You are on page 1of 9

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/235456665

Near-field spectroscopy of silicon dioxide thin films

Article  in  Physical review. B, Condensed matter · October 2011


DOI: 10.1103/PhysRevB.85.075419 · Source: arXiv

CITATIONS READS
98 2,174

9 authors, including:

G. O. Andreev Zhe Fei


University of California, San Diego Iowa State University
25 PUBLICATIONS   5,015 CITATIONS    33 PUBLICATIONS   6,021 CITATIONS   

SEE PROFILE SEE PROFILE

Alexander S McLeod Gerardo Dominguez


University of Minnesota Twin Cities California State University, San Marcos
132 PUBLICATIONS   8,119 CITATIONS    111 PUBLICATIONS   6,592 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

2D Materials where spin orbit coupling produces Dirac fermions View project

Nano-optics of Strained Correlated Electron Oxides View project

All content following this page was uploaded by Michael M Fogler on 31 May 2014.

The user has requested enhancement of the downloaded file.


Near-field spectroscopy of silicon dioxide thin films

L. M. Zhang,1 G. O. Andreev,2 Z. Fei,2 A. S. McLeod,2 G. Dominguez,3


M. Thiemens,3 A. H. Castro Neto,4 D. N. Basov,2 and M. M. Fogler2
1
Department of Physics, Boston University, 590 Commonwealth Avenue, Boston, Massachusetts, 02215
2
Department of Physics, University of California San Diego, 9500 Gilman Drive, La Jolla, California 92093
3
Department of Chemistry, University of California San Diego, 9500 Gilman Drive, La Jolla, California 92093
4
Graphene Research Centre and Department of Physics,
National University of Singapore, 2 Science Drive 3, 117542, Singapore
(Dated: October 25, 2011)
We analyze the results of scanning near-field infrared spectroscopy performed on thin films of
arXiv:1110.4927v1 [cond-mat.mes-hall] 21 Oct 2011

a-SiO2 on Si substrate. The measured near-field signal exhibits surface-phonon resonances whose
strength has a strong thickness dependence in the range from 2 to 300 nm. These observations
are compared with calculations in which the tip of the near-field infrared spectrometer is modeled
either as a point dipole or an elongated spheroid. The latter model accounts for the antenna effect
of the tip and gives a better agreement with the experiment. Possible applications of the near-field
technique for depth profiling of layered nanostructures are discussed.

PACS numbers: 68.37.Uv, 63.22.-mP

I. INTRODUCTION strongly elongated shape of the tip. Such a tip acts as an


optical antenna1–3 that greatly enhances the electric field
Scattering scanning near-field optical microscopy (s- inside the tip-sample nanogap. Unfortunately, analytical
SNOM)1–3 is a powerful tool for probing local electro- models,21,22 that attempt to treat elongated tips do not
magnetic response of diverse materials. The s-SNOM apply to layered substrates. This compels us to study
achieves spatial resolution of 10–20 nm, which is espe- the problem numerically.
cially valuable in the physically interesting infrared re-
gion4,5 where the resolution of conventional spectroscopy
is fundamentally limited by a rather large wavelength λ ∼ tip
5–500 µm. The s-SNOM techniques have been rapidly
advancing,6,7 which enabled their applications to imag-
ing spectroscopy of complex oxides8–13 and graphene.14
The s-SNOM utilizes scattering of incident light by the p
2L
tip of an atomic force microscope (AFM) positioned next
to the probed sample (Fig. 1). The tip couples to the
sample via evanescent waves of large in-plane momenta z E ext
z
Eext
q ∼ 1/a, where a is the tip radius of curvature (a few tens a
θi
of nm). This is why the lateral resolution of the s-SNOM x
x
Eext
is determined primarily by a rather than λ.15–17 ztip rP (q) q
One of the interesting open questions is the depth (z- 1
d1
coordinate) resolution of the s-SNOM probes. Previous
experiments suggested that it is comparable to the lat- 2
eral resolution ∼ a, based on imaging of small sub-surface
particles.18 Surprisingly, our recently near-field measure-
ments of SiO2 thin films have demonstrated that films as
thick as several hundred nm have a response clearly dif-
ferent from that of the bulk material.19 Thus, if instead FIG. 1. (Color online) Schematics of an s-SNOM experiment.
A scanned probe, modeled as a metallic spheroid with length
of particles one has layers, then the s-SNOM is able to
2L and the apex curvature radius a, is positioned distance ztip
detect them at much larger depths. above the sample. The sample contains a film of thickness d1
In this paper these experimental results are re-analyzed and dielectric function 1 , which is deposited on a bulk sub-
and compared with two theoretical models, the conven- strate with dielectric function 2 . The system is illuminated
tional point-dipole approximation1,20 and the spheroidal by infrared field E ext at an angle of incidence θ. Scattering of
model. The former is very simple to implement but is this radiation by the tip creates evanescent waves with large
also very crude. Predictably, it yields a bulk-like re- in-plane momenta q ∼ 1/a. The experiment measures the to-
sponse of the s-SNOM signal as soon as the SiO2 film tal radiating dipole p of tip, which is determined by multiple
thickness exceeds the tip radius, in disagreement with reflections of the evanescent waves between the tip and sam-
the experiment. A plausible reason for shortcomings of ple. The reflections off the sample are characterized by the
the point-dipole model is its failure to account for the coefficient rP (q, ω).
2

