U.S. flag

An official website of the United States government

NCBI Bookshelf. A service of the National Library of Medicine, National Institutes of Health.

König K, Ostendorf A, editors. Optically Induced Nanostructures: Biomedical and Technical Applications. Berlin: De Gruyter; 2015 Jun 23.

Cover of Optically Induced Nanostructures

Optically Induced Nanostructures: Biomedical and Technical Applications.

Show details

Chapter 4Design and fabrication of near- to far-field transformers by sub-100 nm two-photon polymerization

, , , , , , , , and .

In recent years, technological progress in nanotechnology has pushed structure sizes to its limits. As an example, in the semiconductor industry, structures well below 100 nm are routinely produced. The characterization of such structures is a demanding and very important task. Classical microscopy methods do not allow direct imaging in this regime because of the Abbe diffraction limit. Nevertheless, characterization of sub-wavelength structures in the far-field is possible using interferometric Fourier transform scatterometry (IFTS) combined with numerical simulation in a feedback loop. Here, we show that the resolution of this method can be considerably enhanced by use of additional plasmonic nanoantennae structures which transform scattering near-field information into the optical far-field. These structures were realized using different photofabrication approaches.

4.1. Introduction

An electromagnetic near-field around metallic nanostructures is associated with the excitation of localized surface plasmons. The realization of plasmonic near-field to far-field transformers in the optical region between 400 and 800 nm requires structures of sub-wavelength dimensions with a pronounced optical near-field and a strong light scattering ability. When the size of the objects is smaller than the optical wave-length λ, the structures start to scatter the localized plasmonic excitations into the far-field. The enhanced information, now available in the far-field, enables detection of structural feature sizes of sub-wavelength critical dimensions with optical methods. The principle idea of this sub-wavelength far-field characterization of groove and grating structures, providing nanometric resolution, is shown in Fig. 4.1 (a). Additional light scattering from optical near-fields, carrying information about sub-wavelength structural features, into the far-field is accomplished by rod-like nanoantennae.

Fig. 4.1. Principle idea of near-field to far-field transforming plasmonic structures (a). Schematics structures with added plasmonic features as a plasmonic grating (c) or scattering near-field nanoantennae (b).

Fig. 4.1

Principle idea of near-field to far-field transforming plasmonic structures (a). Schematics structures with added plasmonic features as a plasmonic grating (c) or scattering near-field nanoantennae (b).

Fabrication and far-field characterization of sub-wavelength grating structures are presented in this work. Measurements of the scattered light with and without rod nanoantennae from gold-covered samples have been performed and compared with numerical results. Sensitivities to structural dimension of 50 nm have been obtained, using light with a wavelength of 632.8 nm. The fabrication approach for both the sub-wavelength grating structures and the near-to-far-field (NTFF) transformation nanoantennae is based on high-resolution two-photon polymerization (2PP). Repeatable realization of line and dot structures with dimensions down to 50 nm is presented.

4.2. Fabrication approaches for near- to far-field transformation optics

For a demonstration of enhanced optical imaging below the diffraction limit, high-resolution grating structures (see Fig. 4.1 (b & c)), consisting of periodic sequences of lines of different widths have been fabricated with and without NTFF structures. These structures are described in detail in Section 4.3. Examples of test structures with minimum line widths down to 230 nm fabricated in a positive tone photo resist S1805 by microscopic projection photolithography (MPP) [1] are shown in Fig. 4.2.

Fig. 4.2. Structures under consideration for near-field to far-field transformation of plasmonic structures. Upper left: grating structure with sub-wavelength ridge widths of 350 nm fabricated by MPP. Upper right: grating with polymer pillars between the grating ridges as NTFF structures. Bottom left: image of the complete 100 µm × 100 µm structure. Bottom right: achievable line width of 230 to 250 nm by MPP.

Fig. 4.2

Structures under consideration for near-field to far-field transformation of plasmonic structures. Upper left: grating structure with sub-wavelength ridge widths of 350 nm fabricated by MPP. Upper right: grating with polymer pillars between the grating (more...)

The NTFF transforming structures consist of nanoantennas with dimensions of less than 100–300 nm. In order to achieve high flexibility in the realization of the structures and high resolutions down to the sub-100 nm range for both the fabrication of grating NTFF structures, the process of two-photon polymerization (2PP) was applied. Especially for the creation of high aspect ratio NTFF structures, the potential of real 3D structuring using 2PP was exploited.

