Open Access
28 October 2022 Microsphere assistance in interference microscopy with high numerical aperture objective lenses
Author Affiliations +
Abstract

Various attempts have been discussed to overcome the lateral resolution limit and thus to enlarge the fields of application of optical interference microscopy. Microsphere-assisted microscopy and interferometry have proven that the imaging of structures well below Abbe’s resolution limit through near-field assistance is possible if microspheres are placed on the measured surface and utilized as near-field assisting imaging elements. The enhancement of the numerical aperture (NA) by the microspheres as well as photonic nanojets was identified to explain the resolution enhancement, but also whispering gallery modes and evanescent waves are assumed to have an influence. Up to now, to the best of our knowledge, there is no complete understanding of the underlying mechanisms and no model enabling to examine ideal imaging parameters. This contribution is intended to clarify how much the lateral resolution of an already highly resolving Linnik interferometer equipped with 100 × NA 0.9 objective lenses can be further improved by microspheres. Our simulation model developed so far is based on rigorous near-field calculations combined with the diffraction-limited illumination and imaging process in an interference microscope. Here, we extend the model with respect to microsphere-assisted interference microscopy providing a rigorous simulation of the scattered electric field directly above the sphere. Simulation and experimental results will be compared in the three-dimensional spatial frequency domain and discussed in context with ray-tracing computations to achieve an in-depth understanding of the underlying mechanism of resolution enhancement by the microsphere.

1.

Introduction

Due to the ongoing trend toward miniaturization high-resolution optical imaging and three-dimensional (3D) microscopy is highly relevant in certain fields of science and technology. This is particularly true for interference microscopy, one of the most established techniques for micro- and nano-structure measurement. As the lateral resolution Λmin follows the Abbe criterion

Eq. (1)

Λmin=λ2NA.
The conventional way of improving the lateral resolution capabilities of microscopes is to increase the numerical aperture NA=nsinθmax of the objective lenses and to reduce the wavelength λ of light. For air as the surrounding medium with the refractive index n=1, the NA solely depends on the maximum angle θmax, which is the maximum angle of incidence with respect to the optical axis and the maximum scattering or reflection angle that is captured by the microscope objective (MO) lens. In recent years, microsphere assistance has been proposed to improve the lateral resolution capabilities of conventional optical bright-field microscopes.1 Furthermore, the use of microcylinders instead of microspheres has been reported.24

A similar approach demonstrates what is called the super-resolving behavior of liquid-immersed microspheres5 and points out the advantages compared to other resolution enhancement techniques.6 Microsphere assistance was also successfully applied in context with confocal microscopy.7 In addition, the illumination conditions with microsphere support and dark field microscopy were examined.8,9

More recently, microsphere assistance was also applied in white-light interference microscopy.10 For phase-shifting interferometry, it was shown that with the support of near-field information provided by microspheres, it is possible to extend the resolution limit for interferometric height profile measurements.11,12 NAs in the range of 0.3 to 0.85 were used, enabling access to high-frequency image information through the improvement of the optical resolution. Thus, high-spatial-frequency surface height information could be obtained by white-light and phase-shifting interferometers in both, Linnik as well as Mirau configurations.1315

To explain the effect of the resolution enhancement, photonic nanojets are often referred to.1618 Their impact on resolution it is also discussed in detail by Darafsheh.19 Nanojets describe the focus of light on the backside of a microsphere illuminated with a plane wave from the top. On the scale of microspheres, this focus is characterized by its high intensity and narrow waist.20 Numerous papers have been published, studying the behavior and engineering of photonic nanojets.21,22 Also, the role of evanescent waves and whispering gallery modes has been considered.23,24 Detailed studies were also made on the resolution capabilities of Mie particles, which are in contact with the surface under investigation.25 For further details it is referred to a recent review paper that gives an overview of the state-of-the-art in microsphere-assisted microscopy.26

Since incoherent Koehler illumination is applied in conventional bright-field and in interference microscopy, the microspheres are illuminated by multiple plane waves incident under various angles. Annular illumination of the outer region of microspheres turned out to affect the achieved resolution enhancement.27,28 Nevertheless, until now there is no complete and widely accepted explanation of the resolution enhancement by microspheres in conventional microscopy and interferometry. For this reason, further analysis of the underlying physical principles is of predominant interest. Analyzing interference microscopes in the 3D spatial frequency domain by the 3D transfer function (TF) of the imaging system gives physical insight into the relevant transfer characteristics. Sheppard et al.29 introduced a model, which represents the imaging process of confocal microscopes in the 3D spatial frequency domain. This model was applied in further publications to surface profile reconstruction in confocal microscopy30,31 and later introduced as the foil model in coherence scanning interferometry (CSI).32,33 We recently extended this model by treating the reference mirror in the same way as the object’s surface34 and pointed out the analogy with the 3D TF of a bright-field reflection microscope.35 We further recognized that the transfer characteristics of 3D microscopes strongly depend on the scattering characteristics of the surface (single-point scatterers, mirror-like surfaces, or diffraction gratings) and the spectral characteristics of the light source.36

Analyzing the interferometric measurement data in the 3D spatial frequency domain, the influence of microspheres on the transfer behavior of the optical system could already be shown.28,37,38 In the following, the transfer characteristics of a high-resolution Linnik interferometer with and without microsphere assistance are pointed out by 3D spatial frequency domain analysis of 3D image stacks of periodic grating structures. A comparison of the results exhibits that the transfer characteristics of microspheres described in the 3D frequency domain are closely related to the angular ranges of incident, reflected, and diffracted light rays. Thus, ray-tracing computations of light propagation through and inside the microsphere give further insight. Finally, experimental results obtained from gratings of different periods are compared to results of simulations based on rigorous finite element method (FEM) computations.

2.

