ArticlePDF Available

Pump-probe Kelvin-probe force microscopy: Principle of operation and resolution limits

Authors:
  • Lochgelly High School

Abstract and Figures

Knowledge on surface potential dynamics is crucial for understanding the performance of modern-type nanoscale devices. We describe an electrical pump-probe approach in Kelvin-probe force microscopy that enables a quantitative measurement of dynamic surface potentials at nanosecond-time and nanometer-length scales. Also, we investigate the performance of pump-probe Kelvin-probe force microscopy with respect to the relevant experimental parameters. We exemplify a measurement on an organic field effect transistor that verifies the undisturbed functionality of our pump-probe approach in terms of simultaneous and quantitative mapping of topographic and electronic information at a high lateral and temporal resolution.
Content may be subject to copyright.
Pump-probe Kelvin-probe force microscopy: Principle of operation and resolution
limits
J. Murawski, T. Graupner, P. Milde, R. Raupach, U. Zerweck-Trogisch, and L. M. Eng
Citation: Journal of Applied Physics 118, 154302 (2015); doi: 10.1063/1.4933289
View online: http://dx.doi.org/10.1063/1.4933289
View Table of Contents: http://scitation.aip.org/content/aip/journal/jap/118/15?ver=pdfcov
Published by the AIP Publishing
Articles you may be interested in
High potential sensitivity in heterodyne amplitude-modulation Kelvin probe force microscopy
Appl. Phys. Lett. 100, 223104 (2012); 10.1063/1.4723697
Nonuniform doping distribution along silicon nanowires measured by Kelvin probe force microscopy and
scanning photocurrent microscopy
Appl. Phys. Lett. 95, 092105 (2009); 10.1063/1.3207887
Potential shielding by the surface water layer in Kelvin probe force microscopy
Appl. Phys. Lett. 80, 1459 (2002); 10.1063/1.1455145
Electrical characterization of an operating Si pn-junction diode with scanning capacitance microscopy and Kelvin
probe force microscopy
J. Appl. Phys. 90, 443 (2001); 10.1063/1.1375803
Imaging of a silicon pn junction under applied bias with scanning capacitance microscopy and Kelvin probe force
microscopy
Appl. Phys. Lett. 77, 106 (2000); 10.1063/1.126892
[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
138.251.31.218 On: Wed, 28 Oct 2015 11:24:27
Pump-probe Kelvin-probe force microscopy: Principle of operation
and resolution limits
J. Murawski, T. Graupner, P. Milde,
a)
R. Raupach, U. Zerweck-Trogisch,
b)
and L. M. Eng
Institute of Applied Physics, Technische Universit
at Dresden, D-01062 Dresden, Germany
(Received 6 July 2015; accepted 3 October 2015; published online 15 October 2015)
Knowledge on surface potential dynamics is crucial for understanding the performance of modern-
type nanoscale devices. We describe an electrical pump-probe approach in Kelvin-probe force
microscopy that enables a quantitative measurement of dynamic surface potentials at nanosecond-
time and nanometer-length scales. Also, we investigate the performance of pump-probe Kelvin-
probe force microscopy with respect to the relevant experimental parameters. We exemplify a
measurement on an organic field effect transistor that verifies the undisturbed functionality of our
pump-probe approach in terms of simultaneous and quantitative mapping of topographic and elec-
tronic information at a high lateral and temporal resolution. V
C2015 AIP Publishing LLC.
[http://dx.doi.org/10.1063/1.4933289]
I. INTRODUCTION
Kelvin-probe force microscopy (KPFM) has proven a
reliable tool for measuring local contact potential differences
(CPDs) on the nanometer length scale.
13
Despite numerous
advances since its first appearance,
4,5
the quantitative investi-
gation of dynamic CPDs with KPFM at the nanosecond-time
scale remains a challenge. Yet, knowing the temporal devel-
opment of surface potentials is crucial for understanding the
underlying electrodynamics in nanoscale structures, such as
the switching behavior of nanodevices or integrated circuits.
Nunes et al. were the first to introduce a junction-mixing
scanning tunneling microscope (STM) enabling time-resolved
tunneling current measurements down to the picosecond-time
regime.
6,7
At the same time, Bridges et al. started to comple-
ment electrostatic force microscopy with a heterodyne nullify-
ing technique in order to resolve digital voltage signals in
bit-lines close to the Gbit/s data rate with sub-nanosecond pre-
cision.
8,9
This approach already enabled quantitative electro-
dynamic investigations at the sub-nanosecond-time scale with
an atomic force microscope (AFM). A decade later Coffey
and Ginger coined the term time-resolved electrostatic force
microscopy by evaluating the frequency shift of an oscillating
tip upon externally induced stimuli in the scanning condi-
tion.
10
With this method, they obtained both spatial and tem-
poral information and were able to record decay rates in the
millisecond-time regime. More recent works include quantita-
tive CPD measurements recorded from organic or ferroelectric
thin films using standard KPFM setups.
1113
Here, the band-
width of the control loop allows for a temporal resolution in
the millisecond range.
In this work, we describe the modification of a
frequency-modulated KPFM (FM-KPFM) setup. As in the
early junction-mixing approaches, we take advantage of the
non-linear electric force interaction between tip and sample,
which is independent of the cantilever dynamics and the
control electronics’ bandwidth.
69
We establish a pump-
probe scheme by switching the low-frequency modulation
voltage Vmod synchronously to pump pulses that cause a tem-
poral change of the local CPD. Here, we demonstrate that
pump-probe Kelvin-probe force microscopy (pp-KPFM) pre-
serves the full functionality of standard FM-KPFM, i.e., a
high spatial resolution and undisturbed sample topography
measurements, while additionally providing the desired tem-
poral resolution needed for the quantitative investigation of
surface-bound periodic processes.
In Section II, we give an introduction into the physics of
pp-KPFM. The technical realisation is then introduced in
Section III. Finally, the experimental parameter space and re-
solution limits are discussed in great detail in Sections IV
and V, respectively.
II. THE PHYSICS OF PUMP-PROBE KELVIN PROBE
FORCE MICROSCOPY
Kelvin-probe force microscopy is based on minimizing
the electric force caused by the local contact potential differ-
ence VCPD between tip and sample. Minimization is achieved
by applying a compensation voltage Vdc to the tip, equaling
VCPD. Usually, the temporal behavior of VCPD(t) is of less
importance, hence nullifying the mean value
VCPD within
10–100 ms is sufficient. In order to enhance the detection
sensitivity, the electric force is modulated by a sinusoidal
voltage VmodðtÞ¼vmod sinðxmod tÞat frequency xmod. These
three voltages cause a total electric force Fel between tip and
sample that equals
3,14
Fel ~
r;t
ðÞ
¼1
2
@C~
r
ðÞ
@zVdc þVCPD ~
r
ðÞ
þVmod t
ðÞ

2;(1)
where Cdenotes the tip-sample capacitance, and zis the tip-
sample distance. For KPFM, only the term at frequency
xmod is relevant
Fxmod
el ~
r;t
ðÞ
¼@C~
r
ðÞ
@zVdc þVCPD ~
r
ðÞ

Vmod t
ðÞ
:(2)
a)
Electronic mail: peter.milde@tu-dresden.de.
b)
Also at Institut f
ur Luft- und K
altetechnik, Bertolt-Brecht-Allee 20,
D-01309 Dresden, Germany.