4.2.1. Two-photon fabrication

Growing interest in downscaling optical components, photonic and plasmonic devices, as well as metamaterial structures together with the demand for their integration has pushed the development of micro- and nanofabrication methods [2, 3]. One method for the generation of 3D structures in rapid prototyping is UV stereolithography, developed in 1986 by Hull [4]. This technology, although flexible and cost-efficient, is limited in resolution to several micrometers and in fabrication speed [5]. To overcome these limitations, nonlinear light-matter interactions can be used in true 3D structuring. The method of nonlinear laser structuring has been demonstrated to be a promising candidate regarding flexibility, cost efficiency, and the ability to generate arbitrary complex 3D structures.

The process of nonlinear laser writing, as used in this study, is based on two-photon absorption (TPA). TPA was theoretically investigated by Göppert-Mayer in 1930 and 1931 [6, 7]. Kaiser and Garret demonstrated it experimentally in 1961 after the invention of the laser by two-photon excitation of CaF2:Eu2+ [8]. These investigations of nonlinear optical and especially two-photon processes became possible with the advent of the laser. But only since the development of commercially available ultrashort pulsed titanium:sapphire lasers in the 1990s have two-photon absorption processes been routinely applied in microscopy and microfabrication.

The potential of TPA for 3D laser structuring was suggested in 1992 by Wu et al. with the idea for applications in microelectronics [9]. 3D microstructures with a resolution down to 500 nm have already been demonstrated. Three-dimensional microfabrication was later introduced by Maruo et al. in 1997, using a standard polymer SCR-500 [10]. The potential for subwavelength resolution has been pointed out. Since laser radiation can be focused directly into the material volume and is only absorbed in the focal region of the beam, the process speed can be much faster compared to conventional 3D stereolithography, and the fabricated structures can have a much higher complexity. The term two-photon polymerization (2PP), which has been given to this technology, was introduced in 2003, when this technique was applied to radical polymerization of a new class of organic-inorganic hybrid materials [11]. This method was used independently one year later in 2004 by Deubel [12] and Serbin [13] to realize 3D photonic crystals. Today, TPA is also used with polymer systems which can be cationically polymerized (such as the epoxy-based SU8 polymer), or which undergo chemical bond cleavage, like the commercial photoresists used in microelectronics. A detailed overview of this technology up to 2004 is given in [14]. The developments of 2PP technology within the last 10 years are reviewed in [15].

In the following section, the focus is on radical polymerization. The radicals that lead to polymerization are formed in the triplet state of certain photoinitiator molecules upon single- or multiphoton excitation. Typical radical photoinitiators undergo intersystem crossing from the excited singlet to the triplet state on a time scale of the order of 100 ps [16], whereas a direct decay of the excited state to the ground state would lead to fluorescence. In 2PP, the photoinitiator in a (pre-) polymer resin is excited by the simultaneous absorption of two photons. The absorption probability thus depends on the square of the laser intensity. As a consequence, ultrafast laser systems have commonly been used to provide high peak intensities at low average power. The ultrashort laser pulses are focused through a microscope objective such that the photons are confined in space and time. The laser power in 2PP applications is set to a level that the intensity of the light is only high enough in a small region of the focal volume of the beam. In most implementations of 2PP, TPA initiates crosslinking of polymer molecules. By moving the laser focus through the volume of the resin, arbitrary complex, computer generated three dimensional (3D) polymeric [1722] and high resolution 2D plasmonic structures can be created [2326].

Due to the optical nonlinearity of TPA, an intensity threshold exists for 2PP. 2PP can thus create volume elements (voxels) which are considerably smaller than the wavelength of the laser beam, e.g. 800 nm for titanium:sapphire lasers. Although intrinsically capable of producing arbitrary narrow line widths and heights, the 2PP process has encountered a persistent limit of about 100 nm due to potential instabilities of the laser beam power, the positioning systems used, and radical diffusion [15]. Different strategies have been followed to overcome this limit. The first approach was based on radical quenchers to avoid the diffusion of radicals out of the focal volume where the laser intensity is above the 2PP threshold [27]. A further step was the use of more efficient photoinitiators which lead to local crosslinking of the polymer faster than the created radicals can diffuse into the surrounding material regions [28]. With these two strategies, lateral widths of laser-written polymer lines on glass substrate surfaces of 100 nm and 80 nm respectively were obtained. Recently, processes based on the principle of stimulated emission depletion (STED) of fluorophore molecules have been applied [2931]. Here, the widths of laser-written lines on glass surfaces have been reduced by efficient de-excitation of the excited photoinitiator molecules using a second laser beam with ring-shaped intensity distribution and either similar [29] or different [30, 31] wavelengths. Feature sizes of 40 nm [29] and 65 nm [31], respectively, have been demonstrated.

In the bulk volume, however, this approach has also been limited thus far to 100 nm lateral line width realized in photonic crystals [31]. A drawback of this technological approach is its sensitivity to misalignments. The smallest reliable 3D line widths of 65 nm in photonic crystals [32] and of 45 nm for supported single lines [33] to date were obtained using either highly efficient specialized photoinitiators or ultrashort pulse durations down to 10 fs respectively. To this date, there is no common simple recipe for the reliable generation of 3D structures with dimensional features in the sub-100 nm range. One reason for this is the decreasing stability of the polymerized structures.