Experimental Setup

The Linnik interferometer sketched in Fig. 1(a) and displayed in Fig. 1(c) is used to record interference images at certain height positions during a depth scan. This results in the so-called 3D image stack. Since our intention is to compare the resolution enhancement introduced by a microsphere in an already highly resolving microscope, we use two high-resolution MO lenses with 100× magnification and an NA of 0.9, which still provide a working distance of 1 mm. A scientific CMOS camera records the image stack. For illumination, a royalblue light-emitting diode emitting at a center wavelength of λ=440  nm (Luxeon REBEL Color Line, spectral half-width 20 nm) arranged in a Koehler illumination setup and a transverse magnetic (TM) polarizer are utilized.

Fig. 1

(a) Scheme of the Linnik interferometer setup with the microsphere; (b) enlarged section showing a specimen with a microsphere on top creating a virtual image plane; and (c) photograph of the experimental setup in the laboratory.

JOM_2_4_044501_f001.png

The depth scan is carried out using a precision piezo stage moving the object under investigation axially. Small height steps of typically 20 nm between two consecutive image frames are chosen to obtain a high number of sample points of the interference signals, which allows low-pass filtering to further reduce signal noise. The resulting image stack is analyzed pixel by pixel using envelope and phase evaluation algorithms to reconstruct the 3D topography of the surface.39 It should be noted that the width of the envelope of the interference signals in this high NA Linnik interferometer configuration is dominated by the longitudinal spatial coherence of the focused light rather than the temporal coherence due to the spectral width of the blue LED. This effect is further described by Abdulhalim.40

When placing microspheres on the surface to be measured, these spheres are put on the object’s surface in a liquid emulsion for practical reasons. After the liquid has evaporated, the measurement can be carried out. Throughout this study, SiO2 microspheres with a diameter of 7 to 10  μm are being used. With the application of microspheres in the imaging path, an axial shift of the focus occurs and creates a virtual image plane as it is sketched in Fig. 1(b).

Our experiments included different diameters and materials of the spheres. It turned out that for MO lenses of high NA the number parameter combinations of microspheres suitable for microsphere-assisted interferometry is limited. For diameters larger than 20  μm, it was not possible to obtain a sharp image. Similar problems occurred at higher refractive indices. Based on these experimental results, a microsphere made of SiO2 with relatively small diameters of 5 to 15  μm is a good choice for the high aperture experimental setup. However, there may be further parameter combinations working well too.

3.

Analysis in the Spatial Frequency Domain

We assume the scattering geometry41 according to Fig. 2(a), where kin is the wave vector of a plane wave incident under an angle θin, and kr and ks are wave vectors of reflected and scattered waves propagating under the angles θr and θs, respectively. For a grating with a period Λmin corresponding to the Abbe limit according to Eq. (1), the situation is outlined in Fig. 2(b). For the incidence angle θin=θmax, the zeroth-order diffracted wave with wave vector kr propagates under the angle θr=θmax, whereas the scattering angle of the minus first-order diffracted wave is θs=θmax. The scattered electric field Us(q) under the Fraunhofer far-field condition can be calculated using the Kirchhoff formulation29,41 with respect to a microscope in reflection mode assuming a perfectly reflecting surface s(x,y) as

Eq. (2)

Us(q)=1AAei(qxx+qyy+qzs(x,y))dxdy,=1A+A(x,y)ei(qxx+qyy+qzs(x,y))dxdy,=1AF{A(x,y)}*F{eiqzs(x,y)},
where a monochromatic plane wave of wavelength λ and wavenumber k0=2π/λ is incident under the angle θin, and the vector q=kskin defines a point in the 3D Fourier domain. The scattered far-field Us(q) is normalized in a way that for a smooth surface and perpendicular incidence, i.e., θin=0, the amplitude in the specular direction becomes unity. The area A(x,y) of integration in Eq. (2) corresponds to the field of view of the microscope, F{} represents the Fourier transform, and * the convolution symbol. s(x,y)=s(x) applies for a surface textured in one dimension only. If the area A is large enough, Eq. (2) represents the two-dimensional (2D) Fourier transformation with respect to x and y of the phase object exp(iqzs(x,y)). If, for simplicity s(x,y)=s(x) and considering wave vectors kin and kin only in the xz-plane, which then equals the plane of incidence and observation, the incident and scattered wave vectors are given as

Eq. (3)

kin=k0(sinθin0cosθin)andks=k0(sinθs0cosθs).

Fig. 2

Scattering geometry with (a) wave vectors of incident, reflected and scattered waves in the xz-plane and (b) wave vectors of reflected and -1st order diffracted wave vectors in xz-plane for a plane wave incident on a grating under the maximum angle θmax restricted by the NA of a objective lens.

JOM_2_4_044501_f002.png

Thus, the vector q results in

Eq. (4)

q=(qxqyqz)=k0(sinθssinθin0cosθs+cosθin).

According to Eq. (4), the situation shown in Fig. 2(b) leads to

Eq. (5)

|qx|=2k0sinθmax=2k0NA=2πΛmin.
The maximum lateral spatial frequency collected by a microscope lens of given NA value. Equation (5) is directly related to the Abbe resolution limit according to Eq. (1). The coordinates in q-space follow the Ewald sphere construction shown in Fig. 3. Due to the NA of the microscope, the possible directions of the wave vectors of incident and scattered light are limited by the angle θmax. According to Fig. 3(a), this results in the two spherical caps, which need to be correlated to calculate the 3D TF H(q) of the instrument.35 The outer boundary of the resulting Ewald-limiting-sphere is plotted in Fig. 3(b), where the contributions of the specular reflection are located on the qz-axis and the outer spherical shell of radius 2k0 represents the back scattered light. It is worthy to note that the qx-value is related to the spatial frequency component of a surface, whereas the corresponding qz-values represent the spectral range of the corresponding interference signals. The whole construction shows rotational symmetry with respect to the qz axis. Figure 3(c) shows a 2D cross sectional view in the qxqz-plane of H(q), which holds for plane mirror-like surfaces and diffraction gratings.36 Once Us(q) and H(q) are known, the 3D spatial frequency representation of the interference intensity equals the product:

Eq. (6)

ΔI˜(q)=Us(q)H(q),
and the interference image stack ΔI(x,y,z) can be calculated via inverse 3D Fourier transform:

Eq. (7)

ΔI(x,y,z)Re{F1{ΔI˜(q)}}.