(a) 5 (b) 6 (c) 5


300nm 300nm 300nm
105nm 5 105nm 105nm
4 22nm 4 22nm
22nm

|s3 (SiO2 )/s3 (Si)|

|s3 (SiO2 )/s3 (Si)|


s3 (SiO2 )/s3 (Si)

18nm 4 18nm 18nm


3 2nm 2nm 3 2nm
3 50nm 50nm
2 2
2
1 1 1

0 0 0
900 1000 1100 1200 1300 900 1000 1100 1200 1300 900 1000 1100 1200 1300
ω (cm−1) ω (cm−1) ω (cm−1)

FIG. 2. (Color online) (a) Main panel: Measured infrared near-field spectra for several SiO2 film thicknesses. The quantity
plotted is the absolute value s3 of the third harmonic of the scattering amplitude normalized by that for the Si wafer. (b)
Theoretical results for the spheroid model with a = 30 nm and L = 15a. (c) Theoretical results for the point-dipole model with
a = 30 nm and b = 0.75a.

To make the calculations tractable, we follow exam- monic oscillations


ples in the literature23–25 and model the tip as a metallic
spheroid of total length 2L  a, see Fig. 1. As shown ztip (t) = z0 + ∆z (1 − cos Ωt) , (1)
below, this gives results in a much better agreement with
the experiment in terms of both the frequency and the where ∆z = 50 nm typically. In order to suppress un-
thickness dependence of the near-field signal. We at- wanted background and isolate the part of the signal
tribute the origin of the more gradual film-thickness de- scattered by the probe tip, the signal is demodulated.
pendence in the spheroidal model to the aforementioned Namely, we extracted the absolute values sn (ω) and
“antenna effect.” The magnitude of this effect is deter- phases φn (ω) at tapping harmonics
mined by the material response over length scales ranging
from a to 2L, and so it truly saturates only when the film ZT
iφn dt inΩt 2π
thickness becomes much larger than 2L. sn e = e s(ω, t) , T = . (2)
T Ω
The remainder of the paper is organized as follows. In 0
Sec. II we summarize the experimental procedures and
results. In Secs. III and IV we discuss the two theo- The experimental results for the spectra are shown in
retical models and compare their predictions with the Fig. 2(a). These spectra were intended to be taken in
measurements. Concluding remarks are given in Sec. V. the tapping mode, i.e., for zero z0 . However, experimen-
tally z0 can be determined only up to an additive con-
stant ∼ 1 nm. Therefore, we measured z0 -dependence of
s3 (the approach curves) shown in Fig. 3(a) and selected
II. EXPERIMENT the largest observed s3 . Our results are in a qualitative
agreement with previous experimental study,18 which re-
To make the paper self-contained we summarize the ported approach curves for SiO2 at a few discrete fre-
results of our recent experiments19 in this Section. We quencies and film thicknesses.
investigated commercially available calibration gratings, The data points in Fig. 2(a) represent the normalized
which contain strips or islands of SiO2 thermally grown amplitude s3 (SiO2 )/s3 (Si), where s3 (SiO2 ) and s3 (Si) are
on Si. The manufacturer specified thicknesses of the SiO2 the raw third-order demodulation signals averaged over
layer spanned the range d1 = 2, 18, 22, 108, and 300 nm. the entire SiO2 and Si areas, respectively. The statistical
A combination of CO2 and tunable quantum cascade uncertainty of these averaged data traces is about 2%.
lasers (Daylight Solutions) allowed us to cover the For each thickness studied, the normalized amplitude
frequency range between 890 cm−1 and 1250 cm−1 . The s3 (SiO2 )/s3 (Si) exhibits several maxima. The main peak
near-field data were collected using a Neaspec system. is situated at ω ≈ 1130 cm−1 . The key aspect of the
The measured s-SNOM signal represents the elec- data is a rapid decrease in the normalized amplitude of
tromagnetic field backscattered by the probe and the this peak as the thickness is reduced. A trace of this
scanned sample. The complex amplitude s(ω, t) of the resonance can be reliably identified even for the 2-nm
backscattered field varies periodically with the tapping thick SiO2 film. Another notable feature is the growing
frequency Ω ∼ 40 kHz as the distance ztip between the strength and frequency shift of the secondary peaks on
sample and the nearest point of the tip undergoes har- the high-ω side of the main peak as d1 is decreased.
3

