ArticlePDF Available

Status and Review of Two-Dimensional Carrier and Dopant Profiling Using Scanning Probe Microscopy

Authors:

Abstract

An overview of the existing two-dimensional carrier profiling tools using scanning probe microscopy includes several scanning tunneling microscopy modes, scanning capacitance microscopy, Kelvin probe microscopy, scanning spreading resistance microscopy, and dopant selective etching. The techniques are discussed and compared in terms of the sensitivity or concentration range which can be covered, the quantification possibility, and the final resolution, which is influenced by the intrinsic imaging resolution as well as by the response of the investigated property to concentration gradients and the sampling volume. From this comparison it is clear that, at present, none of the techniques fulfills all the requirements formulated by the 1997 Semiconductor Industry Association roadmap for semiconductors [National Technology Roadmap for Semiconductors (Semiconductor Industry Association, San Jose, CA, 1997)]. Most methods are limited to pn-junction delineation or provide a semiquantitative image of the differently doped regions. However, recent comparisons have shown that the techniques can provide useful information, which is not accessible with any other method. © 2000 American Vacuum Society.
Status and review of two-dimensional carrier and dopant profiling using
scanning probe microscopy
P. De Wolf,a) R. Stephenson, T. Trenkler, T. Clarysse, and T. Hantschel
IMEC, Kapeldreef 75, B-3001 Leuven, Belgium
W. Vandervorst
IMEC, Kapeldreef 75, B-3001 Leuven, Belgium and KU Leuven, INSYS, Kard. Mercierlaan 92,
B-3001 Leuven, Belgium
Received 28 March 1999; accepted 21 September 1999
An overview of the existing two-dimensional carrier profiling tools using scanning probe
microscopy includes several scanning tunneling microscopy modes, scanning capacitance
microscopy, Kelvin probe microscopy, scanning spreading resistance microscopy, and dopant
selective etching. The techniques are discussed and compared in terms of the sensitivity or
concentration range which can be covered, the quantification possibility, and the final resolution,
which is influenced by the intrinsic imaging resolution as well as by the response of the investigated
property to concentration gradients and the sampling volume. From this comparison it is clear that,
at present, none of the techniques fulfills all the requirements formulated by the 1997
Semiconductor Industry Association roadmap for semiconductors National Technology Roadmap
for Semiconductors Semiconductor Industry Association, San Jose, CA, 1997兲兴. Most methods are
limited to pn-junction delineation or provide a semiquantitative image of the differently doped
regions. However, recent comparisons have shown that the techniques can provide useful
information, which is not accessible with any other method. © 2000 American Vacuum Society.
S0734-211X0001201-4
I. INTRODUCTION
The 1997 U.S. Roadmap for Semiconductors from the
Semiconductor Industry Association SIAdefined the needs
for nanometer-scale measurement of carrier concentration
profiles for the next decade.1These needs are summarized in
Table I for a 0.25, 0.18, and 0.13
m technology. Clearly,
there is a demand for sub-10-nm resolution, combined with
sufficient sensitivity down to the 1e15 atoms/cm3leveland
high-quantification accuracy over a dynamic range of
1e15– 1e20 atoms/cm3. The need for such an extreme spa-
tial resolution as well as the applicability towards standard
devices has spurred the development of numerous two-
dimensional 2Dcarrier profiling tools. Until today, more
than 20 different methods have been developed for this pur-
pose. These techniques can roughly be divided into four cat-
egories: i2D techniques which are based on a widely used
one-dimensional 1Dtechnique such as secondary ion mass
spectrometry SIMS兲共Ref. 2兲关imaging SIMS,32D SIMS,4,5
tomography SIMS,6and lateral SIMS Ref. 7兲兴 or spreading
resistance profiling SRP.8iiElectron microscopy-based
techniques including field-effect scanning electron micros-
copy FE-SEM兲共Refs. 9 and 10and electron holography.11
iiiInverse modeling techniques.12 ivScanning probe mi-
croscopy SPM-based techniques. This article is limited to
SPM-based techniques. All SPMs are based on the ability to
position various types of probes in very close proximity with
extremely high precision to the sample under investigation.
These probes can detect electrical current, atomic and mo-
lecular forces, electrostatic forces, or other types of interac-
tions with the sample. By scanning the probe laterally over
the sample surface and performing measurements at different
locations, detailed maps of surface topography, electronic
properties, magnetic or electrostatic forces, optical character-
istics, thermal properties, or other properties can be obtained.
The spatial resolution which can be obtained is only limited
by the sharpness of the probe tip, the accuracy with which
the probe can be positioned, the condition of the surface
under study, and the nature of the force being detected. The
resolution can vary from a few angstroms to tens or hundreds
of nanometers. This extremely high spatial resolution makes
SPM the ideal candidate for a general applicable 2D carrier
profiling tool. SPM-based 2D dopant profiling methods in-
clude various scanning tunneling microscopy techniques
STM, dopant selective etching, scanning capacitance mi-
croscopy SCM, Kelvin probe force microscopy KPM,
scanning spreading resistance microscopy SSRM, and
scanning surface harmonic microscopy SSHM. In general,
all SPM-based 2D profiling techniques are being applied on
the cross section of the semiconductor structure under inves-
tigation. Earlier reviews of this topic have been given by
Subrahmanyan in 1992,13 Vandervorst et al.14 and Dagata
and Kopanski in 1995,15 Yu in 1996,16 and Vandervorst and
co-workers in 1997 Ref. 17and 1998.18 The basic charac-
teristics of the methods described in this article are summa-
rized in Table II. Table II displays the type of probe used and
the measured physical quantity.
II. PRINCIPLES OF THE DIFFERENT TECHNIQUES
A. Scanning tunneling microscopy STM
The scanning tunneling microscope is a very surface-
sensitive SPM technique which requires, however, a conduc-
aPresent address: Digital Instruments, 112 Robin Hill Road, Santa Barbara,
CA 93117; electronic mail: dewolf@di.com
361 361J. Vac. Sci. Technol. B 181, JanÕFeb 2000 0734-211XÕ2000Õ181Õ361Õ8Õ$15.00 ©2000 American Vacuum Society
tive surface, such that STM measurements on Si can nor-
mally not be operated in air due to the presence of the native
oxide. It is clear that if the native oxide is removed through
sample heating to relative high temperatures 1000 °C, the
dopant distribution is disturbed. Therefore, one has to use in
situ cleavage to generate a fresh sample cross section, which
is, however, only possible along certain crystal directions.19
The structure of interest must thus be oriented exactly in that
direction, limiting the flexibility of the technique. Addition-
ally, it is not always possible to determine the carrier profile
with respect to the mask edge and mask shape since the STM
cannot image the oxide position due to the requirement of a
conductive surface. Despite these basic limitations, cross-
sectional STM offers several methods for high-resolution
dopant profiling.
1. Dopant atom counting
The most direct STM-based dopant profiling method is to
simply count the number of dopant atoms appearing in or
nearthe surface atomic plane. Examples are given by
Johnson and co-workers,20–23 who clearly resolved indi-
vidual Be and Zndopant atoms in a cleaved 110surface
of Ga. In this work, the negatively charged dopant atoms are
attractive to holes and appear in the STM image as protru-
sions. From the size and shape of the features, dopants at the
surface and one atomic layer below the surface could be
distinguished from each other and from dopants further be-
low the surface. So far, dopant atom counting has not been
reported for Si substrates. In this context it is worth noting
that the capability of imaging individual dopant atoms is
TABLE I. Most important requirements for 2D carrier profiling as given in the SIA Roadmap for Semiconductors
Ref. 1.
Design rule 0.25
m 0.18
m 0.13
m
Accuracy of carrier
concentration 5% 5% 4%
Measurement repeatability 5% 5% 4%
Spatial resolution 5 nm 3 nm 2 nm
Carrier concentration range
atoms/cm3
1e15– 1e20 1e15– 1e21 1e15– 1e21
TABLE II. Summary of the different scanning probe microscopy techniques which can be used for 2D carrier
profiling of semiconductor devices. The ‘‘mode’’ reflects the scanning mode which is being used to control the
movement of the probe (NCnoncontact; Ccontact).
