC 2Cht B C h VDC C ht B Cinh VCPD Cht B C inh VCPD
dz

(3.15)

Where C h and C inh are the capacitances gradients in the z directions (explicitly defined in
the Appendix). The minimum force-gradient condition sets
relation:

t
Ch B Cinh
VDC () =
VCPD ()

t
C
B
C

h
h
We define: = Cht B Ch ; = Cht B Cinh .

dFz

dz

= 0, extracting the

(3.16)

From Eq. (3.14) and Eq. (3.16) we obtain the AM-PSF and FM-PSF of the probe-sample
[19]

Chapter 3 Electrostatic Model

system:

PSF(AM) =
; PSF(FM) =

( )

(3.17)

3.3 Point Spread Function Integration


When scanning an equipotential substrate, the KPFM measurement outputs the exact
surface potential. Therefore, integration of the PSFs over the entire x-y surface must be
equal to one. Thus, the discrete PSFs presented in Eq. (3.17) must be normalized to one.
Elias et al show that AM-PSF is indeed normalized for an infinite surface (However,
numerical limitations yield lower value for the sum of the AM-PSF, as will be discussed
later in Chapter 4). We provide here an appropriate confirmation for the FM-PSF.
Let N and M be the number of probe elements and the number of sample surface
elements, respectively. Then, the mathematical representation of the requirement above
is:
d
H t
dz inh I
MX1 = 1
d
(Hh )
dz

(3.18)

Where 1 is a vector in the length of the sample surface elements whose elements are
all ones.

By substituting and into Eq. (3.18) we left to show that:


d
d
Cht B Cinh I MX1 = Cht B Ch
dz
dz

(3.19)

By writing Cinh = G1 D in the left side and Ch = G1 I 1 in the right side, Eq. (3.19)
becomes:

t
1
1
C
B
G
D

=
Cht
B G
I 1
1
h

[20]

(3.20)

Chapter 3 Electrostatic Model

Meaning,
1 = 1
D 1 + D

(3.21)

are the gradients of and D in the z


Where is defined in Eq. (3.20), and D
t =
direction, respectively, defined as Q
Appendix).

dQt
dz

=
and D

dD
dz

(explicitly defined in the

The sum of every row in D is the summation of the potential induced by a homogenous
dipole layer on a specific probe element. This term is given by2:

Di = Dij =
j

2 SSample ((

zi dxdy

xi x )2 + (yi y )2 +

3
zi2 )2

=1

R2
2 +1
zi

(3.22)

Where (xi , yi , zi ) are the coordinates of the probe element, and is an integrated

area of a disc in a radius of R. Thus, by integrating an infinite area we get for each ith
probe element:

Thus,

lim D = 1

(3.23)

D 1 = 1

(3.24)

rows we get:
As for the summation of D

i = D
ij =
D
j

R2

(R2 + zi2 )2

(3.25)

The derivation of the latter expression is followed in the Appendix.


For each probe element i we then obtain:

Thus,

= 0
lim D

[21]

(3.26)

Chapter 3 Electrostatic Model

1 = 1
D

(3.27)

Where 1 is a vector in the length of the probe elements whose elements are all zeros.

Eq. (3.24) and Eq. (3.27) satisfy the condition in Eq. (3.21), hence the FM-PSF in Eq.
(3.17) is normalized.

3.4 Model Validation


In order to validate our model and to test its accuracy we compare it to the analytical
model proposed by Hudlet et al.6. Their electrostatic system is constructed of a grounded
sample and a probe composed of tip alone, without the cantilever. They estimate that the
force between the probe and the sample is:
F = 0 V 2

R2 (1 sin0 )
z + R(1 sin0 )
Rcos2 0 /sin0

+ k 2 ln
1+
z[z + R(1 sin0 )]
H
z + R(1 sin0 )

Where k 2 =

lntan(0 /2)

(3.28)

Where V is the voltage between the tip and the sample, R is the tip apex radius, z is the

tip-sample distance, H is the length of the tip cone and 0 is the half-aperture angle of the
cone, as described in Fig. 12.

[22]

Chapter 3 Electrostatic Model

Fig. 12: Tip-infinite flat surface system and characteristic dimensions6.

In order to evaluate the force-gradient between the tip and the sample the derivative of
the expression in Eq. (3.28) was calculated:
dF
R2 (1 sin0 )2z + R(1 sin0 )
1
Rcos2 0 /sin0
2

= 0 V 2
+
k
+
dz
z 2 [z + R(1 sin0 )]2
z + R(1 sin0 ) [z + R(1 sin0 )]2

(3.29)

Next, we simulated the above system for different tip-sample distances ranging from 1nm
to 50nm, and for each distance we calculated the force and the force gradient by using
Eq. (A.6) and Eq. (3.15). Fig. 13 compares the force and force-gradient in our model to
the analytical expressions in Eq. (3.28) and Eq. (3.29).

[23]

Chapter 3 Electrostatic Model

Fig. 13: Comparison between the force and force-gradient calculated in our model to
analytical expressions, as a function of the tip-sample distance. The solid lines (red)
indicate analytical values and the marked lines (green) indicate calculated values from
our model. (a)Force vs. tip-sample distance. (b)Gradient-force vs. tip-sample distance.
The insets in (a) and (b) describe the percentage of the absolute error from the force and
force-gradient, respectively. (c)Loglog plot of the force vs. tip-sample distance.
(d)Loglog plot of the force-gradient vs. tip-sample distance. For both the calculation and
simulations we considered: R = 30nm, H = 3.88m and 0 = 10.

All the calculations take into consideration only the component of the force and force-

gradient. Therefore, we substitute 2(VCPD VDC )VAC instead of V 2 in Eq. (3.28) and

Eq. (3.29) and assume: VDC = 0.5V ; VAC = 1V ; VCPD = 1V.

Overall, we observe very good agreement between the two models. The maximal error
for the force is 15% and the error for the force-gradient never exceeds 5% for tip-sample

distances over 3nm. The relatively high error for the force-gradient in 1-2[nm] might
arise from coarse meshing of the tip or due to insufficient resolution.

[24]

Chapter 4 The Effect of Probe Geometry on PSF

4 The Effect of Probe Geometry on PSF


In this chapter we inspect how the distributions of the force and force-gradient on the
probe affect the AM-PSF and FM-PSF. In addition, we show the influence of the
cantilever on the PSFs.

4.1 Force/Force-gradient Distributions


h
We separate the expression for the force in Eq. (A.6) to a homogenous part, Fz,
, which
inh
in not depended on VCPD , and an inhomogeneous part, Fz,
, which is depended on VCPD :
h
Fz,
= 2VAC VDC Cht B Ch

inh
Fz,
=

2VAC Cht

(4.1)

B Cinh VCPD

In the same analogy, the homogenous part and inhomogeneous parts of the force-gradient
are:
dFz h
d
= 2VAC VDC Cht B Ch
dz
dz
inh

dFz
dz

= 2VAC

d
Cht B Cinh VCPD
dz

(4.2)

The probe under study is composed from a conical tip enclosed with a spherical cap and a
tilted cantilever. The geometric parameters of the probe are presented in Fig. 14:

[25]

Chapter 4 The Effect of Probe Geometry on PSF

Fig. 14: (a)Cross-section of a tip with a spherical apex radius , cone length , half

aperture angle 0 , connected to a cantilever with length, width and thickness of L, W and

t, respectively. (b)Cross section of a probe-sample system composed from a cantilever

tilted in degrees towards the sample and a tip-sample distance of d. All the analyses

were carried out using the parameter values: L = 225m, W = 40m, t = 7m,
l = 13m, R = 30nm and 0 = 17.5.