(a) (b) 1.2 (c) 1.2


1 ω = 1066cm−1 ω = 1066cm−1 ω = 1066cm−1
ω = 1126cm−1 ω = 1126cm−1 ω = 1126cm−1
ω = 1158cm−1 1 ω = 1158cm−1 1 ω = 1158cm−1
0.8

s3 (ztip )/s3 (0)

s3 (ztip )/s3 (0)


0.8 0.8
s3 (z0 )/s3 (0)

0.6
0.6 0.6

0.4 0.4 0.4

0.2 0.2 0.2

0 0 0
0 20 40 60 0 20 40 60 0 20 40 60
z0 (nm) ztip (nm) ztip (nm)

FIG. 3. (Color online) Approach curves. (a) Experimental data for the 105-nm thick SiO2 . Theoretical results for the spheroid
(b) and the point-dipole models (c) using the same parameters as in Fig. 2.

Since the response of Si is frequency independent in where z-axis momenta kjz are defined by
our experimental range, the frequency dependence of the r
spectra in Fig. 2(a) originates from that of SiO2 . We ω2
attribute the maxima of s3 (SiO2 )/s3 (Si) to the phonon kjz = j − q2 , Im kjz ≥ 0 . (5)
c2
modes localized at the air-SiO2 interface.6 These reso-
nances occur in the frequency region between the bulk Equation (3) is valid for arbitrary q. In the near-field
transverse and longitudinal modes of SiO2 (the outer case where q is large and kjz ' iq, it simplifies to
dashed lines in Fig. 4 below).
The results of our theoretical calculations for the nor- ∗ − 0 2 + 1 tanh qd1
rP (q, ω) ' , ∗ ' 1 . (6)
malized scattering amplitude are presented in the re- ∗ + 0 1 + 2 tanh qd1
maining panels of Figs. 2 and 3. They are discussed and
compared with the experimental findings in the following Assuming all j are q-independent, the effective dielectric
Sections. function ∗ (q, ω) depends on q only via the product qd1
in this limit. Therefore, rP (q, ω) for one thickness d1 can
be obtained from another by rescaling q. As discussed
in Sec. I and shown in more detail below, the most im-
III. RESPONSE FUNCTIONS AND portant momenta are q ∼ 1/a where a ∼ 30 nm is the
COLLECTIVE MODES
tip radius. Therefore, we can get an approximate under-
standing of the system response by examining the behav-
The sample is modeled as a two-layer system. The ior of rP (q, ω) as a function of ω at fixed qd1 ∼ d1 /a. This
first layer with dielectric function 1 (ω) occupies the slab behavior is dictated by the spectrum of surface collective
−d1 < z < 0. The second layer with dielectric function modes, as follows.
2 (ω) occupies the half-space z < −d1 . The half-space In general, surface modes correspond to poles of the
z > 0 (“layer 0”) is filled with air (dielectric constant 0 = response functions rX . Function rP given by Eq. (6) can
1). The fundamental response functions of the system are have up to two poles at each qd1 , see, e.g., Ref. 26. They
the reflection coefficients rX (q, ω), which are functions are defined by the following condition on 1 (ω):
of in-plane momentum q, frequency ω, and polarization s
X = S or P . The domain of definition of rX (q, ω) is 0 + 2 (0 + 2 )2