Fig. 3

(a) Ewald sphere construction based on two spherical caps defined by numerous wave vectors kin and ks with angles θin and θs below the maximum angle θmax corresponding to the NA of the objective lens; (b) Ewald-limiting-sphere resulting from the correlation of the two caps; and (c) cross-section of the 3D TF for an interference microscope with NA=0.9 and λ=440  nm.36

JOM_2_4_044501_f003.png

Cross sections of experimentally obtained spatial frequency representations ΔI˜(q) for different rectangular silicon phase gratings (RS-N standard by SiMETRICS) are shown in Fig. 4. Figure 4(a) shows the result for a rectangular grating of 6-μm period and 192-nm peak-to-valley (PV) amplitude, whereas Figs. 4(b) and 4(c) belong to gratings of 400 nm (b) and 300-nm period (c), both with a PV amplitude of 140 nm. The shape of the Ewald-limiting-sphere can be clearly recognized in Fig. 4(a). Figures 4(b) and 4(c) show only three diffraction orders, the zeroth order at qx=0 and the 1st and -1st orders, which are located at higher qx values for the shorter period.

Fig. 4

Cross-section of the 3D spatial frequency representation of an interference image stack obtained without a microsphere for (a) a rectangular grating with 6-μm period and 192-nm height difference (RS-N standard by SiMETRICS); (b) a rectangular grating with 0.4-μm period and 140-nm height difference; and (c) a rectangular grating with 0.3-μm period and 140-nm height difference.

JOM_2_4_044501_f004.png

4.

Ray-tracing Results for a Microsphere

We already demonstrated experimentally that a phase grating with a minimum period length of 230 nm is resolved by the Linnik interferometer mentioned above in combination with microsphere assistance.42 This period is slightly below the Abbe resolution limit of 244 nm, which holds for this interferometer without a microsphere. This section is intended to study, how light rays are refracted, reflected and focused by the microsphere to find out plausible mechanisms for the resolution enhancement. Since we use microspheres of at least 7-μm diameter (r=3.5  μm), the Mie-parameter 2πk0r/λ is at least 50 and a physical description based on geometrical optics is a satisfying approximation to study basic effects.43 Figure 5 shows some results of ray tracing computations assuming a microsphere of silica with a refractive index of 1.4655. In Fig. 5(a), a bundle of parallel light rays propagates along the vertical axis and hits the microsphere. The refracted light propagates through the microsphere, is refracted again and shows nearly grazing incidence with respect to a horizontal line located directly under the microsphere, where the object is typically placed. Due to the large angle of nearly 90 deg (defined with respect to the z-axis) of these rays refracted twice, the effective NA of the microsphere as an optical imaging element is close to unity. There are additional situations as shown in Figs. 5(b)5(d), where the light never reaches a measuring object located underneath the microsphere. These ray paths lead to additional interference components although they are not affected by the surface of the measuring object. In Fig. 5(b), the incident light rays show a relatively large angle of incidence with respect to the z-axis. However, all angles are below the maximum angle θmax defined by the NA of the system. The rays travel horizontally through the microsphere. Due to the symmetry of the arrangement the scattering angle equals the angle of incidence, i.e. θs=θin. Hence, for these rays qx=0 and the corresponding qz-values are relatively small. In Fig. 5(c), parallel rays are incident on the left hand side. These rays are refracted at the boundary of the microsphere and are then internally reflected at the bottom of the sphere. After an additional refraction, they propagate in air and include relatively small angles with the z-axis, thus the corresponding qx-values are low and qz-values high. This is the arrangement belonging to the rainbow and, indeed, the rainbow ray43 can be identified as the ray on the right hand side, which corresponds to the maximum scattering angle. The rays close to the rainbow ray form a caustic, and thus, in these regions high intensity values occur. A similar situation is shown in Fig. 5(d), which shows symmetry with respect to the z-axis, since all rays are internally reflected at the coordinate x=0, z=r inside the microsphere and, therefore, qx=0 for these rays. Figure 5(d) demonstrates that the incident and scattered rays again include rather small angles with the z-axes, such that no total internal reflection inside the microsphere occurs. Light incident at higher angles is no longer reflected at x=0, z=r. Therefore qz is quite high for the rays plotted in Fig. 5(d).

Fig. 5

Ray-tracing computations for a silica microsphere of refractive index nr=1.4655 at λ=440  nm, (a) bundle of rays incident under an angle θin=0; (b) rays traveling horizontally through the microsphere; (c) internally reflected bundle of rays incident under an angle θin=0; and (d) incident rays internally reflected at the position x=0, z=r inside the sphere.

JOM_2_4_044501_f005.png

5.

Experimental Results

The interferometric measurement data is acquired through a stepwise depth scan with a step height of 20 nm performed by a piezo scanner. This results in a measuring process known from CSI, however, additionally utilizing microsphere assistance. The measurement object is the RS-N resolution standard by SiMETRICS. Mainly the 300- and 600-nm grating structures with nominal depths of 140 and 160 nm, respectively, are examined.

