Gratings are widely used for coupling into and out of evanescent and propagating electromagnetic modes, which are otherwise not accessible due to their large in-plane wave vector. A precise description of the optical response requires the knowledge of the grating geometry. Here, we present an investigation of the optical properties of dielectric gratings of sub-micron periodicity fabricated on a multilayer structure, which supports Bloch surface waves, by means of spectroscopic ellipsometry. Taking into account non-idealities, such as the finite spectral bandwidth, in the modeling process is shown to be a necessity for an accurate description of the observed spectra. The grating geometry determined from the analysis of ellipsometry data agrees very well with atomic force microscopy scans. Thus, our ellipsometric model is corroborated.

Guided photonic modes in transparent structures have long inherent propagation lengths due to their evanescent nature outside of the allowed regions, i.e., total internal reflection at the interfaces or photonic bandgaps, which can be exploited for low-loss signal transport1 and integrated optics.2 In planar geometries, their dispersion lies beyond the light line, E p h = c k , thus they cannot couple to the ambient field. Therefore, many approaches for coupling light in photonic waveguides have been developed.3 One approach that allows to excite and detect such modes is the use of (one-dimensional) diffraction gratings.4 In fact, by a suitable choice of grating dimensions, this mechanism effectively enlarges the in-plane component of the wavevector such that the guided modes can be coupled to vacuum modes, which is otherwise known as a type of Rayleigh–Woods anomaly.5 In the case of shallow gratings (height λ) the wavevector shift is approximately linear and depends on the periodicity L of the grooves.

Efficient coupling to guided modes using diffraction gratings typically requires a groove spacing similar to the wavelength, which can make an optical characterization using microscopy rather challenging. An alternative for grating characterization is the measurement and description of its optical response and an analysis of the corresponding optical modes. Toward this end, a precise knowledge of the optical properties of the involved materials and the geometrical properties of the layer is necessary. Spectroscopic ellipsometry (SE) is a well-established technique for the determination of optical properties and the layer structure of samples,6 though it is typically used for planar stratified media.

Recently, ellipsometry was used in order to investigate gratings on high quality GaAs,7 Si,8–10 and other substrates.11 Furthermore, the spectral bandwidth of the measurement instrument is often neglected while only few studies take this into account12,13 or found it to have only a small effect.14 Compared to thin film samples and substrates, multilayer samples with shallow gratings exhibit a high number of very narrow modes that are often absent in experimental data.15 

Here, we demonstrate the investigation of the optical response of shallow dielectric gratings on a multilayer structure, which consists of a distributed Bragg reflector (DBR) with a top layer that supports the formation of Bloch surface waves (BSWs), a photonic guided mode confined to the sample surface.16 These modes, which are associated with, e.g., long-range lateral propagation17 and exceptional sensitivity,18–20 can be excited using gratings. We show that even for such a complex system spectroscopic ellipsometry allows us to accurately determine the geometrical properties of the grating, which agree very well with those obtained from atomic force microscopy (AFM).

The sample was produced using pulsed laser deposition (PLD). The laser (Lambda Physik LPX Pro, excitation wavelength of 248 nm, pulse energy of 600 mJ) is focused on a (rotating) target material, creating a plasma plume that expands toward a c-sapphire substrate (Crystec, 10 × 10 mm 2) for layer deposition. The multilayer was grown at room temperature for a smooth surface and consists of a 150 nm yttria-stabilized zirconia (YSZ) top layer followed by a DBR made of 7 layer pairs of 90 nm Al 2 O 3 and 72 nm YSZ. The oxygen partial pressure and pulse repetition rate were 0.02 Pa and 15 Hz for the Al 2 O 3 layers and 0.2 Pa and 15 Hz for the YSZ layers.

The fabrication of nanostructures in dielectric or insulating materials, like YSZ, is a challenging task and was done here by producing a gold hard mask using focused ion beam (FIB) milling and subsequent dry etching. The Au layer, which also reduced FIB related charging effects,21 had a thickness of 150 nm and was produced by sputter deposition (30 W, 90 s, 100 SCCM Ar). For FIB milling (FEI Nova 200 NanoLab), we used a Ga +-beam current of 100 pA at 30 kV, dwell time 1 μ s, dose per volume 200 pC/ μ m 2, and magnification 3000 ×. Next, the sample was dry-etched at 15 °C using an inductively coupled plasma (ICP) with 30 SCCM CF 4 at 1.3 Pa and 150 W and 300 W HF and ICP power, respectively. This resulted in an etch rate of approximately 0.6 nm/s. After this process, residual gold was removed in a potassium-iodide (KI) solution bath (40 g I  + 80 g KI in 600 ml H 2O). Three gratings of different nominal geometries were fabricated, i.e., L { 400 , 400 , 450 } nm and h { 30 , 40 , 40 } nm. In all cases, the grating region spanned 50 × 50 μ m 2.