understood to include nonradiative modes q > 0 ω/c. 1 (ω) = − ± − 0 2 . (7)
It is known from previous studies that the s-SNOM signal
2 tanh qd1 4 tanh2 qd1
is dominated by the P -polarized waves. In our two-layer
model their reflection coefficient is given by a Fresnel-like At large qd1 , where tanh qd1 = 1, this condition yields
formula 1 (ω) = −0 or 1 (ω) = −2 , which correspond to modes
localized at the upper 0–1 and the lower 1–2 interfaces,
∗ k0z − 0 k1z respectively. Actually, the latter “pole” has vanishingly
rP (q, ω) = , (3) small residue because evanescent waves do not reach the
∗ k0z + 0 k1z
lower interface at qd1 = ∞. There is no q-dispersion
2 k1z − 1 k1z tanh ik1z d1 and no coupling of the two modes in this limit. The
∗ (q, ω) = 1 , (4)
1 k2z − 2 k1z tanh ik1z d1 dispersion appears at finite qd1 , where the two modes
4

become mixed. In particular, we find and 4(b),(c). However, the relation between rP (q, ω) and
the measured s-SNOM signal is nontrivial. For example,
qd1
1 (ω) ' − , “0–1” (8a) the frequency positions of the maxima in Im rP (q, ω) and
−1
+ −1
0 2 those in s3 (SiO2 )/s3 (Si) differ by as much as 40 cm−1 .
0 + 2 We also suspect that there may be some slight differences
'− “1–2” (8b)
qd1 between the optical constants of thick films we assume in
our calculations and those of the small SiO2 structures
at qd1  1. At finite q, both interfaces participate in
we probe by the s-SNOM. This is the likely reason why
generating these excitations. The labels “0–1” and “1–
the crossing point of the experimental curves occurs near
2” are for convenience: they indicate at which interface
1060 cm−1 rather than 1036 cm−1 predicted by both our
a given dispersion branch is ultimately localized as q in-
models, cf. Fig. 2.
creases. At qd1 = 0, the “0–1” and “1–2” branches are
characterized by 1 (ω) = 0 and 1 (ω) = −∞, which cor-
respond, respectively, to the bulk longitudinal and trans- Developing a reliable procedure for inferring rP (q, ω)
verse phonon frequencies ωLO and ωTO . from s3 remains a challenge for the theory. The next sec-
If we try to apply this formalism to real materials, we tion presents our current approach towards this ultimate
face the problem that Eq. (7) has no solutions for real goal.
ω because the dielectric functions have finite imaginary
parts. This is why in practice the collective mode spectra
are usually defined differently. They are identified with
the maxima of dissipation, i.e., Im rP . The number of
these maxima can be fewer than the total allowed number
of the modes because some of them can be overdamped. (a) 15
Re ǫ
Similarly, we define ωLO and ωTO as the frequencies that Im ǫ
correspond to the maxima of −Im −1 10
1 (ω) and Im 1 (ω).
To see what kind of spectra are realized in our system,
ǫ(ω) 5
we use our ellipsometry data for 1 (ω) [Fig. 4(a)] and
Eq. (6) to compute rP for several values of qd1 . The 0
plot of these quantities as a function of ω is presented
in Fig. 4(c). Three maxima on each curve in the region -5
of primary interest ω > 1000 cm−1 are apparent. They
(b)
exist already at qd1 = ∞, and so all of them belong to 2
the upper (air-SiO2 ) interface. In fact, we do not expect
Re rP (q, ω)

sharp modes at the lower (SiO2 -Si) interface because the


1
dielectric function of Si is quite large 2 ≈ 11.7 in the
studied range of ω. The lowest value of Re 1 ≈ −5.0 is
not sufficient to compensate 2 and generate “1-2” modes, 0
cf. Eq. (7).
The main peak of Im rP at qd1 = ∞ defines the -1
surface phonon frequency of SiO2 ωSP ≈ 1164 cm−1 . (c) 3 qd1 = 0.05
There also exist secondary peaks at ω ≈ 1100 cm−1 and qd1 = 0.5
qd1 =1
ω ≈ 1220 cm−1 . Their evolution as a function of qd1
Im rP (q, ω)