Fig. 15(a) and Fig. 15(b) show the distribution of the force and force-gradient,
respectively, along the probe presented in Fig. 14 with a tip-sample distance of
d = 5[nm] and cantilever tilt angle of = 20. The probe was divided into eleven

segments where each bar corresponds to a distinct segment of the probe specified as
follows (from left to right): The tip apex, the tip sphere, the lower and upper parts of the

cone (each has vertical length of 5nm) and seven cantilever segments with an equal
length. The left axes in Fig. 15 outline the contribution of each probe section to the total
homogenous force/force-gradient and the right axes describe the relative mesh area of
each segment out of the total surface area on the probe. The insets in Fig. 15(a) and
Fig. 15(b) present the inhomogeneous force and inhomogeneous force-gradient
distributions, respectively, along the probe using the same segments. The force and forcegradients were calculated for a 192nm by 192nm square sample, with equipotential
surface forming VCPD = 1V between the probe and the sample, Kelvin voltage of
VDC = 1V and modulation voltage of VAC = 1V.
[26]

Chapter 4 The Effect of Probe Geometry on PSF

Fig. 15: Distribution along the probe of the homogeneous and inhomogeneous parts of
the (a)force and (b)force-gradient. Left axes - relative contribution of the homogeneous
force/force-gradient on different section of the probe. Right axes - relative area of each
segment of the probe out of the total area of the probe. The insets in (a) and (b) outline
the distributions of the inhomogeneous force and inhomogeneous force-gradient,
respectively, along the probe. The probe is divided into eleven distinct segments defined
as follows (from left to right): the tip apex, the tip sphere, the lower part of the cone
(vertical length of 5m), the upper part of the cone (vertical length of 5m) and seven

cantilever segments each with a lateral length of 26.7m. The simulations were
calculated for = 20 and probe-sample distance of d = 5nm.
[27]

Chapter 4 The Effect of Probe Geometry on PSF

One can observe from Fig. 15(a) that the cantilever contributes 22% of the homogeneous
force and therefore affects greatly the absolute measured CPD. In addition, almost all the
inhomogeneous force (99%) stems from the tip sphere and the cone. Thus, the effect of
the cantilever of the resolution in AM-KPFM can be neglected2. On the other hand, it can
be observed from Fig. 15(b) that the vast majority of the force-gradient (homogenous and
inhomogeneous) resides within the tip sphere (98%), the rest of the force-gradient stems
from the bottom cone (2%), and we observe zero contribution from the upper cone and
the cantilever. Therefore, it indicates that in FM-KPFM the cantilever has no influence on
the measured CPD. Since most of the homogeneous and inhomogeneous force-gradients
stem from the tip apex (81%), whose vertical length is 5nm, very high resolution is
expected in FM-KPFM with measured CPD close to the real CPD. The significant role of
the tip apex in FM-KPFM measurements is consistent with the short-ranged behavior of
the force-gradient, which decays faster than the force.

4.2 Properties of AM-PSF And FM-PSF


The differences between the minimum force/force-gradient conditions presented in the
previous chapter and the force and force-gradient distributions yield two distinct PSFs
with different characteristics. In this subsection we distinguish between the properties of
the AM- and FM-PSF.
The two dimensional AM-PSF and FM-PSF for the probe presented in Fig. 14 are shown
in Fig. 16.

Fig. 16: Two dimensional PSFs of (a)AM-PSF and (b)FM-PSF. The PSFs were
calculated for = 20 and tip-sample distance of d = 5nm.
[28]

Chapter 4 The Effect of Probe Geometry on PSF

The sums of the AM-PSF and the FM-PSF components are 0.5856 and 0.9999,
respectively. The FM-PSF converges to a total sum of 1, as we prove in the previous
chapter. However, integration over the AM-PSF yields a smaller value since the
simulations cannot be performed on an infinite sample due to computational limitations.
Instead, we considered a finite square sample (192nm by 192nm) which is not sufficient
to satisfy Eq. (3.23). It should be noted that the sum of the AM-PSF components
approaches 1 as the tip-sample distance is shortened, since most of the inhomogeneous
force stems from the tip-sphere and by lowering the tip-sample distance the tip-sphere
sees more of the sample and less from the truncated area. In Chapter 6 we present a
method to extrapolate the AM-PSF into an infinite size.
Next, we inspect the influence of the force and force-gradient distributions on the KPFM
measurements. Fig. 17 presents a comparison between the one dimensional FM-PSFs ()

and AM-PSFs (). The FM-PSF peak is greater than the AM-PSF by a factor of 3.5
(Fig. 17(a)) which indicates higher CPD contrast in FM-KPFM measurements than in
AM-KPFM. Moreover, the full width at half maximum (FWHM) of the FM-PSF is
smaller than the AM-PSF by a factor of 1.25, indicating better CPD resolution when

conducting FM-KPFM. The probe is symmetric along the x-axis and therefore both the
AM-PSF and FM-PSF are symmetric along the x-axis (Fig. 17(b)). In the y-axis the
symmetry does not exist, both the AM-PSF and FM-PSF exhibit higher components at
negative y values (Fig. 17(c)). The asymmetry of the PSFs is due to the inhomogeneous
force and force-gradient which stem from the tip-sphere, as shown in Fig. 15. Since the
cantilever is tilted and clamped at positive y, the tip-sphere and the tip cone are mostly
located in negative y values, thus contributing more to the PSF components at these
values. Log scale was used in Fig. 17(b),(c) to highlight the change of the PSFs in the
positive and negative x,y axes, which is unnoticeable from Fig. 17(a) . The log of the
FM-PSF is truncated in order to prevent divergence where the FM-PSF reaches zero.

[29]

Chapter 4 The Effect of Probe Geometry on PSF

Fig. 17: Comparison between one dimensional FM-PSFs(i)(red) and AM-PSFs(ii)(blue).


(a)One dimensional PSFs along the y-axis. The horizontal lines represent the width at
half maximum, the solid and dashed lines correspond to the FM- and AM-PSFs,
respectively. (b)Log of the PSFs along the x-axis. (c)Log of the PSFs along the y-axis.
The simulations were performed with = 20 and tip-sample distance of d = 5nm.
[30]

Chapter 4 The Effect of Probe Geometry on PSF

4.3 The Cantilever Effect


Elias et al.2 examined the AM-PSFs generated from probes with and without the
cantilever. They observed that the cantilever reduces the PSF values but has no influence
on the FWHM. Consequently, they concluded that the cantilever alters the absolute CPD
values but does not affect the AM-KPFM resolution.
In this section we inspect the influence of the cantilever on FM-KPFM measurements.
The effect of the cantilever on the AM- and FM-PSFs is demonstrated in Fig. 18 for two
different tip-sample distances with and without the cantilever, represented by the solid
and dashed lines, respectively. For the probe with the cantilever we simulated a tilt angle
of = 20. At a tip-sample distance of 5nm the peak value of the full probe AM-PSF is

decreases by a factor of 1.65 compared to the tip AM-PSF (Fig. 18(a)), whereas the FM-

PSF is unchanged in the presence of the cantilever (Fig. 18(c)). For a tip-sample distance

of 30nm the peak is attenuated by a factor of 4.12 for the AM-PSF(Fig. 18(b)), whereas

the FM-PSF peak value is slightly decreased by 15% (Fig. 18()); this small attenuation

of the full probe (tip+cantilever) FM-PSF is resulted from a higher portion of

homogeneous force-gradient (25%) stems on the tilted tip cone and not from the
cantilever. The horizontal arrows represent the FWHM for the four cases demonstrate the
conclusion that the cantilever hardly affects the measurement resolution.