0021-8979/2015/118(15)/154302/8/$30.00 V
C2015 AIP Publishing LLC118, 154302-1
JOURNAL OF APPLIED PHYSICS 118, 154302 (2015)
[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
138.251.31.218 On: Wed, 28 Oct 2015 11:24:27
The modulated force Fxmod
el becomes zero either when
the compensation voltage matches the contact potential
difference (Vdc ¼VCPD—constituting the standard way of
operating KPFM) or whenever the modulation voltage itself
becomes zero (Vmod ¼0). For pp-KPFM, we employ the lat-
ter case by multiplying the modulation voltage Vmod with the
function
Pðxreptþudel Þ
¼1udel xrept2pkudel þus;k2Z
0 else;
((3)
where xrep is the repetition rate of the resulting electric
pulses, us¼xrep sis the pulse phase-width of pulse width
s, and udel ¼xrep tdel is the phase-offset of the pulses at the
time delay tdel relative to the beginning of the repetition
period Trep. Hence, we generate probe pulses within a sinu-
soidal envelope, as shown in Fig. 1(a).
Due to this pulse-amplitude modulation, we excite our
investigated system electrically only in the on-time window
Trep kþtdel tTrep kþtdel þs;k2Z, and only sys-
tem responses that fall into this on-time window remain de-
tectable. Furthermore, when we synchronize the repetition
frequency xrep to a periodic pump signal [as schematically
shown in Fig. 1(c)], the time-sensitive pp-KPFM not only
detects the static CPD
VCPD but also is equally able to moni-
tor the time-dependent system response ~
VCPD convolved
with the probe pulse P.
In order to quantify the time-dependent CPD ~
VCPD,the
electric force between tip and sample Fel needs to be mini-
mized in the above-stated time window. Conveniently, this
force leads to a modulation of the AFM’s cantilever reso-
nance frequency which manifests as a sideband of the canti-
lever oscillation signal at frequencies x06xmod, as shown
in the inset of Fig. 1(b). Demodulating the sideband signal
by a lock-in amplifier (LIA) renders the following error
signal X:
X/vmodfg½Vdc þ
VCPDð~
rÞ þ ½ ~
VCPDð~
r;tÞPðudelÞg ;(4)
where vmod denotes the amplitude of the modulation voltage,
g¼s=Trep is the duty cycle of the function P, and
½~
VCPDð~
r;tÞPðudelÞstands for the convolution of the time-
dependent potential ~
VCPD with the probe pulse function Pat
a specific phase offset udel
~
VCPD ~
r;t
ðÞ
P

udel
ðÞ
¼1
Trep ðt0þTrep
t0
~
VCPD ~
r;xrept

Pxreptþudel
ðÞ
dt:(5)
Finally, a Kelvin control loop nullifies the detected error
signal Xby applying the compensation voltage
Vdc ~
r;udel
ðÞ
¼
VCPD ~
r
ðÞ
þ1
g
~
VCPD ~
r;t
ðÞ
P

udel
ðÞ

:(6)
This way, besides the time-independent contact potential dif-
ference
VCPDð~
rÞalso the time-dependent CPD ~
VCPDð~
r;tÞis
accessible. The temporal resolution is determined by the
convolution term in Eq. (5) and is, therefore, limited to the
pulse width s. A schematic of the here-described pp-KPFM
set-up is shown in Fig. 1(b).
For very small duty cycles g, Eq. (6) can be approxi-
mated by a Taylor series as
Vdcð~
r;udelÞ¼½
VCPDð~
rÞþ ~
VCPDð~
r;udelÞ
¼VCPDð~
r;udelÞ:(7)
FIG. 1. (a) Schematic of the pump-probe scheme. Probe pulses with a sinusoidal envelope (yellow dashed line) are synchronized to a square wave pump signal
(green solid line). The zoom-in shows a single repetition cycle Trep. (b) Sketch of the signal path in the pp-KPFM set-up. The modulation voltage is mixed
with the square wave signal P. The resulting probe signal enables time-resolved surface potential measurements. Both the probe signal as well as the control
voltage of the pp-KPFM loop are summed up and applied to the tip. The spectrum of the resulting cantilever motion is shown in the inset on the top left. (c)
Schematic of the tip-sample system used for all characterization measurements. The red cross on the sample indicates that the position of the tip was kept fixed
during these experiments.
154302-2 Murawski et al. J. Appl. Phys. 118, 154302 (2015)
[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
138.251.31.218 On: Wed, 28 Oct 2015 11:24:27
III. DUAL CONTROL LOOP pp-KPFM
According to Eq. (7), our pump-probe approach enables
the Kelvin control loop to match and detect the time-
dependent potential ~
VCPDð~
r;tÞwithin the mentioned on-time
window Trep kþtdel tTrep kþtdel þs;k2Z. The
compensation voltage Vdc, however, is applied continuously
throughout the whole repetition period Trep. This implies that
although the potential is compensated correctly within the
on-time window, the time-averaged potential ^
VCPDð~
rÞmay
significantly deviate from the time-resolved value at udel ,
thus inducing cross-talk from the pp-KPFM compensation
voltage Vdc onto the measured topography.
Fig. 2depicts this effect. A square-wave voltage is
applied to a gold sample at a frequency of xrep ¼32 kHz,
while the phase delay udel of the probe pulse is steadily
increased, as outlined in Fig. 1(a). In this pp-KPFM point-
measurement, the tip stays at one point above the sample and
is not scanned (see Fig. 1(c)). The measured height appears
to be smallest and least disturbed when the delay reaches the
slopes of the square-wave voltage where the time-averaged
CPD ^
VCPD coincides with the time-resolved CPD ~
VCPD and
is, therefore, exactly compensated: Vdc ¼~
VCPD ¼^
VCPD.
At all other phase delays, the measured height appears
falsely larger due to an additional electric force arising from
the mismatch of applied compensation voltage Vdc and time-
averaged CPD ^
VCPD.
In order to minimize the induced topographic artefacts,
we employ a second control loop in parallel to the time-
resolved control loop, as schematically depicted in Fig. 3.
This second standard KPFM control loop compensates
solely the time-averaged potential ^
VCPDð~
rÞand, hence, mini-
mizes any time-averaged electric force that might adversely
affect the topography measurement. The first control loop in
turn remains sensitive to any time-dependent signal and
applies its compensation voltage only within the desired on-
time windows defined by the function P. The dual control
loop approach thus minimizes the overall topography error
while preserving time-resolved CPD measurements.
Notably, the two KPFM control loops need to operate at
their own and independent modulation frequency, which we
set to xavg for the time-averaging Kelvin control loop.
Furthermore, we denote the two modulation amplitudes with
vmod and vavg. Also the compensation voltage splits into the
two terms ~
Vdc and ^
Vdc, stemming from the time-sensitive
and the time-averaging control loop, respectively.
Note that only the modulation voltage Vmod and the
time-dependent compensation voltage ~
Vdc are pulse-
FIG. 2. Topographic height zas a function of the phase delay udel
¼xdel tdel of the probe pulse at the slope of the square wave pump signal
(green dashed line). The probe pulse width is represented by the tanned rec-
tangle in the background. When using a single control loop for pp-KPFM
(blue solid line), the minimal height occurs at the slope of the pump signal,
since the average electrostatic force is minimized only at this phase delay.
With the dual control loop pp-KPFM (purple dashed-dotted line), the correct
height value is always measured regardless of the phase delay.
FIG. 3. Sketch of the signal path in the
dual control loop pp-KPFM set-up.
The output of the time-sensitive pp-
KPFM loop, i.e., the sum of modula-
tion and control voltage, is mixed with
the square-wave signal P. The result-
ing probe signal enables time-resolved
surface potential measurements. A sec-
ond Kelvin control loop with Kelvin
modulation frequency xavg is used in
order to minimize the average electro-
static force. Both the probe signal as
well as the output of the second Kelvin
loop are summed up and applied to the
tip. The spectrum of the resulting can-
tilever motion is shown in the inset on
the top left.