2 qd1 = 1.5
comply with the general scheme outlined above. As qd1 qd1 =2
qd1 =∞
decreases, all the three peaks loose strength, as expected,
because the amount of SiO2 diminishes. The lower-ω 1
secondary peak redshifts, moving towards ωTO , and then
quickly disappears. This agrees with the SiO2 -Si res-
0
onance being highly damped. The higher-ω secondary 900 1000 1100 1200 1300
peak becomes dominant at qd1 < 0.5 and demonstrates ω (cm −1)
a systematic shift towards ωLO , see Fig. 4(c).
A notable feature of Fig. 4(b) is the clustering of
the crossing points of the different curves near ω = FIG. 4. (Color online) (a) Dielectric function of bulk SiO2
1036 cm−1 . This is the frequency where the dielectric as a function of frequency from ellipsometry. (b) The real
function of SiO2 is the closest to that of Si, 2 ≈ 11.7. As and (c) the imaginary parts of the near-field reflection co-
a result, the two layers act almost as one bulk material, efficient, rP (q, ω) for several qd1 . In all the panels three
so that rP (ω) is approximately thickness-independent. dashed lines indicate the transverse optical phonon frequency
There is a qualitative correspondence between the fea- ωTO ≈ 1074 cm−1 , the surface optical phonon frequency
tures displayed by the reflection coefficient rP and the ωSP ≈ 1164 cm−1 , and the longitudinal optical phonon fre-
observed near-field signal s3 (SiO2 )/s3 (Si), cf. Figs. 2(a) quency ωLO ≈ 1263 cm−1 .
5

IV. TIP-SAMPLE INTERACTION (a) 2.5


Si
2
Both radiative and nonradiative waves may play sig- SiO2 +Si

|1 + rP |2
nificant roles in the s-SNOM experiment.23 The radiative 1.5
SiO2
modes magnify the signal by a certain far-field factor 1
(FFF) F (qs , ω), where qs = (ω/c) sin θ is the momen-
tum of these modes for the angle of incidence θ. The 0.5
nonradiative modes influence the effective polarizability 0
χ(ω, ztip ) of the tip, i.e., the ratio of its dipole moment (b) 0.5
pz and the external electric field Eext z
. Altogether the

arg (1 + rP )2 /π
demodulated s-SNOM signal sn eiφn can be written as

¤
0
sn eiφn ∝ χn Eext sin 2θ F (qs , ω) , (9)
ZT

£
dt inΩt 
χn (ω) = e χ ω, ztip (t) . (10)
T -0.5
0 900 1000 1100 1200 1300
ω (cm −1)
Below we discuss the FFF and the tip polarizability sep-
arately.
FIG. 5. (Color online) (a) The absolute value and (b) the
phase of the far-field factor computed as a function of fre-
A. Far-field factor quency for the incidence angle θ = 45◦ . The black trace is for
bulk Si substrate, the blue one is for bulk SiO2 substrate, the
green one is for 300 nm thick SiO2 followed by bulk Si. The
The FFF for an infinite layered system is given by27,28 meaning of the dashed lines is the same as in Fig. 4.

F (qs , ω) = [1 + rP (qs , ω)]2 . (11)


dipole.28,33,34 The actual tip shape in our experiment is
As shown in Figs. 5(a), for d1 = 300 nm SiO2 film, the close to a rounded pyramid.
absolute value of the FFF has a maximum near ω TO ≈ The point-dipole approximation is the simplest one
1074 cm−1 and a suppression near ω LO ≈ 1272 cm−1 . and it has been used extensively for modeling s-SNOM
For thinner films, these features are less pronounced. The experiments, including those performed on multilayer
main maximum of s3 , which is the main focus of our anal- systems.14,28 The point-dipole model has two adjustable
ysis, is away from both ω TO and ω LO . It is essentially parameters: the polarizability a3 of the effective dipole
unaffected by the FFF. Still, if FFF were to be included and its position b with respect to the bottom of the tip.
in the calculation in the form prescribed by Eq. (11), it The results obtained following the standard analysis14,28
would produce a visible hump of s3 (ω) near ω TO and are shown in Fig. 2(c) using a = 30 nm and b = 0.75a.
a dip near ω LO . These features are not present in the We see that even for this rather large a the point-dipole
experimental data, Fig. 2(a). A better agreement with model does not reproduce the observed strong depen-
the experiment is obtained if F (qs , ω) is set to a con- dence of s3 on thickness at d1 > 22 nm.
stant, which is what we do here. We rationalize this de- The discrepancy can be seen more clearly in Fig. 6,
cision by noting that the SiO2 layer in the actual samples where the height of the peak in s3 (SiO2 )/s3 (Si) corre-
does not extend over the entire x–y plane but occupies sponding to the surface phonon is plotted as a func-
only small sub-wavelength regions. Therefore, the FFF tion of d1 . For the point dipole model the curve flat-
is dominated by the ω-independent response of Si. tens at d1 ∼ b. In contrast, the experimentally observed
s3 (SiO2 )/s3 (Si) maximum continues to rise with d1 . The
point-dipole model also predicts a very steep approach
B. Point-dipole model of the tip curve, Fig. 3(c), in poor agreement with the measure-
ments.
The effective tip polarizability χ(ω, ztip ) is the most The physical origin of the saturation of the thickness
important factor on the right-hand side of Eq. (9) and dependence in Fig. 2(c) is easy to understand. One can
it is also the most difficult one to compute. This quan- think about the near-field coupling between the point
tity is dictated by the near-field coupling between the dipole and the sample in terms of the method of images.
tip and the sample. For irregular tip shapes it can be For a dipole positioned at zpd = ztip + b, the image is
calculated only numerically. However, previous s-SNOM concentrated at the depth zpd below the surface. There-
studies demonstrated that acceptable results can often be fore, films of thickness larger than zpd would act as a bulk
obtained if the tip is approximated by a spheroid,23–25 a material. Another way to arrive at the same conclusion
small sphere,24,27,29–32 a “finite” dipole,6,21 or a point is to notice that the characteristic range of momenta of
6