Two approaches have been taken to overcome the persistent limitations for the production of sub-100 nm structures inside the resin volume and on dielectric (glass) surfaces: the use of additional crosslinking molecules to increase the stability of fabricated 3D structures in zirconium-based inorganic-organic hybrid polymer, and careful investigation of the actually achievable resolution in the commercially available small molecule acrylate resin E-Shell 300 using the 2PP technique without further specialized ingredients.

4.2.2. 3D two-photon polymerization with additional crosslinker

The influence of the crosslinker dipentaerythritol penta-methacrylate (DPMA) on the achievable dimensions of supported isolated polymer lines was investigated in the experiments.

4.2.2.1. Materials

A zirconium-based inorganic-organic hybrid polymer (Zr-Hypo) was used as basis material for the investigations. The material has been demonstrated to exhibit only minor volume shrinkage, which is favorable for high resolution structuring [34]. Synthesis of the zirconium-based inorganic-organic hybrid polymer was carried out by a sol-gel-process [35]. The commonly available radical photoinitiator 4,4′-bis (dimethylamino) benzophenone (Bis) was used for initiation of the polymerization process. A diagram of the synthesis protocol is shown in Fig. 4.3.

Fig. 4.3. Synthesis of zirconium-based inorganic-organic hybrid polymer.

Fig. 4.3

Synthesis of zirconium-based inorganic-organic hybrid polymer.

Methacryloxypropyltrimethoxysilane (MAPTMS, Sigma-Aldrich) was hydrolyzed with 0.1 M hydrochloric acid (HCl, Sigma-Aldrich). Methacrylic acid and zirconium isopropoxide were mixed separately in a molar ratio of 4:1 and stirred for 30 min.

The chelated zirconium isopropoxide was subsequently added to the hydrolyzed silane precursor, and a small amount of water added. The molecular ratio of MAPTMS to ZPO was set to 5:1 (20 %). The mixture was stirred for another 30 min. At the end, 0.5 wt % of the photoinitiator was added. Before structuring, the pre-polymer was drop-casted onto a microscope cover glass slide with a thickness of 150 µm and heated for 1 h to 100 °C. A mixture of 50 vol % isopropanol (Iso, Sigma Aldrich) and 50 vol % 4-methyl-2-pentanone (methyl isobutylketone, MIBK, Sigma Aldrich) was used to develop the realized structures, i.e. to dissolve the non-polymerized portion of the material. The samples were immersed in this mixture for 1 h. Line widths in woodpile crystal structures of 320 nm have been demonstrated with this material [34].

The performance of this material was compared to the same material with an addition of 15% crosslinker DPMA for the realization of sub-100 nm structures. The simple change in the synthesis protocol is shown in Fig. 4.4. The amount of DPMA was added to the basis material described above just before addition of the photoinitiator, which again was 0.5 wt % Bis.

Fig. 4.4. Synthesis of DPMA-amplified Zr-Hypo.

Fig. 4.4

Synthesis of DPMA-amplified Zr-Hypo.

2PP of these two materials was initiated using the Ti:Sapphire based high-peak power femtosecond laser oscillator Femtosource scientific XL500 (Femtolasers) with a repetition rate of 5 MHz, a central wavelength of 800 nm, and a pulse width smaller than 50 fs. The laser pulses were spatially focused into the bulk volume of the photosensitive material by a microscope objective lens. A 100× immersion oil microscope objective with a numerical aperture of 1.4 was used to obtain the best focus. The laser beam was focused from the back of the cover glass plate. In order to generate the test structures, the focus was scanned through the polymer volume by moving the sample with a computer controlled positioning system (Physik Instrumente PI, 3x M-126.DG for x-, y-, and z-axes). The scanning speed of the laser beam was kept at 1 mm/s. The laser beam focusing in this configuration is fixed.

For the development of the realized structures, i.e. to dissolve the non-polymerized portion of the material, a mixture of 50 vol % isopropanol (Iso, Sigma Aldrich) with 50 vol % 4-methyl-2-pentanone (methyl isobutylketone, MIBK, Sigma Aldrich) was used. The samples were immersed in this mixture for 1 h.