The gratings were pre-characterized using atomic force microscopy in a contact mode (XE 150) with scan areas ( 1.5 × 1.5 μ m 2, 512 × 512 pixels) and a 3XC-NA cantilever with an aluminum-coated tip and spring constant of 9 N m 1. As the sample is highly insulating, significant drift was observed during the scans, which was corrected using a b-spline interpolation implemented in Gwyddion software.22 

Reflectivity maps R ( E , θ ) were obtained by illuminating the sample with collimated light from a Xe-lamp. An image of the reflected light in the back focal plane of a Mitutoyo objective ( 50 ×, NA = 0.4) was dispersed in a spectrometer (Horiba Jobin Yvon iHR320, 300 grooves/mm) and recorded with a CCD camera (Horiba Jobin Yvon Symphony Open STE). The reflectivity data were calibrated using a Si wafer with a 8 nm oxide layer.

For sensitive measurements that allow reliable quantitative analysis of sample geometries, it is preferable to measure spectroscopic ellipsometry as it is based on measuring the ratio of reflection coefficients, i.e., ρ, instead of reflection intensities, which are prone to large experimental uncertainties and do not take into account phase information. Thus, we measured spectroscopic ellipsometry data for the grating region with a commercially available imaging ellipsometer (Accurion EP4) averaging the spectra over the patterned area in the spectral range of 2–4 eV (310–625 nm) with an energy spacing of 10 meV. We used an UV objective ( 7 × magnification, NA = 0.2). This results in a spatial resolution of 1 μ m, which is larger than L, and thus the averaged optical response of the grating was determined. Furthermore, the angular spread of ± 1 ° is negligible compared to the influence of the bandwidth and a focus scan was performed over the grating areas. A knife edge was placed in the beam path to suppress the signal from reflections at the backside of the sample. Measurements were taken at 45 ° and 55 ° angle of incidence (AOI) and the measured region of interest was adjusted to match the grating area for each angle, individually. Furthermore, the finite bandwidth of the instrument was taken into account, which is described in more detail below.

The dielectric function (DF) of the different materials was determined beforehand for single thin films of comparable thickness as in the multilayer, which were deposited under similar deposition parameters, using a commercial dual rotating-compensator spectroscopic ellipsometer (RC2 of J. A. Woollam company, 193–1690 nm). In the spectral range considered in this work, the ellipsometer has a wavelength spacing of 1 nm and bandwidth <2.5 nm and the materials can be sufficiently described by a transparent Cauchy function, i.e.,
(1)
where A, B, and C are the Cauchy parameters. The corresponding parameters are summarized in Table I. The film thicknesses (of the unpatterned sample) were determined using a commercial spectroscopic ellipsometer.
TABLE I.

Cauchy parameters for the model DF of the different materials.

Cauchy parametersAB (μm2)C (μm4)
YSZ 2.130 1.5 × 10−2 7.0 × 10−4 
Al2O3 1.663 4.1 × 10−3 1.1 × 10−4 
Sapphire 1.753 5.9 × 10−3 2.8 × 10−5 
Cauchy parametersAB (μm2)C (μm4)
YSZ 2.130 1.5 × 10−2 7.0 × 10−4 
Al2O3 1.663 4.1 × 10−3 1.1 × 10−4 
Sapphire 1.753 5.9 × 10−3 2.8 × 10−5 
In order to model the effect of the grating layer, the whole system was described in terms of rigorous coupled-wave analysis (RCWA), which we implemented in MATLAB following Ref. 5. Within this method, a matrix representation of Maxwell’s equations is used, where the fields are expanded in terms of a finite sum of N harmonics, i.e., exp [ i r ( k + G m ) ]. Here, r , k , and G m correspond to the in-plane components of the real space vector, the wavevector, and the reciprocal lattice vector, respectively. The wavevector is connected to the photon energy, E, and angle of incidence θ by
(2)
and in the case of a one-dimensional grating of periodicity ( L), the lattice vector is given by
(3)
where m denotes the diffraction order. The minimum number of N, which has to be taken into account, depends on the specific structure.11 In this work, N = 12 was sufficient to reach convergence.
The dielectric function of a rectangular profile in reciprocal space reads
(4)
which couples different diffraction orders m , m . Here, ε g and ε r correspond to the DF of the grooves and ridges, respectively, and d denotes the width of the grooves (cf. Fig. 1).
FIG. 1.

