Open Access
25 August 2020 Spatial and spectral mode mapping of a dielectric nanodot by broadband interferometric homodyne scanning near-field spectroscopy
Jinxin Zhan, Wei Wang, Jens H. Brauer, Lukas Schmidt-Mende, Christoph Lienau, Petra Groß
Author Affiliations +
Abstract

We investigate the optical properties of nanostructures of antimony sulfide (Sb2S3), a direct-bandgap semiconductor material that has recently sparked considerable interest as a thin film solar cell absorber. Fabrication from a nanoparticle ink solution and two- and three-dimensional nanostructuring with pattern sizes down to 50 nm have recently been demonstrated. Insight into the yet unknown nanoscopic optical properties of these nanostructures is highly desired for their future applications in nanophotonics. We implement a spectrally broadband scattering-type near-field optical spectroscopy technique to study individual Sb2S3 nanodots with a 20-nm spatial resolution, covering the range from 700 to 900 nm. We show that in this below-bandgap range, the Sb2S3 nanostructures act as high-refractive-index, low-loss waveguides with mode profiles close to those of idealized cylindrical waveguides, despite a considerable structural disorder. In combination with their high above-bandgap absorption, this makes them promising candidates for applications as dielectric metamaterials, specifically for ultrafast photoswitching.

1.

Introduction

Stibnite (antimony sulfide, Sb2S3) is currently gaining increasing attention as an emerging photovoltaic material.1,2 Considerable efforts are being undertaken to advance the fabrication and structural characterization25 of these materials in attempts to improve their photovoltaic performance.68 Annealing at moderate temperatures of typically above 250°C activates crystallization of the initially amorphous material and decreases the bandgap energy from 2.24 to 1.73 eV while maintaining a high above-bandgap absorption coefficient of >5×104  cm1.2 This makes the semiconductor Sb2S3 a material which not only can be synthesized from earth-abundant and nontoxic constituents but is strongly absorbing across the entire visible spectrum. The combination of a direct bandgap and a high absorption coefficient leads to considerable photoconductivity, enabling applications in optoelectronics.9 Power conversion efficiencies of up to 7.5% have been demonstrated for an Sb2S3-sensitized heterojunction solar cell.7

For future applications, it is particularly beneficial that comparatively simple and cheap ink-based methods can be used to create Sb2S3 nanoparticles in large-area thin films.10,11 As recently demonstrated, nanopatterning via direct electron beam lithography (EBL) writing enables the creation of 2D and 3D nanometric structures of various shapes.9 The availability of easy structuring, together with the possibility of changing the refractive index on an ultrafast time scale by injecting free carriers, makes the material an interesting candidate for the realization of photoswitchable metasurfaces.12 For this, a high optical quality of Sb2S3 nanostructures is an important prerequisite that has not yet been investigated. Since the structural characterization of the bulk material shows polycrystalline grains with typical sizes of tens of nm, and the optical spectra show a sizeable sub-bandgap absorption, this may indicate a high crystal defect density.2 As such, a nanoscale optical characterization of individual nanoparticles seems particularly desirable.

For these reasons, we report here a first experimental study of the optical properties of Sb2S3 nanostructures. Specifically, we aim at probing sub-bandgap waveguiding in nanodots fabricated with EBL to explore their potential as dielectric metamaterials. For this, we spatially image the optical modes of such nanoparticles using a novel, spectrally broadband scattering-type scanning near-field optical microscopy (sSNOM)1315 technique. Raster-scanning a sharp scattering or collecting tip that is brought into the sample’s near field has in the past been applied to mapping the evanescent field distribution of a multitude of different samples, such as waveguides and microresonators,16,17 photonic crystal microcavities,18 single plasmonic nanoparticles,1923 and, more recently, all-dielectric nanostructures.24,25 Ideally, the recorded intensity distribution provides a map of the local optical density of states (LDOS),21,2631 with a spatial resolution of 20 nm or below.14,3234 The expansion of sSNOM to broad-bandwidth spectroscopy in the visible spectral range, however, is challenging. Unless advanced, virtually background-free techniques such as plasmonic nanofocusing31,3537 are used, this requires efficient background suppression based on tip–sample distance modulation and near-field enhancement using homodyne or heterodyne mixing.3843

Here, we solve these challenges by detecting tip-modulated near-field spectra after homodyne-mixing with a reference field using a monochromator followed by a fast CCD line camera. The camera line readout rate of 210 kHz is sufficiently fast to enable extraction of complete spectra at up to the fourth harmonic of the tip–sample modulation frequency. We obtain background-free near-field spectra in the spectral range from 720 to 900 nm. The SNOM maps recorded across the Sb2S3 nanoparticles reveal that they act as high-refractive index, dielectric waveguides, despite their irregular surface apparent in structural studies. They support waveguide-like modes with the electric field strongly confined in the nanodots and mode profiles matching those of idealized cylindrical waveguides, which makes them promising structures for applications in nanophotonics and specifically creates interest in future studies of their nonlinear properties.

2.

Experimental Setup and Methods

The experimental setup is a home-built sSNOM incorporated in the sample arm of a Michelson interferometer, as depicted schematically in Fig. 1. Light from a titanium:sapphire laser (Femtolasers Rainbow), with a spectral bandwidth of around 150 nm and centered at a wavelength of 820 nm (see the inset in the lower left of Fig. 1), is used to illuminate the sample. The power is set to 12 mW in front of the interferometer setup, and the direction of the electric field vector of the linearly polarized light can be adjusted using a half-wave plate (HWP). The light passes a beam splitter (BS) where 20% is reflected and is then focused on the sample using a reflective microscope objective (MO, Beck Optronics Solutions, model 5003-000) with NA=0.4 and with a working distance of 14.5 mm. Using a reflective MO minimizes dispersion that would otherwise result in a varying fringe spacing in the recorded spectral interferograms. After the BS and the reflective MO, <1  mW is incident on the sample and tip. For these input powers, laser-induced tip- and/or sample-heating effects were not observed. 80% of the incident light is transmitted at the BS and serves as a reference field ER in our measurements. For this, the power is adjusted by a variable gray filter before it is reflected back to the BS by a mirror (MR) that forms the reference arm of the Michelson interferometer. The reference arm length is adjusted to be shorter than the sample arm by 100  μm in order to enable spectral interference with a convenient fringe spacing and is then kept constant during the measurement. To access the near-field region, a sharply etched gold nanotaper is brought in close proximity to the sample surface. Scattered light is collected in the backward direction by the MO. The part that is transmitted through the BS is overlapped with the reference field ER, and the resulting superposition is collected for detection.