154302-3 Murawski et al. J. Appl. Phys. 118, 154302 (2015)
[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
138.251.31.218 On: Wed, 28 Oct 2015 11:24:27
amplitude-modulated, while the modulation voltage Vavg of
the second, time-averaging control-loop as well as its com-
pensation voltage ^
Vdc keep their integrity. In total, the volt-
age applied to the tip reads as Vtip ¼ð~
Vdc þVmodÞP
þ
Vdc þVavg, as illustrated in Fig. 3.
This dual-loop approach results in two distinct error sig-
nals Xmod and Xavg at their respective frequencies xmod and
xavg, as indicated by the inset in Fig. 3
Xmod /gf~
Vdc þ^
Vdc þ
VCPDð~
rÞg þ ½ ~
VCPDð~
r;tÞPðudelÞ;
(8)
Xavg /g
~
Vdc þ^
Vdc þ
VCPDð~
rÞþ ^
VCPD:(9)
In analogy to our previous considerations, these two
error signals are nullified when each control loop applies its
respective compensation voltage ~
Vdc or ^
Vdc
~
Vdc ~
r;udel
ðÞ
¼ ^
Vdc ~
r
ðÞ
þ
VCPD ~
r
ðÞ
þ1
g
~
VCPD ~
r;t
ðÞ
P

udel
ðÞ

;
(10)
^
Vdcð~
r;udelÞ¼½g
~
Vdcð~
r;udelÞþ
VCPDð~
rÞþ ^
VCPDð~
rÞ :
(11)
For small duty cycles, the convolution term in Eq. (10) again
can be approximated by a Taylor series. Rearranging Eqs.
(10) and (11), we find
~
Vdc ~
r;udel
ðÞ
¼ 1
1g
~
VCPD ~
r;udel
ðÞ
^
VCPD ~
r
ðÞ

;(12)
^
Vdc ~
r;udel
ðÞ
¼
VCPD ~
r
ðÞ
þ1
1g
^
VCPD ~
r
ðÞ
g
1g
~
VCPD ~
r;udel
ðÞ
:(13)
Since both ~
Vdc and ^
Vdc compensate more than the time-
dependent or the time-averaged CPD only, we further com-
pute from Equations (12) and (13) the time-averaged CPD
^
V
dcð~
rÞ¼ ^
Vdcð~
r;udelÞþg
~
Vdcð~
r;udelÞ
¼½
VCPDð~
rÞþ ^
VCPDð~
rÞ ;(14)
as well as the time-resolved CPD
~
V
dcð~
r;udelÞ¼ ^
Vdcð~
rÞþ ~
Vdcð~
r;udelÞ
¼½
VCPDð~
rÞþ ~
VCPDð~
r;udelÞ
¼VCPDð~
r;udelÞ:(15)
The purple dashed-dotted line in Fig. 2illustrates a
point-measurement similar to the previous one, this time
recorded with the dual control loop pp-KPFM setup. As
shown, no cross-talk between electric signal and measured
height is discernible over the whole period.
To exemplify the cross-talk between the two control
loops, we used the same sample and applied a square-wave
voltage as in the previous experiments with the exception
that the repetition frequency was chosen to xrep ¼20 MHz.
Fig. 4(a) depicts the data as recorded by the two control
loops. As shown, the time-dependent signal is centered
around 0 V in accordance to Eq. (12). Furthermore, there
is a distinct cross-talk from the time-dependent CPD
visible in the time-averaged signal. Finally, the undisturbed
time-resolved CPD has been retrieved by computation after
Eq. (15).
IV. EXPERIMENTAL PARAMETER SPACE
A. Duty cycle and modulation voltage amplitude
Eq. (6) predicts the contact potential Vdcðudel Þto be
strongly dependent on the duty cycle g. This behavior was
calculated after Eq. (6) in Fig. 5(a) for the measurable con-
trast DV=Vrep of a square waveform of 50% duty cycle and
amplitude Vrep. In order to verify this prediction, we applied
FIG. 4. Measurement of a 20 MHz square-wave signal with an amplitude of
2 V using a 4 ns-wide pulse with 1 V amplitude as mixing function P. (a)
Overview over the compensation voltages of the time-sensitive control loop
~
Vdc (blue dashed-dotted line), the time-averaging control loop ^
Vdc (purple
dashed line), and the detected time-dependent surface potential ~
VCPD (or-
ange solid line) as a function of the delay time tdel .~
Vdc is partly visible in
^
Vdc, while ~
Vdc itself is shifted by ^
VCPD compared to ~
VCPD.~
VCPD itself is cal-
culated after Equation (15). (b) Oscilloscope measurement of the pump
square-wave signal (green dashed-dotted line) and of the probe pulse signal
P(yellow dashed line). The orange solid line displays the pp-KPFM
measurement.
154302-4 Murawski et al. J. Appl. Phys. 118, 154302 (2015)
[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
138.251.31.218 On: Wed, 28 Oct 2015 11:24:27
a square-wave voltage with repetition frequency xrep
¼100 kHz and amplitude Vrep ¼0:5 V to a highly p-doped
silicon sample for different duty cycles gand mapped the as-
deduced waveform by varying the phase delay udel in our
pp-KPFM set-up over a full repetition period Trep. For this
experiment, the tip stayed stationary at one point above the
sample. From the measured Vdcðudel Þ, we located the upper
and lower voltage plateau as well as their difference, which
reveals the contrast DV=Vrep.
Fig. 5(a) plots the resulting voltage contrast DV=Vrep as
a function of duty cycle g. Generally, measurement and
theory match excellently, especially with respect to the con-
trast loss at higher duty cycles. Surprisingly, for very small
duty cycles, a logarithmic decrease reduces the measured
contrast. This observation was reproduced on different sam-
ples with different measurement parameters, and also using
different setups. Although the origin of this deviation is not
entirely understood, it is reasonable to assume that the side-
band amplitude drops below the noise level for small duty
cycles, making stable pp-KPFM operation at a useful band-
width impossible.
The severe reduction in error signal magnitude is indi-
cated in Fig. 5(b), where the output voltage Xof the time-
sensitive LIA is shown as a function of the compensation
voltage Vdc for different duty cycles g. For decreasing duty
cycles, the sensitivity @X
@Vdc reduces linearly with gfollowing
the prediction of Eq. (4).
Since the error signal Xin Eq. (4) depends also on the
modulation voltage amplitude vmod, we repeated the previous
experiment varying vmod only. The inset of Fig. 5(a) depicts
the resulting signal contrast DV=Vrep as a function of duty
cycle gfor various modulation voltage amplitudes vmod.As
already mentioned, the signal contrast declines logarithmi-
cally for small duty cycles. Yet, larger modulation ampli-
tudes vmod seem to shift the onset of the decline towards
smaller duty cycles, supporting the above-stated assumption
that an insufficiently low error signal Xmight be causing the
decline.
Hence, our data suggest to employ larger modulation
voltage amplitudes vmod at small duty cycles g. Yet, the mod-
ulation voltage amplitude should be chosen carefully and
accordingly to the investigated system, for increasing the
amplitude can simultaneously enhance parasitic electric
interactions between tip and sample.
15
Finally, we face a trade-off between high temporal reso-
lution, i.e., small duty cycles g, and a sufficiently large error
signal Xin the experiment. We thus recommend to keep the
duty cycle above 0.1% in order to mitigate contrast loss to
less than 3 dB. Moreover, we also advise to adjust the repe-
tition frequency xrep in accordance to the desired temporal
resolution sin order to sustain a reasonable error signal am-
plitude X.
B. Repetition frequency
In theory, there are no limits with respect to the repeti-
tion frequency xrep. Experimentally, the repetition frequency
and any resonance frequency of the KPFM setup should be
incommensurate, especially with respect to the tip resonance
x
0
and the modulation frequencies xmod and xavg.
