Open Access
22 May 2015 Optical microlithography on oblique and multiplane surfaces using diffractive phase masks
Author Affiliations +
Abstract
Micropatterning on oblique and multiplane surfaces remains a challenge in microelectronics, microelectromechanics, and photonics industries. We describe the use of numerically optimized diffractive phase masks to project microscale patterns onto photoresist-coated oblique and multiplane surfaces. Intriguingly, we were able to pattern a surface at 90 deg to the phase mask, which suggests the potential of our technique to pattern onto surfaces of extreme curvature. Further studies show that mask fabrication error of below 40-nm suffices to conserve pattern fidelity. A resolution of 3  μm and a depth-of-focus of 55  μm are essentially dictated by the design parameters, the mask generation tool, and the exposure system. The presented method can be readily extended for simple and inexpensive three-dimensional micropatterning.

1.

Introduction

Microstructures on oblique and nonplanar surfaces enable unique functionalities in photonics,1,2 electronics,3 and microelectromechanics,4,5 and provide a broad array of interesting applications in high-gain antennas,6 radio-frequency identification devices,7 metamaterials,8 and transformation optics.9 For instance, combining diffractive microstructures atop a curved refractive surface can minimize aberrations in lenses, in a more compact way compared to the conventional methods by adaptive optics.10 In addition, micropatterning on the sidewalls of implantable neural probes could potentially lead to an effective approach of recording three-dimensional (3-D) neural signals.11,12 Meanwhile, shape modification in the vertical direction of microfluidic channels and 3-D integration may significantly enhance their performances.5,13

Optical projection lithography (OPL) on planar silicon substrates is the workhorse of the semiconductor industry due to its high throughput, resolution, and accuracy.14 In OPL, a photomask pattern is imaged with demagnification onto a planar photoresist layer that coats the silicon substrate. In general, it is difficult to utilize OPL to pattern nonplanar or oblique surfaces due to the limitations of the imaging optics. These limitations can be avoided by lens-less lithography that utilizes computer-generated holograms (CGHs) to directly project patterns onto the photoresist surface.15 Thus far, these approaches only project the pattern onto a single plane surface. In this paper, we extend this technique by designing diffractive optics that can manipulate the intensity of light in 3-D space, and thereby allow for patterning onto nonplanar and oblique surfaces. An alternative approach for lithography on nonplanar and oblique surfaces is to utilize a flexible template that contains a master pattern and apply this template conformally over the substrate. The pattern may be transferred via an imprint process1618 or simply by exposure to ultra-violet (UV) light through the template.19 These approaches require contact with the substrate surface, which increases the potential for damage, increases defects, and reduces yield. Furthermore, these approaches perform well only for surfaces of small curvature since conformal contact is necessary. In contrast, our approach can be applied to surfaces of extreme curvature as illustrated by patterning of a substrate that is placed orthogonal to the diffractive mask.

2.

Lithography Principle

Our approach is schematically explained in Fig. 1. A spatially collimated, temporally coherent uniform UV beam illuminates the mask [see Fig. 1(a)]. The sample to be exposed is placed at a specified distance behind the mask, where the light intensity distribution in three dimensions is controlled. The mask is designed using an enhanced direct-binary-search algorithm,2026 where the optimization objective is to maximize intensity within prescribed patterns in multiple planes or within a 3-D volume. This numerical technique was successfully implemented in designing various diffractive-optical elements.27,28 In addition, the intensity uniformity within the target image has to be taken into account. Note that in optical lithography, the (positive-tone) photoresist serves as a nonlinear recording medium, where regions receiving energy higher than a threshold are selectively dissolved away in a developer. Hence, the mask only needs to ensure that the desired regions receive energy (which corresponds to light intensity multiplied by the exposure time) above a certain threshold (defined by the sensitivity of the photoresist at the exposure wavelength). The resulting mask comprised an array of discrete pixels, either in one dimension (along the X direction) or two dimensions (on the XY plane), where each pixel in the array applies a phase shift to the incident light. This resembles a traditional CGH, which is usually used to generate complex beams or images.29 The phase shift of each pixel is controlled during optimization. The array of optimal phase shifts is implemented as an array of pixels with varying heights, which are uniformly quantized by a unit height Δh. In its 1-D form [Fig. 1(a)], Δx denotes its uniform pixel size. The diffractive mask was fabricated into a polymer layer using grayscale lithography21,22,26,30 on a laser-pattern generator. Since the resolution of this process was limited to 3μm, we constrained all our mask designs to pixels of size 3μm or larger. The maximum thickness of the polymer layer H (and hence, of each pixel in the array) was chosen so as to achieve the maximum phase shift of 2π. For a polymer refractive index of 1.76 and illumination wavelength of 325 nm, this corresponded to 430 nm.