Fig. 1

Experimental setup for broad-bandwidth interference SNOM (BISNOM). Light from a broad-bandwidth titanium:sapphire laser (see the spectrum in the inset in the lower left) is focused onto the sample by an MO. The polarization is controlled by an HWP. A sharply etched gold tip is brought close to the sample to scatter light from the near-field region to the far field. The scattered light is collected by the MO. The larger part of the incident light (80%) is transmitted through a BS to the reference arm of a Michelson interferometer, adjusted in power by a variable gray filter, reflected (MR), and superimposed with the light from the sample arm at the BS (see the spectral interferogram in the inset in the lower right). The light is detected either by an APD or by a monochromator followed by a fast CCD line camera.

AP_2_4_046004_f001.png

The near-field probe is a sharply etched single-crystalline gold nanotip with an apex radius of curvature of 10  nm. The tip is brought in close vicinity to the sample surface and the tip–sample distance is modulated at a frequency f26  kHz and with an amplitude of 12 nm. The minimum tip–sample distance is kept at 4 nm. Here, we set the light polarization to s, i.e., the electric field vector is along the y axis such that it lies within the sample plane and is perpendicular to the tip axis.

We employ two different detection methods. In the first case, for quasimonochromatic measurements, the reference arm is blocked and the signal backscattered from the focus region is detected with an avalanche photodiode (APD, Hamamatsu, model C5331-02). It is further processed by a lock-in amplifier (Zurich Instruments HFLI), with the tip modulation frequency as the reference signal. The signal demodulated at the fundamental tip modulation frequency as well as those demodulated at its second, third, and fourth harmonics are obtained and stored. We label these spectrally unresolved signals produced by the lock-in amplifier with S1f, S2f, S3f, and S4f. In this case, with the reference arm blocked, the signal detected by the APD is constituted of light scattered out of the near field by the sharp tip, ENF, and of the unwanted background field EB scattered from the sample and from the tip shaft. Together, this results in a voltage proportional to |ENF+EB|2. Demodulating the signal at higher-order multiples of the tip modulation frequency makes the signal sensitive to the near-field contribution by recording an interference pattern between ENF and EB.39

In the second case, for spectrally resolved measurements, the reference arm is unblocked and the output of the Michelson interferometer is steered to the monochromator (Princeton Instruments, IsoPlane-160) equipped with a fast line camera (e2V AViiVA EM4 with 512 pixels). The spectra measured now constitute a mixture of near field, background, and reference field:

Eq. (1)

S(λ)|ENF+EB+ER|2.

In front of the monochromator, the power detected from the sample arm alone amounts to 10  μW, and the power detected from the reference arm alone is 46  μW. The reference arm length is adjusted such that spectral interference fringes emerge with a spacing of 3 to 5 nm (see the inset at the lower right in Fig. 1). Equation (1) can be written as

Eq. (2)

S(λ)|ENF+EB|2+|ER|2+2Re[(ENF+EB)·ER]·cos(k·ΔL),
where the wavevector k=2πλ1 and the path length difference between sample and reference arm is ΔL. By separating the modulated part of the measured spectra, the term Re{(ENF+EB)·ER} is isolated and the complexity of Eq. (2) is considerably reduced.

The remaining problem of suppressing EB with respect to ENF is solved by the fast line camera behind the monochromator. The line camera has a maximum readout rate of 210  kHz>8·f, and we typically record 60,000 consecutive spectra with the maximum rate. In postprocessing, the recorded values are considered separately, i.e., for each wavelength component of the spectrum S(λ), the line camera records a time series of 60,000 values. A Fourier series expansion is performed around multiples of the tip modulation frequency. For each pixel, the so-determined Fourier coefficients are quantities analogous to the signals generated by the lock-in amplifier in the monochromatic measurements. Assembling the respective Fourier components for all camera pixels results in spectra S1f(λ) to S4f(λ), demodulated at the first to fourth harmonic of the tip modulation frequency.44 As we will show below, for higher-order demodulation the background field EB is strongly suppressed in comparison with the near-field ENF, such that these spectra produce Re{ENFER}.

We investigate individual antimony sulfide (Sb2S3) nanoparticles created on top of an Sb2S3 film by EBL.9 Our samples are annealed at 300°C to form crystalline thin films and nanostructures. This procedure shifts the bandgap from 2.24 to 1.73 eV, and it reduces the above-gap absorption coefficient from 1.8×105  cm1 at 450 nm to 7.5×104  cm1 at 550 nm.2 Here, we study their optical properties at wavelengths between 700 and 900 nm, well below the bandgap. In this spectral region, Sb2S3 has relatively low losses (absorption <2×104  cm1 at 800 nm)2 and a high index of refraction, with a positive real part varying from 3.2 at 700 nm to 2.9 at 900 nm, which means that, in our experiments, it can basically be considered as a lossless dielectric.

Here, we cover a flat, 140-nm-thick film of crystalline Sb2S3 with Sb2S3 nanoparticles, aiming at a cylindrical structure that should guide the incident light into the underlying thin film. The nanoparticles are created by EBL9 on top of the flat film in a regular order, with a spacing of 2  μm [Fig. 2(a); more details on the sample preparation can be found in the Supplemental Materials]. The spacing is chosen sufficiently large to avoid dipolar couplings between adjacent nanoparticles such that each nanodot may be considered as an isolated nanoparticle. They are typically dome-shaped, with a round or slightly elliptical cross section of 350  nm diameter and a height of 150  nm. Figure 2(a) shows a scanning electron microscope (SEM) image of the nanoparticles. A higher resolution image of an individual nanoparticle [Fig. 2(b)] reveals a slightly elliptical shape and a somewhat rough surface. The center of the particle seems to have collapsed to some extent after the annealing processes following EBL. These features (ellipticity, surface roughness, and indentation) vary slightly from one nanoparticle to the next.

Fig. 2

SEM images of the sample. (a) Sb2S3 nanodots are created on a flat, compact Sb2S3 film by EBL. The nanodots are regularly spaced by 2  μm, and their shape varies slightly. (b) The SEM image of an individual nanodot shows a slight ellipticity and a dark shadow in the center where the material has collapsed after EBL and annealing.

AP_2_4_046004_f002.png

3.

Results and Discussion

Before recording the BISNOM spectra, we perform preparatory experiments to validate the near-field contrast and spatial resolution of the optical signals in our near-field microscope. For this, we use a quasimonochromatic spectrum for excitation, created by inserting a 40-nm bandwidth interferometric filter centered at a wavelength of 900 nm in front of the BS, and the APD and lock-in amplifier for detection. The demodulated signals S1f, S2f, S3f, and S4f are generated by the lock-in amplifier and are recorded during the measurement.