To exemplify the broad frequency range that is accessi-
ble via pp-KPFM, we repeated the former experiment for
different signal frequencies ranging from 10 Hz <xrep
<1 MHz. As expected, the full signal contrast is obtained
for all frequencies xrep, as evident from Fig. 6. However, as
indicated by the insets, measurements at low signal frequen-
cies may misrepresent the applied square-wave signal and,
hence, increase measurement errors. This stems from various
system resonances of the setup at these frequencies cross-
talking into the measured signal, as well as from the similar-
ity of signal and modulation frequency. At duty cycles above
g¼5%, correct and stable pp-KPFM measurements over the
entire repetition frequency range are possible with the upper
frequency limit given only by the signal generator in use.
C. Temporal resolution limit
The ultimate temporal resolution of pp-KPFM is deter-
mined by the convolution term in Eq. (6) and, therefore, only
limited to the probe pulse width s. In order to study this limit
of pp-KPFM, we applied a square waveform with a signal
frequency of 20 MHz and a signal amplitude of Vrep ¼2Vto
a copper sample and probed the pp-KPFM response with a
4 ns-wide pulse, which is the smallest possible pulse width in
our current electronic setup. Fig. 4(b) depicts the signal
FIG. 5. (a) Contrast DV=Vrep of a square-wave signal, as measured by pp-
KPFM as a function of probe duty cycle g. The green solid line represents
the calculated contrast, whereas dots represent measured results. Overall, the
measured contrast falls in line with our predictions, except for duty cycles
below 1 %. For these g, a logarithmic decrease sets in, which is exemplified
in the inset for different modulation voltage amplitudes vmod : The lower the
modulation amplitude the earlier the decline sets in. The yellow dashed lines
depict a schematic of the corresponding probe signal at small and large duty
cycles, respectively. (b) Error signal Xas a function of compensation voltage
Vdc at different probe duty cycles g. As shown in the inset, for falling gwe
observe a linear decrease of the sensitivity @X
@Vdc, i.e., the slope of the error
signal.
154302-5 Murawski et al. J. Appl. Phys. 118, 154302 (2015)
[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
138.251.31.218 On: Wed, 28 Oct 2015 11:24:27
shapes as measured with an oscilloscope, revealing a full-
width-at-half-maximum of 4:5 ns for the probe pulse and a
rise time of 4 ns for the square waveform used here.
As shown in Fig. 4(b), the measured pp-KPFM signal is
blurred with a total rise time of 10 ns which arises from three
contributions: (i) the 4 ns rise time of the pump signal, (ii)
the convolution of the pump signal with the 4 ns-wide probe
pulse, and (iii) another 2 ns due to the wiring setup of the
sample. Although they contribute to the blurring of the meas-
ured signal, the rise time of the pump signal as well as the
wiring per se pose no limit to pp-KPFM but are merely
sampled and reproduced here. Therefore, we can conclude
that the intrinsic temporal resolution of our setup is indeed
limited by the probe pulse width s. The temporal resolution
might be enhanced by utilizing a faster pulse generator.
Nevertheless, even when operated at the actual temporal
resolution limit, pp-KPFM renders the true signal amplitude
and reproduces the overall signal shape, which enables quan-
titative inspection of a multitude of electrodynamic proc-
esses on the nanosecond-time and nanometer-length scale.
V. THE LATERAL RESOLUTION OF pp-KPFM
In order to verify the ability to acquire topographic and
time-resolved CPD data simultaneously, we performed pp-
KPFM scans on a pentacene-based bottom gate organic
field-effect transistor (OFET) in coplanar geometry possess-
ing interdigitating gold contacts as source and drain electro-
des
16
with a channel length of L¼5lm. The scan area was
chosen from the edge of the source electrode to the corner of
the drain electrode, as indicated in Fig. 7(a).
Fig. 7(b) provides the topography of the scan area dis-
playing the two elevated electrodes at the left hand-side and
at the bottom right corner, as well as the channel region in-
between. The gold electrodes are covered with small penta-
cene grains, while the channel region is coated with larger
pentacene grains featuring scattered taller bulk phase grains.
The displayed topography was acquired simultaneously to
and unimpaired by the time-resolved CPD data in a single
pass scan, demonstrating the functionality and stability of
the dual control loop pp-KPFM.
A fixed voltage VGS was applied to the bottom gate elec-
trode in order to generate a fixed background charge-carrier
density; the source electrode was grounded while a square-
wave voltage was applied to the drain electrode with an
amplitude of VDS ¼2 V at a frequency of xrep ¼50 kHz.
The CPD development was captured by taking 40 frames
with a temporal resolution of s¼2ls over the full repetition
period of 20 ls. A selection of four frames is shown in
Figs. 7(c)–7(f) (Multimedia view).
The onset of the repetition period Trep at tdel ¼0lsis
shown in Fig. 7(c). Most prominently, the drain electrode
FIG. 6. Contrast DV=Vrep as a function of probe duty cycle gat different rep-
etition frequencies xrep. For very low repetition frequencies, the contrast
error increases at low duty cycles due to cross-talk with various setup
resonances. For duty cycles above g¼5%, measurements are possible over
the entire frequency range.
FIG. 7. (a) Schematic of the investi-
gated OFET structure with interdigitat-
ing gold electrodes. The red frame
indicates the position of the studied
area. (b) Topography of the investi-
gated OFET with the source electrode
on the left hand-side, a corner of the
drain electrode at the bottom right, and
the channel region in between, all cov-
ered with 30 nm pentacene. The
dashed lines indicate the borders of the
electrodes in all images. (c)–(f) 2D pp-
KPFM measurements at designated
time delays tdel of the probe pulse rela-
tive to the beginning of the 20 ls pump
cycle as indicated by the respective
insets. The coloured isopotential lines
highlight the evolution of the CPD
within the OFET channel. (Multimedia
view) [URL: http://dx.doi.org/10.1063/
1.4933289.1]
154302-6 Murawski et al. J. Appl. Phys. 118, 154302 (2015)
[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
138.251.31.218 On: Wed, 28 Oct 2015 11:24:27
exhibits a 2 V lower CPD than the source electrode. Between
the electrodes, the surface potential drops according to the
charge-carrier distribution inside the channel region with a
notable drop close to the drain electrode, as indicated by the
accumulated isopotential lines. This drop hints at an insuffi-
cient charge-carrier density and space charge limited trans-
port.
16
8ls after switching, the picture changes in Fig. 7(e),
where now a more equidistant distribution of the isopotential
lines is observed, pointing to additional charge-carriers
injected from the source electrode.
Fig. 7(d) displays the time-resolved CPD shortly after
grounding the drain electrode. Although the gold electrode
has returned to zero potential now, the channel region still
exhibits significant excess charge-carriers, resulting in a non-
equilibrium potential distribution within the channel region.
It takes another 8 ls for the channel to completely deplete all
excess charge-carriers before the cycle restarts, as shown in
Fig. 7(f).
16
As observed in all pp-KPFM data recorded from this
OFET structure, the CPD step at the edge of the left elec-
trode demonstrates a lateral CPD resolution that is limited
solely by the 20 nm radius of the AFM tip.
VI. SUMMARY
We introduced a novel method to measure periodical
time-dependent contact potential differences quantitatively
on the nanometer-length scale by implementing a pump-
probe scheme into Kelvin-probe force microscopy. We have
shown that the resulting pump-probe-driven Kelvin control
loop can cause topographic artefacts, and presented a way
how to circumvent these by employing a second control loop
that minimizes the average CPD of the investigated time-
dependent CPD. Furthermore, we explored the parameter
space of pp-KPFM and demonstrated the technique’s stabil-
ity over a wide range of parameters while simultaneously
pointing to the boundaries of stable operation. We concluded
that the temporal resolution is only limited by the probe
pulse width s, which in turn is only bound to the bandwidth
of the employed pulse generator. Temporal resolutions even
in the sub-nanosecond-time regime are in reach with an
appropriate pulse generator. Moreover, we demonstrated the
unimpaired acquisition of topographic data and verified the
lateral resolution of pp-KPFM to the tip size. Future work
will extend this pump-probe method to other types of pump
mechanisms, as, for example, the optical excitation of
(photo-)voltage.