[31]

Chapter 4 The Effect of Probe Geometry on PSF

Fig. 18: One dimensional AM- and FM-PSFs simulated for two different tip-sample
distances d with and without the cantilever, represented by the solid (red) and dashed

(blue) lines, respectively. (a)AM-PSF for d=5nm. (b)AM-PSF for d=30nm. (c)FM-PSF
for d=5nm. (d)FM-PSF for d=30nm. The horizontal lines represent the FWHM of the
PSFs - solid lines for the full probe (tip+cantilever) and dashed lines for the tip only. The
simulations were performed with the probe illustrated in Fig. 14 with = 20 for the full

probe.

[32]

Chapter 5 Tip-Sample Distance Effect

5 Tip-Sample Distance Effect


In this chapter we analyze the effect of the tip-sample distance on the measurement
parameters in AM- and FM-KPFM. To this end, we calculate PSFs for the probe in
Fig. 14 for tip-sample distances ranging from 1nm to 100nm. Tilt angle of = 15 was
set for all simulations. Fig. 19(a) shows the FWHM dependence on the tip-sample
distance for both AM-PSF (dashed line) and FM-PSF (solid line). One can observe that
FM-PSF has smaller FWHM values than AM-PSF for all tip-sample distances, which
indicates a superior resolution of FM-KPFM over AM-KPFM. Fig. 19(2) shows the PSF
peak value (maximal PSF value) as a function of the tip-sample distance. It is observed
that the FM-PSF (solid line) demonstrates much higher peak value than the AM-PSF
(dashed line). For example, at tip-sample distance of 30nm, which is frequently used in
ambient KPFM, the FM-PSF peak is higher by an order of magnitude than the AM-PSF
peak. Since contrast of CPD images correlates with the PSF peak value, we conclude a
superior contrast in FM-KPFM compared to AM-KPFM.

Fig. 19: (a)Full width at half maximum of both AM-PSF (dashed line) and FM-PSF
(solid line) as a function of the tip-sample distance. (b)PSF peak value (maximal PSF
value) as a function of the tip sample distance for AM-PSF (dashed line) and FM-PSF
(solid line). All simulation were performed with the probe in Fig. 14 for = 15.
[33]

Chapter 5 Tip-Sample Distance Effect

Fig. 20 presents the homogenous force (blue line) and homogenous force-gradient (green
line) as a function of the tip-sample distance. The left axis corresponds to the force
whereas the right axis corresponds to the force-gradient. The force-gradient demonstrates
much steeper descent than the force; from 1nm to 100nm the force-gradient decreases by
4 orders of magnitude whereas the force decreases merely by 1 order of magnitude.
1

These curves are in good agreement with the long-range behavior of the force (Fz 2 )
dF

and the short-range behavior of the force-gradient ( dzz 3 ).


r

Fig. 20: Homogenous force (left axis) and homogenous force-gradient (right axis) vs. the
tip-sample distance. The force and force-gradient are marked with blue and green lines,
respectively.

[34]

Chapter 5 Tip-Sample Distance Effect

From Eq. (4.1) and Eq. (4.2) we derive:


inh
Fz,
VCPD
=
PSF(AM)

h
Fz,
VDC

dFz inh
VCPD
dz
= PSF(FM)
h
VDC
dFz
dz

(5.1)

(5.2)

The force and force-gradient were calculated for VDC = 1V , VAC = 1V and an

equipotential sample with VCPD = 1V, therefore the sums over the PSFs equals:
PSF(AM) =

inh
Fz,
h
Fz,



() =

(5.3)

(5.4)

inh
h
In AM-KPFM and FM-KPFM we should theoretically obtain Fz,
= Fz,
and
dFz inh
dz

dFz h
dz

, respectively. However, due to computational limitations mentioned in

Section 4.2 we cannot calculate the inhomogeneous force and force-gradient for infinite
surfaces. Therefore, a 192nm 192nm sample area was simulated beneath the tip.

Hence, summation over the PSFs gives the portion of the inhomogeneous force/forcegradient component in the simulated area. Fig. 21 presents the summation over the PSFs
as a function of the tip-sample distance. The left axis corresponds to the AM-PSF
whereas the right axis corresponds to the FM-PSF. It is observed that the summation over
the AM-PSF decreases rapidly whereas the summation over the FM-KPFM maintains
above 90% for conventional tip-sample distances (1nm to 50nm). Thus, less attenuation
of the CPD signal is expected with FM-KPFM since most of the detected force-gradient
stems from the area beneath the tip.

[35]

Chapter 5 Tip-Sample Distance Effect

Fig. 21: Summation over the AM-PSF (left axis) and FM-PSF (right axis) vs. the tipsample distance. All PSFs were simulated for 192nm 192nm sample area.

Summation on the simulated PSF also provides quantitative information about the
attenuation factor. If we assume a square feature of 192nm 192nm with a constant

CPD of VCPD surrounded by an infinite substrate with a constant CPD of Vsub , then the
KPFM measures:

+ Vsub 1 PSF

VDC = VCPD PSF

(5.5)

is the simulated PSF on a 192 192 sample.


Where
Consequently, the real CPD on the feature can be extracted with:
VCPD =

VDC Vsub 1 PSF

PSF

(5.6)

The next chapter demonstrates a more rigorous method to reconstruct the real CPD on a
feature.

[36]

Chapter 6 Image Reconstruction

6 Image Reconstruction
The following chapter describes an algorithm to reconstruct the actual CPD of a sample
from the KPFM measurements. In the first section we overview the CPD image
degradation resulting from the KPFM measurement and explain the deconvolution
process. In the second section we validate our reconstruction algorithm with a calibration
sample. The last section shows several CPD reconstruction results.

6.1 Outline of the Deconvolution Method


The effect of the measuring AFM probe in KPFM is very large, especially in AM-KPFM
where the measured forces are long range. The probe averaging effect degrades the
spatial resolution but more important largely change the magnitude of the measured
surface potential. The CPD measured on flat surfaces is modeled by a linear shiftinvariant system where the impulse response is the Point Spread Function (PSF). Fig. 22
illustrates the KPFM measurement:

Fig. 22: Input and Output signals in the KPFM system

[37]

Chapter 6 Image Reconstruction

The measured surface potential, VDC , is the sum of the convolution VCPD PSF with the

system noise, :

VDC = VCPD PSF +

(6.1)

Where * denotes convolution. Once the PSF and the noise statistics are obtained,
reconstruction of the KPFM measurements is possible using deconvolution. One might
consider using the straight-forward inversion:
VCPD = PSF 1 (VDC )

(6.2)

Because the PSF is a low pass filter (LPF), the noise is amplified greatly at high
frequencies. Therefore, Eq. (6.2) must be replaced with an optimized filter. We choose to
deconvolve the KPFM data using the Wiener filter, since it is straightforward, easy to
implement and in most cases gives good results. The Wiener filter is an optimal estimator
in term of minimum square error (MSE). Wiener deconvolution is performed in the
frequency domain by32:

PSF (u, v)
(
)
VDC (u, v)
VCPD u, v =
1
|PSF(u, v)|2 +
SNR(u, v)

Wiener Filter

Where (u, v) are the coordinates in the frequency domain, SNR(u, v) =

(6.3)

Sf(u,v)

Sn (u,v)

is the

signal-to-noise ratio with Sf (u, v) = |VCPD (u, v)|2 and Sn (u, v) = |N(u, v)|2 as the power
spectrums of VCPD and the noise, respectively. We estimate the power spectrum of VCPD
by the power spectrum of VDC 33. PSF (u, v) is the complex conjugate of PSF(u, v).