4.2.2.2. Experimental results

The writing speed was fixed at 1 mm/s. The applied average laser power varied within the range of 0.1–10 mW. The line widths and heights were measured depending on the applied average laser power using high resolution SEM images. The line widths and heights measured for the two materials Zr-Hypo and DPMA-Zr-Hypo are shown in Fig. 4.5. Note that the applied average laser power in the case of the non-amplified Zr-Hypo has to be multiplied by a factor of 10. At first glance it is obvious that there is an abrupt cut-off for the achievable widths and heights towards low laser powers in all cases. This cut-off is due to the stability of the materials. Below a certain laser power, and thus a certain achieved polymerized material strength, the structures do not withstand the development process. The structures are destroyed by the evaporating solvent mixture. Further, it can be seen that the smallest lines can indeed be achieved with amplified DPMA-Zr-Hypo for a minimum applicable average laser power of 0.5 mW, whereas the smallest lines were realized for the Zr-Hypo for an average laser power of 1 mW.

Fig. 4.5. Supported line widths (left) and heights (right) achieved for the two polymers Zr-Hypo (diamonds) and DPMA-Zr-Hypo (circles).

Fig. 4.5

Supported line widths (left) and heights (right) achieved for the two polymers Zr-Hypo (diamonds) and DPMA-Zr-Hypo (circles).

The results and the achieved minimum line widths and heights for Zr-Hypo and DPMA-Zr-Hypo are summarized in Fig. 4.6. In the non-amplified material, the smallest reproducible line width achieved was around 150 nm. Smaller polymerized structures have principally been observed in the online process observation, but were washed away during development.

Fig. 4.6. Summary of the structuring of Zr-Hypo and amplified DPMA-Zr-Hypo. Lines of less than 82.5 nm in width were reproducibly fabricated under the influence of an additional crosslinker.

Fig. 4.6

Summary of the structuring of Zr-Hypo and amplified DPMA-Zr-Hypo. Lines of less than 82.5 nm in width were reproducibly fabricated under the influence of an additional crosslinker.

The reproducible smallest line width achieved in the amplified polymer containing 15 % DPMA was measured to be 82.5 nm according to the SEM images. It should be noted that the structures had been isotropically covered with a gold layer of at least 10 nm thick by sputtering in a gas discharge with a background gas pressure of 10 mbar for SEM inspection [36]. Under these conditions the achieved minimum line width can be estimated to be smaller than 62 nm. In further experiments, free-standing polymer structures with high aspect ratios of 10:1 were also obtained. The smallest diameter of a rod-like nanoantenna has been measured to be 40 nm.

4.2.3. 2D two-photon polymerization of E-Shell 300 on dielectric surfaces

2PP has also been investigated with materials which are cheap and easily available commercially. One of those materials is E-Shell 300, which is commonly used for 3D printing and hearing aid fabrication. The material has a high flexural strength up to 88.4 MPa and is suitable for the fabrication of structures with dimensions below 100 nm. The photoinitiator used in the commercial product is Irgacure 127, at a concentration of 5 wt %. Initial experiments have shown that the material can directly be structured by 2PP using frequency doubled ytterbium:potassium-gadolinium-tungstate laser radiation with 200 fs pulse duration. A laser power variation was performed to quantify the material properties and the structural sizes obtainable with 2PP. The exposure time for the fabrication of single voxels was fixed to 20 ms. A typical result is shown in Fig. 4.7. In the first test of the material variable voxel sizes ranging from 200 nm down to 50 nm were obtained.

Fig. 4.7. 2PP with E-Shell 300: fabrication of voxels for material characterization. Upper left: large array of pillar structures. Upper right: power variation for a fixed exposure time of 20 ms. Minimum voxel diameters of 53 nm have been measured. A grating structure with 100 nm line width and 350 nm period is shown bottom left. By careful slow scanning and low laser powers of 1 mW, line widths of 58 down to as low as 50 nm respectively (bottom right) were obtained.

Fig. 4.7

2PP with E-Shell 300: fabrication of voxels for material characterization. Upper left: large array of pillar structures. Upper right: power variation for a fixed exposure time of 20 ms. Minimum voxel diameters of 53 nm have been measured. A grating structure (more...)

Line structures have also been fabricated with the same material. Gratings consisting of polymer lines with diameters of 100 nm and below have been realized, as demonstrated in Fig. 4.7. The writing speed was set to 10 µm/s and the laser power adjusted to 2 mW.

By carefully adjusting the laser focus with respect to the substrate surface, line widths of 58 nm down to 50 nm were obtained for applied laser powers of 1 mW, and slow scanning speeds of 5 µm/s and 7 µm/s respectively. At these dimensions, even very small fluctuations in the laser beam power and the positioning stages have huge influences on the structuring results. These fluctuations are slightly visible in the images in Fig. 4.7. They have no influence on structures with a critical dimension greater than 100 nm however.

Gratings with and without NTFF nanoantennae have been realized from E-Shell 300. The results are shown in Fig. 4.8. The period of the grating was chosen to be 450 nm. A line width of 95 nm and a height of 70 nm for the grating ridges were obtained for a scanning speed of 7 µm/s and a laser power of 2 mW in front of the focusing microscope objective. The NTFF nanoantennae with 500 nm period are cylindrical pillars with a rounded top of 110 nm diameter and a height of 150 nm.