the relevant nonradiative waves is q <∼ 1/zpd . Since rP 6


depends on q through the term tanh qd1 [Eq. (6)], the de-
pendence of the near-field coupling on d1 should saturate 5

max s3 (SiO2 )/s3 (Si)


at d1 >
∼ zpd ∼ b.
4
C. Spheroid model of the tip
3
Experiment
The lack of saturation in the observed s-SNOM signal
a = 30nm pdp
as a function of d1 at d1  a indicates that evanescent 2 a = 50nm pdp
waves with momenta q  1/a also play an important role
in the near-field coupling between the tip and the sam- a = 30nm L = 15a
ple. This is a signature of models in which the tip has a 1
0 100 200 300
finite extent in space 2L  a, see Fig. 1. Although such d (nm)
models are certainly more realistic than a point-dipole
approximation, there has not been a systematic study of
how the results would depend on the exact shape of the
FIG. 6. (Color online) The thickness dependence of the s3
tip. Given some initial success of the point-dipole ap- peak for different tip models. The circles represent the point-
proximation, we speculate that a suitable simple shape dipole calculations, one for a = 30 nm (blue) and the other for
can provide a good compromise between increase in com- a = 50 nm (green). The diamonds are for the spheroid model
putational effort and ability to capture relevant physics. with a = 30 nm and L = 15a, the same as in Figs. 2(b). The
To test this idea, we model the tip as an elongated black squares are derived from the experimental data shown
metallic spheroid positioned above a two-layer medium. in Fig. 2(a) after some smoothing over fluctuations.
This follows a tradition in the literature wherein sim-
ilar models were considered23–25 for the case of bulk
substrates. In Ref. 21 an analytical formula for the sphere above a dielectric half-space.27,35 We verified that
spheroidal tip was also proposed, based on heuristic ar- the two methods give identical results.
guments. However, it cannot be easily extended to the Substituting the computed polarizability χ into
q-dependent rP we study here. Instead, our calculations Eqs. (9) and demodulating per Eq. (10), we obtain ap-
are done numerically. They involve only two essential proach curves. Figure 3(b) illustrates that some approach
approximations. One is neglecting retardation, which is curves are nonmonotonic near the resonances. In cal-
justified is the length 2L of the spheroid is smaller than culating the s-SNOM amplitude s3 we choose ztip that
λ. The other one is neglecting the finite skin depth of corresponds to the largest s3 because this is how it was
the metal (Pt-Ir alloy) covering the tip. Due to compu- done in the experiments. The results for the normalized
tational difficulties involved, this issue is left for future amplitude are plotted in Fig. 2(b).
investigation. The spheroid model has two adjustable parameters:
The calculations were performed in two ways. First is the apex radius of curvature a and the half-length L.
the standard boundary-element method. In this method When L = a the spheroid becomes a sphere. In this
we divide the entire tip — assuming azimuthal symmetry case the spheroid model gives results similar to the point-
— into a large number (typically, 200) of small cylindri- dipole model, i.e., Fig. 2(c). As the ratio L/a increases,
cal segments. We assume that different segments interact the differences appear. However, once L/a exceeds ten,
by Coulomb interaction as coaxial rings. The interaction the normalized signal s3 (SiO2 )/s3 (Si) does not change
of each segment with itself is defined in such a way that much at d1 ≤ 300 nm. Therefore, for long spheroids we
the polarizability of the tip in the absence of the sample effectively have only a single adjustable parameter, a.
coincides with the known analytical result for the prolate Remarkably, the thickness dependence of the s3 peak for
spheroid. The effect of the sample is included by adding the spheroid model matches the experiment extremely
ring-ring interactions mediated by reflected electrostatic well (Fig. 6).
fields. This is accomplished by numerical quadrature over
the product of rP (q, ω) and suitable form-factors. This
is the most time-consuming step of the simulation. Af- V. CONCLUSIONS
ter the interaction kernel is generated in this way, it is
straightforward to solve numerically for the dipole mo- In this paper we analyzed the results of experimen-
ment of the tip induced by a unit external field, which is tal study of amorphous SiO2 films on Si obtained by
the desired polarizability χ(ω, ztip ). scanning near-field optical spectroscopy.19 We discussed
We also developed a second numerical method of com- the collective mode spectra of such structures and com-
puting χ (to be described elsewhere), based on an expan- pared measurements with two theoretical calculations.
sion of the electric field in ellipsoidal harmonics. This The first is based on a conventional approximation in
alternative method is similar to that used for a metallic which the tip of the scanned probe is modeled as a point
7