A schematic of the sample geometry (not to scale). The coordinate system defined by the plane of incidence is given by x , y , and z, compared to x , y, and z defined by the grating. Note that both systems share the same z axis and the rotation around the z axis is described by the angle φ.

FIG. 1.

A schematic of the sample geometry (not to scale). The coordinate system defined by the plane of incidence is given by x , y , and z, compared to x , y, and z defined by the grating. Note that both systems share the same z axis and the rotation around the z axis is described by the angle φ.

Close modal
The optical response of the system can then be described in terms of a ( 4 N × 4 N ) scattering matrix S _ _ which connects the incoming and outgoing amplitudes of electromagnetic waves in the ambient, A a i and A a o, and substrate, A s i and A s o, i.e.,
(5)
The in-plane components of the electric field, E = ( E x , G 1 E x , G N E y , G 1 E y , G N ) T and equivalently of the magnetic field, H , in the jth layer are connected to the amplitudes by a ( 4 N × 4 N ) matrix F _ _ j via
(6)
Consequently, this also allows the determination of the out-of plane components. For specular reflection ( m = 0), the reflection coefficients of the transverse electric (s) and transverse magnetic (p) polarization can be derived from the ratio of the reflected and incoming amplitudes in the ambient ( j = 0) by computing
(7)
Note that the reflection coefficients contain contributions from several diffraction orders m. This allows us to calculate the spectroscopic angles ( Ψ , Δ ) defined by
(8)
As described in Ref. 5, one type of Rayleigh–Woods anomaly appears as sharp dips in the transmission spectra (peaks in the reflectivity), corresponding to quasi-guided modes. They arise due to diffractive coupling of evanescent to the vacuum modes, i.e., the aforementioned enhancement of the wavevector k , in the presence of the grating. For shallow gratings, this situation can be described approximately by
(9)

In the formalism presented above, the response at a given wavelength is not affected by the neighbored one. However, experimental setups exhibit a finite spectral bandwidth ( δ E) so that the measured response is also affected by the response of the neighbored energies. This plays an important role if the feature response is smaller than the spectral bandwidth. In this case, it will cause an additional broadening in the optical response spectrum. In the following, we expand the numerical method presented above to include the spectral bandwidth, δ E, mainly caused by the monochromator, arising from the finite width of its entrance slit. For the instrument used here, δ E as a function of energy is shown in Fig. 2(a).

FIG. 2.

(a) Bandwidth of the ellipsometer in meV as a function of photon energy provided for each measurement by the device. The jump above 3 eV can be attributed to a change of the detector. (b) Gaussian weights w i for different numbers of points used in the convolution.

FIG. 2.

(a) Bandwidth of the ellipsometer in meV as a function of photon energy provided for each measurement by the device. The jump above 3 eV can be attributed to a change of the detector. (b) Gaussian weights w i for different numbers of points used in the convolution.

Close modal
The effect of the spectral bandwidth is taken into account by performing a convolution. In doing so, n w points are equidistantly distributed in the energy range [ E 0 δ E , E 0 + δ E ], and each point is connected to a weight factor w i. These factors are normalized in such a way that i n w w i = 1 holds and describes the magnitude of the optical response at this energy to the entire spectrum at E 0. The measured quantity f ( E ) is then given by
(10)
In our case, the determined quantities f ( E ) are Ψ and Δ and as weighting function we use the Gaussian function. It should be noted that the bandwidth is a source for depolarization, which for this system is < 10 % for most of the spectral range, but reaches up to 40 % at narrow resonances, caused by the presence of the grating, cf. supplementary material. This effect is considered in our model and the experimental data is valid for depolarizing samples, though it is not possible to measure the degree of depolarization directly with our device.