Technique Mode Probe Measured
quantity
Scanning tunneling microscopy/
spectroscopy
STM/STS
STM Metallic
needle No. doping atoms
IVspectra
Selective etchingatomic force
microscopy NC-AFM Ultrasharp Si Topography
after chemical
etch
Scanning capacitance microscopy/
spectroscopy
SCM/STS
C-AFM Metal-coated
Si
or metallic
Depletion
capacitance
CVspectra
Scanning spreading resistance
microscopy SSRM
C-AFM Diamond-
coated Si
or diamond
Electrical
resistance
IVspectra
Kelvin probe force microscopy
KPM
NC-AFM Metal-coated
Si
or metallic
Electrostatic
potential
electric field
Scanning surface harmonic
microscopy
SSHM
STM Metallic
needle with
microwave
cavity
Depletion
capacitance
362 De Wolf
et al.
: Status and review of 2D carrier profiling 362
J. Vac. Sci. Technol. B, Vol. 18, No. 1, JanÕFeb 2000
associated with the special properties of cleaved GaAs110
surfaces where the intrinsic surface states are outside the
bulk band gap.23 Although this method provides the ultimate
resolution, its sensitivity is rather poor. Since only the atoms
in the top two atomic layers are revealed, a high dopant
concentration is required to be able to determine the dopant
concentration adequately.
2. Scanning tunneling spectroscopy (STS) and
current imaging tunneling spectroscopy (CITS)
The work using STM as a dopant profiler has primarily
been focused on pn-junction delineation by detecting differ-
ences in tunneling current characteristics for n- and p-type
material. Feenstra and co-workers carried out the first de-
tailed imaging and spectroscopic studies of GaAs pn junc-
tions, observing electronically induced topography.24–26
Current–voltage spectroscopy allowed the n-type, p-type,
and depleted regions to be identified unambiguously. The
STM current dependence on dopant type and concentration
within semiconductors is due to tip-induced bandbending at
the surface.27
Kordic and co-workers performed the first cross-sectional
STM studies of Si pn junctions exposed by cleaving in air
and UHV.28–30 The location of the junctions could be re-
solved to within 30 nm using a potentiometric technique in
UHV scanning tunneling potentiometry STP兲兴: A forward
or reverse bias was applied across the pn junction, and dif-
ferences in tunneling current measured in p-type, depleted,
and n-type regions for various bias voltages applied to the pn
junctions then revealed the electronic structure of the biased
junctions.
Yu and co-workers31,32 used current imaging tunneling
spectroscopy CITS. In this technique a constant-current to-
pographic scan with the current stabilized at a fixed value I0
for a specific voltage V0is performed over the sample sur-
face and, at each point, a current–voltage spectrum is mea-
sured. Variations in electronic structure across the sample
surface produce corresponding variations in the current–
voltage spectra; these spatial variations can be revealed by
plotting the current measured at specific bias voltages other
than V0—so-called current images. CITS under UHV condi-
tions on cleaved, hydrogen-passivated cross-sectioned Si
metal–oxidesemiconductor MOSstructures made it pos-
sible to image the source and drain junction profiles with a
spatial resolution on the order of 10 nm.32 Because the
current–voltage spectra differed significantly for the p-type,
n-type, and depleted regions, current images generated from
the spatially resolved tunneling spectra were able to reveal
the profiles of the pn junctions. A problem to extend this
work further to quantitative dopant profile information, is the
dependence of the tunneling current on the Fermi level and
bandbending rather than on the carrier concentration directly.
In addition, the surface disorder problems form a significant
limitation for the STM approach.
In conclusion, the different STM-based cross-sectional
carrier or dopingprofiling methods have low performance
on Si. This difficulty can be ascribed to the difficult cross-
section preparation, combined with the fact that the STM is a
very surface-sensitive technique. First, Si 001wafers are
more difficult to cleave than III–V wafers; and second, the
as-cleaved 110cross-sectional surface is atomically disor-
dered with its electronic structure dominated by dangling-
bond states. Despite these complications, 2D junction delin-
eation and qualitative 2D nanometer-scale carrier profiling of
cross-sectioned Si-based structures have been achieved by
several research groups. The applicability of STM or scan-
ning tunneling spectroscopy STS兲兴 as a quantitative carrier
profiling tool is complicated and, therefore, limited.
B. Dopant selective etching or staining
The method of chemical staining or etching of doped Si
layers has been known since the early days of semiconductor
processing. Staining techniques use selective deposition of a
metal such as Cu, Au, Ag, or Pt on one side of the junction
by an electrochemical displacement reaction from a metal-
ion-based solution. Etching techniques use mixtures of HF
and an oxidizing agent to preferentially etch regions with a
high carrier concentration. The impurity-sensitive etching so-
lutions typically consist of a mixture of HF, HNO3, and ei-
ther H2OorCH
3COOH. The sample topography after the
staining or etching step is a measure for the 2D carrier profile
and can be imaged using transmission electron microscopy
TEM,32–34 SEM,35 STM,36,37 or atomic force microscopy
AFM.38 In a final step, the measured topography profiles
have to be converted to electrical carrier distributions.
Among the advantages of using AFM compared with STM
in performing topographic measurements on selectively
etched cross sections is that AFM makes it possible to mea-
sure the surrounding structures oxides, metals, etc.since
this is not possible with the STM, which requires a conduct-
ing sample surface. Also, AFM is less sensitive to the de-
tailed electronic structure of the etched surface and to pos-
sible contamination of the sample or tip surfaces. Both the
STM and AFM method can suffer from tip–sample convo-
lution leading to an incorrect image of the etched region. The
regions with a high carrier concentration gradient are particu-
larly sensitive to this effect.
Independently of the technique which is used to measure
the sample topography, the accuracy of etching and staining
techniques is always limited by the reproducibility of factors
such as surface preparation, the concentration of the staining
or etching solution, etching time, the volume and the agita-
tion of the solution on the sample, temperature, and light
illumination. The effects of these factors can be minimized
by careful sample preparation and precise control of the etch-
ing factors. In this context, different approaches have been
used by several groups. A promising method to obtain longer
etching times and a better general control of the etching
makes use of an electrochemical etching procedure with po-
tentiostatic control.39
An important drawback of the etching techniques until
now is that analysis suffers from a poor understanding of the
etching process and the correlation between carrier concen-
tration and the observed topography. Two approaches have
363 De Wolf
et al.
: Status and review of 2D carrier profiling 363
JVSTB-MicroelectronicsandNanometer Structures
been employed so far for this calibration. The first calibra-
tion method uses a 1D doping profile adjacent to the 2D area
of interest. The 1D doping profile may be measured by SIMS
or SRP or be numerically simulated. After etching, the 1D
etched profile is characterized and related to the known 1D
carrier or dopingdistribution. A 2D doping map is obtained
by assuming that the dependence of the etching rate on car-
rier concentration obtained for the 1D region is also correct
for the complete 2D area. The second calibration method
includes direct measurement of the etching rate as a function
of carrier concentration using homogeneously doped bulk or
epitaxially grown Si samples. Again, it is assumed that under
similar etching conditions temperature, concentrations,
light, time, etc.the etching rate for a particular spot depends
exclusively on the carrier concentration at this point. How-
ever, it has been found that the etching also depends on the
carrier concentration gradient.40
In summary, the applicability of the selective etching
method to 2D carrier profiling is limited because of ipoor
understanding of the etching process, iipoor control of the
etching conditions, and iiilack of a good general quantifi-
cation procedure. Despite these limitations, the method re-
mains valuable for fast qualitative analysis.