dipole. In the second the tip is treated as an elongated hancement have momenta ranging from q ∼ 1/a down
spheroid, significantly improving agreement with the ex- to q ∼ 1/L. Therefore, one may expect that the depen-
periment. dence of the s-SNOM signal on the thickness d1 of the
We explain the qualitative difference between the two top layer would saturate only when d1 ∼ L. Our simu-
models as follows. An important physical ingredient lations provide direct evidence for this claim. Therefore,
missing in the point-dipole model is the enhancement of we think that the spheroid model holds a great promise
the electric field near the apex of the tip — the antenna as an analysis tool for near-field experiments. It captures
effect. This phenomenon is well-known from classical a lot of physics relevant to the near-field interaction while
electrostatics. The enhancement of the field is controlled remaining computationally fast.
primarily by the ratio of the total length of the tip 2L The strong experimentally observed thickness depen-
(actually, the smaller of 2L and λ) and the apex radius dence of the near-field signal19 indicates that s-SNOM is
of curvature ∼ a. The point-dipole model has been suc- capable of not only high lateral resolution but can also
cessful in the past without this enhancement factor only probe the system in the third dimension. However, the
on account of the normalization procedure. Instead of response of a layered system is different from those con-
absolute sn , one usually reports sn normalized to some taining small subsurface particles18 . We hope that ex-
reference material such as Au or in our case, Si. This perimental and theoretical approaches presented in this
way, one eliminates any possible frequency dependence paper may be of use for accurate depth profiling of vari-
of the source radiation, but at the same time cancels the ous dielectric and metallic nanostructures.
part of the signal scaling with tip size. For a stratified The work at UCSD is supported by ONR, AFOSR,
sample this cancellation is imperfect because the the field NASA, and UCOP. AHCN and LMZ acknowledge
enhancement depends also on the dielectric response of DOE grant DE-FG02-08ER46512 and ONR grant MURI
the sample, which is a function of momentum q. For N00014-09-1-1063. We thank F. Keilmann and R. Hil-
a tip of length 2L, harmonics relevant for the field en- lenbrand for illuminating discussions.