Fig. 1

(a) Schematic of the cross-sectional view of a one-dimensional (1-D) diffractive phase mask. (b) Schematic of microlithography on multiple planes by a 1-D mask, (d) and a two-dimensional (2-D) mask. Intensity distributions U’ are generated at distances separated by Δd. (c) Target exposure images at three planes: nine lines at z=d1=d0=80mm with 30-μm linewidth and 60μm spacing; three lines at z=d1=d0+Δd=81mm with 90μm linewidth and 180μm spacing; five lines at z=d1=d0+2Δd=82mm with 60μm linewidth and 120μm spacing. (e) Target exposure images with 180μm×180μm period at four planes: letter “U” at z=d1=d0=5.0mm; letter “T” at z=d1=d0+Δd=5.3mm; letter “A” at z=d1=d0+2Δd=5.6mm; letter “H” at z=d1=d0+3Δd=5.9mm. Linewidth is 18μm.

JM3_14_2_023507_f001.png

Two mask designs in 1-D [Figs. 1(b) and 1(c)] and 2D [Figs. 1(d) and 1(e)] were optimized. The light intensity patterns at multiple planes (or in 3-D volume) U(x,y,d) can be derived based on the transmission function of the mask T(x,y), which describes phase modulation. In the 1-D case, the desired patterns are three groups of periodic lines. Linewidths of 30, 90, 60μm and spacings of 60, 180, and 120μm are designated at z=d1=d0=80mm, z=d2=d0+Δd=81mm and z=d2=d0+2Δd=82mm, respectively [see Fig. 1(c)]. The mask is Lx=3mm long (1000 pixels). Each pixel is quantized into 32 levels such that Δh=13.9nm. The 2-D target patterns are “U,” “T,” “A,” and “H” letters separated by a gap of Δd=0.3mm with an initial distance of d0=5mm [see Fig. 1(e)]. 18-μm wide lines are used to draw the patterns. Since they are periodic, each unit cell has dimension of Lx×Ly=180μm×180μm (60×60pixels). The 2-D phase mask has square pixel with Δx=Δy=3μm and unit height of Δh=6.8nm (64 levels). Ridges in Fig. 1(c) and red parts in Fig. 1(e) stand for places to be exposed in positive-tone photoresist. Note that XY and XY coordinates are employed in the mask space and the image space, respectively.

3.