C. Scanning capacitance microscopy SCM
In the scanning capacitance microscope, the sample or
the metallic AFM tipis covered with a thin dielectric layer,
such that the tip–sample contact forms a metalinsulator
semiconductor MIScapacitor, whose capacitance–voltage
(CV) behavior is determined by the local carrier concen-
tration of the semiconductor sample. If no dielectric layer is
used, the tip–sample contact forms a metalsemiconductor
MSstructure and one has a so-called Schottky contact
SCM.41,42 By monitoring the capacitance variations as the
probe scans across the sample surface, one can measure a 2D
carrier concentration profile. Since the total tip–sample ca-
pacitance is large compared to the capacitance variations due
to different carrier concentrations, one usually measures the
capacitance variations and not the absolute capacitance val-
ues. Note that no signal is measured if the probe is posi-
tioned over a dielectric or metallic region since these regions
can not be depleted. Most SCMs are based on contact-mode
AFM with a conducting tip, and an essentially independent
capacitance measurement in parallel.43 The capacitance be-
tween the tip and sample is measured by using a high-
frequency capacitance sensor, based on a 915 MHz oscillator
driving a resonance circuit which is tuned in part by the
external capacitance to be measured. The capacitance detec-
tion limit is as small as 1e-19 F in a 1 kHz bandwidth trans-
lating into a noise level of 3e-21F/
Hz). An extensive re-
view of SCM for 2D dopant profiling is given by Williams.44
The SCM is usually operated in one of the following two
modes.
1. Differential-capacitance (open-loop) mode
In the open-loop mode, an ac bias typically, 0.2–2 V,
10–100 kHzis superimposed on a dc sample bias 2–2
V, while the tip is at dc ground. The ac bias can alternately
deplete and accumulate the semiconductor surface region.
The modulated surface capacitance changes Cunder the
probe tip are registered using a lock-in technique, simulta-
neously with the topographical data, while the probe is
scanned across the surface. When using large ac voltages
several volts, this setup measures Cacross the entire
CVcurve, and is almost independent on shifts in the flat-
band voltage caused by oxide or surface charges. When
smaller ac bias voltages are used, the differential capacitance
dC/dV is measured.
2. Closed-loop mode
In the closed-loop mode, the magnitude of the ac bias
voltage applied to the sample is adjusted by a feedback loop
to maintain a constant capacitance change as the tip is
scanned across the sample at constant force.45,46 The
feedback-controlled magnitude of the ac bias voltage is re-
corded. The main advantage of the closed-loop mode is the
fact that the capacitance and, consequently, the depletion
width is kept constant, whereas the depletion width might
become very large 1
mfor lowly doped regions in the
differential capacitance mode, leading to a loss in spatial
resolution.
Since in SCM the measured capacitance signal is propor-
tional to the tip interaction area, shrinkage of the tip size will
improve the spatial resolution, but also reduce the sensitivity
of the capacitance measurement. A dynamic range of
1e14– 1e20 has been demonstrated on a special calibration
structure.47 The sensitivity is a function of the carrier con-
centration and can be increased by reducing the thickness of
the oxide layer. The response of SCM is not necessarily
monotonic with concentration but does depends on the ap-
plied experimental conditions as well.48 Several groups have
published qualitative 1D and 2D Refs. 49–51open- or
closed-loop SCM images, which were compared with TCAD
simulations and conventional 1D carrier profiling results.
The resolution of these images is on the order of 10–20
nm.52
Much of the recent work related to SCM is focused on the
theoretical interpretation and quantitative conversion of the
measured signals into carrier profiles. In an ideal flatMOS
capacitor, the dopant concentration is easily extracted from
the variation of capacitance with voltage. The situation is
more complex for SCM since there are stray fields between
the probe shaft and the sample surface, the sample has a
nonconstant carrier concentration, and the quality of the sur-
face is largely unknown surface charges, contaminants, ox-
ide quality, etc.. Several models were presented to set up a
quantification procedure. In the simplest model, the quantifi-
cation of the SCM images is achieved using SIMS or SRP
data in conjunction with contour mapping software. Hereby,
it is assumed that the dependence of the SCM signal on
carrier concentration obtained for the 1D line is also correct
for the complete 2D area.
In a quasi-1D analytical model, the tip is modeled as a
metallic sphere placed in an insulating dielectric medium
364 De Wolf
et al.
: Status and review of 2D carrier profiling 364
J. Vac. Sci. Technol. B, Vol. 18, No. 1, JanÕFeb 2000
near a Si surface with a sphere–Si gap just equal to the
experimental measured oxide thickness.49 The Si surface is
divided into annular regions. The insulator capacitance be-
tween the sphere and each annular region, and the Si deple-
tion capacitance for each annular region are calculated and
summed to give the total tip–sample capacitance. This ap-
proximate analytical model provides a means to rapidly cal-
culate the CVrelation as a function of tip radius, dielectric
constant, gap distance, and carrier density. Note that a con-
stant carrier density is assumed in these calculations, the ef-
fect of a carrier gradient in the lateral directionis not yet
included.
In an alternative model the CVrelation is calculated
using the solution of the three-dimensional 3DPoisson
equations for various tip–sample bias voltages, tipsample
gap distances, carrier concentrations, and oxide
thickness.49,53,54 The probe is modeled as a cone with a hemi-
spherical tip end to include stray-field effects. The quality of
the sample surface, although known to be an important pa-
rameter, was not included in the calculations. Algorithms
based on interpolation within the results of this database
have then been developed, to convert a measured SCM pro-
file into a quantitative carrier profile. Kopanski and
co-workers49,53 treat each measured capacitance point inde-
pendently and thus ignore the effect of carrier gradients,
whereas Yu, Griffin, and Plummer54 include the effect of the
dopant gradient in their simulations.
It needs to be pointed out that the behavior of the SCM
signal in the open- or closed-loop modeat pn junctions is
not well understood and is not yet included in the calcula-
tions, although pronounced effects have been observed.49
Calculations and experiments have shown that the tip–
sample bias ac and dchas a pronounced effect on the pres-
ence, position, and extent of extra contour lines in the vicin-
ity of the junction, complicating precise junction delineation
and data quantification at pn junctions.55,56 Several groups
have proposed and implemented promising approaches to
overcome this problem. Timp et al. use heavily doped Si tips
rather than metal-coated tips.57 In this case, not only the
sample, but also the tip itself, is being depleted and accumu-
lated. The resulting SCM images persist over a wide tip–
sample bias range. The delineation of the highly doped re-
gions such as the poly-Si gateand the metal regions
becomes now clearer and junction positions appear less in-
fluenced by the bias voltage used. However, now the exact
concentration of the tip material enters as an extra parameter
in the quantification in particular, in those cases where car-
rier concentration in the sample and tip are equivalent, inter-
pretation becomes less transparent. Edwards et al. present
another approach:58 scanning capacitance spectroscopy
SCS. In SCS the CVcurve is measured for every position
of the probe in the 2D profile. The shape of the CVcurve
allows one to distinguish between n-type, p-type, and deple-
tion regions and facilitates data quantification. Both
approaches—Si tips, and SCS—result in easier data interpre-
tation, in particular, at pn junctions. However, quantification
of the data into carrier concentration values still requires 3D
simulations in order to determine the exact impact of carrier
concentration gradients and pn junctions on the observed
SCM images. In conclusion, the SCM is a promising tool for
quantitative 2D carrier profiling with nanometer resolution.
The spatial resolution 10–20 nmand dynamic range
(1e15– 1e20 atoms/cm3) are good. The major challenges in
SCM include: surface preparation including the formation of
a good dielectric oxide,59 and the quantification and calibra-
tion methodologies. At present, the technique is widely used
in a qualitative manner and only limited by a lack of a gen-
erally applicable quantification routine.
D. Scanning surface harmonic microscopy SSHM
The SSHM consists of a STM with a microwave cavity,
in which a microwave signal is applied across the tip–sample
tunneling gap.60 The nonlinear tip–sample MIS capacitance
Cresults in higher-order harmonics in the tunneling current.
The second- and third-harmonic signals are proportional to
the first and second derivatives of C. The driving frequency
is chosen such that the second- or third-harmonic frequency
corresponds to the resonance frequency of the cavity typi-
cally, 2.7 GHz, making it detectable.61,62 The capacitance C
is a measure for the local active carrier concentration in the
semiconductor and the insulator quality in the same way as
in the SCM technique. Bourgoin, Johnson, and Michel have
applied the SSHM technique to delineate the qualitative car-
rier profile of Si pn junctions with 5 nm resolution.61 In
general, the signal-to-noise ratio is lower compared to the
SCM method and quantification of the data is faced with
similar problems as the SCM technique.