VII. COMPLEMENTARY INFORMATION
A. Experimental setup
Our pp-KPFM experiments were carried out with a
custom-made AFM. All scans were performed in non-
contact AFM (nc-AFM) mode using a digital phase-locked
loop with a bandwidth of 500 Hz to control the frequency
shift Dxat a constant cantilever oscillation amplitude
17
and
at room temperature. OFET measurements were carried out
with the same nc-AFM under a nitrogen atmosphere in order
to avoid degradation of the organic materials.
As depicted in Fig. 3, the signal path was altered slightly
in order to implement pp-KPFM into our existing FM-KPFM
setup. The first LIA demodulates the tip oscillation at x
0
and
supplies the following LIAs with the necessary side-band
information to demodulate the signals at xmod and xavg ,
respectively. Between the modulating sine generator of the
time-sensitive LIA and the AFM tip, a pulse generator and a
signal switch were added. The pulse generator provides tran-
sistor-transistor-logic pulses Pthat trigger the signal switch,
which in turn switches the modulation voltage Vmod as well
as the compensation voltage ~
Vdc added beforehand. This
results in a pulsed modulation voltage on top of a likewise
pulsed compensation voltage.
In order to compensate the average electrostatic error orig-
inating from the time-sensitive control loop, a third LIA was
employed operating at a separate modulation frequency xavg
followed by a conventional KPFM loop. The resulting com-
pensation voltage ^
Vdc as well as the second modulation voltage
Vavg are then added unaltered to the pulse-modulated signal
from the pp-KPFM branch and finally applied to the tip.
In all measurements, we employed tips from the
Olympus OMCL-AC240TM series. All LIAs are supplied by
Stanford Research with the first LIA being an SR844 and the
second two LIAs an SR830 each. For parameter space meas-
urements, a Stanford Research DG535 pulse generator to-
gether with a Mini Circuits ZASWA-2–50DR signal switch
were utilized, whereas the waveform generator was a
Tektronix AFG3252. For the temporal resolution limit mea-
surement, the same waveform generator was employed for
generating both the pump signal and the probe pulses. The
signal switch for the dual control loop pp-KPFM was built
around an Analog Devices ADL5391 and positioned in close
proximity (<50 mm) to the AFM tip in order to reduce wir-
ing influences. Signals different from the pp-KPFM measure-
ments were acquired directly at the output of the waveform
generator using a LeCroy wavepro 900 oscilloscope with an
attached Yokogawa PB500 probe. The gate voltage in the
OFET measurement was supplied by a Keithley 2602 source
meter, while the square wave pump voltage as well as the
probe pulses were again provided by a Tektronix AFG3252.
B. Large duty cycles
It is noteworthy that very large duty cycles grestore the
temporal resolution capability of the pump-probe approach
in analogy to what is known as Babinet’s principle in
optics.
18
The effect is best described by introducing the
function
Pðxreptþudel Þ¼1Pðxreptþudel Þ;(16)
with the complementary pulse width s¼Trep s. The error
signal Xin Eq. (4) then can be rewritten as
X/vmodfgðVdc þ
VCPDð~
rÞÞ þ ^
VCPDð~
rÞ
½~
VCPDð~
r;tÞPðudelÞg ;(17)
where ^
VCPDð~
rÞdenotes the time average of ~
VCPD integrated
over one repetition period Trep, and ½~
VCPDð~
r;tÞPðudelÞis
154302-7 Murawski et al. J. Appl. Phys. 118, 154302 (2015)
[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
138.251.31.218 On: Wed, 28 Oct 2015 11:24:27
the convolution of ~
VCPD with the function Pat a phase off-
set udel. Again nullifying X, the Kelvin control loop applies
the compensation voltage
Vdc ~
r;udel
ðÞ
¼
VCPD ~
r
ðÞ
þ1
g
^
VCPD ~
r
ðÞ
1
g
~
VCPD ~
r;t
ðÞ
P

udel
ðÞ
:(18)
In contrast to Eq. (6) where the value of the convolution
term was proportional to g, in this case the convolution term
scales proportionally to the complementary duty cycle
g¼ð1gÞ. This behavior reduces the absolute measured
signal contrast with increasing duty cycle g.
Furthermore, for g¼1, the convolution term vanishes
Vdcð~
r;udelÞ¼½
VCPDð~
rÞþ ^
VCPDð~
rÞ ;(19)
such that only the time-independent contact potential differ-
ence
VCPDð~
rÞand the time-averaged potential ^
VCPDð~
rÞremain
detectable, i.e., the result for standard KPFM is restored.
To illustrate these consequences, the solid line in
Fig. 5(a) represents the expected signal contrast DV=Vrep
for a 50% duty cycle square-wave signal plotted as a func-
tion of the probe duty cycle g. It becomes evident that the
penalty for an increased error signal strength Xat large g
manifests in the loss of signal contrast.
ACKNOWLEDGMENTS
We gratefully acknowledge financial support by the
German Science Foundation through the Cluster of
Excellence “Center for Advancing Electronics Dresden,” the
Grant No. ZE 891/1-1, and through the RTG 1401/2. The
authors thank Dr. Moritz Philipp Hein and Dr. Sylvia Nicht
for providing OFET samples.
1
M. R. Weaver and D. W. Abraham, J. Vac. Sci. Technol., B 9, 1559
(1991).
2
M. Nonnenmacher, M. P. O’Boyle, and H. K. Wickramasinghe, Appl.
Phys. Lett. 58, 2921 (1991).
3
U. Zerweck, C. Loppacher, T. Otto, S. Grafstr
om, and L. M. Eng, Phys.
Rev. B 71, 125424 (2005).
4
S. Kitamura and M. Iwatsuki, Appl. Phys. Lett. 72, 3154 (1998).
5
W. Melitz, J. Shen, A. C. Kummel, and S. Lee, Surf. Sci. Rep. 66,1
(2011).
6
G. Nunes, Jr. and M. R. Freeman, Science 262, 1029 (1993).
7
G. M. Steeves, A. Y. Elezzabi, and M. R. Freeman, Appl. Phys. Lett. 72,
504 (1998).
8
G. E. Bridges, R. A. Said, and D. J. Thompson, Electron. Lett. 29, 1448
(1993).
9
G. E. Bridges, J. Vac. Sci. Technol., B 13, 1375 (1995).
10
D. C. Coffey and D. S. Ginger, Nat. Mater. 5, 735 (2006).
11
C. Melzer, C. Siol, and H. von Seggern, Adv. Mater. 25, 4315 (2013).
12
E. Strelcov, S. Jesse, Y.-L. Huang, Y.-Ch. Teng, I. I. Kravchenko, Y.-H.
Chu, and S. V. Kalinin, ACS Nano 7, 6806 (2013).
13
M. Beu, K. Klinkm
uller, and D. Schlettwein, Phys. Status Solidi 211, 1960
(2014).
14
S. Sadewasser and M. Lux-Steiner, Phys. Rev. Lett. 91, 266101 (2003).
15
T. Glatzel, S. Sadewasser, and M. Lux-Steiner, Appl. Surf. Sci. 210,84
(2003).
16
A comprehensive discussion of the OFET will be given in a follow-up
publication. Here, the OFET serves as a showcase for demonstrating the
dual control loop pp-KPFM capabilities.
17
C. Loppacher, M. Bammerlin, F. M. Battiston, M. Guggisberg, D. M
uller,
H. R. Hidber, R. L
uthi, E. Meyer, and H.-J. G
untherodt, Appl. Phys. A 66,
S215 (1998).
18
H. G. Booker, J. Inst. Electr. Eng., Part IIIA: Radiolocation 93(4), 620
(1946).