Fig. 4.8. Grating and NTFF structures fabricated by high-resolution 2PP. Upper left: overview image of an exemplary 20 µm × 20 µm grating with nanorod antennae. Upper right: magnified image showing the regularity of the structure. Bottom left: side view, period of NTFF nanoantennae is 500 nm. Bottom right: front view, line width is 95 nm, line height is 70 nm. The diameter of the NTFF structures is 110 nm and height 150 nm.

Fig. 4.8

Grating and NTFF structures fabricated by high-resolution 2PP. Upper left: overview image of an exemplary 20 µm × 20 µm grating with nanorod antennae. Upper right: magnified image showing the regularity of the structure. Bottom (more...)

4.3. Fourier scatterometry on gratings with designed near-field structures

The characterization of sub-lambda structures with a fast non-destructive optical method is a common metrology task.

One of the most widely used techniques for these tasks is the scatterometry method [37]. Scatterometry is a method for analysis of the diffraction spectrum from a periodic arrangement of nanostructures. It is used to reconstruct the unknown structure parameters by comparison of measured and simulated spectra, which is why it belongs to the field of so-called model-based metrology methods [38].

There have been various attempts to improve scatterometric sensitivity, mainly by increasing the illumination wavelength range used, variation of the incidence angle or a high numerical aperture illumination, as well as combinations with other measurement methods, such as for example white light interferometry [38]. While all of these methods increase the sensitivity of the measurements by increasing the information content of the measurement itself, a different approach was taken in the work presented. The simulation branch needed for all scatterometric measurements can itself be exploited to design an optimized scatterometric sensitivity inherent to the structure. While this may not be generally applicable due to practical constraints, it is conceivable for alignment, testing or calibration targets.

While we have already shown some promising results based on simulations [39], we now want to verify these results with experimental data. We apply the so-called Fourier scatterometry method [38] to detect the backscattered light of polymer test line gratings fabricated as described in Section 4.2. We then analyze the influence of added near-field structures on sensitivity to line width. In contrast to the simulation study mentioned, we used a brightfield illumination with a broad range of incident directions containing much more information. In addition to new adapted simulations we show a first experimental implementation and verification of the result obtained previously.

4.3.1. Simulation method

Simulations of the Fourier scatterometry technique are performed with our simulation tool, MicroSim [40]. It is based on the rigorous coupled wave analysis method [41] for diffraction from arbitrary three-dimensional structures and includes some improved convergence methods [42].

The extended illumination pupil is modeled by planar waves emerging from a quadratic grid, which is sampled with equidistant points corresponding to the different incident angles. Due to the symmetry of the structure studied, only one-quarter of the pupil was simulated. This quarter was sampled by a 99 × 99 grid. The points are defined by their NA coordinates, NAx and NAy. For each of these points, the diffracted spectrum from the sample is calculated. This high sampling is needed due to the very fine features which can be found in the pupil images caused by higher diffraction orders of the illuminated gratings. A wavelength of 617 nm was assumed.

Finally, the calculated fields for every incidence angle were coherently superposed to obtain the resulting pupil images. The analyzed periodic grating was modeled with a staircase approximation. The width of the lines (CD) was varied and the sensitivity towards those changes was obtained from the difference in the corresponding pupil images. These sensitivities were compared to the ones calculated after the introduction of near-field structures (nanorods) between the lines of the gratings.

4.3.2. Experimental setup

We built a Fourier scatterometry setup which will be explained in the following section. It was completely redesigned and based on a setup used in previous works [39]

We start with the illumination path. A red LED with a wavelength of 617 nm is used as light source. We use a microscope objective (10×) to magnify and depict the end of the polished fiber on the back focal plane of the microscope objective used to illuminate the sample. A polarizer allows selection of s or p polarized illumination. The image of the fiber covers the complete aperture of the backfocal plane of the objective, producing a homogenous intensity distribution. We use a microscope objective with a NA of 0.95 (250× magnification); this means that the sample is illuminated with plane waves covering the angles from 0 to 72°. The backscattered light from the analysed sample is again collected by the same objective, as our microscope works in reflection mode. A beamsplitter is used to image the backfocal plane of the objective and the image plane of the sample at the same time. The backfocal plane is imaged with the help of a Bertrand lens on a CCD camera, while the image plane is imaged with help of a matched tube lens. Both lenses are specially designed and aberration -corrected for the microscope objective used.