E. Kelvin probe force microscopy KPM
Kelvin probe force microscopy and the related scanning
Maxwell-stress microscopy, are high-resolution and highly
sensitive potential imaging methods.63 A conductive tip is
scanned in the noncontact AFM mode while an ac voltage
frequency fis applied to it. This voltage results in an elec-
trostatic field which causes an oscillation of the cantilever at
the same frequency. This force disappears when the dc po-
tential difference between tip and sample is zero. Thus, by
observing the amplitude of the cantilever oscillation at fre-
quency fby a lock-in technique, and nulling it by changing
the dc bias voltage on the tip, the sample’s surface potential
can be measured. The maximum sensitivity is obtained when
fcorresponds to one of the resonance frequencies of the can-
tilever. In order to separate the height-control signal and the
voltage signal, the first cantilever resonance peak is usually
employed for the tip height control while fis taken equal to
the second resonance peak on a dual resonant probe.63 This
mode was used to measure the 2D potential distribution in-
side Si device structures.64–66 The measured electrochemical
potential difference between the probe tip and sample surface
is dependent on the carrier concentration-related work-
function difference, and can thus be used as a measure for
the local carrier concentration, although the sensitivity is
limited. This mode of KPM has been applied successfully for
qualitative 2D carrier profiling of Si structures.66,67 The tech-
365 De Wolf
et al.
: Status and review of 2D carrier profiling 365
JVSTB-MicroelectronicsandNanometer Structures
nique is sensitive to changes in carrier concentration from
1e15 to 1e20 atoms/cm3with a spatial resolution of about
100 nm. However, the sensitivity to small concentration
changes and the application towards quantitative profiling
are limited by surface charges on the sample and calibration
of the KPM technique against absolute doping concentration
standards remains to be demonstrated.
F. Scanning spreading resistance microscopy
SSRM
In SSRM the electrical resistance is measured between the
conductive probe tip and a large current-collecting back con-
tact while the probe is scanned in the contact mode across
the cross section of the Si device. When the applied force
exceeds a certain threshold force, the measured resistance is
dominated by the spreading resistance.68 As for the conven-
tional 1Dspreading resistance profiling method, the
spreading resistance depends inverse proportionally on the
local carrier concentration underneath the probe–silicon con-
tact. If one uses small bias voltages about 100 mV, the
displacements as observed in SCM imaging are minimized.
Contact potential differences between tip and surface can
possibly alter this situation. Junction positions can be as-
signed to a single point as their positions show up as a peak
in the resistance profile. On Si structures, high forces typi-
cally, a few
Nare required in order to penetrate the native
oxide and to establish a stable electrical contact. As standard
AFM probes deform at these high forces, one needs to use
doped diamond or diamond-coated Si probes. If lower forces
are being used, the measured resistance is no longer domi-
nated by the spreading resistance but by the contact resis-
tance of the tip–sample contact. In this case, qualitative car-
TABLE III. Intercomparison of two-dimensional doping Dand carrier Cprofiling methods (NAnot available).
Method Ref. Resol.
nm
Range
cm3
Conc.
resol. D/
CQuanti-
fiable Comments and problems
SPM techniques
SCM 43–5910 1e15– 1e20 Power C Limited Uncertainties at junctions, poor
quantification procedure
SSHM 60–625 NA Power C No No quantification procedure
STM-atom
counting 20–23Atomic 1e18– 1e20 Linear D Yes Only on GaAs, not on Si
STM-STS/
CITS 24–26
31,32
10 NA Log. C Limited Only junction delineation and type n
or pidentification
STM-STP 27–3010 NA Limited C Limited Only junction delineation
KPM 66,67100 1e15 1 e20 Limited C Limited Poor quantification procedure, stray-
fields limit the resolution
SSRM 68–7320 1e15– 1e20 Linear C Yes Availability diamond probes
Chemical etch
AFM/STM 37–3910–20 1e171e20 Limited C Limited Difficult to quantify, poor
reproducibility
1D-based techniques
Imaging SIMS 3100 NA Linear D Yes Sensitivity limited by target volume
2D-SIMS 4,530–50 1e16– 1e21 Linear D Yes Special structures required
2D-tomography
SIMS 650 NA Linear D Yes Special structures required, complex
sample preparation
Lateral SIMS 75–10 Dose Linear D Yes Only the lateral dose distribution is
measured
2D-SRP 8100 1e15 1 e21 Linear C Yes Special structures required
EM techniques
Chemical etch
SEM/TEM 32–3520 1e17– 1e21 Limited C Limited Only qualitative
FE-SEM 9,1010–20 4e16– 1e21 Limited Limited Robust model for quantification is not
available
E-holography 111–10 1e17–1e20 Limited C Limited
Inverse modeling techniques
Inv. modeling
with CV
12NA NA C Yes Resolution and accuracy are unknown,
long calculation times
366 De Wolf
et al.
: Status and review of 2D carrier profiling 366
J. Vac. Sci. Technol. B, Vol. 18, No. 1, JanÕFeb 2000
rier profiling remains possible, although the sensitivity to
small changes in the carrier concentration is lost.69
The quantification of the SSRM resistance data into car-
rier concentration values is possible using an n- and p-type
calibration curve in combination with an algorithm which
corrects for the effect of nearby layers with different carrier
concentrations.70 This quantification method is simple and
fast and does not require one to measure the 1D in-depth
profile by SIMS or SRP. The SSRM method has been ap-
plied on various InP Ref. 71and Si Refs. 68, 72, and 73
device structures. These measurements include fully quanti-
tative 2D profiles of both nMOSFETs and pMOSFETs
metal–oxidesemiconductor field-effect transistorsand al-
low extraction of values for effective channel lengths and
two-dimensional diffusion effects depending on gate
lengths.73 Reasonable agreement is also obtained when com-
pared to data extracted from device characteristics.74 The
data indicate a dynamic range of 1e15– 1e20 atoms/cm3,a
spatial resolution of about 25 nm, and also show that SSRM
can be applied on arbitrary device structures. At present, the
SSRM method is mainly limited by the availability of con-
ductive diamond or diamond-coated probes.71
III. INTERCOMPARISON
Various features of the methods described above are sum-
marized in Table III. The SPM-based 2D carrier profiling are
compared with alternative methods based on 1D techniques,
electron microscopy, or inverse modeling. The most impor-
tant features are spatial resolution in nanometers, dynamic
range in atoms/cm3, sensitivity defined here as the ratio of
the change in the instrument response to a corresponding
change in carrier concentration, quantification ability, and
applicability to real devices without needing special test
structures. The listed resolution and dynamic range corre-
spond to the best specifications found in the literature. Note
that different definitions for spatial resolution and sensitivity
or dynamic rangeare used by different groups, complicat-
ing this intercomparison. If no value is found in the litera-
ture, the specification is labeled as ‘‘not available’’ NA.
The ability to transform the measured raw data values into a
fully quantitative 2D doping or carrier profile is labeled as
‘‘yes,’’ ‘‘no,’’ or ‘‘limited.’’ The concentration resolution is
labeled as ‘‘linear’’ if the instrument response is proportional
to the carrier concentration, ‘‘logarithmic’’ if the instrument
response is proportional to the logarithm of the carrier con-
centration, and ‘‘power’’ if there is a power-law relation be-
tween the carrier concentration and the instrument response,
or ‘‘limited.’’
Based on this intercomparison, it is clear that none of the
available techniques fulfills all the requirements formulated
in Table I and can at the same time be applied on any arbi-
trary semiconductor structure. Most techniques are limited to
pn-junction delineation or provide a qualitative image of the
differently doped regions. Whereas all SPM-based methods
can be applied on arbitrary devices, all 1D-based techniques
require a special test structure, making them less flexible as
compared to the SPM-based methods.
The authors believe that although the SSRM and the SCM
techniques are still in a development or optimization stage,
they are the leading candidates to become a general appli-
cable 2D carrier profiling tool. Both techniques have a reso-
lution close to the values asked for, can be applied on arbi-
trary structures, and have the required dynamic range.
Quantification accuracy and reproducibility are still issues
which need to be determined with much more detail.