Figure 3(a) shows approach curves recorded above a flat Sb2S3 film, i.e., the demodulated signals as a function of tip–sample distance. The approach is stopped when the tuning fork oscillation amplitude is reduced by 2.5% [see the inset of Fig. 3(a)]. This position, which corresponds to a minimum tip–sample distance of about 4 nm, is later taken as the set point for lateral sSNOM scans. The blue, green, and red curves in Fig. 3(a) show the optical signals S1f, S2f, and S4f, respectively, when the sample is retracted from the tip. To retain clarity in Fig. 3(a), S3f is not shown, as its shape is very similar to that of S2f. As the tip is retracted, the signal amplitude decays quickly. This is reflected by all three curves shown in Fig. 3(a), where the strong peaks at the minimum tip–sample distance of 4 nm decrease quickly with increasing tip–sample distance. Compared to the near field, the background field’s dependence on the tip–sample distance is weaker. As a result, the first-order demodulated signal S1f, which mainly probes background scattering from the tip shaft, remains at a rather high level for large tip–sample distances. Fluctuations arise from finite mechanical instabilities and/or laser noise. The higher demodulation orders, however, measure an interference between the near-field signal and the background. Here, a measurable signal is only seen for tip–sample distance of <10  nm. The signal amplitude in contact decreases when going from 2f to 4f due to the demodulation of the signal recorded with finite tapping amplitudes.41

Fig. 3

Quasimonochromatic SNOM measurements of an individual nanoparticle using a 40-nm bandwidth laser spectrum centered at 900 nm for excitation and the APD and lock-in amplifier for detection. (a) Optical signal demodulated at the first (blue curve), second (green curve), and fourth harmonic (red curve) of the tip modulation frequency as a function of tip–sample distance. The higher-order demodulated signals demonstrate an improved near-field contrast. The inset shows the simultaneously recorded tuning fork amplitude. (b) A topographical map and (c)–(e) maps of the optical signals S1f, S2f, and S4f, respectively. All maps show a ring-shaped intensity distribution. (f) Cross cuts through the topographical map (blue curve) and the S4f signal (black curve) along dashed lines in (b) and (e).

AP_2_4_046004_f003.png

To further evaluate the optical image capabilities of the near-field microscope, we then scanned a single Sb2S3 nanodot with the quasimonochromatic measurement method while the tuning fork amplitude was kept constant. The heights of the sample and the demodulated signals S1f to S4f were recorded over an area of 1  μm by 1  μm. The sample was scanned linewise in the x direction with a step size of 4 nm and with a step size of 16 nm between the lines, i.e., in the y direction. The topographic image displayed in Fig. 3(b) shows that this nanodot was almost circular, with lengths of 360 and 380 nm of the short and the long axis of an ellipse, respectively, and with a height of about 150 nm. The simultaneously recorded maps of the signals S1f, S2f, and S4f are displayed in Figs. 3(c)3(e), respectively. They all display a clear optical signal drop in the shape of a ring over the particle. In the following, we investigate several of the observed features one by one.

The near-field images of the nanodot in Fig. 3 show a rather smooth surface and sharp signal change at the edge of the nanoparticle. The maps of the S1f and S2f signals display some variation across the surrounding Sb2S3 film, which may be caused by residual background interference. The S1f and S2f maps both show a strong and sharp signal increase at the left-hand edge of the nanoparticle, which is probably caused by the grazing incidence of the illuminating laser light. The spatial resolution of the optical images is better than 20 nm (see the Supplemental Materials).

The best near-field contrast is reached with the S4f signal [Fig. 3(e)]. The optical signal measured across the particle is reduced in the ring-shaped area with respect to the signal measured on the film by roughly 90%. This difference can be understood as follows. While the tip is within a small distance to the flat film, incident light and light reflected from the film couple to the polarizability component of the tip dipole that is oriented along the polarization direction of the incident laser field. The laser is polarized perpendicular to the tip axis to avoid strong resonant excitation of the longitudinal tip polarizability. The y-polarized tip dipole induces an image dipole in the sample and the near-field coupling between the tip dipole and the image dipole enhances the field that is scattered from the tip to the detector. In the flat film region, the near-field contrast thus results from the coupling of the tip and image dipoles, as introduced by Keilmann.40 For such a planar sample, the LDOS is spatially and spectrally homogeneous, and the near-field signal probes the dielectric function of the bulk material.39,40 As the tip reaches the edge of the particle, it no longer couples to an extended film, but to a nanoparticle of finite size. Hence, the LDOS of the sample is dominated by the resonant optical modes of the nanoparticle. As will be discussed in more detail below, the SNOM images essentially provide a spatial map of the profile of these modes.

We find that the near-field signal rises steadily from the low-signal region near the particle rim toward the center of the nanodot. The over-all symmetry is roughly radial. To show this more clearly, cuts through the topographic image and through the S4f map [along the dashed lines in Figs. 3(b) and 3(e)] are plotted together in Fig. 3(f). The topography contour (blue curve) is a smooth cone, with its maximum leaning somewhat to the right. We thus interpret the central feature in the SNOM images as the spatial extent of the optical modes of the nanoparticle that are excited by the external laser light and probed by their coupling to the near-field tip. The dome-like shape of the observed mode is reminiscent of fundamental fiber-optical modes and suggests that waveguide modes that are excited in the Sb2S3 nanoparticle are of a similar shape.

In order to investigate the nanodot modes further, we calculate the solutions of the vector Helmholtz equation for a weakly guiding, cylindrical waveguide, as it is routinely applied to fiber-optical waveguides.45,46 Taking a long cylinder of 200 nm radius, the refractive index of Sb2S3 n1=3.04 at a wavelength of 800 nm, and n2=1 for the surrounding medium air, there exist four solutions for guided waves. Adopting the terminology that was introduced for fiber modes in 1971 by Gloge,45 these are the LP01, LP11, LP02, and LP21 modes, which are strongly localized within the waveguide. Here, the first index indicates the number of pairs of nodes of the electric field strength in the azimuthal coordinate, and the second index gives the number of nodes in the radial direction, including the approach toward zero for large radii.

We assume that the LDOS in the nanodot region is mainly given as a sum over these bound modes with eigenvectors em(r) and eigenfrequencies ωm.26 Each of these modes is characterized by a finite damping rate γm that describes the temporal decay of the electromagnetic field emitted by this mode after impulse excitation. In the limit of sufficiently weak damping, the LDOS may be phenomenologically written as

Eq. (3)

ρ(r,ω)=m|em(r,ω)|2δg(ωωm).

Here, δg(ωωm)=2iωmπ(ωm2ω22iωγm) with ωm=ωm2γm2 is a generalized Dirac delta function. Its integral along the frequency axis gives unity. Quite generally, the damping can be induced by the finite lifetime of the mode m, T1m, and possible pure dephasing processes with pure dephasing time T2m* resulting from stochastic fluctuations of the environment. Phenomenologically, the damping rate can then be taken as γm=(2T1m)1+(T2m*)1 and the mode can be characterized by a complex resonance frequency ω˜m=ωmiγm. The contribution of unbound radiation modes is neglected since their overlap with the nanodot core is small. For the perfect cylindrical waveguide considered here, for each bound mode a continuum of modes that are propagating along the fiber axis exists and the LDOS is spectrally flat.