Furthermore, the number of points ( n w) should be large enough for the convolution to obtain a smooth curve, i.e., the spectral point spacing should be smaller than the mode broadening, cf. Fig. 2(b). However, too many points will strongly increase the computation time and will not change the calculated spectrum. Here, n w = 25 points were sufficient for a good description of the spectra. Note that computation is not increased by a factor n w as there is a considerable overlap of the distribution of wavelengths for neighboring sampling energies, and observables for these wavelengths have to be computed only once.

For the modeling of the samples patterned with a grating, all film thicknesses were taken from the analysis obtained before the patterning process, except for the layer containing the grating ( d S) as initial parameters. For this layer, we allow a thickness variation of about ± 5 nm, due to local inhomogeneities. Consequently, free parameters of the model were L, η, h, d S, and φ, an azimuthal rotation of the sample orientation (cf. Fig. 1).

The AFM topography for the different gratings with nominal pitch L = 400 nm and two different heights and L = 450 nm is shown in Fig. 3. The relevant geometric parameters were extracted as follows: L is given as the distance between periodic peaks of the first derivative, h as the distance between peaks of the height histogram with bin sizes of 1 nm and η as the ratio of the number values below half the peak-to-valley maximum and the number of pixels per row. The obtained values are summarized in Table II. These results reveal the successful fabrication of homogeneous gratings with nearly rectangular cross section, though slightly slanted sidewalls, as well as a smooth ridges and grooves in the sample surface. The absolute median deviation is less than 5 %, except for the first groove of the grating with L = 450 nm [Fig. 3(c)]. This we attribute to droplets in the mask or the grating. Note, we suspect that the sharp features at the edge of the ridges are measurement artifacts as they arise on the other side of the ridges for the opposite scan direction.

FIG. 3.

AFM scans in 1.5 × 1.5 μ m 2 areas of the gratings with nominal pitch and height L = 400 nm, h = 30 nm in (a), L = 400 nm, h = 40 nm in (b), L = 450 nm, h = 40 nm in (c). The respective height profiles along the x-direction in (d)–(f). The gray area represents the obtained maximum and minimum values, whereas the black solid line corresponds to the median.

FIG. 3.

AFM scans in 1.5 × 1.5 μ m 2 areas of the gratings with nominal pitch and height L = 400 nm, h = 30 nm in (a), L = 400 nm, h = 40 nm in (b), L = 450 nm, h = 40 nm in (c). The respective height profiles along the x-direction in (d)–(f). The gray area represents the obtained maximum and minimum values, whereas the black solid line corresponds to the median.

Close modal
TABLE II.

Model fit parameters and geometric grating parameters as obtained from spectroscopic ellipsometry, also when excluding the bandwidth in the modeling (δE = 0) and AFM measurements. The uncertainties in the ellipsometric data were estimated by considering 10% deviation from the best mean squared error. The more shallow grating is denoted by *.

Nom. L (nm)400*400450
L (nm) AFM 409 ± 3 408 ± 4 459 ± 4 
 SE 404 ± 1 402 ± 1 450 ± 2 
 SE (δE = 0) 399 ± 2 398 ± 2 450 ± 2 
h (nm) AFM 28 ± 2 45 ± 3 43 ± 3 
 SE 30 ± 2 41 ± 2 39 ± 2 
 SE (δE = 0) 26 ± 2 36 ± 3 34 ± 5 
η AFM 0.26 ± 0.06 0.29 ± 0.06 0.30 ± 0.05 
 SE 0.25 ± 0.02 0.30 ± 0.01 0.33 ± 0.02 
 SE (δE = 0) 0.21 ± 0.02 0.28 ± 0.02 0.33 ± 0.04 
φ(°) SE 7 ± 2 6 ± 1 7 ± 1 
 SE (δE = 0) 9 ± 1 9 ± 4 7 ± 1 
dS (nm) SE 149 ± 1 150 ± 1 150 ± 1 
 SE (δE = 0) 148 ± 1 150 ± 2 150 ± 2 
Nom. L (nm)400*400450
L (nm) AFM 409 ± 3 408 ± 4 459 ± 4 
 SE 404 ± 1 402 ± 1 450 ± 2 
 SE (δE = 0) 399 ± 2 398 ± 2 450 ± 2 
h (nm) AFM 28 ± 2 45 ± 3 43 ± 3 
 SE 30 ± 2 41 ± 2 39 ± 2 
 SE (δE = 0) 26 ± 2 36 ± 3 34 ± 5 
η AFM 0.26 ± 0.06 0.29 ± 0.06 0.30 ± 0.05 
 SE 0.25 ± 0.02 0.30 ± 0.01 0.33 ± 0.02 
 SE (δE = 0) 0.21 ± 0.02 0.28 ± 0.02 0.33 ± 0.04 
φ(°) SE 7 ± 2 6 ± 1 7 ± 1 
 SE (δE = 0) 9 ± 1 9 ± 4 7 ± 1 
dS (nm) SE 149 ± 1 150 ± 1 150 ± 1 
 SE (δE = 0) 148 ± 1 150 ± 2 150 ± 2 