IV. CONCLUSIONS
At present there are numerous SPM-based methods for
2D carrier profiling of Si and otherdevice structures. These
include several scanning tunneling microscopy modes, scan-
ning capacitance microscopy, Kelvin probe microscopy,
scanning spreading resistance microscopy, and dopant selec-
tive etching. All of these methods can be directly applied on
the cross section of an arbitrary device and do not require
special test structures. The methods differ in the range of
carrier concentrations which can be mapped, the spatial reso-
lution, and the quantification possibility. The features of the
different methods can be summarized as follows:
iThe different STM-based methods have a rather low
performance on Si due to the difficult cross-section prepara-
tion, combined with the fact that the STM is a very surface-
sensitive technique. STM-based carrier profiling is, there-
fore, limited to 2D junction delineation and qualitative 2D
imaging. At present, the applicability of STM as a quantita-
tive carrier profiling tool is not possible.
iiThe applicability of the selective etching method fol-
lowed by AFM imagingis limited because of a poor under-
standing of the etching process, a poor control of the etching
conditions, and a lack of a general quantification procedure.
However, the etching method is a very valuable tool for fast
qualitative analysis of the 2D carrier profile.
iiiBoth the SCM and SSRM are promising tools for
quantitative 2D carrier profiling with nanometer resolution.
The spatial resolution 10–20 nmand dynamic range
(1e15– 1e20 atoms/cm3) are good. Some further research
and development is required such that these methods fulfill
the requirements set by the 1997 SIA roadmap for semicon-
ductors. This future work includes the development and fab-
rication of reliable probes for SSRM, and the development of
routines to interprete and convert the SCM images into quan-
titative carrier concentration values.
1National Technology Roadmap for Semiconductors Semiconductor In-
dustry Association, San Jose, CA, 1997.
2M. G. Dowsett, Proceedings of the SIMS XI Conference, edited by G.
Gillen, R. Larea, J. Bennet, and F. Stevie Wiley, Chichester, 1997,p.
157.
3S. R. Bryan, W. S. Woodward, R. W. Linton, and D. P. Griffis, J. Vac.
Sci. Technol. A 3, 2102 1985.
4M. G. Dowsett and G. A. Cooke, J. Vac. Sci. Technol. B 10,3531992.
5V. A. Ukraintsev, P. J. Chen, J. T. Gray, C. F. Machala, L. K. Magel, and
M.-C. Chang, J. Vac. Sci. Technol. B, these proceedings.
6S. H. Goodwin-Johansson, M. Ray, Y. Kim, and H. Z. Massoud, J. Vac.
Sci. Technol. B 10,3691992.
7R. Van Criegern, F. Jahnel, R. Lange-Gieseler, P. Pearson, G. Hobler, and
A. Simionescu, J. Vac. Sci. Technol. B 16, 386 1998.
8V. Privitera, W. Vandervorst, and T. Clarysse, J. Electrochem. Soc. 140,
262 1993.
367 De Wolf
et al.
: Status and review of 2D carrier profiling 367
JVSTB-MicroelectronicsandNanometer Structures
9R. Turan, D. D. Perovic, and D. C. Houghton, Appl. Phys. Lett. 69, 1593
1996.
10D. Venables, H. Jain, and D. C. Collins, J. Vac. Sci. Technol. B 16, 362
1998.
11W. D. Rau, F. H. Baumann, H. H. Vuong, B. Heinemann, W. Hoppner, C.
S. Rafferty, H. Rucker, P. Schwander, and A. Ourmazd, Tech. Dig. Int.
Electron Devices Meet., 713 1998.
12N. Khalil, J. Faricelli, C. L. Huang, and S. Selberherr, J. Vac. Sci. Tech-
nol. B 14,2241996.
13R. Subrahmanyan, J. Vac. Sci. Technol. B 10, 358 1992.
14W. Vandervorst, T. Clarysse, P. De Wolf, L. Hellemans, J. Snauwaert, V.
Privitera, and V. Raineri, Nucl. Instrum. Methods Phys. Res. B 96, 123
1995.
15J. A. Dagata and J. J. Kopanski, Solid State Technol. 38,911995.
16E. T. Yu, Mater. Sci. Eng. 17,1471996.
17W. Vandervorst, T. Clarysse, P. De Wolf, T. Trenkler, T. Hantschel, and
R. Stephenson, Future Fab. Int. 4,2871997.
18W. Vandervorst, T. Clarysse, P. De Wolf, T. Trenkler, T. Hantschel, R.
Stephenson, and T. Janssens, in Proceedings of the International Work-
shop on Semiconductor Characterization: Present Status and Future
Needs, edited by D. Seiler AIP, New York, 1998.
19Y.-C. Kim, M. J. Nowakowaski, and D. N. Seidman, Rev. Sci. Instrum.
67, 1922 1996.
20M. B. Johnson, O. Albrektsen, R. M. Feenstra, and H. W. M. Salemink,
Appl. Phys. Lett. 63, 2923 1993.
21M. B. Johnson, H. P. Meier, and H. W. M. Salemink, Appl. Phys. Lett.
63, 3636 1993.
22H. W. M. Salemink, M. B. Johnson, O. Albrektsen, and P. Koenraad,
Solid-State Electron. 37, 1053 1994.
23K.-J. Chao, A. R. Smith, A. J. McDonald, D.-L. Kwong, B. G. Streetman,
and C.-K. Shih, J. Vac. Sci. Technol. B 16, 453 1998.
24R. M. Feenstra, E. T. Yu, M. Woodall, P. D. Kirchner, C. L. Lin, and G.
D. Pettit, Appl. Phys. Lett. 61,7951992.
25S. Gwo, A. R. Smith, K.-J. Chao, C. K. Shih, K. Sadra, and B. G. Street-
man, J. Vac. Sci. Technol. A 12, 2005 1994.
26R. M. Silver, J. A. Dagata, and W. Tseng, J. Vac. Sci. Technol. A 13,
1705 1995.
27R. M. Feenstra and J. A. Stroscio, J. Vac. Sci. Technol. B 5, 923 1987.
28S. Kordic, E. J. Van Loenen, D. Dijkkamp, A. J. Hoeven, and H. K.
Moraal, J. Vac. Sci. Technol. A 8, 549 1990.
29S. Kordic, E. J. Van Loenen, and A. J. Walker, IEEE Electron Device
Lett. 12,4221991.
30M. B. Johnson and J.-M. Halbout, J. Vac. Sci. Technol. B 10, 508 1992.
31E. T. Yu, M. B. Johnson, and J.-M. Halbout, Appl. Phys. Lett. 61, 201
1992.
32E. T. Yu, K. Barmak, P. Ronsheim, M. B. Johnson, P. McFarland, and
J.-M. Halbout, J. Appl. Phys. 79,21151996.
33R. Alvis, B. Mantiply, and M. Young, J. Vac. Sci. Technol. B 14, 452
1996.
34S. S. Neogi, D. Venables, Z. Ma, and D. M. Maher, J. Vac. Sci. Technol.
B16,4711998.
35L. Gong, S. Petersen, L. Frey, and H. Ryssel, Nucl. Instrum. Methods
Phys. Res. B 96,1331995.
36T. Takigami and M. Tanimoto, Appl. Phys. Lett. 58, 2288 1991.
37M. Tanimoto and T. Takigami, Ultramicroscopy 4244, 1381 1992.
38G. Neubauer, M. Lawrence, A. Dass, and T. J. Johnson, Mater. Res. Soc.
Symp. Proc. 286,2831992.
39T. Trenkler, W. Vandervorst, and L. Hellemans, J. Vac. Sci. Technol. B
16,3491998.
40V. A. Ukraintsev, R. McGlothin, M. A. Gribelyuk, and H. Edwards, J.
Vac. Sci. Technol. B 16,4761998.
41M. Morooka and S. Fukasako, Jpn. J. Appl. Phys., Part 1 35, 3686 1996.
42Y. Li, J. N. Nxumalo, and D. J. Thomson, in Proceedings of the 4th
International Workshop on Ultra Shallow Junctions, Raleigh, NC, edited
by M. Current, M. Kump, and G. McGuire 1997, p. 63.1.
43Y. Huang and C. C. Williams, J. Vac. Sci. Technol. B 12, 369 1994.
44C. C. Williams, Annu. Rev. Mater. Sci. 29,4711999.