While performing CSI measurements, the height information of the specimen’s surface is encoded in the phase of the interference signals. To gain a better understanding of the phase modulation occurring in interferometric measurement, xz-cross sections of (offset-free) interference images acquired through a depth scan are shown in Fig. 6. The results shown in Figs. 6(a) and 6(b) are obtained from the 300- and 600-nm grating structure of the RS-N standard using microsphere assistance, i.e., placing a microsphere directly on the grating structure. For comparison, Fig. 6(c) shows how the image stacks recorded for a mirror instead of a grating leads to highest contrast of the modulation in z-direction of the interference signals centered around the virtual image plane. It is worthy to note that the position on the z-axis is related to the range of the depth scan, which is typically 5- to 10-μm long. The virtual image plane is placed approximately in the middle of this range and z=0 denotes the starting position of the scan range. In x-direction, the phase modulation introduced by the surface of the measured grating structures is visible in Figs. 6(a) and 6(b). The interference signals can be analyzed using envelope- and phase retrieving algorithms to reconstruct the grating structure of the object as previously shown.42 In addition, the resolution enhancement through the imaging process also an additional magnification is introduced by the microspheres.

Fig. 6

xz-cross sections of offset-reduced interference image stacks acquired with microsphere assistance via a depth scan using royal blue illumination of rectangular silicon gratings (RS-N standard) with period lengths of (a) 300 nm; (b) 600 nm; and (c) a mirror surface for comparison.

JOM_2_4_044501_f006.png

For comparability with the results shown in Secs. 3 and 4, the spatial frequency domain representation of the interference data is analyzed as introduced beforehand. First, the data are preprocessed by means of Blackman-windowing as well as zero-padding to exclude contributions occurring besides the microsphere and to improve the spatial frequency resolution, respectively. The data are further offset reduced. Cross sections of the spatial frequency representation in the qxqz-plane are shown in Fig. 7 accordingly.

Fig. 7

qxqz-cross-sections of 3D spatial frequency representations of the interference image stacks corresponding to (a) the 300-nm grating structure; (b) the 600-nm grating; and (c) the mirror surface.

JOM_2_4_044501_f007.png

As it is shown in Eq. (5) and illustrated in Fig. 4, for a one-dimensional grating structure imaged by an interference microscope, the diffraction orders are visible as corresponding sharp lines at qx=const. in the 3D spatial frequency representation. The additional magnification factor introduced by the microsphere was determined to be M=1.4 for our setup. Thus, for the grating of 300-nm period the qx-value corresponding to the first-order diffraction maximum is shifted by the mircosphere from qx,a±21  μm1 in Fig. 4(c) to qx,a±15  μm1 according to Fig. 7(a), since the period length Λ is multiplied by M. Consequently, the 600-nm grating magnified by the microsphere leads to first-order diffraction maxima at qx,a±7.5  μm1 as it is displayed in Fig. 7(b). Finally, for the mirror no diffraction order except for the zero-order located at qx=0 appears, as shown in Fig. 7(c).

Further, the transfer behavior of the microsphere-assisted setup differs compared to an interference microscope without a microsphere. The size of the field of view limited by the microsphere significantly influences the intensity distribution in the 3D spatial frequency domain. According to Eq. (2), the field of view corresponds to the area of integration described by A(x,y). This leads to a convolution in the 3D spatial frequency domain of the diffraction pattern of the grating with an Airy-disk function (implying a circular shaped field of view). The smaller the field of view A(x,y) the broader is this Airy-disk and the corresponding blurring of the diffraction maxima. As stated by Sheppard,44 broadening of diffraction orders due to a small field of view has a significant influence on the resolution capabilities of a system. Due to optical aberrations introduced by the microsphere and the Blackman window used to extract the relevant lateral and axial range in the spatial domain, the frequency response is additionally affected. As a result, a broadening of the discrete intensity pattern, i.e., the sharp lines at certain qx-values initiated by diffraction at the grating structure, occurs. This becomes apparent by comparison of Figs. 7(a), 7(b) and 4. In addition, the light diffracted by the grating, additional intensity contributions in Fig. 7 can be attributed to the microsphere itself, e.g., the intensity maximum at qx=0, qz15  μm1, which corresponds to the situation according to Fig. 5(b) and the higher frequency rippling, which is a consequence of the increased scattered light intensity under the rainbow angle shown in Fig. 5(c).

6.

Rigorous Simulations

To investigate further influences on the transfer behavior of the microspheres as well as the influence of photonic nanojets, rigorous simulations were performed, which are presented in the following.

6.1.

Simulation of the Complete Imaging Process

The measured data are compared to simulation results of the imaging process, where the light-surface interaction considering the microsphere is based on rigorous FEM computation of the electric field distribution. The transfer characteristics of the interference microscope as well as the phase shifts introduced by the depth scan are considered by filter operations using Fourier optics modeling. The combined model enables a full 3D simulation of illumination and diffraction at 2D periodic surfaces and provides accurate results compared with CSI measurements.45,46 In this study, the model is extended considering a microsphere with a radius of r=2.5  μm and a refractive index of n=1.5, which is placed directly on the specimen. For computational reasons, microspheres are arranged in a periodic manner with a period length Lx=13.2  μm. Due to computational and time constraints, the simulation is performed on 2D surface structures and the microsphere is approximated by a microcylinder, as cylindrical microelements were shown to enhance the lateral resolution capabilities too.4,47,48 Since we are interested in general effects of microsphere-induced resolution enhancement, it is reasonable not to use exactly the same configuration for measurement and simulation.