Besides the illumination and imaging path of the Fourier scatterometry microscope, a reference arm for a Linnik type setup is also included in the setup. It includes a matched objective with identical NA as the imaging objective and allows imaging of the interfering backfocal planes of both objectives. This type of measurement is detailed in [39] but will not be used in this setup. The reference path is blocked with an absorbing plate. The CAD construction of the setup as well as the schematic setup showing the detailed light paths can be found in Fig. 4.9.

Fig. 4.9. CAD design of the setup (top) and scheme of the experimental setup showing the different beam paths (bottom).

Fig. 4.9

CAD design of the setup (top) and scheme of the experimental setup showing the different beam paths (bottom).

A typical measurement consists of searching the grating to be analyzed on the sample using classical imaging microscopy and focusing directly on top of the structure and then recording the pupil plane images for both polarizations (s and p). Using structures which only vary in width and comparing the images allows extraction of the sensitivity of the signal towards changes of the line width.

4.4. Experimental results

First the structures are characterized with an electron microscope to get an impression of the different structure parameters. The images are shown in Section 4.2. This data is used for the modelling of the structure to be used by the simulations. We define the sensitivity towards a parameter as the change in the pupil plane intensity compared to the absolute changes of the parameter.

The analyzed structures consist of polymer test line gratings fabricated as described in Section 4.2. The period of the lines is 1800 nm, while the height is 500 nm. There are gratings with different line widths (250 and 350 nm). The complete grating is covered with a 20 nm layer of gold.

Pupil images for these structures were taken for both s- and p-polarization. The resulting images and the difference in pupil plane intensity can be found in Fig. 4.10 together with the results of the same measurements and simulations with added nanorods between the lines. In the case with rods, the linewidths are 380 nm and 420 nm, meaning that the difference is smaller than for the lines without the rods (250 to 350 nm). The rod itself has a radius of approx. 175 nm and a height of 500 nm. Again the structures are covered with 20 nm of gold.

Fig. 4.10. Resulting pupil images (simulated and measured) for s- and p-polarization for gratings with and without nanorods between the lines, as well as the corresponding differences when the CD of the gratings is varied. For the lines without nanorods the compared CD values are 250 and 350 nm. For the lines with nanorods the values are 380 and 420 nm. The radius of the rods is 175 nm. The difference in intensity is a direct indication of the sensitivity towards the change in line width (CD).

Fig. 4.10

Resulting pupil images (simulated and measured) for s- and p-polarization for gratings with and without nanorods between the lines, as well as the corresponding differences when the CD of the gratings is varied. For the lines without nanorods the compared (more...)

Differences between measured and simulated pupil images can have different reasons. The most important is the simplified simulation model which had to be used due to the very large calculation times. The grating was modelled with perfectly steep walls; neither roundings nor line edge roughness was taken into account. The SEM pictures and microscope pictures show that this is a very rough assumption. Additionally, till now we have not taken any aberrations of the optical elements used in the setup into account. The LED light source is modelled as perfectly homogenous and with a very narrow band of only 1 nm wavelength range, while realistically 20 nm should be assumed. Each wavelength would need one simulation and the global simulation time for that would increase drastically. Additionally, the agreement between simulation and experiment is worse in the case of added nanorods. This could indicate that a more sophisticated modelling of the nanorods is needed, which would lead to significantly increased computation time.

The measured and simulated results show comparable absolute differences in pupil intensity for the gratings both with and without rods. It must be noted, however, that the line widths analyzed do not agree for both cases. In the case of the lines without rods, the line width differs from approx. 250 to 350 nm, a difference of 100 nm. While the widths for the lines with rods are much closer; 380 compared to 420 nm, a difference of only 40 nm. The results show almost identical differences in intensity. For that the system is much more sensitive for the case of a grating with rods. This proves the improved sensitivity obtained by adding scattering near-field structures.

4.5. Conclusions

Based on the assumption of higher scatterometric sensitivity of structures with added near- to far-field transformers designed by previous simulation-based studies, we have now provided a first experimental verification for measuring critical dimensions of gratings with resolutions as low as 50 nm. 2PP methods have been developed, improved, and successfully applied for the fabrication of sub-wavelength structures. The experimental results supported by corresponding rigorous simulations show the expected sensitivity gain in the scatterometric signatures.

Acknowledgments

The authors acknowledge contributing work by Tim Fischer and Ayman El-Tamer for the preparation of 3D sub-100 nm structures.