45Y. Huang, C. C. Williams, and J. Slinkman, Appl. Phys. Lett. 66,344
1995.
46Y. Huang, C. C. Williams, and M. A. Wendman, J. Vac. Sci. Technol. A
14, 1168 1996.
47T. Clarysse, M. Caymax, P. De Wolf, T. Trenkler, W. Vandervorst, J. S.
McMurray, J. Kim, C. C. Williams, J. G. Clark, and G. Neubauer, J. Vac.
Sci. Technol. B 16,3941998.
48R. Stephenson, A. Verhulst, P. De Wolf, M. Caymax, and W. Vander-
vorst, Appl. Phys. Lett. 73,25971998.
49J. J. Kopanski, J. F. Marchiando, D. W. Berning, R. Alvis, and H. E.
Smith, J. Vac. Sci. Technol. B 16, 339 1998.
50J. S. McMurray, J. Kim, C. C. Williams, and J. Slinkman, J. Vac. Sci.
Technol. B 16, 344 1998.
51W. Timp, M. L. O’Malley, R. N. Kleiman, and J. P. Garno, Tech. Dig.
Int. Electron Devices Meet., 555 1998.
52Y. Huang, C. C. Williams, and H. Smith, J. Vac. Sci. Technol. B 14,433
1996.
53J. F. Marchiando, J. J. Kopanski, and J. R. Lowney, J. Vac. Sci. Technol.
B16, 463 1998.
54G. Yu, P. Griffin, and J. Plummer, Tech. Dig. Int. Electron Devices
Meet., 717 1998.
55R. N. Kleiman, M. L. O’Malley, F. H. Baumann, J. P. Garno, and G. L.
Timp, Tech. Dig. Int. Electron Devices Meet., 671 1997.
56C. J. Kang, C. K. Kim, J. D. Lera, Y. Kuk, K. M. Mang, J. G. Lee, K. S.
Suh, and C. C. Williams, Appl. Phys. Lett. 71,15461997.
57W. Timp, M. L. O’Malley, R. N. Kleiman, and J. P. Garno, Tech. Dig.
Int. Electron Devices Meet., 555 1998.
58H. Edwards, R. Mahaffy, C. K. Shih, R. McGlothlin, R. San Martin, M.
Gribelyuk, R. S. List, and V. A. Ukraintsev, Appl. Phys. Lett. 72, 698
1998.
59V. A. Ukraintsev, F. R. Potts, R. M. Wallace, L. K. Magel, H. Edwards,
and M.-C. Chang, Proceedings of the International Conference on Char-
acteristics and Metrology for ULSI Technology 1998.
60B. Michel, W. Mitzutani, R. Schierle, A. Jarosh, W. Knop, H. Benedick-
ter, W. Bachtold, and H. Rohrer, Rev. Sci. Instrum. 63, 4080 1992.
61J.-P. Bourgoin, M. B. Johnson, and B. Michel, Appl. Phys. Lett. 65, 2045
1994.
62F. Bordoni, L. Yinghua, B. Spataro, F. Feliciangeli, F. Vasarelli, G. Car-
darilli, B. Antonini, and R. Scrimaglio, Meas. Sci. Technol. 6, 1208
1995.
63M. Nonnenmacher, M. P. O’Boyle, and H. K. Wickramasinghe, Appl.
Phys. Lett. 58, 2921 1991.
64A. Kikukawa, S. Hosaka, and R. Imura, Appl. Phys. Lett. 66,3510
1995.
65A. Kikukawa, S. Hosaka, and R. Imura, Rev. Sci. Instrum. 67, 1463
1996.
66M. Tanimoto and O. Vatel, J. Vac. Sci. Technol. B 14, 1547 1996.
67A. K. Henning, T. Hochwitz, J. Slinkman, J. Never, S. Hoffmann, P.
Kaszuba, and C. Daghlian, J. Appl. Phys. 77,18881995.
68P. De Wolf, T. Clarysse, W. Vandervorst, L. Hellemans, Ph. Niedermann,
and W. Ha
¨nni, J. Vac. Sci. Technol. B 16,4011998.
69J. N. Nxumalo, D. T. Shimizu, and D. J. Thomson, J. Vac. Sci. Technol.
B14, 386 1996.
70P. De Wolf, T. Clarysse, and W. Vandervorst, J. Vac. Sci. Technol. B 16,
320 1998.
71P. De Wolf, M. Geva, T. Hantschel, W. Vandervorst, and R. B. Bylsma,
Appl. Phys. Lett. 73, 2155 1998.
72P. De Wolf, R. Stephenson, S. Biesemans, Ph. Jansen, G. Badenes, K. De
Meyer, and W. Vandervorst, Tech. Dig. Int. Electron Devices Meet., 559
1998.
73P. De Wolf, W. Vandervorst, H. Smith, and N. Khalil, J. Vac. Sci. Tech-
nol. B, these proceedings.
74T. Trenkler, T. Hantschel, R. Stephenson, P. De Wolf, W. Vandervorst, L.
Hellemans, A. Malave
´,D.Bu
¨chel, E. Oesterschulze, W. Kulisch, P. Nie-
dermann, T. Sulzbach, and O. Ohlsson, J. Vac. Sci. Technol. B, these
proceedings.
368 De Wolf
et al.
: Status and review of 2D carrier profiling 368
J. Vac. Sci. Technol. B, Vol. 18, No. 1, JanÕFeb 2000
... These junctions' positions depend on the doping of the n-type drift layer and on the electrical activation of aluminum and phosphorous implants employed for the formation of the body and source, respectively. Moreover, other factors can affect the device characteristics, e.g., the off-cut angle of the 4H-SiC crystals along the (11)(12)(13)(14)(15)(16)(17)(18)(19)(20) direction, the shape of the hard masks used for selective ion implantation doping, etc. As an example, the lateral straggling of implanted Al in 4H-SiC has been observed to depend on the crystallographic orientation [8] and can result in asymmetric p-type doping profiles. ...
... Usually, SCM is used for the semiconductor carrier profile and for the determination of the p-n junctions [16,17]. However, the cross-sectional methodologies (TEM, SEM, etc.) suffer from a lack of statistical relevance due to the fact that the information comes from a limited volume fraction of the device (~1 µm in depth) cross-section. ...
Article
Full-text available
In this paper, a two-dimensional (2D) planar scanning capacitance microscopy (SCM) method is used to visualize with a high spatial resolution the channel region of large-area 4H-SiC power MOSFETs and estimate the homogeneity of the channel length over the whole device perimeter. The method enabled visualizing the fluctuations of the channel geometry occurring under different processing conditions. Moreover, the impact of the ion implantation parameters on the channel could be elucidated.
... The SCM measurement results for the calibration sample ( Figure S4) and the non-doped nanowire sample ( Figure S5) discussed in this article are given in the Supporting Information. The other mode called scanning spreading resistance microscopy (SSRM) could be a better-adapted alternative tool for the electrical characterization of nanostructures with resolution down to 10 nm in the air 26 and 1 nm in vacuum. 27 Indeed, the repeatability of SSRM measurements in ZnO under ambient conditions is higher than the SCM measurements due to uncontrollable thickness of the water layer which acts as a dielectric. ...
... The AFM advanced modes have been implemented in the routine characterization of microelectronics since their development in the late 80s. Electrical modes such as scanning capacitance microscopy (SCM) and scanning spreading resistance microscopy (SSRM) are used to image dopant distribution and concentration within semiconductors [32]. ...