[38]

Chapter 6 Image Reconstruction

6.2 Reconstruction Process


6.2.1 PSF Expansion
As was shown in Section 4.2, summation over the AM-PSF components yields result
smaller than 1 due to the long-ranged electrostatic force. Therefore, we must extrapolate
the AM-PSF beyond the simulated surface.
Away from the origin, the two dimensional AM-PSF illustrated in Fig. 16(a) converges to
the expression:
|x| 50nm
|y| 50nm

PSF(x, y) = A1 x P1

PSF(x, y) =

A2 y P2 ,

A3 y P3 ,

y>0

(6.4)

y<0

Where A1 = A1 (y) , P1 = P1 (y) are positive numbers which vary for each y line and

A2 = A2 (x) , A3 = A3 (x) , P2 = P2 (x) , P3 = P3 (x) are positive numbers which vary for

each x line. 2 and 2 are related to the cantilever side whereas 3 and 3 are related to
the opposite side.

Considering the above expressions, we calculated all the parameters in Eq. (6.4) by
observing log(PSF) vs. log(x) and log(PSF) vs. log(y), as shown in Fig. 23(a). Once

all the parameters are obtained, we extrapolated the PSF for all x and y lines, as
illustrated by the blue arrows in Fig. 23(b). Next, we observe that near the corners the PSF
converges to the expression:

|r| 70nm

PSF(r) =

Ac r Pc ,
At r Pt ,

Where r = x 2 + y 2 s.t. |x| = |y|

y>0

y<0

(6.5)

, , , are positive numbers. r is the distance from the origin in a diagonal

direction,

as

illustrated

by

the

[39]

red

squares

in

Fig.

23(b).

Chapter 6 Image Reconstruction

Fig. 23: (a)Loglog plot of the AM-PSF vs. x and y lines. A linear relation is observed far
from the origin. (b)Expanding the AM-PSF to an infinite area. First, the x and y lines are
extrapolated (illustrated by blue arrows). Next, extrapolation is performed on the
diagonal direction (illustrated by red arrows). In the final step, the remaining pixels on
the plane (black domains) are interpolated using 2 pixels which were previously
extrapolated. An example for this interpolation is mark with a green area.
[40]

Chapter 6 Image Reconstruction

By calculating the parameters in Eq. (6.5) as demonstrated earlier, we extrapolate the PSF
to the corners. The red arrows delineate the direction of this extrapolation. The PSF
components on the remaining areas of the plane (black domains) are interpolated by
including two pixels which were extrapolated in the preceding steps. An example of such
interpolation is illustrated by the green area in Fig. 23(b). The two pixels used for this
interpolation are marked with red and blue filled squares.
The above algorithm was tested on 2 AM-PSFs generated from a full probe
(tip+cantilever) and a probe composed from only a tip, in both cases the tip-sample
distance was d = 2[nm], the tilt of the cantilever was = 15, all additional geometric

parameters are given in Fig. 14(a). By expanding the PSFs, the summation over the PSF
of the tip was increased from 0.98132 (for 192nm X 192nm area) to 1.00107 (for

37.5m X 37.5m area) and the summation over the PSF of the full probe was increased
from 0.83189 (for 5m X 5m area) to 0.99824 (for 0.3mm X 0.3mm area).

Alternatively, deconvolution can be performed by subtracting the average CPD on the


sample, Vsub , which is usually the CPD value of the substrate outside the scan area. Then,

reconstruction is performed using the following relation between VCPD Vsub and

VDC Vsub (Mathematical formation can be found in the Appendix):


VDC
V

sub

Deconvolution input

PSF
= (VCPD Vsub ) PSF

PSF
Deconvolution output

Where PSF denotes summation over the PSF.

[41]

(6.6)

Chapter 6 Image Reconstruction

6.2.2 Calibration Sample Measurements


We validate our reconstruction algorithm by scanning a calibration sample assembled of
Ni electrodes under predefined bias, and then deconvoluting by using an effective PSF
and acquiring the system noise statistics.

Thin lines of were patterned on an isolating layer of 2 using e-beam lithography.

Next, we mount the sample onto an AFM (NTMDT, Eindhoven, Netherlands) for
measuring topography and CPD in ambient. The topography of the sample was measured
via contact mode, and is shown in Fig. 24(a). From a cross section of the topography
(white line) we observe a height difference of 50nm (Fig. 24 (b)). The CPD of the sample
was measured via AM-KPFM twice, once where the substrate and the Ni electrodes share

common ground (Fig. 24 (c)) and second where the Ni contacts were biased (Fig. 24 (e)).

The left and right Ni contacts were biased with a voltage of 0.5V and 0.5V,

respectively. All KPFM measurements were conducted in lift-mode with 5nm lift height.
Conductive TiN coated tips (NTMDT) with a 1st resonance frequency of ~160kHz were
used. Cross sections of the surface potentials with and without external bias are shown in
Fig. 24(f) and Fig. 24(d), respectively. One can observe that due to the averaging effect
of the probe, the measured CPD contrast in the biased sample is 0.58V instead of 1V.

It should be noted that the actual CPD values are in opposite sign to our measured CPD
values since all the KPFM measurements were conducted by applying Kelvin voltage
(VDC ) on the probe34. We assume that the measured CPD is a convolution result of the

actual CPD and a PSF despite the existence of the topography observed in Fig. 24(b). Our
assumption was validated by Baier et al35 and Sadewasser et al36 who measured AMKPFM on flat surfaces and on surfaces with topography step. They observed almost no
change in CPD distributions for both cases. Simulations on these surfaces also confirmed
very little topography influence on the averaging effect.

[42]

Chapter 6 Image Reconstruction

Fig. 24: (a)Contact mode topography of the sample: two Ni contacts on a layer of SiO2 .
(b)Height profile (as indicated in (a)). (c)CPD image of the sample. (d)CPD profile (as
indicated in (c)). (e)CPD image of the sample while the left Ni contact was biased with
0.5V and the right Ni contact was biased with -0.5V, related to the SiO2 . (f)CPD profile

(as indicated in (e)). From this profile one can observe an almost linear slope in the
surface potential. A small deviation upwards is visible due to the higher CPD value on
the SiO2 compared to the Ni. All KPFM measurements were conducted with AM using
Lift-Mode where the lift height was 5nm.

[43]

Chapter 6 Image Reconstruction

In order to evaluate the system noise, we mask features in Fig. 24(c) to analyze the CPD
statistics of the pure SiO2 substrate (Fig. 25(a)). The measured CPD on the substrate is

shown in Fig. 25(b). The histogram in Fig. 25(c) shows the distribution of the SiO2 CPD

which is a Gaussian distribution with an expected value and variance of 0.5132V and
3.23 105 V 2 , respectively. The noise is obtained by subtracting the mean value.

Fig. 25(d) presents the autocorrelation of the noise in the marked area of Fig. 25(a). The
delta-function-like feature in the center indicates that white noise can accurately describe
the system; therefore we have used an Additive White Gaussian Noise (AWGN) in the
deconvolution process.

Fig. 25: Evaluation of the statistics of the noise. (a)Sections containing only clean SiO2
were framed for distinguishing the noise from the CPD signal. (b)The CPD of the SiO2
substrate taken from (a). (c)Histogram outlines the distribution of the substrate CPD
(bars) and a normal fit (dashed-line) with mean and variance of 0.5132V and

3.23 105 V 2, respectively. (d) Autocorrelation of the noise extracted from the marked
area in (a).

[44]

Chapter 6 Image Reconstruction

The theoretical peak of the autocorrelation function of the AWGN is given by32:
= 2

(6.7)

Where and are the number of pixels in the x and y axes, respectively, and 2 is the

variance of the noise. By substituting the values of , and 2 we obtain a theoretical

= 0.34427 2 , which is in excellent agreement with = 0.34432 2 obtained from


Fig. 25(d).