The LDOS for a cylindrical waveguide of 200 nm radius is plotted in Fig. 4(a) for a wavelength of 800 nm. In this picture, the corresponding eigenvectors are twofold polarization degenerate for light field polarization along the x and y axes, while the z axis is the direction of propagation of the guided waves. An SNOM measurement carried out by scanning a point dipole p oriented along u^ across a structure with localized modes probes the projected LDOS,13,26

Eq. (4)

ρu^(r,ω)=2ωπc2Im[u^·G(r,r,ω)·u^],
where, for weakly dissipative systems, Green’s function can be written as26

Eq. (5)

G(r,r,ω)c2mem*(r)em(r)ωm2ω22iωγm.

Fig. 4

(a) Local density of states in an Sb2S3-nanodot with a circular cross section of a 200-nm radius at a wavelength of 800 nm. (b) An x-oriented tip dipole is moved along the x axis from the left rim to the center over the surface of the elliptical nanodot. Below is shown how the dipole orientation changes relative to the surface, and the red arrows at the bottom indicate the projection on the surface. (c) The projected LDOS as probed with an x-oriented tip dipole shows minima at the left and right rims, corresponding to the small projection of the tip-dipole onto the surface at these edges. (d) Using a y-oriented tip dipole results in a 90-deg rotated image, and (e) a z-oriented dipole particularly probes the edge of the rim. (f) The asymmetric SNOM map calculated for a wavelength of 800 nm and (g) for comparison, the measured S4f map [same as Fig. 3(e)]. The main maximum as well as the asymmetric side lobes on the left and lower right and the near-zero signal regions at the top and bottom are well recreated. (h) Horizontal and vertical cuts (upper and lower graphs, respectively) of the measured SNOM map (black circles) together with the calculated SNOM signals (black curve), calculated for the nanodot curvature shown by the blue curves.

AP_2_4_046004_f004.png

In principle, if a tip closely representing a point dipole was scanned across a flat surface with such a waveguide mode profile (such as the flat top of a pillar of Sb2S3 surrounded by air), the resulting SNOM map should be an image of the calculated LDOS shown in Fig. 4(a). In our experiment, however, the surface of the nanodot is not flat, but it is more truthfully described by a flattened half-ellipse with circular base of 200-nm radius and a shorter height of 140 nm. When scanning a point dipole across the nanodot, the point dipole projection onto its surface changes continuously. We have calculated the projection of a tip dipole ptip of a fixed orientation onto the surface as a function of the tip position r and have substituted the point dipole orientation u^ in Eq. (4) by this projection (more information on these calculations can be found in the Supplemental Materials). Figure 4(b) shows in a sketch how the projection onto the surface (the red arrow) increases as an x-oriented tip dipole is moved along the central x axis from the rim of the particle to its center. When scanning an x-oriented dipole along the central y axis, however, the projection onto the surface does not change. In total, the projected LDOS then resembles the LDOS shown in Fig. 4(a) but has minima at the left and right sides [see Fig. 4(c)]. Analogously, when scanning a y-oriented dipole across the nanodot, the projected LDOS appears rotated by 90 deg [Fig. 4(d)]. In contrast, a z-oriented tip dipole creates the largest projection close to the particle rim, resulting in the image shown in Fig. 4(e).

In our experiment, the light was incident from a direction in the xz plane, with an angle of 20  deg to the negative x axis and 70  deg to the z axis, with the field polarization along the y axis. However, as the tips are etched and then glued onto the tuning fork manually, their orientation relative to the sample is with some error margin (10  deg). Thus, an excitation of the z-oriented as well as the x-oriented tip dipoles can hardly be avoided due to uncertainties in the alignment of the tips, and an excitation of the three components on the diagonal of the polarizability tensor was considered. Finally, the radiation of the tip dipole, locally enhanced by its coupling to the waveguide modes into the far field (i.e., the detector, located at position Rdet in the direction of the incident light), is calculated to yield the simulated SNOM map shown in Fig. 4(f). The SNOM map displays a distinct radial asymmetry, which is likely to arise from the particular orientation of the tip dipole. The simulation using the projection of the tip dipole on an elliptical surface explains in particular the rather narrow central maximum in the SNOM map as well as the side lobes to the left and lower right and the near-zero signal strength at the upper and lower rim [compare S4f map in Fig. 4(g)]. Figure 4(h) compares the simulated (black curves) and measured S4f signals (black circles) along a horizontal and a vertical cut (upper and lower graph, respectively). These cuts have been calculated for the LDOS shown in Fig. 4(a), but for differently curved surfaces (indicated by the blue curves), in order to account for the slightly asymmetric shape of the nanoparticle. One can see that neither simulated nor measured values reach zero at the left and right rim of the nanodot (see x-cut), but they do so at the upper and lower rim (y-cut). Even though the simulations reproduce most of the experimental observations, some minor discrepancies remain. Specifically, the plateau formed around the center of the measured SNOM map and a very sharp ring around this central plateau are not well reproduced. These differences may be the effect of the finite disorder seen in Fig. 2(b) on the local field enhancement. In summary, the simulations show that the shape of the SNOM images resembles the projected LDOS given by the bound modes of a circular waveguide with a dome-like surface topography.

In the following, we will employ spectrally resolved near-field measurements of an Sb2S3 nanodot to study these modes in more detail. We can approach this in two ways. First, the waveguide modes proposed above provide a solution for a continuum of wavelengths. At a fixed spatial position, these modes can be excited across the complete spectral range covered by the broad-bandwidth laser, and the scattered spectra therefore should show negligible spectral variation in our experiments. This is different in case that resonant modes of the nanodot are excited, showing distinct spectral resonances. We will next present spectra recorded at different positions on the nanodot and the film to distinguish between both cases.

In order to verify that the BISNOM measurements probe a pure near-field contrast, we record near-field spectra during an approach over the flat Sb2S3 film. Figures 5(a) and 5(b) compare the spectra demodulated at the fundamental tip modulation frequency S1f(λ), and spectra demodulated at the fourth harmonic, S4f(λ), for a few tip–sample distances between 3 and 12 nm. The S1f(λ) spectrum [Fig. 5(a)] increases in intensity when the sample is removed from the tip, and the spectral shape changes. Both effects can be ascribed to interference of the reference field with a strong background signal scattered from the tip. In contrast, the S4f(λ) spectrum [Fig. 5(b)] decays strongly as the sample is removed from the tip; as the distance is increased by 12 nm, the signal mostly vanishes. During the process, the spectral shape is retained. In order to show that the measured S4f(λ) spectrum represents in good approximation the near-field spectrum, we have calculated the latter.