Thesis
The continuous miniaturization and complexity of devices have pushed existing nano-characterization techniques to their limits. The correlation of techniques has then become an attractive solution to keep providing precise and accurate characterization. With the aim of overcoming the existing barriers for the 3D high-resolution imaging at the nanoscale, we have focused our research on creating a protocol to combine time-of-flight secondary ion mass spectrometry (ToF-SIMS) with atomic force microscopy (AFM). This combination permits the correlation of the composition in 3-dimensions with the maps of topography and other local properties provided by the AFM. Three main results are achieved through this methodology: a topography-corrected 3D ToF-SIMS data set, maps of local sputter rate where the effect of roughness and vertical interfaces are seen and overlays of the ToF-SIMS and AFM advanced information. The application fields of the ToF-SIMS and AFM combined methodology can be larger than expected. Indeed, four different applications are discussed in this thesis. The procedure to obtain the topography-corrected 3D data sets was applied on a GaAs / SiO2 patterned structure whose initial topography and composition with materials of different sputter rates create a distortion in the classical 3D chemical visualization. The protocol to generate sputter rate maps was used on samples with structured and non-structured nano-areas in order to study the possible ToF-SIMS sputtering artefacts, especially the geometric shadowing effect. Finally, we have explored the combination of ToF-SIMS analysis with three AFM advanced modes: piezoresponse force microcopy (PFM), scanning capacitance microscopy (SCM) and scanning spreading resistance microscopy (SSRM). Specifically, two main applications were studied: the chemical modification during electrical stress of a piezoelectic thin film and the recovery of initial electrical characteristics of a sample subjected to Ga implantation during FIB preparation. Technical aspects of the methodology will be discussed for each application and the perspectives of this combination will be given.
Article
Full-text available
Solar cells are complex devices, being constituted of many layers and interfaces. The study and the comprehension of the mechanisms that take place at the interfaces is crucial for efficiency improvement. This paper applies Kelvin probe force microscopy (KPFM) to study materials and interfaces with nanometer scale imaging of the surface potential in the dark and under illumination. KPFM measurements are highly sensitive to surface states and to the experimental measurement environment influencing the atomic probe operating conditions. Therefore, in order to develop a quantitative understanding of KPFM measurements, we have prepared a dedicated structured sample with alternating layers of InP:S and InP:Fe whose doping densities were determined by secondary-ion mass spectroscopy. We have performed KPFM measurements and shown that we can spatially resolve 20 nm thick InP layers, notably when performed under illumination which is well-known to reduce the surface band-bending.
Article
Full-text available
Scanning probe microscopy (SPM) allows the spatial imaging, measurement, and manipulation of nano and atomic scale surfaces in real space. In the last two decades, numerous advanced and functional SPM methods, particularly atomic force microscopy (AFM), have been developed and applied in various research fields, from mapping sample morphology to measuring physical properties. Herein, we review the recent progress in functional AFM methods and their applications in studies of two-dimensional (2D) materials, particularly their interfacial physical properties on the substrates. This review can inspire more exciting application works using advanced AFM modes in the 2D and functional materials fields.
Thesis
Recent advances in fabrication techniques have led to the successful nano-engineering of semiconductor heterostructures with nanometer-scale structures. Such heterostructures make it possible to tune band gap energies and control carrier confinement for a range of electronic and optoelectronic applications, including solar cells, light emitters, and quantum-computing applications. However, the nanoscale morphologies (i.e., sizes, shapes, and local compositions) of complex heterostructures and their influence on electronic characteristics are not fully understood. To address these issues, it is essential to probe materials on the nanoscale using advanced experimental and computational tools. Therefore, this dissertation focuses on investigating the effects of nanostructure morphologies on the electronic structure of epitaxially-grown semiconductor materials that employ alloying and/or low-dimensional structures (such as quantum dots). In particular, we investigate GaAsNBi and BiSbTe alloys, and InAs/GaAs and GaSb/GaAs quantum dots (QDs) using nanoscale experimental probes in conjunction with self-consistent Schrödinger-Poisson simulations using nextnano. First, we demonstrate an approach to examine apparent stoichiometry in GaAs-based alloys and nanostructures using local electrode atom probe (LEAP) tomography, in conjunction with Rutherford Backscattering Spectrometry (RBS) and high-resolution x-ray diffraction (HRXRC). Using the LEAP conditions identified for achievement of near-stoichiometry in GaAs, we investigate local N and Bi compositions in GaAsNBi alloys and local Si concentrations in the vicinity of Si-doped InAs/GaAs QD superlattices. For the GaAsNBi alloys, LEAP-determined average Bi compositions correlate with those determined using RBS. These are the first known studies that use LEAP to directly measure N and Bi compositions for GaAsNBi films. For Si-doped InAs/GaAs QD superlattices, 3D LEAP data reveals laterally and vertically inhomogeneous Si incorporation, with clusters of Si throughout the layers. Using the local In, Ga, As, and Si compositions from 3D LEAP data as input into Schrodinger-Poisson simulations, we find that electrons are predicted to be localized near both the QDs and the Si clusters. Furthermore, we determined the distribution of compositions within Ga(As)Sb quantum dots (QDs), clusters, and circular arrangements of smaller QDs, termed QD-rings (QDRs) using LEAP. Sizes, shapes, and compositional gradients are used as input into self-consistent Schrödinger-Poisson simulations to compute confinement energies for individual nanostructure types. The computed confinement energies and the measured photoluminescence emission energies increase from QDs to QD-rings to 2D layers, enabling direct association of nanostructure morphologies with the optical properties of the GaSb/GaAs multilayers. This is the first known work that uses measured compositional gradients as input into 8 x 8 k·p calculations for Ga(As)Sb/GaAs nanostructures, opening opportunities for tailoring emission energies for near to far-infrared optoelectronics by varying the QD morphology. Finally, we have investigated the bulk and local electronic states in (Bi1-xSbx)2Te3 alloys using scanning tunneling spectroscopy (STS) and magnetoresistance (MR) measurements. STS reveals both the Fermi level and Dirac point located inside the bulk bandgap, indicating bulk-like insulating behavior with accessible topological surface states (TSSs). STS and reveals a transition from n-type to p-type conduction at x ≈ 0.6. We use a two-channel analysis of MR data to differentiate the charge carrier types for surface and bulk transport; we conclude that surface transport is dominated by electrons and bulk transport is dominated by holes. Prior to this work, direct detection of topological surface states in BiSbTe systems has been achieved only for T<10 K.
Article
Full-text available
This work addresses the need for a comprehensive methodology for nanoscale electrical testing dedicated to the analysis of both “front end of line” (FEOL) (doped semiconducting layers) and “back end of line” (BEOL) layers (metallization, trench dielectric, and isolation) of highly integrated microelectronic devices. Based on atomic force microscopy, an electromagnetically shielded and electrically conductive tip is used in scanning microwave impedance microscopy (sMIM). sMIM allows for the characterization of the local electrical properties through the analysis of the microwave impedance of the metal–insulator–semiconductor nanocapacitor (nano-MIS capacitor) that is formed by tip and sample. A highly integrated monolithic silicon PIN diode with a 3D architecture is analysed. sMIM measurements of the different layers of the PIN diode are presented and discussed in terms of detection mechanism, sensitivity, and precision. In the second part, supported by analytic calculations of the equivalent nano-MIS capacitor, a new multidimensional approach, including a complete parametric investigation, is performed with a dynamic spectroscopy method. The results emphasize the strong impact, in terms of distinction and location, of the applied bias on the local sMIM measurements for both FEOL and BEOL layers.
Thesis
Cette thèse s’intéresse à la caractérisation des propriétés électriques des dispositifs photovoltaïques (PV) par deux techniques de microscopie à sonde locale : la microscopie à force atomique à pointe conductrice (c-AFM) et la microscopie à sonde de Kelvin (KPFM). Elle commence par une étude sur le silicium cristallin, et plus spécifiquement sur l’influence des états de surface sur la mesure KPFM. Cette dernière a été réalisée à l’obscurité et sous éclairement dans le but d’extraire le photovoltage de surface (SPV). Cette étude expérimentale a été complétée avec de la simulation numérique. Dans une deuxième étude nous avons caractérisé des dispositifs PV à nanofils. Il s’agit de jonctions radiales PIN en silicium amorphe hydrogéné réalisées sur des nanofils de silicium cristallin fortement dopé qui ont été étudiées par KPFM et SPV. Nous avons ainsi pu montrer que sur des dispositifs à nanofils recouverts d’ITO, la mesure de SPV permet de mesurer la tension de circuit-ouvert (Voc). Cette même mesure appliquée sur des nanofils uniques (sans contact ITO en face avant) est fortement influencée par l’ombrage de la pointe AFM et par les états de surface de la couche en silicium amorphe. Enfin nous nous sommes intéressés aux contacts passivants (poly-Si/SiOx/c-Si). Les caractérisations KPFM ont révélé des cartographies de potentiel de surface très homogènes en l’absence de la couche enterrée de SiOx, alors que de nombreuses zones de diamètre inférieur au micron présentent des valeurs de potentiel de surface plus faibles lorsque la couche de SiOx est incorporée dans la structure. Ces résultats semblent compatibles avec la présence d’inhomogénéités structurelles de taille nanométrique (pinholes) dans la couche de SiOx.