With respect to Koehler illumination, it can be assumed that the illumination of the specimen is composed of individual incoherent partial plane waves. These illuminate the specimen from all angles which cover the aperture of the objective. To realize the conical illumination, discretization is performed over these angular values and rigorous simulations of the near field are conducted using plane wave illumination for each discrete incident angle. For each simulated near field, the far-field expansion and Fourier optics modeling of the imaging processes in the microscope are performed. Afterwards, the interference intensity distribution considering the reference field is calculated for each set of incident angles and the final incoherent summation of the results is done.46

Figure 8 shows extracts of offset reduced interference image stacks in the presence of a microsphere simulated with TM polarized monochromatic light of λ=440  nm and NA=0.9. Due to the high NA value, the light source can be assumed as monochromatic, since the influence of temporal coherence is negligible for single-colored LEDs. The underlying structure is a sinusoidal grating with PV height of 25 nm and period lengths corresponding to those of the measured results shown in Fig. 6. To avoid additional effects such as multiple scattering and edge diffraction, the height is chosen to be smaller compared to the measured profile.

Fig. 8

xz-cross sections of offset-reduced interference image stacks obtained by rigorous simulation.46 The periods of the underlying sinusoidal grating structure of 25-nm PV amplitude correspond to the RS-N resolution standard with period length of (a) 300; (b) 600 nm; and (c) results of a flat mirror surface for comparison.

JOM_2_4_044501_f008.png

Considering Figs. 6 and 8, significant key characteristics can be compared. In both figures, the grating structure appears in the phase modulation of the interference signals. For the experimental results, the phase modulation is more pronounced in amplitude, since a rectangular grating (RS-N standard) was used. Similar to the measurement result, the simulated interference signals obtained from a flat mirror [Fig. 8(c)] do not show a grating-dependent lateral phase modulation. Detailed comparison of simulated and measured results exhibits deviations. These are mainly due to the fact that results measured using a sphere are compared to simulation results obtained from a cylinder because of the computational effort mentioned above. Further, the sphere is arbitrarily placed on the grating and thus at different positions with respect to the grating structure in simulations. Additionally, the spheres have a different radius compared to the cylinder leading to further deviations. Nevertheless, the qualitative comparison between simulated and experimental results demonstrates that in principle the simulation model reflects the relevant physical mechanisms in microsphere-assisted interferometry.

To analyze the interference signals in more detail, Fig. 9 shows the representation in q-space. Comparing simulation and measurement results, the diffraction orders are blurred in the measurement results. The intensity obtained from a plane mirror [Fig. 9(c)] shows an additional periodic modulation of the intensity distribution for larger qz and small qx values. This is probably a consequence of the rainbow effect explained in Sec. 4, since further simulation results turn out that the modulation period behaves anti-proportional to the microsphere’s diameter. This phenomenon is not as clearly observed in the measurement results obtained from a plane mirror, but the differences can be explained by the use of microcylinders in the simulation instead of spheres.

Fig. 9

qxqz-cross-sections of 3D spatial frequency representations of the simulated interference image stacks corresponding to (a) the 300-nm sinusoidal grating; (b) the 600-nm grating; and (c) the mirror surface.

JOM_2_4_044501_f009.png

However, a slight modulation can be observed even in the measurement results, especially in the case of Figs. 7(a) and 7(b). It is worthy to note that due to the sinusoidal grating structure with a relatively low PV amplitude of 25 nm, the intensity of the blurred diffraction orders is not as pronounced as in the measurement results. Furthermore, in both, measurement and simulation, an intensity contribution at qx=0 for small values of qz is visible with and without a grating structure. This can be assigned to the rays traveling horizontally through the microcylinder as it is shown in Fig. 5(b) and confirmed by simulations assuming a microcylinder in free space. The high intensity for large qz values and qx0 follow from the rays depicted in Fig. 5(d), where rays of oblique incidence and scattering angles inside the sphere are refracted such that above the sphere smaller angles and hence larger values of qz result.

In sum, the rigorous simulation model reproduces the major effects occurring in measurement results. Artefacts observed in the 3D spatial frequency domain can be assigned to general cases of ray tracing. Thus, the q-space representation is shown to be quite useful to analyze the physical mechanism introduced by microspheres. Furthermore, diffraction orders belonging to the measured grating structure under the sphere lead to smaller qx and larger qz values due to the magnification of the sphere as supposed by Hüser and Lehmann.28,42 This effect is now confirmed by simulations of grating structures of different period.

6.2.

Studies on Photonic Nanojets

Rigorous simulations were performed to study the influence of photonic nanojets. In this case, the results are obtained with 2D FEM simulations of the electromagnetic field distribution. Only perpendicular incidence of a monochromatic plane wave under an angle θin=0 is considered (λ=440  nm). The phenomenon of photonic nanojets occurring on the back surface of microspheres or microcylinders was widely studied.19,49 However, these investigations do not include the configuration of an interferometer in reflection mode as it is studied here. Therefore, an additional simulation with a microcylinder on a grating (period length 300 nm) is carried out. The intensity distributions obtained with and without the grating are shown in Fig. 10.

Fig. 10

Simulation results for a microcylinder illuminated by a monochromatic plane wave for the cases (a) with a diffraction grating (300-nm period length) directly below the cylinder and (b) in air surrounding.

JOM_2_4_044501_f010.png

When comparing the results from Figs. 10(a) and 10(b), the photonic nanojet obviously disappears for the case of the grating located directly below the microcylinder. The specimen is placed at the location, where the nanojet in air occurs. Hence, the generation of a nanojet will be disturbed, and instead, an interaction of the corresponding electromagnetic waves with the specimen’s surface takes place. Furthermore, the case of near grazing incidence below the microcylinder shown in Fig. 5(a) arises. The light diffracted from the grating can contribute to the imaging process and thus high aperture imaging occurs even if MO lenses of low NA are used.

7.