As mentioned above, the electromagnetic field of the BSW mode decays on both sides of the surface interface at z = 0 nm, which is exemplarily shown in Fig. 4(a). Using diffraction gratings allows these modes to couple to the vacuum states since the dispersion is shifted inside the vacuum light cone, bounded by E = c k . Hence, BSW can be observed as peaks in reflection spectra for finite θ. In Figs. 4(b) and 4(c), the measured and simulated TE-polarized reflectivity maps are shown for the structure patterned with a grating with L = 400 nm and very good qualitative agreement is found. The dispersion of the BSW, which is symmetric around θ = 0 ° corresponding to k < 0 and k > 0, can be seen for diffraction orders m = 1 and m = 2. At the crossing points of the dispersions, an anti-crossing behavior can be found, most prominently for m = 1 at θ = 0 °, which agrees with findings in the literature, e.g., Ref. 23. The width of the resulting spectral gap depends on the geometry of the grating, especially on η and h and its position is determined mostly by L 1. A comparison of the reflectivity maps for different grating geometries can be seen in the supplementary material.

FIG. 4.

(a) Exponentially decaying field distribution of BSW superimposed on the dielectric function profile (for fixed energy) of the sample, (b) and (c) show measured and simulated reflectivity maps of the grating area with L = 400 nm and h = 30 nm, respectively. Narrow dispersion lines of high reflectivity correspond to quasi-BSW modes of different diffraction order m.

FIG. 4.

(a) Exponentially decaying field distribution of BSW superimposed on the dielectric function profile (for fixed energy) of the sample, (b) and (c) show measured and simulated reflectivity maps of the grating area with L = 400 nm and h = 30 nm, respectively. Narrow dispersion lines of high reflectivity correspond to quasi-BSW modes of different diffraction order m.

Close modal

In order to gain a deeper insight into the optical properties, e.g., layer thicknesses and change of the polarization, we performed spectroscopic ellipsometry. In Fig. 5, the Ψ and Δ spectra for the unpatterned surface as well as for those patterned with different gratings for an angle of incidence of 55 ° are shown, respectively. The data for angle of incidence of 45 ° can be found in the supplementary material. For the analysis, the spectra measured at an angle of incidence of 45 ° and 55 ° were analyzed simultaneously.

FIG. 5.

Ψ [(a)–(d)] and Δ [(e)–(h)] spectra for different lattice constants and grating geometries AOI = 55 °. Open symbols correspond to experimental data and are scaled in size with the instrument bandwidth, green solid line marks the model fit including the bandwidth, while the red line corresponds to the same spectrum computed without bandwidth, i.e., δ E = 0. Note that L = 400 nm corresponds to the more shallow grating.

FIG. 5.

Ψ [(a)–(d)] and Δ [(e)–(h)] spectra for different lattice constants and grating geometries AOI = 55 °. Open symbols correspond to experimental data and are scaled in size with the instrument bandwidth, green solid line marks the model fit including the bandwidth, while the red line corresponds to the same spectrum computed without bandwidth, i.e., δ E = 0. Note that L = 400 nm corresponds to the more shallow grating.

Close modal

In Figs. 5(a) and 5(e), the model data (green line) are displayed along with the experimental data (black circles whose size grows proportionally to the bandwidth δ E) for the unpatterned sample, which follows the well-known behavior of a DBR. The stop band with a central photon energy around 2 eV has a width of around 0.4 eV. Outside the stop band, typical Fabry–Pérot oscillations can be seen, which are due to multiple reflections at layer interfaces and damped with increasing absorption in the system for increasing photon energies.