1
F. Keilmann and R. Hillenbrand, Phil. Trans. Roy. Soc. O. G. Shpyrko, and D. N. Basov, Phys. Rev. B 83, 165108
London, Ser. A 362, 787 (2004). (2011).
2 14
L. Novotny and B. Hecht, Principles of Nano-Optics Z. Fei, G. O. Andreev, W. Bao, L. M. Zhang, Z. Zhao,
(Cambridge University Press, 2006). G. Dominguez, M. Thiemens, M. M. Fogler, A. H. Castro
3
F. Keilmann and R. Hillenbrand, “Nano-optics and Near- Neto, C. N. Lau, F. Keilmann, and D. N. Basov, Nano
field Optical Microscopy,” (Artech House, Norwood, 2009) Lett. (in press).
15
Chap. 11: Near-Field Nanoscopy by Elastic Light Scat- K. Lai, M. B. Ji, N. Leindecker, M. A. Kelly, , and Z. X.
tering from a Tip, pp. 235–266, edited by A. Zayats and Shen, Rev. Sci. Instrum. 78, 063702 (2007).
16
D. Richards. A. J. Huber, F. Keilmann, J. Wittborn, J. Aizpurua,
4
D. N. Basov, R. D. Averitt, D. van der Marel, M. Dressel, and R. Hillenbrand, Nano Lett. 8, 3766 (2008), pMID:
and K. Haule, Rev. Mod. Phys. 83, 471 (2011). 18837565.
5 17
D. N. Basov and A. V. Chubukov, Nat. Phys. 7, 272 (2011). R. L. Olmon, P. M. Krenz, A. C. Jones, G. D. Boreman,
6
S. Amarie and F. Keilmann, Phys. Rev. B 83, 045404 and M. B. Raschke, Opt. Express 16, 20295 (2008).
18
(2011). T. Taubner, F. Keilmann, and R. Hillenbrand, Opt. Ex-
7
F. Huth, M. Schnell, J. Wittborn, N. Ocelic, and R. Hil- press 13, 8893 (2005).
19
lenbrand, Nat. Mat. 10, 352 (2011). G. O. Andreev et al., “Infrared nanooptics of ultrathin
8
M. M. Qazilbash, M. Brehm, B.-G. Chae, P.-C. Ho, G. O. materials,” in preparation.
20
Andreev, B.-J. Kim, S. J. Yun, A. V. Balatsky, M. B. R. Hillenbrand, T. Taubner, and F. Keilmann, Nature
Maple, F. Keilmann, H.-T. Kim, and D. N. Basov, Science 418, 159 (2002).
21
318, 1750 (2007). A. Cvitkovic, N. Ocelic, and R. Hillenbrand, Opt. Express
9
H. Zhan, V. Astley, M. Hvasta, J. A. Deibel, D. M. Mittle- 15, 8550 (2007).
22
man, and Y.-S. Lim, App. Phys. Lett. 91, 162110 (2007). K. Moon, E. Jung, M. Lim, Y. Do, and H. Han, Opt.
10
M. M. Qazilbash, M. Brehm, G. O. Andreev, A. Frenzel, Express 19, 11539 (2011).
23
P.-C. Ho, B.-G. Chae, B.-J. Kim, S. J. Yun, H.-T. Kim, J. A. Porto, P. Johansson, S. P. Apell, and T. López-Rı́os,
A. V. Balatsky, O. G. Shpyrko, M. B. Maple, F. Keilmann, Phys. Rev. B 67, 085409 (2003).
24
and D. N. Basov, Phys. Rev. B 79, 075107 (2009). J. Renger, S. Grafström, L. M. Eng, and R. Hillenbrand,
11
A. Frenzel, M. M. Qazilbash, M. Brehm, B.-G. Chae, B.-J. Phys. Rev. B 71, 075410 (2005).
25
Kim, H.-T. Kim, A. V. Balatsky, F. Keilmann, and D. N. R. Esteban, R. Vogelgesang, and K. Kern, Opt. Express
Basov, Phys. Rev. B 80, 115115 (2009). 17, 2518 (2009).
12 26
K. Lai, M. Nakamura, W. Kundhikanjana, M. Kawasaki, B. Prade, J. Y. Vinet, and A. Mysyrowicz, Phys. Rev. B
Y. Tokura, M. A. Kelly, and Z.-X. Shen, Science 329, 190 44, 13556 (1991).
27
(2010). S. V. Sukhov, Ultramicroscopy 101, 111 (2004).
13 28
M. M. Qazilbash, A. Tripathi, A. A. Schafgans, B.-J. Kim, J. Aizpurua, T. Taubner, F. J. G. de Abajo, M. Brehm,
H.-T. Kim, Z. Cai, M. V. Holt, J. M. Maser, F. Keilmann, and R. Hillenbrand, Opt. Express 16, 1529 (2008).
8

29 33
R. W. Rendell and D. J. Scalapino, Phys. Rev. B 24, 3276 R. Hillenbrand and F. Keilmann, Phys. Rev. Lett. 85, 3029
(1981). (2000).
30 34
P. K. Aravind and H. Metiu, J. Phys. Chem. 86, 5076 T. Taubner, F. Keilmann, and R. Hillenbrand, Nano Lett.
(1982). 4, 1669 (2004).
31 35
P. K. Aravind and H. Metiu, Surf. Sci. 124, 506 (1983). G. Ford and W. Weber, Physics Reports 113, 195 (1984).
32
R. Ruppin, Phys. Rev. B 45, 11209 (1992).

View publication stats

You might also like