Discussion

To analyze the resolution enhancement, we compare both, experimental and simulated interference image stacks in the spatial and the three-dimensional spatial frequency domain. This methodology provides advantages as it explains the occurring phenomena based on light rays traveling under different angles with respect to the optical axis and thus enables comparisons with ray-tracing computations and rigorous simulations as well. In 3D Fourier space effects occurring through the microsphere itself independently of the specimen can be separated from effects introduced by the interaction of light and the specimen via the microsphere. Microspheres shift the intensity diffraction maxima of a grating to lower spatial frequencies. As a consequence, the central wavelength of the resulting interference signals is significantly reduced if microspheres are used. In combination with ray-tracing and rigorous simulation results, we conclude that the most dominant effect, which arises from the microspheres, can be viewed as an effective enlargement of the NA of the optical system.

8.

Conclusion

Although microsphere-assisted resolution enhancement is frequently used in optical imaging and 3D microscopy, the physical mechanism behind this phenomenon is still not completely understood. The most frequently mentioned explanation attempts refer to the enhancement of the NA, the collection of evanescent waves, the photonic nanojet effect, and the excitation of whispering gallery modes.

The methodologies presented in this article can be applied to several configurations of microsphere-assisted interferometry and microscopy. Analysis in the 3D spatial frequency domain gives valuable insight into the transfer behavior of optical systems with different NA and other optical properties.

The rigorous simulation model represents a complete computational approach of the imaging process through the microsphere. This method enables a more specific look on relevant mechanisms responsible for the transfer behavior of microsphere-assisted interferometry. With this model, e.g., the influences of whispering gallery modes and evanescent waves on the imaging process can be elaborated further.

Since we use 3D microscopy in reflection mode, we suppose direct reflection and diffraction of propagating waves to be dominant and thus implications of evanescent waves and whispering gallery modes to be negligible. Therefore, our investigation suggests that the most likely mechanism is the enhancement of the NA, which is close to 1 in case of microsphere assistance, in combination with the rather limited field of view under the microsphere. The fact that higher magnifications than M=1.4 could not be obtained with our experimental configuration is one of the facts that led us to conclude that the numerical aperture must be increased effectively by the microsphere and finally limits the resolution of microsphere-based interferometry. For our microscope with an NA of 0.9 a resolution enhancement of 11% is achieved using microspheres. Since the occurrence of photonic nanojets also relies on the coherent superposition of waves propagating under higher angles with respect to optical axis the physical origins of nanojets and NA enhancement are closely related to each other.

Acknowledgments

The partial support of this project by the DFG (German Research Foundation) (Grant No. LE992/15-1) is gratefully acknowledged.

References

1. 

Z. Wang et al., “Optical virtual imaging at 50 nm lateral resolution with a white-light nanoscope,” Nat. Commun., 2 (1), 1 –6 https://doi.org/10.1038/ncomms1211 NCAOBW 2041-1723 (2011). Google Scholar

2. 

A. Darafsheh, Y. Li and V. N. Astratov, “Super-resolution microscopy by dielectric microcylinders,” in Int. Conf. Transparent Opt. Netw., 100 –102 (2013). https://doi.org/10.1109/ICTON.2013.6602907 Google Scholar

3. 

A. Darafsheh and D. Bollinger, “Systematic study of the characteristics of the photonic nanojets formed by dielectric microcylinders,” Opt. Commun., 402 270 –275 https://doi.org/10.1016/j.optcom.2017.06.004 OPCOB8 0030-4018 (2017). Google Scholar

4. 

J. N. Monks et al., “Spider silk: mother nature’s bio-superlens,” Nano Lett., 16 (9), 5842 –5845 https://doi.org/10.1021/acs.nanolett.6b02641 NALEFD 1530-6984 (2016). Google Scholar

5. 

A. Darafsheh et al., “Optical super-resolution by high-index liquid-immersed microspheres,” Appl. Phys. Lett., 101 (14), 141128 https://doi.org/10.1063/1.4757600 APPLAB 0003-6951 (2012). Google Scholar

6. 

A. Darafsheh et al., “Advantages of microsphere-assisted super-resolution imaging technique over solid immersion lens and confocal microscopies,” Appl. Phys. Lett., 104 (6), https://doi.org/10.1063/1.4864760 APPLAB 0003-6951 (2014). Google Scholar

7. 

Y. Yan et al., “Microsphere-coupled scanning laser confocal nanoscope for sub-diffraction-limited imaging at 25 nm lateral resolution in the visible spectrum,” ACS Nano, 8 1809 –1816 https://doi.org/10.1021/nn406201q ANCAC3 1936-0851 (2014). Google Scholar

8. 

S. Perrin et al., “Illumination conditions in microsphere-assisted microscopy,” J. Microsc., 274 (1), 69 –75 https://doi.org/10.1111/jmi.12781 JMICAR 0022-2720 (2019). Google Scholar

9. 

S. Perrin et al., “Transmission microsphere-assisted dark-field microscopy,” Phys. Status Solidi - Rapid Res. Lett., 13 (2), 1800445 –4 https://doi.org/10.1002/pssr.201800445 (2019). Google Scholar

10. 

F. Wang et al., “Three-dimensional super-resolution morphology by near-field assisted white-light interferometry,” Sci. Rep., 6 (1), 24703 https://doi.org/10.1038/srep24703 SRCEC3 2045-2322 (2016). Google Scholar

11. 

S. Perrin et al., “Microsphere-assisted phase-shifting profilometry,” Appl. Opt., 56 (25), 7249 –7255 https://doi.org/10.1364/AO.56.007249 APOPAI 0003-6935 (2017). Google Scholar

12. 

A. Leong-Hoi et al., “High resolution microsphere‐assisted interference microscopy for 3D characterization of nanomaterials,” physica. status solidi. (a) , 215 (6), 1700858 https://doi.org/10.1002/pssa.201700858 (2018). Google Scholar

13. 

P. C. Montgomery et al., “Sub-diffraction surface topography measurement using a microsphere-assisted Linnik interferometer,” Proc. SPIE, 10329 1032918 https://doi.org/10.1117/12.2270223 PSISDG 0277-786X (2017). Google Scholar

14. 