References

1.
Love JC, Wolfe DB, Jacobs HO, Whitesides GM. Microscope Projection Photolithography for Rapid Prototyping of Masters with Micron-Scale Features for Use in Soft Lithography. Langmuir. 2001;17:6005–6012.
2.
Köhler M, Fritzsche W. Nanotechnology. 2nd Edition. Wiley; 2007.
3.
Madou M. Fundamentals of Microfabrication and Nanotechnology. Boca Raton: CRC Press; 2012.
4.
Hull CW. Apparatus for Production of Three-Dimensional Objects by Stereolithography. 4,575,330. U.S. Patent. 1982
5.
Choi JW, MacDonald E, Wicker R. Multi-material microstereolithography. International Journal of Advanced Manufacturing Technology. 2010;49:543–551.
6.
Göppert Mayer M. Über Elementarakte mit zwei Quantensprüngen, Dissertation. Univ. Göttingen; 1930.
7.
Göppert Mayer M. Über Elementarakte mit zwei Quantensprüngen. Annalen der Physik. 1931;9:273–294.
8.
Kaiser W, Garrett CGB. Two-photon excitation in CaF2:Eu2+ Phys. Rev. Lett. 1961;7:229–231.
9.
Wu ES, Strickler JH, Harrell WR, Webb WW. Two-photon lithography for microelectronic application. Proceedings Article, Proc. SPIE. 1992;1674:776–782.
10.
Maruo S, Nakamura O, Kawata S. Three-dimensional microfabrication with twophotonabsorbed photopolymerization. Optics Letters. 1997;22:132–134. [PubMed: 18183126]
11.
Serbin J, Egbert A, Ostendorf A, Chichkov BN, Houbertz R, Domann G, Schulz J, Cronauer C, Fröhlich L, Popall M. Femtosecond laser-induced two-photon polymerization of inorganic organic hybrid materials for applications in photonics. Optics Letters. 2003;28:301–303. [PubMed: 12659425]
12.
Deubel M, von Freymann G, Wegener M, Pereira S, Busch K, Soukoulis CM. Direct laser writing of three-dimensional photonic-crystal templates for telecommunications. Nature Materials. 2004;3:444–447. [PubMed: 15195083]
13.
Serbin J, Ovsianikov A, Chichkov BN. Fabrication of woodpile structures by two-photon polymerization and investigation of their optical properties. Optics Express. 2004;12:5221–5228. [PubMed: 19484080]
14.
Sun H-B, Kawata S. Two-Photon Photopolymerization and 3D Lithographic Microfabrication. Advances in Polymer Science. 2004;170:169–273.
15.
Malinauskas M, Farsari M, Piskarskas A, Juodkazis S. Ultrafast laser nanostructuring of photopolymers: a decade of advances. Physics Reports. 2013;533:1–31.
16.
Colley CS, Grills DC, Besley NA, Jockusch S, Matousek P, Parker AW, Towrie MI, Turro NJ, Gill PMW, George MW. Probing the reactivity of photoinitiators for free radical polymerization: time-resolved infrared spectroscopic study of benzoyl radicals. Journal of the American Chemical Society. 2002;124:14952–14958. [PubMed: 12475337]
17.
Schizas C, Melissinaki V, Gaidukeviciute A, Reinhardt C, Ohrt C, Dedoussis V, Chichkov BN, Fotakis C, Farsari M, Karalekas D. On the design and fabrication by two-photon polymerization of a readily assembled micro-valve. International Journal of Advanced Manufacturing Technology. 2010;48:435–441.
18.
Koch J, Bauer T, Reinhardt C, Chichkov BN. Recent Advances in Laser Material Processing. Elsevier; 2006. Recent progress in direct write 2D and 3D photofabrication technique with femtosecond laser pulses, Chapter 8; pp. 472–495.
19.
Claeyssens F, Hasan EA, Gaidukeviciute A, Achilleos DS, Ranella A, Reinhardt C, Ovsianikov A, Shizhou X, Fotakis C, Vamvakaki M, Chichkov BN, Farsari M. Three- Dimensional Biodegradable Structures Fabricated by Two-Photon Polymerization. Langmuir. 2009;25:3219–3223. [PubMed: 19437724]
20.
Passinger S, Saifullah MS, Reinhardt C, Subramanian KRV, Chichkov BN, Welland ME. Direct 3D Patterning of TiO2 Using Femtosecond Laser Pulses. Advanced Materials. 2007;19:1218–1221.
21.
Sakellari I, Gaidukeviciute A, Giakoumaki A, Gray D, Fotakis C, Farsari M, Vamvakaki M, Reinhardt C, Ovsianikov A, Chichkov BN. Two-photon polymerization of titanium-containing sol-gel composites for three-dimensional structure fabrication. Applied Physics A. 2010;100:359–364.
22.
Terzaki K, Vasilantonakis N, Gaidukeviciute A, Reinhardt C, Fotakis C, Vamvakaki M, Farsari M. 3D conducting nanostructures fabricated using direct laser writing. Optical Materials Express. 2011;1:58–597.
23.