Deconvolution of Fig. 24(e) requires the PSF of the probe in the exact tip-sample
distance that was used in the KPFM measurement, which is d = dtopo + dlift where dtopo

is the averaged tip-sample distance used for recording the topography trajectory and
dlift = 5nm is the lift height distance for the KPFM measurement. Since dtopo varies for

every scan, more analytical approach is used to determine the exact tip-sample distance.

Since the averaging effect is not prominent when the Ni electrodes share common ground
(Fig. 24(c)), we approximate the actual CPD values on the SiO2 and Ni to be 0.513V and

0.465V, respectively. The KPFM is suitable for measuring only semiconductor or

metallic surfaces18; In practice, the measured CPD on the SiO2 is the CPD of the Si

substrate beneath the oxide layer37.

Fig. 26(a) shows the band diagram of the unbiased sample where Ef (Ni),Ef (Si) and
Ef (tip) indicate the Fermi levels of the Ni, Si and tip, respectively. LVL(tip) and

LVL(sample) indicate the LVLs of the tip and sample, respectively. The band diagram of
the sample under bias is presented in Fig. 26(b). When biased, the CPD of the left and

right Ni electrodes are 0.035[] and 0.965[], respectively. is defined as the CPD
between

the

Ni

and

Si

when

they

are

unbiased.

We

measured

= Ef (Si) Ef (Ni) = 50mV which is different from the theoretical value (= 220),

the deviation might be related to absorption of molecules from the ambient on the sample

surface or due to the small averaging effect of the probe, which is not entirely negligible
in Fig. 24(c). The voltage drop along the is not linear24, however, for simplicity we

assume constant electric field between the two electrodes.

[45]

Chapter 6 Image Reconstruction

Fig. 26: (a)Band diagram of the unbiased sample. (b)Band diagram of the sample where
the left and right Ni electrodes are biased with 0.5V and -0.5V, respectively, related to
the tip. Ef(Si),Ef(Ni),Ef(tip) refers to the Fermi levels of the silicon, Nickel and tip,
respectively. LVL(tip) and LVL(sample) are the LVLs of the tip and sample,
respectively, and = 50meV. The CPD is marked in red. Under bias, the CPD increases
from -0.035V (left electrode) to 0.965V (right electrode).

[46]

Chapter 6 Image Reconstruction

From the theoretical CPD shown in Fig. 26(b) we guess the actual CPD on the surface,
VCPD (Fig. 27 (a)). Due to a voltage drift between the first (unbiased) and second (biased)
scan, we shift all CPD values by 16.5[mV].

Fig. 27: (a)2D image of the theoretical CPD on the biased sample. (b)Measured CPD
(red), theoretical CPD (blue) and convolution of the theoretical CPD with PSFs generated
for tip-sample distance of 6nm (brown), 8nm (black) and 10nm(green). All CPD
linescans are plotted along the white line in (a).

[47]

Chapter 6 Image Reconstruction

The guessed VCPD is then convolved with PSFs at several tip-sample distances

(6nm 10nm) by using Eq. (A.22) in the Appendix:

VDC = (VCPD Vsub ) PSF + Vsub

(6.8)

Fig. 27(b) shows the different convolution results with the theoretical and measured CPD
along the cross-section (white line) marked in Fig. 27(a).
The PSF generated for tip-sample distance of 8nm is our best estimate since the

convolution result at 8nm bears the lowest error from VDC (by L2 norm). A good

indication for our estimate is the contrast of the measured CPD (0.557) which is very
similar to the summation over the PSF at 8nm (0.544).

Finally, after obtaining the PSF and the noise statistics, deconvolution is performed on
VDC (Fig. 24(e)) using Eq. (6.6). Fig. 28 shows the profiles of the measured and
deconvolved CPDs.

Fig. 28: Measured CPD (red), theoretical CPD (black) and deconvolved CPD (blue). All
CPD profiles are related to the linescan indicated in Fig. 24(e).

The deconvolved CPD resembles the measured CPD in shape and preserves a contrast of 1.03,

as expected from the actual CPD on the surface.

[48]

Chapter 6 Image Reconstruction

6.3 Results and Comparison with Measurements


6.3.1 Reconstruction of CdS-PbS Images
CdS-PbS nanorods (NRs) were fabricated on freshly-peeled highly-oriented pyrolytic
Graphite (HOPG) and topography and KPFM images were simultaneously recorded
using single-pass technique in an argon glovebox. The topography was measured in
tapping mode with ~50 (1 resonance frequency) while an AC bias modulation of
300 400 (2 resonance frequency) was added to measure KPFM38. A schematic
of the KPFM tip and sample characteristics is shown in Fig. 29:

Fig. 29: Schematic of KPFM experimental apparatus. The AFM tip used for this setup
has a tip apex of 30nm. The measured NRs are approximately 80nm in length and 4nm in
diameter.

In single pass technique, the Kelvin probe controller minimizes the oscillations of the
cantilevers second resonance, leaving it to oscillate only in the first mechanical
resonance. We define d as the average tip-sample distance and A as the amplitude of the

cantilever oscillation, as illustrated in Fig. 30(a). Since the cantilever in inclined towards
[49]

Chapter 6 Image Reconstruction

the surface by degrees, the movement of the probe is better comprehended in the
rotated axis system , as shown in Fig. 30(b).

Fig. 30: (a)Representation of the probe in the y-z axes. is the angle between the
cantilever and the horizontal axis. Distance d is the average tip-sample distance.
(b)Representation of the probe in the y'-z' axes. A is the oscillation amplitude.

Neglecting the influence of the tip on the cantilever oscillation, the vertical deformation
along the y axis as a function of time is given by39:
Z(y, t) =

A
By
By
By
By
2t
cos cosh + sin sinh cos

2
L
L
L
L
T0

(6.9)

Where B = 1.875, = 0.7341, L is the cantilever length and T0 is the oscillation


period. Thus, the deflection of the tip is:

Z(y = L, t) Acos

2t

T0

Considering the tip position of the cantilever unaffected by the sample (free cantilever) in
Fig. 30(a) to be at (y, z) = (0, d), the tip position in the rotated system will be:
(y , z ) = (dsin(), dcos())

Therefore, when oscillating the tip position over time is:


[50]

Chapter 6 Image Reconstruction

(y(t), z(t)) = dsin() , dcos() Acos

2t

T0

Using the rotation matrix, we describe the oscillations in the y z system:

2t
Acos
sin()
y(t)
cos() sin()
T0

2t =

2t
z(t)
sin() cos() dcos() Acos

T0
d Acos
cos()

T0
dsin()

(6.10)

The KPFM measurement was performed using tapping-mode, therefore we bound


min{ z} = 0 for the tip bottom position. Meaning, A =

cos()

. To take into account the

cantilever oscillations we sampled the position of the oscillating tip 25 times in one
oscillation period:
2n
dtan()cos

y(n)
N

=
2n
z(n)
d 1 cos

Where N = 25 and n increases from 1 to 25. In the measurement = 12 and

d = 16.7nm were used. Since this first mechanical resonance frequency is around

50kHz, whereas the time constant of the Kelvin controller is 1ms, the effective PSF is
calculated by minimizing the averaged electrostatic force on the probe rather than
minimizing the force at each tip-sample distance3.
Fig. 31(a) shows AM-KPFM measurement of CdS-PbS NRs on the HOPG substrate. The
CPD on the substrate is obtained by masking all the NRs in the images, similar to the
method used in Fig. 25(a) for the calibration sample. The measured CPD on the substrate
is presented in Fig. 31(b). The histogram in Fig. 31(c) shows the CPD distribution on the
substrate, which is a Gaussian distribution with an expected value and variance of
0.3862V and 5.4 105 V 2 , respectively; the noise is obtained by subtracting the mean
value. Fig. 31(d) presents the autocorrelation function of the noise in the marked

rectangle area in Fig. 31(a). The peak of the delta-function-like feature in the center
satisfies Eq. (6.7), therefore we deduce that the noise in the system can be considered as
an AWGN.