154302-8 Murawski et al. J. Appl. Phys. 118, 154302 (2015)
[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
138.251.31.218 On: Wed, 28 Oct 2015 11:24:27
... In addition, most time-resolved KPFM modes developed so far rely on the use of a pulsed/modulated illumination chain. This is for instance the case of intensitymodulated KPFM [7][8][9], pump-probe KPFM [4,10,11], or G-mode KPFM [12], to name a few (we refer the readers to review articles for a more comprehensive introduction to AFMbased time-resolved potentiometric and electrostatic modes [13,14]). ...
... Nevertheless, it is prone to capacitive artefacts [15]. Accurate quantitative measurements can be performed by pump-probe KPFM (pp-KPFM), with a time-resolution down to the sub-nanosecond scale [10]. However, pp-KPFM features a severely limited bandwidth. ...
... It cannot be used for an active electrostatic compensation. However, as in the case of pumpprobe KPFM [4,10], it is possible to use a second additional FM-KPFM loop which cancels the "time-averaged" dc part of the potential. This feature will not be used in this work, but its implementation has been validated. ...
Article
Full-text available
We present a new open-loop implementation of Kelvin probe force microscopy (KPFM) that provides access to the Fourier spectrum of the time-periodic surface electrostatic potential generated under optical (or electrical) pumping with an atomic force microscope. The modulus and phase coefficients are probed by exploiting a double heterodyne frequency mixing effect between the mechanical oscillation of the cantilever, modulated components of the time-periodic electrostatic potential at harmonic frequencies of the pump, and an ac bias modulation signal. Each harmonic can be selectively transferred to the second cantilever eigenmode. We show how phase coherent sideband generation and signal demodulation at the second eigenmode can be achieved by using two numerical lock-in amplifiers configured in cascade. Dual-heterodyne KPFM (DHe-KPFM) can be used to map any harmonic (amplitude/phase) of the time-periodic surface potential at a standard scanning speed. The Fourier spectrum (series of harmonics) can also be recorded in spectroscopic mode (DHe-KPFM spectroscopy), and 2D dynamic images can be acquired in data cube mode. The capabilities of DHe-KPFM in terms of time-resolved measurements, surface photovoltage (SPV) imaging, and detection of weak SPV signals are demonstrated through a series of experiments on difference surfaces: a reference substrate, a bulk organic photovoltaic heterojunction thin film, and an optoelectronic interface obtained by depositing caesium lead bromide perovskite nanosheets on a graphite surface. The conclusion provides perspectives for future improvements and applications.
... Kelvin probe force microscopy (KPFM) 1-3 is a powerful tool for studying electrostatic properties in a nanometer [4][5][6][7] or even atomic scale. [8][9][10][11][12] In this method, an AC bias at f m is applied between the tip and the sample, which induces the first-and second-harmonic electrostatic forces, F 1st ele and F 2nd ele , respectively. ...
... Furthermore, we show the dependence of Df m j jand Df 2m j jon V h and V l with h % 304 (Fig. 5), where Df m j j and Df 2m j j should be described as Eq. (6). Both Df m j j and Df 2m j jshow a linear dependence on V h and V l , which is consistent with Eq. (6). ...
... (6). Both Df m j j and Df 2m j jshow a linear dependence on V h and V l , which is consistent with Eq. (6). ...
Article
The use of a heterodyne detection scheme in Kelvin probe force microscopy (KPFM) is an effective way for enhancing the performance of KPFM. However, this detection scheme generally has difficulty in detecting the first- and second-harmonic electrostatic forces simultaneously. To overcome this problem, we propose dual-bias modulation heterodyne frequency modulation KPFM (DM-hetero-FM KPFM), in which dual AC biases at [Formula: see text] are applied between the tip and the sample. DM-hetero-FM KPFM enables us to measure the contact potential difference and capacitance gradient simultaneously at high frequencies (in the MHz range) beyond the bandwidth of phase-lock loop. Moreover, the present method allows us to perform it in the open-loop mode, which is highly desired for performing KPFM on semiconductors or in liquids at high frequencies.
... The mapping of local electrical properties of the interface is essential to further our understanding of corrosion, sensing, solar cells, energy storage devices, and bioelectric interfaces [3][4][5][6][7][8]. Since its first application in 1991 [2], there have been significant developments in the field of KPFM [6,9,10] with significant advances in both temporal [11][12][13][14] and spatial resolution [13,[15][16][17][18][19]. These advances have enabled investigations mapping light-induced surface potential dynamics [20], ferroelectric domains [19], individual quantum dots [21,22], and even submolecular charge distributions [23][24][25][26][27]. ...
... The mapping of local electrical properties of the interface is essential to further our understanding of corrosion, sensing, solar cells, energy storage devices, and bioelectric interfaces [3][4][5][6][7][8]. Since its first application in 1991 [2], there have been significant developments in the field of KPFM [6,9,10] with significant advances in both temporal [11][12][13][14] and spatial resolution [13,[15][16][17][18][19]. These advances have enabled investigations mapping light-induced surface potential dynamics [20], ferroelectric domains [19], individual quantum dots [21,22], and even submolecular charge distributions [23][24][25][26][27]. ...
... OL techniques, by contrast, are feedback-free and can be used to determine the CPD without the need to apply a DC bias [28,29]. OL techniques are increasingly being adopted to enable the mapping of voltage-sensitive materials [30][31][32], to enable investigations of fast electrodynamic processes [11][12][13][14] and to enable measurements in liquid environments (where bias application could lead to stray currents and unwanted electrochemical reactions) [9,[33][34][35]. OL techniques avoid the limitations and artefacts that can arise when using a feedback loop, for example, bandwidth limitations due to the time constant of the feedback loop [29], increased noise [36,37], and electrical crosstalk [38,39]. ...
Article
Full-text available
In this paper, we derive and present quantitative expressions governing the performance of single and multifrequency Kelvin probe force microscopy (KPFM) techniques in both air and water. Metrics such as minimum detectable contact potential difference, minimum required AC bias, and signal-to-noise ratio are compared and contrasted both off resonance and utilizing the first two eigenmodes of the cantilever. These comparisons allow the reader to quickly and quantitatively identify the parameters for the best performance for a given KPFM-based experiment in a given environment. Furthermore, we apply these performance metrics in the identification of KPFM-based modes that are most suitable for operation in liquid environments where bias application can lead to unwanted electrochemical reactions. We conclude that open-loop multifrequency KPFM modes operated with the first harmonic of the electrostatic response on the first eigenmode offer the best performance in liquid environments whilst needing the smallest AC bias for operation.
... En effet, la taille de la sonde détermine la période temporelle durant laquelle la modulation de la force électrostatique est appliquée sur la pointe (cette modulation est utilisée par le lock-in pour déterminer la valeur de V CPD ). En particulier, il peut être démontré que la sensibilité de la mesure pp-KPFM (δV dc ) est donnée par l'expression suivante [111] : ...
... Dans une configuration à deux boucles, une deuxième boucle est ajoutée, en plus de la boucle permettant les mesures résolues en temps, pour détecter et compenser le potentiel de surface moyen à l'origine des artefacts sur la topographie. [111]. Le rectangle jaune représente la sonde et le rectangle en pointillé vert représente l'excitation électrique i.e. la pompe Dans le cadre de ce travail, nous avons implémenté les deux configurations i.e. une boucle et deux boucles. ...