P. Montgomery et al., “3D nano surface profilometry by combining the photonic nanojet with interferometry,” J. Phys.: Conf. Ser., 794 (1), 012006 https://doi.org/10.1088/1742-6596/794/1/012006 JPCSDZ 1742-6588 (2017). Google Scholar

15. 

I. Kassamakov et al., “3D super-resolution optical profiling using microsphere enhanced mirau interferometry,” Sci. Rep., 7 (1), 1 –7 https://doi.org/10.1038/s41598-017-03830-6 SRCEC3 2045-2322 (2017). Google Scholar

16. 

H. Yang et al., “Super-resolution imaging of a dielectric microsphere is governed by the waist of its photonic nanojet,” Nano Lett., 16 (8), 4862 –4870 https://doi.org/10.1021/acs.nanolett.6b01255 NALEFD 1530-6984 (2016). Google Scholar

17. 

A. Maslov and V. Astratov, “Resolution and reciprocity in microspherical nanoscopy: point-spread function versus photonic nanojets,” Phys. Rev. Appl., 11 (6), 064004 https://doi.org/10.1103/PhysRevApplied.11.064004 PRAHB2 2331-7019 (2019). Google Scholar

18. 

B. S. Luk’yanchuk et al., “Refractive index less than two: photonic nanojets yesterday, today and tomorrow [Invited],” Opt. Mater. Express, 7 (6), 1820 https://doi.org/10.1364/OME.7.001820 (2017). Google Scholar

19. 

A. Darafsheh, “Photonic nanojets and their applications,” J. Phys.: Photonics, 3 (2), 022001 https://doi.org/10.1088/2515-7647/abdb05 (2021). Google Scholar

20. 

A. Heifertz et al., “Photonic nanojets,” J. Comput. Theor. Nanosci., 6 (9), 1979 –1992 https://doi.org/10.1166/jctn.2009.1254 (2009). Google Scholar

21. 

C. Rockstuhl et al., “Engineering photonic nanojets,” Opt. Express, 19 (11), 10206 https://doi.org/10.1364/OE.19.010206 OPEXFF 1094-4087 (2011). Google Scholar

22. 

M. S. Kim et al., “Gouy phase anomaly in photonic nanojets,” Appl. Phys. Lett., 98 (19), 191114 https://doi.org/10.1063/1.3591175 APPLAB 0003-6951 (2011). Google Scholar

23. 

S. Zhou et al., “Effects of whispering gallery mode in microsphere super-resolution imaging,” Appl. Phys. B: Lasers Opt., 123 (9), 1 –9 https://doi.org/10.1007/s00340-017-6815-7 (2017). Google Scholar

24. 

Y. Ben-Aryeh, “Increase of resolution by use of microspheres related to complex Snell’s law,” J. Opt. Soc. Am. A, 33 (12), 2284 https://doi.org/10.1364/JOSAA.33.002284 JOAOD6 0740-3232 (2016). Google Scholar

25. 

A. V. Maslov and V. N. Astratov, “Optical nanoscopy with contact Mie-particles: resolution analysis,” Appl. Phys. Lett., 110 261107 https://doi.org/10.1063/1.4989687 APPLAB 0003-6951 (2017). Google Scholar

26. 

A. Darafsheh, “Microsphere-assisted microscopy,” J. Appl. Phys., 131 (3), 031102 https://doi.org/10.1063/5.0068263 JAPIAU 0021-8979 (2022). Google Scholar

27. 

M. X. Wu et al., “Modulation of photonic nanojets generated by microspheres decorated with concentric rings,” Opt. Express, 23 (15), 20096 https://doi.org/10.1364/OE.23.020096 OPEXFF 1094-4087 (2015). Google Scholar

28. 

L. Hueser and P. Lehmann, “Analysis of resolution enhancement through microsphere-assisted interferometry in the 3D spatial frequency domain,” Proc. SPIE, 11782 117820R https://doi.org/10.1117/12.2593296 PSISDG 0277-786X (2021). Google Scholar

29. 

C. J. Sheppard, T. J. Connolly and M. Gu, “Imaging and reconstruction for rough surface scattering in the Kirchhoff approximation by confocal microscopy,” J. Mod. Opt., 40 (12), 2407 –2421 https://doi.org/10.1080/09500349314552431 JMOPEW 0950-0340 (1993). Google Scholar

30. 

J. C. Quartel and C. J. Sheppard, “A surface reconstruction algorithm based on confocal interferometric profiling,” J. Mod. Opt., 43 (3), 591 –605 https://doi.org/10.1080/09500349608232768 JMOPEW 0950-0340 (1996). Google Scholar

31. 

J. C. Quartel and C. J. Sheppard, “Surface reconstruction using an algorithm based on confocal imaging,” J. Mod. Opt., 43 (3), 469 –486 https://doi.org/10.1080/09500349608232758 JMOPEW 0950-0340 (1996). Google Scholar

32. 

J. Coupland et al., “Coherence scanning interferometry: linear theory of surface measurement,” Appl. Opt., 52 (16), 3662 –3670 https://doi.org/10.1364/AO.52.003662 APOPAI 0003-6935 (2013). Google Scholar

33. 

R. Su et al., “Scattering and three-dimensional imaging in surface topography measuring interference microscopy,” J. Opt. Soc. Am. A, 38 (2), A27 https://doi.org/10.1364/JOSAA.411929 JOAOD6 0740-3232 (2021). Google Scholar

34. 

P. Lehmann, M. Künne and T. Pahl, “Analysis of interference microscopy in the spatial frequency domain,” J. Phys.: Photonics, 3 014006 https://doi.org/10.1088/2515-7647/abda15 (2021). Google Scholar

35. 