Fig. 5

Spectrally resolved approach curves measured above the flat Sb2S3 film. (a) The spectra recorded on the Sb2S3 film, demodulated at the fundamental tip modulation frequency S1f(λ) as a function of the tip–sample distance d. Due to background interference, the spectra change their shape as the tip is retracted. (b) The signal demodulated at the fourth harmonic S4f(λ) retains its shape and decreases rapidly as the tip–sample distance increases. (c) The effective tip polarizability calculated for the same tip–sample distances as used in the measurements; and (d) the calculated near-field spectra. The inset shows the reference spectrum that was measured in the experiment and used as the input spectrum for the simulations. The calculated near-field spectra are in good agreement with the measured S4f(λ) spectra.

AP_2_4_046004_f005.png

We model the tip polarizability in the x and y directions (perpendicular to the tip axis) by the polarizability of a gold sphere with radius R=10  nm. A dipole oscillation is excited by an incident field polarized in the y direction. Following Knoll and Keilmann,40 the tip dipole causes a polarization in the material. The field emitted by the image dipole acts back on the tip dipole, effectively enhancing the incident field. This is described by an effective polarizability, which increases in a super-linear fashion as the tip–sample distance is decreased. In Fig. 5(c), the yy matrix element of the effective polarizability (the only nonzero element under excitation with y-polarized light) is plotted as a function of wavelength for the same tip–sample distances d as were measured in the experiment and taking into account the dielectric functions of gold and of Sb2S3. One can see that the effective polarizability increases as the tip–sample distance decreases. The increase toward short wavelengths is caused by the tip resonance, which is far detuned to the blue. This spectral shape is preserved during the approach. We calculate the dipole field component that is radiated toward the detector and mix it with the reference field to yield the near-field spectrum SNF(λ,d) for the different distances d. Here, we have used the measured reference spectrum [shown in the inset of Fig. 5(d)] as both the reference and incident spectrum. The result is shown in Fig. 5(d). The agreement with the measured S4f spectra is excellent, both in spectral shape and in distance dependence.

We now perform a spectrally resolved BISNOM scan of an Sb2S3 nanoparticle using the monochromator and the fast line camera. The laser spectrum used for this particular measurement is the one shown in the setup sketch in Fig. 1. We record complete near-field spectra S1f(λ), S2f(λ), S3f(λ), and S4f(λ) across a 500-nm by 400-nm large area, and with a step size of 12.5 and 20 nm in the x and y directions, respectively. In Figs. 6(a)6(c), we show sets of these four spectra exemplarily for three different positions: in the center of the nanoparticle [Pos. 1, Fig. 6(a)], on the flat Sb2S3 film [Pos. 2, Fig. 6(b)], and on the outer area of the nanoparticle where the near-field signal is strongly reduced [Pos. 3, Fig. 6(c)]. The positions are marked in the SNOM map in the inset in Fig. 6(d), which has been prepared by plotting the spectrally integrated signal S4f(λ) as a function of position on the sample. Within each of the Figs. 6(a)6(c) one can see that the measured S2f(λ), S3f(λ), and S4f(λ) spectra closely resemble each other, while the spectral shape of S1f(λ) clearly deviates from these and furthermore varies between the positions. In agreement with the observation shown in Figs. 5(a) and 5(b), this leads us to conclude that the higher-order demodulated spectra reliably reflect the near-field spectra, while spectral interference of the background and reference fields leads to spectral variations in the S1f(λ) signal. In contrast, the higher-order demodulated spectra seem to hardly vary as the position on the sample is changed. In order to facilitate a direct comparison, the S4f(λ) spectra of the three positions are plotted together in Fig. 6(d). The black curve represents a near-field spectrum calculated from the effective polarizability above a flat Sb2S3 film and using the laser input spectrum shown in Fig. 1. As expected, the S4f(λ) spectrum recorded on the film (Pos. 2, dotted red curve) matches the calculated spectrum well. A similar agreement can be seen for the S4f(λ) spectrum recorded at the center of the nanodot (Pos. 1, dashed red curve). This supports our interpretation of the field profile as that of a mode, which has a broad bandwidth and is spectrally flat within our measurement range. Coupling to the nanodot is nearly wavelength-independent, which is reflected by the close resemblance of the calculated spectra and those measured on the film and the central area of the nanodot. The last spectrum has been measured in the outer area of the nanodot, where the over-all signal strength is markedly reduced (Pos. 3, solid red curve). For a better comparison with the other two spectra, this one is scaled by a factor 2.2. The slightly differing shape can thus be attributed to the lower signal-to-noise ratio at this position. Together, the measurements presented in Fig. 6 verify the assumption of the broad bandwidth and flat spectral shape of the observed fundamental optical mode of the nanodot. The spatial–spectral behavior is similar to that of guided wave modes. No evidence for resonant modes of the nanodot is found.

Fig. 6

Near-field spectra recorded at different positions on an Sb2S3 nanoparticle using the monochromator and fast line camera. (a) The demodulated signals S1f(λ) to S4f(λ) recorded with the tip in close contact above the center of the nanodot; (b) on the film; and (c) above the outer area on the nanodot. (d) Comparison of the three S4f(λ) spectra shown in (a)–(c) and a calculated S4f spectrum (black curve). The inset indicates the three positions where the spectra shown in this figure are recorded.

AP_2_4_046004_f006.png

As a final verification of the spatial–spectral mode properties, we extract spectral near-field maps from the BISNOM data. Since the waveguide modes are spectrally flat, we expect very similar near-field maps for different spectral bands, possibly increasing slightly in diameter with increasing wavelength, as known from fiber modes. In Figs. 7(a)7(g), we have plotted maps created by spectrally integrating the S4f signal over 30-nm wide spectral bands, centered at equally spaced wavelengths from 715 to 895 nm. The contrast in these SNOM maps remains high throughout these seven spectral bands, which is yet another implication of the negligible spectral variation of the waveguide modes. At the short- and the long-wavelength edges, the measurement noise is increased due to the lower spectral power density of the laser. In each of these maps, a ring-shaped signal decrease can be seen clearly. On closer inspection, the characteristic features from the projected LDOS calculation can be found here again, namely the secondary maxima at the left and lower right of the nanodot, and the near-zero signal regions at the top and bottom. Compared to the earlier investigated nanodot (cmp. Fig. 4), here the main maximum appears to be more centered, which indicates that the nanodot is more regularly shaped.

Fig. 7

Spectrally resolved near-field maps of an Sb2S3 nanodot. (a)–(g) Near-field maps recreated from 30-nm spectral bands centered at (a) 715 nm, (b) 745 nm, (c) 775 nm, (d) 805 nm, (e) 835 nm, (f) 865 nm, and (g) 895 nm. (h) Cross cuts through the near-field maps shown in (a) (spectral range 700 to 730 nm, blue symbols) and in (f) (spectral range 850 to 880 nm, red symbols) together with the calculated SNOM maps at 715 nm (blue curve) and 865 nm (red curve).