Thesis
La caractérisation des interfaces est cruciale pour plusieurs types de dispositifs comme les diodes électroluminescentes, les cellules solaires, les photo-détecteurs ou les transistors. Dans le cas particulier des photodétecteurs, les défauts situés aux interfaces sont considérés comme une source de courant d'obscurité, phénomène qu'il est nécessaire de réduire pour améliorer les performances du capteur. Pour surmonter tous les problèmes liés aux défauts, la passivation des interfaces est nécessaire. Nous pouvons citer deux approches de passivation telles que la passivation chimique (c.a.d. diminution de la densité de défaut) et la passivation par effet de champ (c.a.d. diminution de l'activité électrique des défauts).Ce travail de thèse s'inscrit dans le développement de méthodologies de caractérisation des empilements de passivation via l'étude du photo-potentiel de surface. Cette thèse propose, décrit et utilise un ensemble de méthodologies basées sur la microscopie à sonde de Kelvin sous illumination continue et sous illumination pulsée. L’originalité de ces travaux de thèse repose sur des développements méthodologiques ayant conduit à l’implémentation d’une approche de mesure du potentiel de surface résolue en temps permettant la caractérisation directe des dynamiques du photo-potentiel qui, à leur tour, permettent de caractériser les défauts situés aux interfaces. L'objectif de ces travaux est de proposer une nouvelle approche de caractérisation des défauts afin d’en améliorer leurs compréhensions. Une attention particulière sera accordée à l’étude de plusieurs empilements de passivation comme l'Al2O3, l’HfO2, le Ta2O5 et le SiN. Enfin, les limites de cette nouvelle approche concernant son application en industrie seront discutées et une solution alternative sera proposée.
... For example, a method was proposed where the cantilever deflection signal was directly recorded in real time and then the frequency and phase shifts in the time domain were computationally demodulated to investigate the time constants for their response in the sub-µs range [12][13][14]. Another demodulation method was proposed, where pulse trains synchronized with the tip motion were applied [15,16], and the charge dynamics under pulsed laser excitation or electrical voltage application were investigated with a time resolution around µs. Furthermore, the pump-probe method [17][18][19][20][21] is another promising way to realize time-resolved measurements in EFM and KFM. The advantage of the pump-probe method is that fast signal detection and/or large data storage systems are not always required. ...
Article
Full-text available
We have proposed the base-bias-level control method, in which a contact potential difference is always compensated in a similarly way to Kelvin probe force microscopy, applicable to time-resolved electrostatic force microscopy using a pump-probe method. We experimentally acquired temporal waveforms of the electrostatic force signal taken on two SiO 2 / n -type Si samples, one of which was as grown and the other was treated by forming gas annealing, and consequently the effectiveness of the proposed method has been well confirmed.
... In timeresolved experiments, the sample often undergoes multiple global perturbations while the probe measures the local response at a specific nanoscale location before moving on to the next location. To date, tr-EFM and tr-KPFM have been applied to a variety of systems, including optoelectronic materials [16][17][18] , semiconductors for transistors [19][20][21] or memristors 22 , with temporal resolution ranging from 10 s to 1 ps [23][24][25][26][27] . While these approaches provide a detailed record of how the sample's properties evolve at each location over time, they are typically limited to fast processes that are fully reversible within the measurement window. ...
Article
Full-text available
Unraveling local dynamic charge processes is vital for progress in diverse fields, from microelectronics to energy storage. This relies on the ability to map charge carrier motion across multiple length- and timescales and understanding how these processes interact with the inherent material heterogeneities. Towards addressing this challenge, we introduce high-speed sparse scanning Kelvin probe force microscopy, which combines sparse scanning and image reconstruction. This approach is shown to enable sub-second imaging (>3 frames per second) of nanoscale charge dynamics, representing several orders of magnitude improvement over traditional Kelvin probe force microscopy imaging rates. Bridging this improved spatiotemporal resolution with macroscale device measurements, we successfully visualize electrochemically mediated diffusion of mobile surface ions on a LaAlO3/SrTiO3 planar device. Such processes are known to impact band-alignment and charge-transfer dynamics at these heterointerfaces. Furthermore, we monitor the diffusion of oxygen vacancies at the single grain level in polycrystalline TiO2. Through temperature-dependent measurements, we identify a charge diffusion activation energy of 0.18 eV, in good agreement with previously reported values and confirmed by DFT calculations. Together, these findings highlight the effectiveness and versatility of our method in understanding ionic charge carrier motion in microelectronics or nanoscale material systems.
... The Kelvin probe force microscopy (KPFM) method of surface potential measurement is widely used in the study of ferroelectric inorganic materials. The methodological and experimental results are described in a number of papers [1][2][3][4][5][6][7][8][9][10]. It should be added to the conclusions of these papers that the relaxation processes of the polarized domain structure should take into account the noted surface potential [11]. ...
Article
Full-text available
The values of the surface potentials of two sides of films of polyvinylidene fluoride, and its copolymers with tetrafluoroethylene and hexafluoropropylene, were measured by the Kelvin probe method. The microstructures of the chains in the surfaces on these sides were evaluated by ATR IR spectroscopy. It was found that the observed surface potentials differed in the studied films. Simultaneously, it was observed from the IR spectroscopy data that the microstructures of the chains on both sides of the films also differed. It is concluded that the formation of the surface potential in (self-polarized) ferroelectric polymers is controlled by the microstructure of the surface layer. The reasons for the formation of a different microstructure on both sides of the films are suggested on the basis of the general regularities of structure formation in flexible-chain crystallizing polymers.
... Speci cally, tr-EFM or tr-KPFM involve placing or slowly scanning a tip above a speci c location of interest while rapidly perturbing the sample at regular time intervals using an external stimulus (e.g., bias or light). So far, tr-EFM and tr-KPFM have been applied to a variety of systems, including photo-responsive materials [14][15][16] , semiconductors for transistors [17][18][19] or memristors 20 , with temporal resolution ranging from 10s to 1ps [21][22][23][24][25] . While these approaches provide a detailed record of how the sample's properties evolve at each location over time, they are slow and limited to processes that are cyclo-stationary, reversible, or at least reproducible within a short interval window. ...
Preprint
Full-text available
Understanding local dynamic charge processes is essential for developing advanced materials and devices, from batteries and microelectronics to medicine and biology. Continued progress relies on the ability to map electronic and ionic transport phenomena across different time and length scales, encompassing the intrinsic heterogeneities of the material itself (e.g., grain boundaries, domain walls, etc.). To address this challenge, we introduce high-speed Spiral-Scanning Kelvin Probe Force Microscopy (SS-KPFM), which combines sparse spiral scanning and image reconstruction via Gaussian process optimization. SS-KPFM enables functional sub-second imaging rates (≈ 3 fps), which represents a significant improvement over current state-of-the-art and several orders of magnitude over traditional KPFM methods. We apply it to study the spatiotemporal charge dynamics at a LaAlO 3 /SrTiO 3 planar device and charge injection and diffusion dynamics in polycrystalline TiO 2 thin films, providing full 2D Contact Potential Difference (CPD) maps of the surface charge dynamics in a fast and automated fashion.
... However, dual-closed loop and detection of the first-harmonic signal direct in open-loop measurements can be used to avoid this problem. 124,125 In perovskite solar cells, the frequency modulation (FM) KPFM have been used for quantitative surface potential measurements being less sensitive to crosstalk and stray electric field artifacts. 126 KPFM is also used in cross-section in which sample preparation is a critical point, similar to TEM measurements. ...
Article
The performance and stability of metal halide perovskite solar cells strongly depend on precursor materials and deposition methods adopted during the perovskite layer preparation. There are often a number of different formation pathways available when preparing perovskite films. Since the precise pathway and intermediary mechanisms affect the resulting properties of the cells, in situ studies have been conducted to unravel the mechanisms involved in the formation and evolution of perovskite phases. These studies contributed to the development of procedures to improve the structural, morphological, and optoelectronic properties of the films and to move beyond spin-coating, with the use of scalable techniques. To explore the performance and degradation of devices, operando studies have been conducted on solar cells subjected to normal operating conditions, or stressed with humidity, high temperatures, and light radiation. This review presents an update of studies conducted in situ using a wide range of structural, imaging, and spectroscopic techniques, involving the formation/degradation of halide perovskites. Operando studies are also addressed, emphasizing the latest degradation results for perovskite solar cells. These works demonstrate the importance of in situ and operando studies to achieve the level of stability required for scale-up and consequent commercial deployment of these cells.