Exposure Results

In Fig. 2, the 1-D diffractive phase mask was designed to project three groups of lines of varying widths and spacings onto three planes positioned at z=80, 81, and 82 mm, respectively. The target images are summarized in Fig. 1(c). Since these patterns have no variations in the Y direction, they could be exposed onto a plane tilted at 45 deg, instead of three exposures at three planes. In this way, it is also possible to record all the intensity patterns along the Z direction. The optimized phase mask topography is plotted in Fig. 2(b). Figure 2(c) shows an optical micrograph of the fabricated mask along with an atomic-force micrograph of the region delimited by the black rectangle. The multiple height levels and the discrete pixels are clearly visible. The simulated light intensity in the XZ plane from z=78.5 to z=83.5mm is shown in Fig. 2(d). At the design planes corresponding to z=80, 81, and 82 mm, the patterns corresponding to 9 lines (period=60μm), 3 lines (period=180μm), and 5 lines (period=120μm), respectively, are clearly visible. The optical efficiency, η in one plane, is defined as the ratio of the energy within the desired pattern to the total energy incident on the mask. The calculated optical efficiencies are denoted in the figure. The samples for lithography were silicon wafers coated with a 1.3-μm-thick photoresist (Shipley 1813) and mounted on a holder that was placed at 45 deg to the optical axis. The illumination power density at the mask plane was 0.635mW/cm2 and the exposure time was 90 s. The sample was developed in 352 developer for 60 s. Optical micrographs of the patterns corresponding to the regions close to the planes at z=80, 81, and 82 mm [rectangular blocks of yellow-broken lines in Fig. 2(d)] are shown in Figs. 1(e)1(g), respectively. Excellent agreement with the simulation results is seen. The linewidths at three z positions (80, 81, and 82 mm) are 34, 100, and 65μm, respectively, which indicate the deviation of +4, +10, and +5μm. These errors, together with the undesired exposures outside the designated line regions, are partially ascribed to overexposure. The simulated light intensity at three positions is plotted as blue lines beside the micrographs in Figs. 2(e)2(g). By applying a proper threshold, it is possible to achieve clean lines with accurate widths and suppressed noises (black lines). Subsequently, numerical analysis will show how fabrication errors affect the exposure results. Additionally, a simulated 3μm line [Fig. 2(h) representing the green box in Fig. 2(d)] was measured roughly 5μm wide by exposure [Fig. 2(i) representing the green box in Fig. 2(e)].

Fig. 2

Lithography on a 45-deg tilted surface. (a) Schematic of the exposure setup. (b) Height profile of the optimized 1-D phase mask. (c) Optical microscope image of one edge of the fabricated phase mask (inset: AFM measurement of a 50μm×50μm region, and Δx=3-μm pixel size is labeled). (d) Simulated intensity distribution in the XZ plane, where Z is the direction of light propagation. Optical efficiencies at three planes are given. (e)–(g) Optical micrographs of the exposed and developed results at three regions enclosed by yellow blocks in (d). Measured linewidths at (e) z=80, (f) z=81 and (g) z=82mm are 34, 100, and 65μm, respectively. Blue lines are simulated intensity distributions at three planes and black lines represent the estimated exposure outcomes by applying a proper threshold to the simulated patterns. (h) and (i) Magnified views of small areas delimited by the green boxes in (d) and (e), respectively. The labeled 3μm line in (h) is experimentally measured 5μm.

JM3_14_2_023507_f002.png

In the next experiment, another mask was designed to project four letters “U,” “T,” “A,” and “H” onto XY planes corresponding to z=5, 5.3, 5.6, and 5.9 mm, respectively, as illustrated in Fig. 3(a). Figure 3(b) gives the topography of the designed mask. An optical micrograph of the fabricated mask along with an atomic-force micrograph of a small region is shown in Fig. 3(c). The multiple height levels of the square pixels are evident. Simulated light intensity distributions in the XY planes at the four planes are plotted in Figs. 3(d)3(g). The corresponding optical efficiencies are also denoted in the figures. The measured optical intensity at the mask plane was 0.734mW/cm2 and the sample was exposed for 52 s. Optical micrographs of the corresponding exposed and developed patterns are shown in Figs. 3(h)3(k). The experimental results agree very well with the simulation predictions. 21, 20, and 19μm widths are obtained for the 18μm lines by measurements. Arrays of the patterned letters are given by microscope images in Figs. 3(l)3(o). The noise present in the exposure results in Figs. 3(d)3(g) are likely due to both overexposure and mask fabrication errors. Figures 3(p) and 3(q) show the exposure patterns predicted by implementing high (critical exposure) and low (overexposure) thresholds to the simulated light intensity distributions in Figs. 3(d)3(g). Compared to Fig. 3(p), Fig. 3(q) clearly includes more noise and approaches the experimental results in Figs. 3(h)3(k) with better accuracy.