[51]

Chapter 6 Image Reconstruction

Fig. 31: (a)AM-KPFM measurement of CdS-PbS NRs on HOPG substrate. Two NRs
(rod1 and rod2) are selected for further analysis. (b)The CPD of the HOPG substrate
taken from (a). (c)Histogram presents the distribution of the substrate CPD (bars) and a
normal fit (dashed-line) with a mean and variance of 0.3862V and 5.4 105 V 2,
respectively. (d)Autocorrelation of the noise extracted from the marked area (blue) in (a).

[52]

Chapter 6 Image Reconstruction

Fig. 32(b) shows the actual CPD image, compared to the measured raw data (Fig. 32(a)).

Fig. 32: (a)Measured CPD on CdS-PbS NRs. (b)Actual CPD on CdS-PbS NRs obtained
by deconvolution with the Wiener filter.

The color scale of the Y-axis is indicative of the magnitude of CPD, and is thus increased
from Fig. 32(a) to Fig. 32(b), as a result of the deconvolution. The averaging effect of the
measuring probe explains the large attenuation of the signal in the measured CPD
compared to the actual CPD. The attenuation factor varies for each NR, ranging from 2.5
to 3.4 (average of 3.02). This fairly large attenuation factor is mainly due to the size of
the cantilever relative to the size of a typical NR. Fig. 33 shows examples of the effect of
the convolution on the CPD profile over two NRs along their longitudinal axis.

[53]

Chapter 6 Image Reconstruction

Fig. 33: Linescans of the measured (dashed line) and deconvolved (solid line) CPD across the
longitudinal axis of two symmetric PbS-CdS-PbS NRs (inset KPFM image) using the Wiener
filter, effective PSF and noise statistics. (a)CPD along rod1. (b)CPD along rod2. Both NRs are
highlighted in Fig. 31(a).

6.3.2 Reconstruction of Graphene Images


Topography and FM-KPFM measurements of Graphene layers were performed
simultaneously in ultrahigh vacuum (UHV) using single pass technique. For topography
measurements, the oscillation amplitude of the first resonance of the cantilever
(~100kHz) was held constant whereas the CPD was measured by setting a modulation
frequency of 1kHz with a bias amplitude of 200mV40. The time constant of the Kelvin
controller (~1ms) is not small enough to detect the force-gradient on the oscillating

probe (~100kHz) for every tip-sample distance; therefore the effective PSF is calculated
by minimizing the averaged electrostatic force-gradient on the probe rather minimizing
the force-gradient at each tip-sample distance. Thus, the measure potential becomes:

VDC () = VDC (, t)|

lim

1 T dFz
dt=0
T 0 dz

T d
0 dz H tinh dt
= lim
VCPD ()
T d
T
(
)
H
dt

h
0 dz

Effective PSF

[54]

(6.11)

Chapter 6 Image Reconstruction

The Graphene CPD images were taken by using a minimal tip-sample distance of 1nm,
therefore we bound min{ z} = 1nm in Eq. (6.10) to find the exact position of the
oscillating tip in time:

2t

Acos
sin()
T
0
y(t)

=
2t
z(t)
1 + () 1

T0

Where all the parameters are portrayed in Fig. 30. = 13 and A = 3nm 5nm were
used In the measurements. The effective PSF was then calculated in the same manner as
in Section 6.3.1.
Fig. 34(a) and Fig. 34(b) show FM-KPFM measurements of single layers and double
layers of Graphene where the double layers exhibit higher CPD than the single layers
with a measured contrast of about 140mV. The system noise statistics is obtained by
masking all transition areas between single layers and double layers of Graphene, as

shown in Fig. 34(c) and Fig. 34(d). We mark the single layer areas with II,IV and the
double layer areas with I,III,V. Fig. 34(e) and Fig. 34(f) present the CPD distribution on
these areas (blue bars). We distinguish between two separated Gaussians, the left and
right Gaussians correspond to the single layer and double layer areas, respectively. The
expected values and standard deviations of CPD on the layers are obtained by fitting
probability density functions (PDF) of bimodal Gaussian distributions.

[55]

Chapter 6 Image Reconstruction

Fig. 34: Obtaining system noise statistics from CPD measurements. (a) and (b) show raw
CPD measurements of single layers and double layers of Graphene. (c) and (d) present
only pure areas of single layers or double layers of Graphene without the interfacial area
between them. The single layers are marked with II,IV whereas the double layers are
marked with I,III,V. (e) and (f) show the CPD distribution (blue bars) on the areas
presented in (c) and (d), respectively, along with fitted PDFs of bimodal Gaussian fits
corresponding to them (dashed line).

[56]

Chapter 6 Image Reconstruction

The horizontal lines in Fig. 34(e) and Fig. 34(f) demonstrate more sparse CPD
distribution on the layers in Fig. 34(a) than in Fig. 34(b). The exact mean () and
standard deviation () on the layers are summarized in Table 1.
Single layer Graphene

Fig. 34(e)
Fig. 34(f)

407mV
1.182V

15.683mV
7.235mV

Double layer Graphene

277mV
1.029V

15.772mV
7.581mV

Table 1: Expected values () and standard deviations () of the CPD on single and
double layers of Graphene, obtained from Fig. 34(a) and Fig. 34(b). The values were
extracted from the bimodal Gaussian fits in Fig. 34(e) and Fig. 34(f).

Fig. 35(a) and Fig. 35(b) show the actual CPD (red) and the measured CPD (blue) along
the profiles illustrated by the lines in Fig. 34(a) and Fig. 34(b), respectively. From these
profiles one can observe that the averaging effect does not play any role in the
measurements.

Fig. 35: (a) and (b) show linescans of the measured (blue line) and deconvolved (red line)
CPD along the profiles illustrated in Fig. 34(a) and Fig. 34(b), respectively. The
deconvolved CPD values were calculated using the Wiener filter, effective PSF and noise
statistics obtained beforehand.
[57]

Chapter 7 Summary and Conclusions

7 Summary and Conclusions


This work presented an algorithm to reconstruct the actual CPD distribution on a sample
out of raw KPFM measurements. The CPD is recovered by calculating an effective PSF
for the measurement and obtaining the exact noise statistics. We have shown that
measurements conducted in AM-KPFM demonstrate very large averaging effect due to
the presence of the cantilever whereas measurements conducted in FM-KPFM display
CPD values close to the real CPD without any averaging.
In addition, we calculated the influence of the cantilever on the measured CPD. We
showed that for both AM- and FM-KPFM the cantilever does not change the
measurement resolution. However, it was shown that the cantilever has a profound
influence on the absolute measured CPD value in AM-KPFM whereas in FM-KPFM the
presence of the cantilever almost does not change the measured CPD. Furthermore, we
study the tip-sample distance effect in KPFM measurements and found that for
conventional tip-sample distances (1nm-50nm) there is almost no attenuation when
performing FM-KPFM, however in AM-KPFM measurements conducted above 10nm
the measured CPD is derived mainly from the substrate and not from the feature beneath
the tip apex.

The reconstruction algorithm is valid only for flat surfaces or surfaces which exhibit
small topography variations, since it is based on the fact that the system response is
invariant to the probe position. A possible extension to this work is to consider a real
sample with rough surface and correlate the surface potential to the KPFM signal while
taking under consideration the sample topography. A different extension might be a
development of an improved reconstruction algorithm by designing more sophisticated
deconvolution filters instead of the Wiener filter.