P. Lehmann and T. Pahl, “Three-dimensional transfer function of optical microscopes in reflection mode,” J. Microsc., 284 45 –55 https://doi.org/10.1111/jmi.13040 JMICAR 0022-2720 (2021). Google Scholar

36. 

P. Lehmann, S. Hagemeier and T. Pahl, “Three-dimensional transfer functions of interference microscopes,” Metrology, 1 (2), 122 –141 https://doi.org/10.3390/metrology1020009 (2021). Google Scholar

37. 

L. Hüser and P. Lehmann, “Microsphere-assisted interference microscopy for resolution enhancement,” Technisches Messen, 88 (5), 311 –318 https://doi.org/10.1515/teme-2020-0101 (2021). Google Scholar

38. 

L. Hüser et al., “The use of microsphere assistance in interference microscopy with high numerical aperture objective lenses,” Proc. SPIE, 12152 1215204 https://doi.org/10.1117/12.2622892 PSISDG 0277-786X (2022). Google Scholar

39. 

S. Tereschenko, “Digitale analyse periodischer und transienter Messsignale anhand von Beispielen aus der optischen Präzisionsmesstechnik,” Universität Kassel, (2018). Google Scholar

40. 

I. Abdulhalim, “Spatial and temporal coherence effects in interference microscopy and full-field optical coherence tomography,” Annalen der Physik, 524 (12), 787 –804 https://doi.org/10.1002/andp.201200106 ANPYA2 0003-3804 (2012). Google Scholar

41. 

P. Beckmann and A. Spizzichino, The Scattering of Electromagnetic Waves from Rough Surfaces, Artech House Inc., Norwood (1987). Google Scholar

42. 

L. Hüser and P. Lehmann, “Microsphere assisted interferometry with high numerical apertures for 3D topography measurements,” Appl. Opt., 59 (6), 1695 –1702 https://doi.org/10.1364/AO.379222 APOPAI 0003-6935 (2020). Google Scholar

43. 

H. C. van de Hulst, Light Scattering by Small Particles, Dover Publications, Inc( (1981). Google Scholar

44. 

C. J. R. Sheppard, “Resolution and super-resolution,” Microsc. Res. Tech., 80 590 –598 https://doi.org/10.1002/jemt.22834 MRTEEO 1059-910X (2017). Google Scholar

45. 

T. Pahl et al., “Two-dimensional modelling of systematic surface height deviations in optical interference microscopy based on rigorous near field calculation,” J. Mod. Opt., 67 (11), 963 –973 https://doi.org/10.1080/09500340.2020.1801871 JMOPEW 0950-0340 (2020). Google Scholar

46. 

T. Pahl et al., “3D modeling of coherence scanning interferometry on 2D surfaces using FEM,” Opt. Express, 28 (26), 39807 https://doi.org/10.1364/OE.411167 OPEXFF 1094-4087 (2020). Google Scholar

47. 

A. Darafsheh, “Optical super-resolution and periodical focusing effects by dielectric microspheres,” The University of North Carolina at Charlotte, (2013). Google Scholar

48. 

A. Darafsheh et al., “Super-resolution optical microscopy by using dielectric microwires,” Proc. SPIE, 9713 97130U https://doi.org/10.1117/12.2211431 PSISDG 0277-786X (2016). Google Scholar

49. 

Z. Chen, A. Taflove and V. Backman, “Photonic nanojet enhancement of backscattering of light by nanoparticles: a potential novel visible-light ultramicroscopy technique,” Opt. Express, 12 (7), 1214 https://doi.org/10.1364/OPEX.12.001214 OPEXFF 1094-4087 (2004). Google Scholar

Biography

Lucie Hüser has been working as a research assistant and a PhD candidate in the Measurement Technology Group, Department of Electrical Engineering and Computer Science at the University of Kassel since 2018. She received her master’s degree in electrical engineering in 2017 at the University of Kassel. Her main research areas are interference microscopes with high NAs and near-field support in interference microscopy.

Tobias Pahl received his master’s degree in physics in 2018 at the University of Münster. He has been working as a research assistant and PhD candidate in the Measurement Technology Group, Department of Electrical Engineering and Computer Science at the University of Kassel since 2019. His main research interests are interference and confocal microscopes with high NAs and their modeling.

Marco Künne is a research assistant in the measurement technology group of the University of Kassel, faculty of Electrical Engineering and Computer Science Department. He studied nanoscience at the University of Kassel until 2019. He is working in the research field of high-resolution optical interferometry.

Peter Lehmann studied physics at the Universities of Münster and Karlsruhe. He reached the Dr.-Ing. degree at the University of Bremen in 1994 and finished his Habilitation in 2002. From 2001 until 2008, he was employed at an industrial manufacturer of measuring instruments, where he coordinated research activities in optical metrology. Since then he is a full professor (W3) and holds the chair in measurement technology at the faculty of electrical engineering and computer science of the University of Kassel, Germany. His research interests relate especially to interferometric methods in optical metrology.

CC BY: © The Authors. Published by SPIE under a Creative Commons Attribution 4.0 International License. Distribution or reproduction of this work in whole or in part requires full attribution of the original publication, including its DOI.
Lucie Hüser, Tobias Pahl, Marco Künne, and Peter Lehmann "Microsphere assistance in interference microscopy with high numerical aperture objective lenses," Journal of Optical Microsystems 2(4), 044501 (28 October 2022). https://doi.org/10.1117/1.JOM.2.4.044501
Received: 13 June 2022; Accepted: 26 September 2022; Published: 28 October 2022
Lens.org Logo
CITATIONS
Cited by 7 scholarly publications.
Advertisement
Advertisement
KEYWORDS
Microscopy

Objectives

Diffraction gratings

Lenses

Microscopes

3D image processing

3D modeling

Back to Top