AP_2_4_046004_f007.png

Figure 7(h) compares cross cuts through the maps created at the outermost spectral bands at 715 and 865 nm [Figs. 7(a) and 7(f), respectively] with calculated SNOM measurements above an elliptical surface over the LP01, LP11, LP02, and LP21 mode profiles, once calculated for a wavelength of 715 nm and once for 865 nm. The measured values are shown as symbols in Fig. 7(h) and the calculated values as solid curves. As all curves are normalized, the larger extent of the long-wavelength mode is reflected by the slower decrease with a larger radius (red curve). The general shape of the cross cuts through the measured near-field maps agrees well with the simulated projected LDOS. The increase in diameter with wavelength predicted by the modeling cannot easily be seen in the measurements. Seeing this subtle effect is certainly made difficult by the finite signal-to-noise ratio measurement. In addition, it may be obscured by the effect of the structural disorder of the nanodot on its optical eigenmodes. In particular, the imperfections in the center of the dot seen in Fig. 2(b) are likely to affect the mode profile structure near the surface. Such disorder effects are likely to be seen on the left part of the nanodot center. Together, the spectrally resolved near-field measurements presented in Figs. 6 and 7 convincingly demonstrate spatial mapping of the mode profiles of the individual Sb2S3 nanodot. They show that the spatial mode profiles of these dots are similar to those of the lowest-order LP guided modes of the idealized cylindrical waveguide. No evidence for resonant cavity modes of the nanodot is found in these measurements. This suggests that backreflections at the interface between the nanodot and the supporting Sb2S3 film are too weak to result in Fabry–Perot-like standing modes of the dielectric nanodot cavity.

4.

Summary and Conclusions

We have realized a new interferometrically detected scanning near-field optical spectroscopy technique with a broad bandwidth in the near-infrared. The key to this method is a fast CCD line camera capable of recording complete spectra with a readout rate of more than eight times the tip modulation frequency such that near-field spectra demodulated at the fourth order can be obtained in postprocessing. Furthermore, we have incorporated the near-field microscope in a Michelson interferometer for homodyne mixing and boosting the weak near-field signal.

We have demonstrated the potential of broad-bandwidth interferometric near-field spectroscopy by experimentally investigating the optical modes supported by individual Sb2S3 nanodots deposited on an Sb2S3 thin film. The near-field measurements reveal mode profiles reminiscent of low-order optical modes typical for more idealized cylindrical waveguides. These modes are seen in near-field spectra across the entire bandwidth supported by our laser while we find no clear signatures for spectrally sharp Fabry–Perot resonances of the nanodots. These results show that the Sb2S3 nanodots can act as high-refractive-index dielectric waveguides with low losses in a broad spectral range below the bandgap. Together with their high above-bandgap absorption, their high photoconductivity, and the possibility for easy 2D and 3D structuring, this makes them promising candidates as switchable metamaterials. It will be highly interesting to investigate their nonlinear properties in future works.

Acknowledgments

We acknowledge funding by the Deutsche Forschungsgemeinschaft (SPP1391, SPP1839, GRK1885), the Niedersächsisches Ministerium für Wissenschaft und Kultur (LGRK, Nano-Energieforschung), the Korea Foundation for International Cooperation of Science and Technology (K20815000003), and the German-Israeli Foundation (1256). W. W. acknowledges financial support by the China Scholarship Council (CSC 201404910464).

References

1. 

R. Kondrotas, C. Chen and J. Tang, “Sb2S3 solar cells,” Joule, 2 (5), 857 –878 (2018). https://doi.org/10.1016/j.joule.2018.04.003 Google Scholar

2. 

M. Y. Versavel and J. A. Haber, “Structural and optical properties of amorphous and crystalline antimony sulfide thin-films,” Thin Solid Films, 515 (18), 7171 –7176 (2007). https://doi.org/10.1016/j.tsf.2007.03.043 THSFAP 0040-6090 Google Scholar

3. 

P. P. Boix et al., “Hole transport and recombination in all-solid Sb2S3-sensitized TiO2 solar cells using CuSCN as hole transporter,” J. Phys. Chem. C, 116 (1), 1579 –1587 (2012). https://doi.org/10.1021/jp210002c JPCCCK 1932-7447 Google Scholar

4. 

Y. C. Choi et al., “Highly improved Sb2S3 sensitized-inorganic–organic heterojunction solar cells and quantification of traps by deep-level transient spectroscopy,” Adv. Funct. Mater., 24 (23), 3587 –3592 (2014). https://doi.org/10.1002/adfm.201304238 AFMDC6 1616-301X Google Scholar

5. 

S. Messina, M. Nair and P. Nair, “Antimony sulfide thin films in chemically deposited thin film photovoltaic cells,” Thin Solid Films, 515 (15), 5777 –5782 (2007). https://doi.org/10.1016/j.tsf.2006.12.155 THSFAP 0040-6090 Google Scholar

6. 

J. A. Chang et al., “High-performance nanostructured inorganic—organic heterojunction solar cells,” Nano Lett., 10 (7), 2609 –2612 (2010). https://doi.org/10.1021/nl101322h NALEFD 1530-6984 Google Scholar

7. 

Y. C. Choi and S. I. Seok, “Efficient Sb2S3-sensitized solar cells via single-step deposition of Sb2S3 using S/Sb-ratio-controlled SbCl3-thiourea complex solution,” Adv. Funct. Mater., 25 (19), 2892 –2898 (2015). https://doi.org/10.1002/adfm.201500296 AFMDC6 1616-301X Google Scholar

8. 

D.-H. Kim et al., “Highly reproducible planar Sb2S3-sensitized solar cells based on atomic layer deposition,” Nanoscale, 6 (23), 14549 –14554 (2014). https://doi.org/10.1039/C4NR04148H NANOHL 2040-3364 Google Scholar

9. 

W. Wang, P. Pfeiffer and L. Schmidt-Mende, “Direct patterning of metal chalcogenide semiconductor materials,” Adv. Funct. Mater., 30 2002685 (2020). https://doi.org/10.1002/adfm.202002685 AFMDC6 1616-301X Google Scholar

10. 

M. Abulikemu et al., “Colloidal Sb2S3 nanocrystals: synthesis, characterization and fabrication of solid-state semiconductor sensitized solar cells,” J. Mater. Chem. A, 4 (18), 6809 –6814 (2016). https://doi.org/10.1039/C5TA09546H Google Scholar

11. 

W. Wang et al., “Hybrid solar cells from Sb2S3 nanoparticle ink,” Sol. Energ. Mat. Sol. C, 172 335 –340 (2017). https://doi.org/10.1016/j.solmat.2017.07.046 SEMCEQ 0927-0248 Google Scholar

12. 

M. R. Shcherbakov et al., “Ultrafast all-optical tuning of direct-gap semiconductor metasurfaces,” Nat. Commun., 8 17 (2017). https://doi.org/10.1038/s41467-017-00019-3 NCAOBW 2041-1723 Google Scholar

13. 