In the spatial region patterned with a grating, additional narrow peaks are observable in the spectrum. These peaks arise due to a coupling of the evanescent modes to the vacuum states. They correspond to guided modes, analogous to the Fabry–Pérot oscillations mentioned above, and surface modes, both of which lie outside the light line. For the three different gratings, Ψ , Δ spectra are shown in Figs. 5(b)5(d) and Figs. 5(f)5(h), respectively. Note, due to the broken rotational symmetry of our sample in presence of the gratings, mode conversion takes place in general. However, we arranged the sample such that the grating lines are almost perpendicular to the plane of incidence, i.e., ϕ 0 °. Therefore, this effect can be neglected, which was also verified in a limited spectral range by measuring the ( 3 × 4 ) Mueller matrix. The green solid line corresponds to the model fit, taking into account the experimental bandwidth, which yields excellent agreement. For the resulting best-fit parameters, the calculated optical spectrum, assuming a vanishing bandwidth ( δ E = 0), is shown as the red solid line. Good agreement of the energetic position of the peaks is obtained, their shape, i.e., linewidth ( γ), and height, is poorly reproduced. In general, the calculated peaks exhibit a much smaller linewidth and amplitude than experimentally obtained. This effect can be attributed to the effect that the linewidth is much smaller than the bandwidth of the experimental setup [cf. Fig. 2(a)], i.e., γ < δ E, as described above.

Though the spectral information of the modes resulting from the grating agrees for both cases, i.e., δ E = 0 and δ E 0, which was exploited in Ref. 15, it is clear that the bandwidth must be considered in order to accurately describe the measured data and explain the absence of the sharp features resulting from the grating and cannot be neglected as it is typically done when γ > δ E.7,9,14 In Ref. 12, for example, depolarization effects resulting from finite bandwidth were taken into account to characterize asymmetry in nanoimprinted gratings.

Note the remaining deviations between the model and measurement data we attribute to irregularities of the gratings and deviations from a perfect rectangular profile. Assuming the grating layer to be made up of M thin layers with linearly (or parabolically) growing η has a similar broadening effect). Furthermore, deviations are slightly larger starting at around 3.8 eV with the onset of absorption in the YSZ layers.

The features at high energies around 3.8 eV correspond to the excited BSW modes for transverse electric and transverse magnetic polarization, which are absent in the spectrum from the unpatterned region, because BSW are not able to couple to the incident light, as expected. The peak in Ψ gives the position of the T M-polarized BSW, which depends on the shift of the wavevector Δ k introduced by the grating, mainly determined by L. As increasing L corresponds to decreasing Δ k , the BSW modes of the gratings with L = 450 nm appear red-shifted with respect to the gratings with L = 400 nm. The peaks that correspond to the BSW exhibit a larger broadening in the experimental spectrum compared to the calculated one, which was also reported with Ref. 15. These results show that gratings can be used for spatially controlled BSW excitation, which is of interest for BSW-based optical circuits24–27 or strong light–matter interactions.28–30 

A comparison between AFM results and our model fit parameters yields overall very good agreement and a comparison is shown in Table II. The lattice constant L seems to be slightly but systematically underestimated in SE analysis by 1%–2% compared to the value obtained by AFM. The model is also sensitive with regard to the grating height h though relative differences are larger at 10%. These differences may be attributed in part to the difference in the AFM scan range ( 1.5 × 1.5 μ m 2) compared an average over the whole grating ( 50 × 50 μ m 2) in SE measurements. Furthermore, the assumption of a perfectly rectangular grating profile which was used in the model analysis is not strictly fulfilled by the actual grating geometries here. Nonetheless, the groove ratio η shows excellent agreement between the two measurement techniques and the surface thickness (including the grating layer) is constant for all geometries, which indicates very homogeneous films. The azimuthal angle offset is well within the tolerance of centering the sample by eye on the ellipsometer.

We would like to note that if the experimental spectra are modeled excluding the bandwidth, similar grating parameters can be obtained, whose differences to the AFM values are systematically larger along with their uncertainties, cf. Table II. Though this may suffice for some investigations, all the information regarding the spectral broadening of the modes, i.e., the imperfection and inhomogeneity of the grating itself, are lost. Furthermore, we want to point out that neglecting the bandwidth leads to a more unstable minimizing routine that is very sensitive on the initial parameter values as opposed to a more robust minimization if the bandwidth is considered.