Reinhardt C, Kiyan R, Passinger S, Stepanov AL, Ostendorf A, Chichkov BN. Rapid laser prototyping of plasmonic components. Applied Physics A. 2007;89:321–325.
24.
Reinhardt C, Seidel A, Evlyukhin AB, Cheng W, Kiyan R, Chichkov BN. Direct laser-writing of dielectric-loaded surface plasmon-polariton waveguides for the visible and near infrared. Appl. Phys. A. 2010;100:347–352.
25.
Reinhardt C, Evlyukhin AB, Cheng W, Birr T, Markov A, Ung B, Skorobogatiy M, Chichkov BN. Bandgap-confined large-mode waveguides for surface plasmon-polaritons. Journal of the Optical Society of America B. 2013;30:2898–2905.
26.
Reinhardt C, Seidel A, Evlyukhin AB, Cheng W, Chichkov BN. Mode selective excitation of laser-written dielectric-loaded surface plasmon polariton waveguides. Journal of the Optical Society of America B. 2009;26:B55–B60.
27.
Takada K, Sun H-B, Kawata S. Improved spatial resolution and surface roughness in photopolymerization-based nanowriting. Applied Physics Letters. 2005;86:071122.
28.
Xing J-F, Dong X-Z, Chen W-Q, Duana X-M, Takeyasu N, Tanaka T, Kawata S. Improving spatial resolution of two-photon microfabrication by using photoinitiator with high initiating efficiency. Applied Physics Letters. 2007;90:131106.
29.
Li L, Gattass RR, Gershgoren E, Hwang H, Fourkas JT. Achieving λ/20 Resolution by One-Color Initiation and Deactivation of Polymerization. Science. 2009;324:910–913. [PubMed: 19359543]
30.
Fischer J, Wegener M. Three-dimensional direct laser writing inspired by stimulated-emission-depletion microscopy. Optical Materials Express. 2011;1:614–624.
31.
Fischer J, Wegener M. Three-dimensional optical laser lithography beyond the diffraction limit. Laser Photonics Review. 2013;7:22–44.
32.
Haske W, Chen VW, Hales JM, Dong W, Barlow S, Marder SR, Perry JW. 65nm feature sizes using visible wavelength 3-D multiphoton lithography. Optics Express. 2007;15:3426–3436. [PubMed: 19532584]
33.
Emons M, Obata K, Binhammer T, Ovsianikov A, Chichkov BN, Morgner U. Two-photon polymerization technique with sub-50 nm resolution by sub-10 fs laser pulses. Optical Materials Express. 2012;2:942–947.
34.
Ovsianikov A, Shizhou X, Chichkov BN, Farsari M, Vamvakaki M, Fotakis C. Optics Express. 2009;17:2143–2148. [PubMed: 19219118]
35.
Bhuian B, Winfield RJ, O’Brien S, Crean GM. Applied Surface Science. 2006;252:4845–4849.
36.
Ferreras Paz V, Emons M, Obata K, Ovsianikov A, Peterhänsel S, Frenner K, Reinhardt C, Chichkov B, Morgner U, Osten W. Development of functional sub-100 nm structures with 3D two-photon polymerization technique and optical methods for characterization. Journal of Laser Application. 2012;24:042004.
37.
Raymond CJ. Scatterometry for Semiconductor Metrology. In: Diebold AC, editor. Handbook of silicon semiconductor metrology. CRC Press; 2001. pp. 477–513.
38.
Ferreras Paz V, Peterhänsel S, Frenner K, Osten W. Solving the inverse grating problem by white light interference Fourier scatterometry. Nature Light: Science & Applications. 2012;1(11):e36.
39.
Ferreras Paz V, Frenner K, Osten W. Increasing Scatterometric Sensitivity by Simulation Based Optimization of Structure Design. In: Osten W, editor. Fringe 2013 - 7th International Workshop on Advanced Optical Imaging and Metrology. Berlin Heidelberg: Springer; 2014. pp. 345–348.
40.
Totzeck M. Numerical simulation of high-NA quantitative polarization microscopy and corresponding near-fields. Optik - International Journal for Light and Electron Optics. 2001;112(9):399–406.
41.
Moharam MG, Gaylord TK. Rigorous coupled-wave analysis of planar-grating diffraction. Journal of the Optical Society of America. 1981;71(7):811.
42.
Schuster T, Ruoff J, Kerwien N, Rafler S, Osten W. Normal vector method for convergence improvement using the RCWA for crossed gratings. Journal of the Optical Society of America A. 2007;24(9):2880.
© 2015 C. Reinhardt et al., published by De Gruyter.

This work is licensed under the Creative Commons Attribution-NonCommercial-NoDerivs 3.0 License.

Bookshelf ID: NBK321738PMID: 26491776

Views

Similar articles in PubMed

See reviews...See all...

Recent Activity

Your browsing activity is empty.

Activity recording is turned off.

Turn recording back on

See more...