L. Novotny and B. Hecht, Principles of Nano-Optics, Cambridge University Press, Cambridge (2012). Google Scholar

14. 

T. Taubner, R. Hillenbrand and F. Keilmann, “Performance of visible and mid-infrared scattering-type near-field optical microscopes,” J. Microsc., 210 (3), 311 –314 (2003). https://doi.org/10.1046/j.1365-2818.2003.01164.x JMICAR 0022-2720 Google Scholar

15. 

F. Zenhausern, Y. Martin and H. Wickramasinghe, “Scanning interferometric apertureless microscopy: optical imaging at 10 angstrom resolution,” Science, 269 (5227), 1083 –1085 (1995). https://doi.org/10.1126/science.269.5227.1083 SCIEAS 0036-8075 Google Scholar

16. 

M. Balistreri et al., “Tracking femtosecond laser pulses in space and time,” Science, 294 (5544), 1080 –1082 (2001). https://doi.org/10.1126/science.1065163 SCIEAS 0036-8075 Google Scholar

17. 

S. Götzinger et al., “Mapping and manipulating whispering gallery modes of a microsphere resonator with a near-field probe,” J. Microsc., 202 (1), 117 –121 (2001). https://doi.org/10.1046/j.1365-2818.2001.00865.x JMICAR 0022-2720 Google Scholar

18. 

F. Intonti et al., “Spectral tuning and near-field imaging of photonic crystal microcavities,” Phys. Rev. B, 78 (4), 041401 (2008). https://doi.org/10.1103/PhysRevB.78.041401 Google Scholar

19. 

R. Esteban et al., “Direct near-field optical imaging of higher order plasmonic resonances,” Nano Lett., 8 (10), 3155 –3159 (2008). https://doi.org/10.1021/nl801396r NALEFD 1530-6984 Google Scholar

20. 

K. Imaeda, S. Hasegawa and K. Imura, “Imaging of plasmonic eigen modes in gold triangular mesoplates by near-field optical microscopy,” J. Phys. Chem. C, 122 (13), 7399 –7409 (2018). https://doi.org/10.1021/acs.jpcc.8b00678 JPCCCK 1932-7447 Google Scholar

21. 

K. Imura, T. Nagahara and H. Okamoto, “Imaging of surface plasmon and ultrafast dynamics in gold nanorods by near-field microscopy,” J. Phys. Chem. B, 108 (42), 16344 –16347 (2004). https://doi.org/10.1021/jp047950h JPCBFK 1520-6106 Google Scholar

22. 

R. L. Olmon et al., “Near-field imaging of optical antenna modes in the mid-infrared,” Opt. Express, 16 (25), 20295 –20305 (2008). https://doi.org/10.1364/OE.16.020295 OPEXFF 1094-4087 Google Scholar

23. 

M. Rang et al., “Optical near-field mapping of plasmonic nanoprisms,” Nano Lett., 8 (10), 3357 –3363 (2008). https://doi.org/10.1021/nl801808b NALEFD 1530-6984 Google Scholar

24. 

A. Y. Frolov et al., “Near-field mapping of optical Fabry–Perot modes in all-dielectric nanoantennas,” Nano Lett., 17 (12), 7629 –7637 (2017). https://doi.org/10.1021/acs.nanolett.7b03624 NALEFD 1530-6984 Google Scholar

25. 

T. G. Habteyes et al., “Near-field mapping of optical modes on all-dielectric silicon nanodisks,” ACS Photonics, 2 (9), 794 –798 (2014). https://doi.org/10.1021/ph500232u Google Scholar

26. 

R. Carminati et al., “Electromagnetic density of states in complex plasmonic systems,” Surf. Sci. Rep., 70 (1), 1 –41 (2015). https://doi.org/10.1016/j.surfrep.2014.11.001 SSREDI 0167-5729 Google Scholar

27. 

G. C. des Francs et al., “Relationship between scanning near-field optical images and local density of photonic states,” Chem. Phys. Lett., 345 (5–6), 512 –516 (2001). https://doi.org/10.1016/S0009-2614(01)00914-9 CHPLBC 0009-2614 Google Scholar

28. 

K. Joulain et al., “Definition and measurement of the local density of electromagnetic states close to an interface,” Phys. Rev. B, 68 (24), 245405 (2003). https://doi.org/10.1103/PhysRevB.68.245405 Google Scholar

29. 

V. Krachmalnicoff et al., “Towards a full characterization of a plasmonic nanostructure with a fluorescent near-field probe,” Opt. Express, 21 (9), 11536 –11545 (2013). https://doi.org/10.1364/OE.21.011536 OPEXFF 1094-4087 Google Scholar

30. 

C. Chicanne et al., “Imaging the local density of states of optical corrals,” Phys. Rev. Lett., 88 (9), 097402 (2002). https://doi.org/10.1103/PhysRevLett.88.097402 PRLTAO 0031-9007 Google Scholar

31. 

M. Esmann et al., “Plasmonic nanofocusing spectral interferometry,” Nanophotonics, 9 (2), 491 –508 (2020). https://doi.org/10.1515/nanoph-2019-0397 Google Scholar

32. 

E. Bailo and V. Deckert, “Tip-enhanced Raman spectroscopy of single RNA strands: towards a novel direct-sequencing method,” Angew. Chem. Int. Ed., 47 (9), 1658 –1661 (2008). https://doi.org/10.1002/anie.200704054 Google Scholar

33. 

S. F. Becker et al., “Gap-plasmon-enhanced nanofocusing near-field microscopy,” ACS Photonics, 3 (2), 223 –232 (2016). https://doi.org/10.1021/acsphotonics.5b00438 Google Scholar

34. 

R. Zhang et al., “Chemical mapping of a single molecule by plasmon-enhanced Raman scattering,” Nature, 498 82 –86 (2013). https://doi.org/10.1038/nature12151 Google Scholar

35. 

M. Esmann et al., “Vectorial near-field coupling,” Nat. Nanotechnol., 14 (7), 698 –704 (2019). https://doi.org/10.1038/s41565-019-0441-y NNAABX 1748-3387 Google Scholar

36. 

C. C. Neacsu et al., “Near-field localization in plasmonic superfocusing: a nanoemitter on a tip,” Nano Lett., 10 (2), 592 –596 (2010). https://doi.org/10.1021/nl903574a NALEFD 1530-6984 Google Scholar

37. 

C. Ropers et al., “Grating-coupling of surface plasmons onto metallic tips: a nanoconfined light source,” Nano Lett., 7 (9), 2784 –2788 (2007). https://doi.org/10.1021/nl071340m NALEFD 1530-6984 Google Scholar

38. 

E. Ash and G. Nicholls, “Super-resolution aperture scanning microscope,” Nature, 237 510 –512 (1972). https://doi.org/10.1038/237510a0 Google Scholar

39. 