Article
Full-text available
Mechanistic probing of surface potential changes arising from dynamic charge transport is the key to understanding and engineering increasingly complex nanoscale materials and devices. Spatiotemporal averaging in conventional heterodyne detection-based Kelvin probe force microscopy (KPFM) inherently limits its time resolution, causing an irretrievable loss of transient response and higher-order harmonics. Addressing this, we report a wavelet transform (WT)-based methodology capable of quantifying the sub-ms charge dynamics and probing the elusive transient response. The feedback-free, open-loop wavelet transform KPFM (OL-WT-KPFM) technique harnesses the WT's ability to simultaneously extract spatial and temporal information from the photodetector signal to provide a dynamic mapping of surface potential, capacitance gradient, and dielectric constant at a temporal resolution 3 orders of magnitude higher than the lock-in time constant. We further demonstrate the method's applicability to explore the surface-photovoltage-induced sub-ms hole-diffusion transient in bismuth oxyiodide semiconductor. The OL-WT-KPFM concept is readily applicable to commercial systems and can provide the underlying basis for the real-time analysis of transient electronic and electrochemical properties.
Article
Full-text available
Kelvin probe force microscopy is a scanning probe technique capable of mapping the local surface potential or work function on various surfaces with high spatial resolution. This technique can be realized on the basis of either an amplitude-sensitive method or a frequency-modulation method, which are sensitive to the electrostatic force and its gradient, respectively. We present a detailed experimental and theoretical study of the accuracy and resolution provided by the two methods, including the setup for the frequency-modulation technique. Au(111) with a submonolayer coverage of KCl serves as a test sample exhibiting extended sharply bounded areas that differ in work function by an amount well known from ultraviolet photoelectron spectroscopy. The influence of all relevant experimental parameters on the measurement is investigated. The experimental results are compared with the predictions of a numerical simulation based on a realistic model for the tip-sample geometry. Good agreement is found. The experimental analysis allows us to specify the lateral, vertical, and potential resolution that can be achieved with the two methods for a given tip size. Our work clearly proves that the frequency-modulation method is preferable in most applications because it (i) provides much higher lateral resolution, (ii) yields quantitative surface potential values on areas larger than the tip radius, and (iii) is little affected by variations of the tip-sample distance during topographic imaging.
Article
Full-text available
One way to improve imaging in dynamic force microscopy is to increase the cantilever resonance frequency. Higher frequencies require faster electronics. This work presents fast new digital electronics based on the principle of phase-locked loop techniques. First results show very high frequency resolution and very stable imaging.
Article
Full-text available
A method is presented for performing attractive‐mode force potentiometry with submillivolt accuracy and a typical spatial resolution of order 50 nm. The technique permits measurements to be made in air on specimens which may be passivated or oxidized, conducting or semiconducting, with virtually no sensitivity to oxide thickness or character. An initial demonstration is presented showing voltage measurements on a commercial operational amplifier.
Article
Kelvin probe force microscopy at air in the dark and under illumination by visible light was performed for electrodeposited porous ZnO films, which were sensitized by the indoline dye D149. A contact potential difference was measured that confirmed electron injection into the semiconductor matrix and their subsequent stabilization in trap states. A locally widely homogenous signal was observed speaking in favor of a well-crystallized sample. The rise and decay times of the photovoltage in the ms to s timescale could be analyzed by use of individual scan lines or scanned images, respectively. The results are discussed in the context of recent pump–probe spectroscopy results as well as such from photoelectrochemical experiments with such electrodes and implications will be deduced for their use in dye-sensitized solar cells.
Article
A scanning probe microscopy (SPM) based technique for probing local ionic and electronic transport and their dynamic behavior on the 10 ms - 10 s scale is presented. The time-resolved Kelvin Probe Force Microscopy (tr-KPFM) allows mapping surface potential in both space and time domains, visualizing electronic and ionic charge dynamics and separating underlying processes based on their time responses. Here, tr-KPFM is employed to explore the interplay of the adsorbed surface ions and bulk oxygen vacancies and their role in the resistive switching in the Ca-substituted bismuth ferrite thin film.
Article
The temporal evolution of the surface-potential distribution in the channel of pentacene based field-effect transistors is investigated during the charge reversal from the electron to the hole dominated operation. This measurement allows for the determination of the carrier density and electric field dependent hole mobility in the sub-threshold regime of the transistor.
Article
An ultrahigh vacuum scanning Kelvin probe force microscope (UHV SKPM) utilizing the gradient of electrostatic force, was developed based on an ultrahigh vacuum noncontact atomic force microscope (NC-AFM) capable of atomic level imaging, and used for simultaneous observation of contact potential difference (CPD) and NC-AFM images. CPD images of a Si(111) surface with Au deposited, clearly showed the potential difference in phases between 7×7 and 5×2 structures. When Ag was deposited as a submonolayer on the Si(111) 7×7 reconstructed surface, the atomic level lateral resolution was observed in CPD images as well as in NC-AFM topographic images. © 1998 American Institute of Physics.
Article
We present experimental data demonstrating the spatial resolution of an ultrafast junction mixing scanning tunneling microscope (JM-STM). The experiment uses a patterned metal-on-metal (Ti/Au) surface to establish electronic structure contrast on a short length scale. Our measurements achieve a spatio-temporal resolution of 20 nm–20 ps, limited only by the sample properties. The fine spatial resolution proves that the time-resolved signal is generated in the tunnel junction, indicating that atomic resolution should be possible in the JM-STM mode of operation. © 1998 American Institute of Physics.
Article
a b s t r a c t Kelvin probe force microscopy (KPFM) is a tool that enables nanometer-scale imaging of the surface potential on a broad range of materials. KPFM measurements require an understanding of both the details of the instruments and the physics of the measurements to obtain optimal results. The first part of this review will introduce the principles of KPFM and compare KPFM to other surface work function and potential measurement tools, including the Kelvin probe (KP), photoemission spectroscopy (PES), and scanning electron microscopy (SEM) with an electron beam induced current (EBIC) measurement system. The concept of local contact potential difference (LCPD), important for understanding atomic resolution KPFM, is discussed. The second part of this review explores three applications of KPFM: metallic nanostructures, semiconductor materials, and electrical devices.
Article
A half-wave slot in a metal sheet may be used as a resonant aerial in a manner similar to a half-wave dipole. The polar diagrams are the same, but the directions of vibration of the electric and magnetic fields are interchanged. If the slot is driven by means of a transmission line connected between opposite edges at the centre, the input impedance at resonance is about 485 ohms. The relation between resonant dipoles and resonant slots is an example of what in optics is known as Babinet's principle. This principle is established in the form required for electromagnetism and quantitative relations are deduced from it. Arrays of slots may be fed by transmission lines or wave guides to form linear or broadside aerials. Resonant gratings of slots may be used to polarize incident waves or act as band-pass filters. Single resonant slots may be used in wave guides in the same way as paralleltuned circuits are shunted across transmission lines. All devices adopted in connection with wire aerials have complements for slot aerials. Slots and dipoles may be combined to produce aerials with special polarization properties. Slots (with dielectric plugs) may be used in the skin of aircraft to provide dragless aerials. In addition, Babinet's principle may be used to reduce new problems to old ones whose solution is already known. These include problems connected with discontinuities in transmission lines, wave guides with corrugated sides, magnetron cavities, and the leakage of electromagnetic radiation through holes in metal walls. There is a wide variety of problems, both practical and theoretical for which it is always wise to consider the possibility of applying Bubinet's principle.