Within this work, we have demonstrated an ellipsometric modeling approach for one-dimensional gratings on top of a transparent multilayer structure supporting Bloch surface waves. The grating allows incident light to couple to these modes which can be seen as peaks in the specular reflected light. Furthermore, it was shown that the ellipsometer’s spectral bandwidth must be included in the model for an accurate description of the Ψ , Δ spectra, which was accounted for here by using a discrete set of Gaussian weights. As a consequence of the grating coupler, it was possible to observe BSW peaks in the spectra, which are otherwise absent. We compared the model parameters to AFM measurements and found very good agreement for the lattice constant and groove ratio, while the grating height showed relative differences of <10%. The remaining deviations between the model and experimental values and the differences between AFM and SE measurements were attributed to a grating profile that was not perfectly rectangular and the different size of the measurement areas, respectively. These results provide an efficient framework for the optical determination of grating dimensions and characterization of BSW launching regions, while stressing the importance to consider instrument limitations, especially the spectral bandwidth.

See the supplementary material for depolarization data, ellipsometric spectra at AOI = 45 °, and angle dependent reflectivity maps for different grating geometries.

The authors would like to thank M. Hahn for PLD target preparation, Dr. Daniel Splith and Dr. Lukas Trefflich for technical support (all at Leipzig University). An Accurion EP4 imaging ellipsometer was used for the measurements, which was purchased as part of the UltraSpec project [Bundesministerium für Bildung und Forschung (BMBF) German Ministry for Education and Research], Association of German Engineers (VDI)/Association of Electrical, Electronic, and Information Technology (VDE), Project No. 03VP08180). Sebastian Henn acknowledges the graduate school BuildMoNa and funding by the Deutsche Forschungsgemeinschaft (DFG, Project No. STU 647/2-1).

The authors have no conflicts to disclose.

Sebastian Henn: Conceptualization (equal); Formal analysis (lead); Investigation (lead); Methodology (lead); Validation (equal); Visualization (lead); Writing – original draft (lead). Marius Grundmann: Formal analysis (supporting); Funding acquisition (supporting); Resources (lead); Validation (supporting); Writing – review & editing (equal). Chris Sturm: Conceptualization (equal); Formal analysis (supporting); Funding acquisition (lead); Methodology (supporting); Project administration (lead); Validation (equal); Writing – original draft (supporting); Writing – review & editing (equal).

The data that support the findings of this study are available from the corresponding author upon reasonable request.