Fig. 3

Lithography on multiple planes parallel to the mask. (a) Schematic of the exposure setup. (b) Height profile of the optimized 2-D phase mask. (c) Optical microscope image of one corner of the fabricated periodic phase mask (inset: AFM measurement of a 50μm×50μm region, and 3-μm pixel size is labeled). (d)–(g) Simulated intensity distributions of one period on the XY plane. Optical efficiencies are given. (h)–(k) Optical micrographs of the exposed and developed results of one period. Designed 18μm lines have measured linewidths of 21, 20, and 19μm, respectively. (l)–(o) Optical micrographs of the exposed and developed results of the periodic arrays. (d), (h) and (l) are letter “U” at z=5.0mm. (e), (i) and (m) are letter “T” at z=5.3mm. (f), (j) and (n) are letter “A” at z=5.6mm. (g), (k) and (o) are letter “H” at z=5.9mm. Estimated exposure results by applying high (p) and low (q) thresholds.

JM3_14_2_023507_f003.png

In a third experiment, the sample substrate was placed orthogonal to the diffractive mask as illustrated in Fig. 4(a). For simplicity, the same phase mask as in Fig. 3 was used. The simulated light intensity distribution in the XZ plane is shown in Fig. 4(c) and the optical micrograph of the exposed and developed pattern is shown in Figs. 4(d) and 4(e). The pattern corresponds to the lower part of the four characters, i.e., the fat line at the bottom of “U” (z=5mm), the center line of “T” (z=5.3mm), the legs at the bottoms of “A” and “H” (z=5.6mm and z=5.9mm). A cross-sectional schematic is depicted in Fig. 4(b). Note that several periods (spacing of 180μm) of the design in Fig. 4(b) were fabricated on the mask. This resulted in repeated patterns as indicated in Figs. 4(d) and 4(e). The agreement between the two patterns indicates that the diffractive mask is capable of patterning onto surfaces of extreme obliqueness. In this case, the surface is perpendicular to the diffractive mask. The measured laser intensity at the diffractive mask was 0.748mW/cm2 and the exposure time was 30 min. The exposure time is significantly increased compared to Fig. 3 due to the large angle between the light propagation direction and the surface of the photoresist. Nevertheless, patterns with micron-scale fidelity can be achieved. Simulated features of 81 and 30μm [Fig. 4(c)] are 87 and 35μm wide [Fig. 4(e)] because of overexposure and mask fabrication error.

Fig. 4

Lithography on a surface orthogonal to the mask. (a) Schematic of the exposure setup. (b) Schematic illustrating that the exposure plane (pink) is the XZ cross section of the intensity pattern of the phase mask that generates “U” “T” “A” “H” letters. (c) Simulated intensity distribution on the XZ plane [pink surface in (b)]. Optical micrographs of the exposed and developed results of two (d) and four (e) periods. Simulated 81 and 30μm lines in (c) have measured linewidths of 87 and 35μm in (e), respectively.

JM3_14_2_023507_f004.png

Additionally, it is instructive to design a phase mask for exposure on a highly oblique surface. Figure 5 summarizes the numerical results of an optimized 1-D phase mask that is designed to expose three narrow regions (6μm width) spaced by 0.3 and 0.9 mm in X and Z directions, respectively. It contains 1000 pixels of 3μm width; the mask is 3 mm long. It has a maximum height of 600 nm and unit height of 10 nm (61 levels). The first exposure plane is 30 mm away from the mask [Fig. 5(a)]. The simulated light intensity distribution shows highly efficient focus spots at the three designated positions in the XZ plane [Fig. 5(c)]. By following the white dashed line in Fig. 5(c), light intensity received by the 71.6 deg tilted surface is plotted in Fig. 5(d). After applying a proper threshold (green line), the desired exposure pattern can be produced with excellent accuracy.

Fig. 5

Design of a 1-D phase mask for exposure on highly oblique surface. (a) Schematic of exposure on a surface of 71.6 deg angle. Three exposed regions, represented by small black blocks, are separated by 0.9 mm axially and 0.3 mm laterally. (b) Height profile of the optimized 1-D phase mask. (c) Simulated light intensity distribution along X direction and the direction of propagation (Z). White dashed line represents the 71.6 deg tilted surface. (d) Intensity distribution on the 71.6 deg tilted surface [along white dashed line in (c)]. Green line represents exposure threshold.