[58]

Chapter 8 - References

8 References
1
2
3
4
5
6
7
8
9
10
11
12
13

14
15
16
17

18
19
20

Glatzel, T. in 5th International Conference on Noncontact Atomic Force Microscopy.


Elias, G. Modeling and Simulations of Kelvin Probe Force Microscopy. M.Sc thesis, Tel
Aviv University, (2011).
Elias, G. et al. The role of the cantilever in Kelvin probe force microscopy measurements.
Beilstein J. Nanotechnol. 2, 252-260, doi:10.3762/bjnano.2.29 (2011).
Strassburg, E. A fast algorithm to restore the surface potential image from atomic force
microscopy measurements. M.Sc thesis, Tel Aviv University, (2005).
Strassburg, E., Boag, A. & Rosenwaks, Y. Reconstruction of electrostatic force
microscopy images. Rev. Sci. Instrum. 76, doi:10.1063/1.1988089 (2005).
Hudlet, S., Saint Jean, M., Guthmann, C. & Berger, J. Evaluation of the capacitive force
between an atomic force microscopy tip and a metallic surface. Eur Phys J B 2, 5-10
(1998).
Girard, P. Electrostatic force microscopy: principles and some applications to
semiconductors. Nanotechnology 12, 485 (2001).
Nonnenmacher, M., Oboyle, M. P. & Wickramasinghe, H. K. Kelvin Probe Force
Microscopy. Appl Phys Lett 58, 2921-2923 (1991).
Kelvin, L., Fitzgerald, G. & Francis, W. Contact electricity of metals. Phil. Mag 46, 82
(1898).
Weaver, J. M. R. & Abraham, D. W. High-Resolution Atomic Force Microscopy
Potentiometry. J Vac Sci Technol B 9, 1559-1561 (1991).
Jacobs, H. O., Knapp, H. F. & Stemmer, A. Practical aspects of Kelvin probe force
microscopy. Rev. Sci. Instrum. 70, 1756-1760 (1999).
Li, G., Mao, B., Lan, F. & Liu, L. Practical aspects of single-pass scan Kelvin probe force
microscopy. Rev. Sci. Instrum. 83, 113701 (2012).
Ziegler, D. & Stemmer, A. Force gradient sensitive detection in lift-mode Kelvin probe
force microscopy. Nanotechnology 22, doi:Artn 075501 Doi 10.1088/09574484/22/7/075501 (2011).
Kitamura, S. & Iwatsuki, M. High-resolution imaging of contact potential difference with
ultrahigh vacuum noncontact atomic force microscope. Appl Phys Lett 72, 3154-3156
(1998).
Albrecht, T. R., Grutter, P., Horne, D. & Rugar, D. Frequency-Modulation Detection Using
High-Q Cantilevers for Enhanced Force Microscope Sensitivity. J Appl Phys 69, 668-673
(1991).
Ziegler, D. Techniques to quantify local electric potentials and eliminate electrostatic
artifacts in atomic force microscopy Ph.D thesis, ETH, (2009).
Zerweck, U., Loppacher, C., Otto, T., Grafstrom, S. & Eng, L. M. Accuracy and resolution
limits of Kelvin probe force microscopy. Physical Review B 71, 125424, doi:Artn 125424
Doi 10.1103/Physrevb.71.125424 (2005).
Melitz, W., Shen, J., Kummel, A. C. & Lee, S. Kelvin probe force microscopy and its
application. Surface Science Reports 66, 1-27, doi:10.1016/j.surfrep.2010.10.001 (2011).
Loppacher, C. et al. FM demodulated Kelvin probe force microscopy for surface
photovoltage tracking. Nanotechnology 16, S1 (2005).
Sommerhalter, C., Glatzel, T., Matthes, T. W., Jger-Waldau, A. & Lux-Steiner, M. C.
[59]

Chapter 8 - References

21
22
23
24
25
26
27
28
29
30
31

32
33
34

35
36

Kelvin probe force microscopy in ultra high vacuum using amplitude modulation
detection of the electrostatic forces. Appl Surf Sci 157, 263-268, doi:10.1016/s01694332(99)00537-1 (2000).
Charrier, D. S. H., Kemerink, M., Smalbrugge, B. E., de Vries, T. & Janssen, R. A. J. Real
versus Measured Surface Potentials in Scanning Kelvin Probe Microscopy. ACS Nano 2,
622-626, doi:10.1021/nn700190t (2008).
Girard, P., Ramonda, M. & Saluel, D. Electrical contrast observations and voltage
measurements by Kelvin probe force gradient microscopy. Journal of Vacuum Science &
Technology B: Microelectronics and Nanometer Structures 20, 1348-1355 (2002).
Glatzel, T., Sadewasser, S. & Lux-Steiner, M. C. Amplitude or frequency modulationdetection in Kelvin probe force microscopy. Appl Surf Sci 210, 84-89, doi:Doi
10.1016/S0169-4332(02)01484-8 (2003).
Hudlet, S., Saint Jean, M., Roulet, B., Berger, J. & Guthmann, C. Electrostatic forces
between metallic tip and semiconductor surfaces. J Appl Phys 77, 3308-3314,
doi:10.1063/1.358616 (1995).
Hochwitz, T., Henning, A. K., Levey, C., Daghlian, C. & Slinkman, J. Capacitive effects on
quantitative dopant profiling with scanned electrostatic force microscopes. J Vac Sci
Technol B 14, 457-462, doi:Doi 10.1116/1.588494 (1996).
Sadewasser, S., Glatzel, T., Shikler, R., Rosenwaks, Y. & Lux-Steiner, M. C. Resolution of
Kelvin probe force microscopy in ultrahigh vacuum: comparison of experiment and
simulation. Appl Surf Sci 210, 32-36, doi:Doi 10.1016/S0169-4332(02)01475-7 (2003).
Law, B. M. & Rieutord, F. Electrostatic forces in atomic force microscopy. Physical
Review B 66, 035402 (2002).
Jacobs, H. O., Leuchtmann, P., Homan, O. J. & Stemmer, A. Resolution and contrast in
Kelvin probe force microscopy. J Appl Phys 84, 1168-1173, doi:Doi 10.1063/1.368181
(1998).
Machleidt, T., Sparrer, E., Kapusi, D. & Franke, K.-H. Deconvolution of Kelvin probe force
microscopy measurementsmethodology and application. Measurement Science and
Technology 20, 084017 (2009).
Shikler, R. PhD thesis, Tel-Aviv University, (2003).
Rosenwaks, Y., Shikler, R., Glatzel, T. & Sadewasser, S. Kelvin probe force microscopy of
semiconductor surface defects. Physical Review B 70, doi:Artn 085320 Doi
10.1103/Physrevb.70.085320 (2004).
Gonzalez, R. C. & Woods, R. E. Digital Image Processing (3rd Edition). (Prentice-Hall,
Inc., 2006).
Machleidt, T., Sparrer, E., Kubertschak, T., Nestler, R. & Franke, K.-H. Kelvin probe force
microscopy: measurement data reconstruction. 73781C-73781C, doi:10.1117/12.821787
(2009).
Rosenwaks, Y. & Shikler, R. in Scanning Probe Microscopy: Characterization,
Nanofabrication and Device Application of Functional Materials Vol. 186 NATO Science
Series II: Mathematics, Physics and Chemistry (eds PaulaMaria Vilarinho, Yossi
Rosenwaks, & Angus Kingon) Ch. 6, 119-151 (Springer Netherlands, 2005).
Baier, R., Leendertz, C., Lux-Steiner, M. C. & Sadewasser, S. Toward quantitative Kelvin
probe force microscopy of nanoscale potential distributions. Physical Review B 85,
165436 (2012).
Sadewasser, S., Leendertz, C., Streicher, F. & Lux-Steiner, M. C. The influence of surface
topography on Kelvin probe force microscopy. Nanotechnology 20, 505503 (2009).
[60]

Chapter 8 - References
37
38
39
40

Barth, C., Foster, A. S., Henry, C. R. & Shluger, A. L. Recent Trends in Surface
Characterization and Chemistry with High-Resolution Scanning Force Methods.
Advanced Materials 23, 477-501, doi:10.1002/adma.201002270 (2011).
Nanayakkara, S. et al. Built-in potential and charge distribution within single
heterostructured nanorods measured by scanning Kelvin probe microscopy. Under
review (2012).
Young, D. & Felgar, R. P. Tables of characteristic functions representing normal modes of
vibration of a beam. The University of Texas Engineering Research Series 44 (1949).
Held, C., Seyller, T. & Bennewitz, R. Quantitative multichannel NC-AFM data analysis of
graphene growth on SiC(0001). Beilstein J. Nanotechnol. 3, 179-185, doi:Doi
10.3762/Bjnano.3.19 (2012).