1.
D.
Melati
,
A.
Melloni
, and
F.
Morichetti
,
Adv. Opt. Photonics
6
,
156
(
2014
).
2.
E. J.
Murphy
,
Integrated Optical Circuits and Components: Design and Applications
(
CRC Press
,
2020
).
3.
H. S.
Dutta
,
A. K.
Goyal
,
V.
Srivastava
, and
S.
Pal
,
Photonics Nanostruct. Fundam. Appl.
20
,
41
(
2016
).
4.
S. T.
Koev
,
A.
Agrawal
,
H. J.
Lezec
, and
V. A.
Aksyuk
,
Plasmonics
7
,
269
(
2012
).
5.
S. G.
Tikhodeev
,
A.
Yablonskii
,
E.
Muljarov
,
N. A.
Gippius
, and
T.
Ishihara
,
Phys. Rev. B
66
,
045102
(
2002
).
6.
H.
Fujiwara
,
Spectroscopic Ellipsometry: Principles and Applications
(
John Wiley & Sons
,
2007
).
7.
E.
Balogh
,
P.
Basa
,
A.
Suto
,
B.
Powell
,
A.
Bölcskei-Molnár
, and
S.
Biró
,
J. Vac. Sci. Technol. B
37
,
062928
(
2019
).
8.
T.
Ghong
,
S.
Han
,
J.
Chung
,
J.
Byun
,
Y.
Kim
, and
D.
Aspnes
,
J. Korean Phys. Soc.
58
,
1426
(
2011
).
9.
C.
Walder
,
M.
Zellmeier
,
J.
Rappich
,
H.
Ketelsen
, and
K.
Hinrichs
,
Appl. Surf. Sci.
416
,
397
(
2017
).
10.
M.
Foldyna
,
A.
De Martino
,
E.
Garcia-Caurel
,
R.
Ossikovski
,
F.
Bertin
,
J.
Hazart
,
K.
Postava
, and
B.
Drévillon
,
Phys. Status Solidi A
205
,
806
(
2008
).
11.
R.
Antos
,
J.
Pistora
,
J.
Mistrik
,
T.
Yamaguchi
,
S.
Yamaguchi
,
M.
Horie
,
S.
Visnovsky
, and
Y.
Otani
,
J. Appl. Phys.
100
,
054906
(
2006
).
12.
X.
Chen
,
C.
Zhang
,
S.
Liu
,
H.
Jiang
,
Z.
Ma
, and
Z.
Xu
,
J. Appl. Phys.
116
(19),
194305
(
2014
).
13.
T. A.
Germer
and
H. J.
Patrick
, in Metrology, Inspection, and Process Control for Microlithography XXIV (SPIE, 2010), Vol. 7638, pp. 476–485.
14.
T. A.
Germer
,
H. J.
Patrick
, and
R. G.
Dixson
,
AIP Conf. Proc.
1395, 309–313 (2011).
15.
G. A.
Rodriguez
,
J. D.
Ryckman
,
Y.
Jiao
, and
S. M.
Weiss
,
Biosens. Bioelectron.
53
,
486
(
2014
).
16.
P.
Yeh
,
A.
Yariv
, and
A. Y.
Cho
,
Appl. Phys. Lett.
32
,
104
(
1978
).
17.
T.
Perani
and
M.
Liscidini
,
Opt. Lett.
45
,
6534
(
2020
).
18.
A.
Sinibaldi
,
N.
Danz
,
E.
Descrovi
,
P.
Munzert
,
U.
Schulz
,
F.
Sonntag
,
L.
Dominici
, and
F.
Michelotti
,
Sens. Actuators, B
174
,
292
(
2012
).
19.
X.-B.
Kang
,
L.-W.
Wen
, and
Z.-G.
Wang
,
Opt. Commun.
383
,
531
(
2017
).
20.
D.
Ge
,
J.
Shi
,
A.
Rezk
,
C.
Ma
,
L.
Zhang
,
P.
Yang
, and
S.
Zhu
,
Nanoscale Res. Lett.
14
,
1
(
2019
).
21.
F.
Ay
,
K.
Wörhoff
,
R. M.
De Ridder
, and
M.
Pollnau
,
J. Micromech. Microeng.
22
,
105008
(
2012
).
22.
D.
Nečas
and
P.
Klapetek
,
Cent. Eur. J. Phys.
10
,
181
(
2012
).
23.
E.
Descrovi
,
F.
Giorgis
,
L.
Dominici
, and
F.
Michelotti
,
Opt. Lett.
33
,
243
(
2008
).
24.
E.
Descrovi
,
T.
Sfez
,
M.
Quaglio
,
D.
Brunazzo
,
L.
Dominici
,
F.
Michelotti
,
H. P.
Herzig
,
O. J.
Martin
, and
F.
Giorgis
,
Nano Lett.
10
,
2087
(
2010
).
25.
Y.
Augenstein
,
M.
Roussey
,
T.
Grosjean
,
E.
Descrovi
, and
C.
Rockstuhl
,
Photonics Nanostruct. Fundam. Appl.
52
,
101079
(
2022
).
26.
K. R.
Safronov
,
A. A.
Popkova
,
D. I.
Markina
,
A. P.
Pushkarev
,
S. V.
Makarov
,
V. O.
Bessonov
, and
A. A.
Fedyanin
,
Laser Photonics Rev.
16
,
2100728
(
2022
).
27.
M.
Scaravilli
,
A.
Micco
,
G.
Castaldi
,
G.
Coppola
,
M.
Gioffrè
,
M.
Iodice
,
V.
La Ferrara
,
V.
Galdi
, and
A.
Cusano
,
Adv. Opt. Mater.
6
,
1800477
(
2018
).
28.
J.
Nong
,
X.
Xiao
,
F.
Feng
,
B.
Zhao
,
C.
Min
,
X.
Yuan
, and
M.
Somekh
,
Opt. Express
29
,
27750
(
2021
).
29.
S.
Henn
,
M.
Grundmann
, and
C.
Sturm
,
New J. Phys.
23
,
093031
(
2021
).
30.
S.
Pirotta
,
M.
Patrini
,
M.
Liscidini
,
M.
Galli
,
G.
Dacarro
,
G.
Canazza
,
G.
Guizzetti
,
D.
Comoretto
, and
D.
Bajoni
,
Appl. Phys. Lett.
104
,
051111
(
2014
).