R. Hillenbrand, B. Knoll and F. Keilmann, “Pure optical contrast in scattering-type scanning near-field microscopy,” J. Microsc., 202 (1), 77 –83 (2001). https://doi.org/10.1046/j.1365-2818.2001.00794.x JMICAR 0022-2720 Google Scholar

40. 

B. Knoll and F. Keilmann, “Enhanced dielectric contrast in scattering-type scanning near-field optical microscopy,” Opt. Commun., 182 (4–6), 321 –328 (2000). https://doi.org/10.1016/S0030-4018(00)00826-9 OPCOB8 0030-4018 Google Scholar

41. 

N. Ocelic, A. Huber and R. Hillenbrand, “Pseudoheterodyne detection for background-free near-field spectroscopy,” Appl. Phys. Lett., 89 (10), 101124 (2006). https://doi.org/10.1063/1.2348781 APPLAB 0003-6951 Google Scholar

42. 

Y. Sasaki and H. Sasaki, “Heterodyne detection for the extraction of the probe-scattering signal in scattering-type scanning near-field optical microscope,” Jpn. J. Appl. Phys., 39 L321 –L323 (2000). https://doi.org/10.1143/JJAP.39.L321 Google Scholar

43. 

R. Hillenbrand and F. Keilmann, “Complex optical constants on a subwavelength scale,” Phys. Rev. Lett., 85 (14), 3029 –3032 (2000). https://doi.org/10.1103/PhysRevLett.85.3029 PRLTAO 0031-9007 Google Scholar

44. 

J. Brauer et al., “In-line interferometer for broadband near-field scanning optical spectroscopy,” Opt. Express, 25 (13), 15504 –15525 (2017). https://doi.org/10.1364/OE.25.015504 OPEXFF 1094-4087 Google Scholar

45. 

D. Gloge, “Weakly guiding fibers,” Appl. Opt., 10 (10), 2252 –2258 (1971). https://doi.org/10.1364/AO.10.002252 APOPAI 0003-6935 Google Scholar

46. 

F. Mitschke, Fiber Optics, Springer-Verlag, Berlin (2016). Google Scholar

47. 

X. Wang et al., “A fast chemical approach towards Sb2S3 film with a large grain size for high-performance planar heterojunction solar cells,” Nanoscale, 9 (10), 3386 –3390 (2017). https://doi.org/10.1039/C7NR00154A NANOHL 2040-3364 Google Scholar

48. 

S. Schmidt et al., “Adiabatic nanofocusing on ultrasmooth single-crystalline gold tapers creates a 10-nm-sized light source with few-cycle time resolution,” ACS Nano, 6 (7), 6040 –6048 (2012). https://doi.org/10.1021/nn301121h ANCAC3 1936-0851 Google Scholar

49. 

R. L. Olmon et al., “Optical dielectric function of gold,” Phys. Rev. B, 86 (23), 235147 (2012). https://doi.org/10.1103/PhysRevB.86.235147 Google Scholar

50. 

P. Groß et al., “Plasmonic nanofocusing–grey holes for light,” Adv. Phys. X, 2 (2), 297 –330 (2016). https://doi.org/10.1080/23746149.2016.1177469 Google Scholar

51. 

N. Talebi et al., “Excitation of mesoscopic plasmonic tapers by relativistic electrons: phase matching versus eigenmode resonances,” ACS Nano, 9 (7), 7641 –7648 (2015). https://doi.org/10.1021/acsnano.5b03024 ANCAC3 1936-0851 Google Scholar

52. 

B. Piglosiewicz et al., “Carrier-envelope phase effects on the strong-field photoemission of electrons from metallic nanostructures,” Nat. Photonics, 8 (1), 37 –42 (2014). https://doi.org/10.1038/nphoton.2013.288 NPAHBY 1749-4885 Google Scholar

54. 

S. Grésillon et al., “Experimental observation of localized optical excitations in random metal-dielectric films,” Phys. Rev. Lett., 82 (22), 4520 –4523 (1999). https://doi.org/10.1103/PhysRevLett.82.4520 PRLTAO 0031-9007 Google Scholar

55. 

J. Zhong et al., “Strong spatial and spectral localization of surface plasmons in individual randomly disordered gold nanosponges,” Nano Lett., 18 (8), 4957 –4964 (2018). https://doi.org/10.1021/acs.nanolett.8b01785 NALEFD 1530-6984 Google Scholar

Biography

Jinxin Zhan is a PhD student in the Near-Field Photonics Laboratory of Professor Christoph Lienau at University of Oldenburg. Her current interests are focused on broadband interferometric homodyne scanning near-field spectroscopy to study optical resonances of individual nanostructures and coupled plasmonic systems. She graduated in June 2014 at the Shanghai Institute of Optics & Fine Mechanics, Chinese Academy of Sciences.

Wei Wang received his PhD from the Physics Department of Konstanz University in 2019. His works have been focused on fabrication of semiconductor nanostructures, such as nanoparticle, 2D/3D periodic patterns for optoelectronic and photovoltaic applications.

Jens Brauer works as a software developer for Bertrandt and BMW in Munich, Germany. He graduated from University of Oldenburg in 2013 and will defend his PhD thesis in 2020. His research interests are nano optics and near-field imaging.

Lukas Schmidt-Mende: Biography is not available.

Christoph Lienau is a professor in experimental physics at the University of Oldenburg. After receiving his PhD in physical chemistry in Göttingen, he worked with Ahmed H. Zewail at Caltech and then as a scientific staff member at Max Born Institute. In 2006, he became a full professor in physics. He has authored more than 200 publications and is a fellow of OSA. His research interests are in ultrafast, nano, and quantum optics.

Petra Groß is a researcher at the Institute for Physics at University of Oldenburg. She received her PhD from University of Twente in 2003 and her postdoctoral lecture qualification from University of Münster in 2014. Her research interests are in imaging, ultrafast, nonlinear and nano optics, and ultrafast electron microscopy.

CC BY: © The Authors. Published by SPIE under a Creative Commons Attribution 4.0 Unported License. Distribution or reproduction of this work in whole or in part requires full attribution of the original publication, including its DOI.
Jinxin Zhan, Wei Wang, Jens H. Brauer, Lukas Schmidt-Mende, Christoph Lienau, and Petra Groß "Spatial and spectral mode mapping of a dielectric nanodot by broadband interferometric homodyne scanning near-field spectroscopy," Advanced Photonics 2(4), 046004 (25 August 2020). https://doi.org/10.1117/1.AP.2.4.046004
Received: 2 April 2020; Accepted: 5 August 2020; Published: 25 August 2020
Lens.org Logo
CITATIONS
Cited by 5 scholarly publications.
Advertisement
Advertisement
KEYWORDS
Near field

Antimony

Nanoparticles

Near field scanning optical microscopy

Dielectrics

Waveguides

Near field optics

Back to Top