[61]

Chapter 9 - Appendix

9 Appendix
The appendix defines in detail the matrices which were used in this thesis and elaborate
on mathematical formulations and derivation related to this work.
We define i as the location of the ith boundary element of the probes surface. The ij
element of matrix G is given by:
Gij =

1
1
1

ds
40 Sj |i | |i |

(A.1)

Where = (x , y , z ) is the location of jth boundary element of the probes surface and

= (x , y , z ) is the location of the image charge of the probes jth element relative to
an infinite ground plane at z = 0 in the homogenous system. The integral is calculated
over the jth surface element of the probe.

The diagonal of matrix is given by:


Bii =

ds
20 Si

(A.2)

i is the outward unit vector normal to the ith surface element of the probe. The
Where
integral is calculated over the ith surface element of the probe.

The sample surface potential is discretized using uniform square elements. The sides of
the square equal to the resolution of the scan, denoted as . The center of the kth surface

element is located at k = (k x , k y , 0) where k x , k y are integers. The ik element of


matrix D is given by:

Dik =

(i )
1

ds
2 Sk |i |3

The integral is calculated over the kth surface element of the sample.

[62]

(A.3)

Chapter 9 - Appendix

By placing the expression for the surface charge density from Eq. (3.10) to Eq. (3.11) we
find the vertical force on the ith probe element:
2

2
Fzi = 2i Bii = Ch i Vprobe
2Ch i Cinh VCPD Vprobe + Cinh VCPD Bii
i

(A.4)

For extracting the component of the force, Fz, , we use Eq. (3.12) and substitute in
2
= 2VAC VDC .
(A.4): Vprobe, = VAC ; Vprobe,

Therefore, the vertical force on the ith probe element at frequency is:
2

Fz,i = Ch i 2 2Ch i Cinh VCPD Bii

(A.5)

Fz, () = 2VAC VDC ()Cht B Ch Cht B Cinh VCPD ()

(A.6)

Thus, we find the expression for the total force on the probe at frequency :
From the minimum force condition (Fz, = 0) and from (A.6) we derive Eq. (3.14).

The force-gradient on the ith probe element is obtained by differentiating Eq. (A.4):

dCh i
dCh i
dFz
2
Cinh VCPD Vprobe
= 2Ch i
Vprobe
2
dz i
dz
dz
i
2Ch i

d Cinh VCPD
dz

Vprobe + 2 Cinh VCPD

d Cinh VCPD
dz

(A.7)
i

Bii

2
By using Eq. (3.12) and substituting Vprobe, = VAC ; Vprobe,
= 2VAC VDC we get:

d Cinh VCPD
dCh i
dCh i
dFz
i
Cinh VCPD VAC 2Ch i
= 2Ch i
2VDC VAC 2
VAC Bii
dz i
dz
dz
dz
i

dC
We define: C h = dzh and C inh =

dCinh
dz

. Thus, for each probe element:

[63]

(A.8)

Chapter 9 - Appendix

d Cinh VCPD
dFz
i

= 2Ch i Ch i 2VDC VAC 2Ch i Cinh VCPD VAC 2Ch i


VAC Bii
dz i
dz
i

(A.9)

And the total force-gradient on the probe is therefore:

d Cinh VCPD
dFz
= 4Cht B C h VDC VAC 2C ht B Cinh VCPD VAC 2Cht B
VAC
dz
dz

(A.10)

VCPD = VCPD (x, y) is independent in z. Therefore, the force gradient is:

dFz
= 2VAC 2Cht B C h VDC C ht B Cinh VCPD Cht B C inh VCPD
dz

The minimum gradient condition,


and VCPD :
VDC (r) =

dFz

dz

(A.11)

= 0, gives the following relation between VDC

C ht B Cinh + Cht B C inh


2Cht B C h

VCPD (r)

(A.12)

B is related to the probe geometry and it is unaffected by the tip-sample distance. Thus, a
shorter notation for the numerator in Eq. (A.12) is:

d
(A.13)
C t B Cinh
dz h
Furthermore, since B is diagonal: C ht B Ch = Cht B C h . Thus, the denominator in
C ht B Cinh + Cht B C inh =

Eq. (A.12) can be written as:

2Cht B C h = C ht B Ch + Cht B C h =

d
C t B Ch
dz h

Substituting Eq. (A.13) and Eq. (A.14) in (A.12) gives Eq. (3.16).

[64]

(A.14)

Chapter 9 - Appendix

=
We define the gradient of D in the z direction as D
ik =
D

Differentiation the term in Eq. (A.3) gives:


ik =
D

dDik
dz

dD
dz

. Meaning:

|i |2 3zi2
1

ds
|i |5
2 Sk

(A.15)

(A.16)

is therefore:
The sum of each row in D
i = D
ij =
D
j

2 Ssample

1
1
3zi2

ds =
2 Ssample |i |3 |i |5

3zi2

3
5 dx dy
((xi x )2 + (yi y )2 + zi2 )2 ((xi x )2 + (yi y )2 + zi2 )2
1

(A.17)

Solving Eq. (A.17) yields Eq. (3.25).

Let be the physical, continuous Point Spread Function of the probe. The Kelvin
voltage (DC voltage) on the probe satisfies:

VDC = VCPD = (VCPD Vsub + Vsub )

(A.18)

VDC Vsub = (VCPD Vsub )

(A.19)

Since Sample ds = 1 we obtain:

Let PSF be the discrete Point Spread Function calculated by Matlab. PSF is cropped to a
specific area:

[65]

Chapter 9 - Appendix

PSF(x, y) =

(x, y) ,
0 ,

|x|

n
n
res |y| res
2
2

(A.20)

otherwise

Where n is the number of simulated surface points in x- and y-axes and res is the scan
resolution.

If there are no features outside the scan area we can approximate (VCPD (x, y) Vsub ) 0
n

for x and y outside the scan area. Since (, ) values for |x|, |y| > res are

significantly smaller than the values close to the origin, we estimate:

(VCPD Vsub ) PSF (VCPD Vsub )

(A.21)

VDC Vsub = (VCPD Vsub ) PSF

(A.22)

And from Eq. (A.19) we get:

Since deconvolution in Matlab requires a normalized PSF, we divide and multiply the
above expression by PSF to get Eq. (6.6).

[66]



) .(Kelvin Probe Force Miscroscopy
,
) .(tip ,
) (Point Spread Function
.
.


. ,
. .
- ) (nanorods CdS-PbS
) .(UHV

.

,
. ) (Full Width Half Maximum
.
- ) 1 50(
, - 10
.

" "



'

,"

" "

,"

Potrebbero piacerti anche