Article
Dopant concentration sensitive etching of silicon in HF:HNO3:CH3COOH solution was studied using epitaxially grown silicon samples. The study has shown the unstable character of the process, significant time and structure size dependencies of the etching rate, as well as the dependence of the rate on the dopant concentration gradient. The data may be rationalized on the basis of the electrochemical and autocatalytic nature of the reaction. The influence of the dopant gradient and overall device geometry on the etching rate may cause significant inaccuracy of the dopant distribution measurements. © 1998 American Vacuum Society.
Article
To help correlate scanning capacitance microscope measurements of silicon with uniformly doped concentrations, model capacitance curves are calculated and stored in a database that depends on the probe-tip radius of curvature, the oxide thickness, and the dopant density. The oxide thicknesses range from 5 to 20 nm, the dopant concentrations range from 1017 to 1020 cm-3, and the probe-tip radius of curvature is set to 10 nm. The cone-shaped probe is oriented normal to the sample surface, so that the finite-element method in two dimensions may be used to solve Poisson’s equation in the semiconductor region and Laplace’s equation in the oxide and ambient regions. The equations are solved within the semi-classical quasistatic approximation, where capacitance measurement depends only on the charge due to majority carriers, with inversion and charge trapping effects being ignored. Comparison with one-dimensional-related models differs as much as 200% over the given doping range. For shallow gradient profiles satisfying quasiuniformity conditions, the database is used directly to find the doping profile. Converting a 512×512 point image takes less than 2 min. © 1998 American Vacuum Society.
Article
“Lateral secondary ion mass spectrometry (SIMS)” has been demonstrated to be capable, within certain limits, of measuring the lateral straggle and diffusion of dopants under mask edges. In the present work measurements of lateral dose distributions of arsenic and boron in silicon are presented and discussed, and information on the error margins of the method derived. It is shown that the lateral dose distribution of arsenic and boron can be measured down to dose levels of 5×1012cm-2 and 2.6×1013 cm-2, respectively. In repeated measurements, the profile shapes have been reproduced to within a few nanometers, and the profile position relative to the mask edge determined with about ±10 nm error. It is this rather high degree of accuracy that allows Lateral SIMS data to be of value in the calibration and validation of two-dimensional process simulators. An example is given showing the correlation between IMSIL (Technical University Vienna) simulation and experimental data for an as-implanted sample. By comparing experimental data from samples after dopant diffusion and at several topography variants with TMA SUPREM4 simulation, it is shown that an accurate determination of the topography of the implant window in the vicinity of the mask edge is an essential prerequisite of any simulator calibration.© 1998 American Vacuum Society.
Article
The depth dependent carrier density was measured on an arsenic implanted silicon sample using scanning capacitance microscopy (SCM). The capacitance versus voltage scan was performed by applying de biases with a dither ac signal. A strong de bias dependence was observed at the interface of an abrupt junction between n(+) and p. The bias dependent SCM images shaw good agreement with quasi-three dimensional simulations, suggesting that they can be used to map a device structure. (C) 1997 American Institute of Physics.
Article
The application of the conventional spreading resistance profiling (SRP) method on ultrashallow profiles is endangered by the phenomenon of pressure enhanced carrier spilling which results in false dopant and carrier profile differences, the need for large correction factors, and the presence of surfaces states on the bevel. Furthermore, the tedious preparation of the conventional SRP probes requires a lot of expertise. In this work, it is shown that some of these limitations can be resolved by the application of the nano-SRP technique on beveled samples. The use of a single conductive diamond-coated silicon tip mounted on an atomic force microscope (AFM) maintains the strong points of SRP while eliminating the need for probe conditioning. Contact sizes for nano-SRP are a factor of one hundred smaller than in conventional SRP. The small contact size combined with small probe movements allows for considerably larger bevel angles and provides a geometrical resolution as small as 0.5 nm. The smaller contact, i.e., 20–50 nm in diameter, reduces the SRP correction factor by a factor of 20. The much smaller pressure interaction volume reduces the pressure enhanced carrier spilling component. As a consequence, the application of the so-called zero field Poisson contact model becomes more feasible. © 1998 American Vacuum Society.
Article
Quantitative two-dimensional (2D) dopant profiling of a gatelike structure is achieved by scanning capacitance microscopy (SCM) on a cross-sectioned polished silicon wafer. The gatelike structures consist of heavily implanted n+ regions separated by a lighter doped n region underneath a 0.56 μm gate. The SCM is operated in the constant change capacitance mode while scanning with a 37 nm radius tip. The 2D SCM data are converted to dopant density through a physical model of the SCM/silicon interaction. The model parameters are adjusted so that the SCM dopant profile far from the gate edge fits the vertical secondary ion mass spectrometry (SIMS) profile. A 15% error in average accuracy is found between SCM and SIMS profiles evaluated over the dopant range of 1020–1017 cm-3. The same model parameters are used for all points in converting the 2D SCM data, indicating that the accuracy of the full 2D result should be comparable to that of the vertical profile. A direct comparison of the 2D SCM and 2D TSUPREM4 results is made for the first time. The agreement is generally good, but there are some notable differences. © 1998 American Vacuum Society.
Article
Transmission electron microscopy (TEM) was used to characterize image contrast obtained from doping-dependent etching of p-n junctions in silicon. The local variations in crystal thickness give rise to the appearance of thickness fringes which may be interpreted as two-dimensional iso-concentration contours that map the dopant distribution. The samples used for the study consisted of solid source diffusions of boron into substrates of varying resistivities of both n- and p-type. The factors which affect the interpretation of dopant profiles obtained from selective chemical etching of cross section TEM samples is addressed. One-dimensional chemical dopant concentration data were derived from secondary ion mass spectroscopy and one-dimensional carrier concentration data were derived from spreading resistance profiling. © 1998 American Vacuum Society.
Article
We have used scanning tunneling microscopy and spectroscopy (STM/S) to study multiple pn junctions on cross-sectional surfaces of both Si and GaAs devices. The spectroscopy results indicate that pn junctions can be resolved at the nanometer scale by using the two-dimensional STS technique. STM is also used to identify Zn dopants on GaAs(110) surfaces. A detail dopant location identification method is presented. © 1998 American Vacuum Society.
Article
Secondary electron (SE) imaging of semiconductors reveals contrast between n- and p-type areas that can serve as the basis for a two-dimensional dopant profiling technique. In this article, recent experiments that address sensitivity, spatial resolution, calibration methodology, p/n junction effects, and sample preparation issues are reviewed and discussed for boron doped silicon. In addition, several examples of successful applications of SE imaging as a two-dimensional dopant profiling technique are presented. © 1998 American Vacuum Society.
Article
One of the methods to delineate a two-dimensional dopant profile in Si devices is the selective etching technique combined with imaging of the doping dependent topography by atomic force microscopy. A major problem of the etching procedures reported so far is the critical control of the etch duration of only a few seconds as well as a limited sensitivity and dynamic range. We report further improvements of the conventional technique and compare it to an alternative approach. Etching experiments were carried out in an electrolytic cell, whereby the etching is controlled potentiostatically. Better control of the dissolved volume is achieved through current integration. Buffered HF as the electrolyte replaces the difficult nitrogen related chemistry of the classical approach. The etch rates thereby decrease by a factor of 10–100 thus allowing for a longer etch duration. The basic mechanisms, the sensitivity, and the dynamic range of the two techniques have been studied by using homogeneous samples and special test structures for calibration and ultrashallow profiling. Two-dimensional profiling has been carried out on cross sections of 0.25 μm complementary metal–oxide–semiconductor devices with the classical and the electrochemical approaches. Problems of preparation, etching procedure, and imaging techniques are discussed in order to improve the quantification of the selective etching technique. © 1998 American Vacuum Society.