JM3_14_2_023507_f005.png

4.

Discussion

4.1.

Fabrication Error

Since the designed mask generates 3-D light field by introducing spatial phase modulation, it is necessary to pattern microstructures as close to the optimized height distribution as possible. Therefore, it is important to understand how fabrication errors of the diffractive phase mask affects its performance. Figure 6 plots the calculated optical efficiencies where Gaussian noise with zero mean (μ) and various standard deviations (δ) are added to the original design. The efficiencies are reduced with increased standard deviations. Both Figs. 6(a) and 6(b) indicate that errors with standard deviation greater than 100 nm (23% of the maximum height 430 nm) lead to meaningless results where noise overwhelms the signal. With δ=40nm (9% of 430 nm), the 1-D and the 2-D masks have average optical efficiencies decreased from 70% and 60% to 54% and 45%. Insets of Fig. 6(b) include the intensity distribution simulations of the 2-D phase mask with applied errors (δ=5, 40, and 100 nm). 5-nm error has trivial effect on the signal-to-noise ratio. However, with 40-nm error the patterns start to lose their accuracy. This also explains the undesired exposures observed in Figs. 2(e)2(g) and Figs. 3(h)3(k), which occurred outside the designated regions [defined in Figs. 1(c) and 1(e)]. Based upon measurements, the height error in our grayscale lithography is about 30 nm. Hence, it is critical to suppress fabrication errors, especially δ<40nm, by accurate calibration, process parameter optimization, and better condition control.

Fig. 6

Optical efficiencies after adding Gaussian noises with zero mean and standard deviations from 0 to 200 nm to the height distributions of the (a) 1-D and (b) 2-D phase masks. They are calculated at the designated exposure planes and take the average (black lines). Insets of (b): simulated light intensity distributions at four planes separated by 0.3 mm with low noise of δ=5nm (left), medium noise of δ=40nm (middle), and high noise of δ=100nm (right).

JM3_14_2_023507_f006.png

4.2.

Resolution

The spatial resolution by the proposed lithography technique is primarily defined by the fabrication resolution of the phase mask. In this paper, we exploited the Heidelberg microPG101 machine with 3-μm mode write-head for grayscale patterning (Δx=Δy=3μm). Theoretically, the attainable resolution by OPL is defined by C.D.=k1λ/NA, in which λ is the illumination wavelength, NA is the numerical aperture of the projection lens, and k1 is a system-related scale coefficient.3133 For a pixelated phase mask NAλ/2Δx. Usually, k1 takes a value of 0.5, which results in a resolution C.D.Δx. In Figs. 2(h) and 2(i), a simulated 3μm line was measured 5μm wide due to overexposure and the limited resolution of the optical microscope. Similarly, the other measured linewidths are within +15% of the nominal values. Therefore, by optimizing exposure condition and minimizing mask fabrication error, is it possible to approach the predicted resolution. Additionally, smaller features can be achieved once an advanced mask generation tool is utilized (down to <1μm resolution).

4.3.

Defocus

Depth-of-focus (DOF) is another issue considered in OPL systems. Generally, a projection lens has a DOF determined by DOF=k2λ/NA2=4k2Δx2/λ, in which k2 is another system-related factor.3234 Assuming k2=0.5, DOF=55μm for a λ=325nm laser with Δx=3μm. Contrary to conventional 2-D lithography, a shorter DOF is desired in micropatterning on oblique and multiplane surfaces, since more pattern changes are expected within a certain distance. This can be realized by reducing the pixel size of the phase mask and using a long-wavelength light source. The optical efficiencies at various defocus planes are plotted in Fig. 7. The efficiencies drop to 60% at ±500 and ±100μm defocus for both 1-D [Fig. 7(a)] and 2-D [Fig. 7(b)] masks. At the 100μm plane [top inset of Fig. 7(a)], the letters “U” and “T” have well-preserved patterns, while the other two exhibit obvious distortions. On the other hand, “A” and “H” look good, while the first two have worsened shapes at the +100μm plane [bottom inset of Fig. 7(b)]. Thus, in exposure experiments, it is crucial to control the gap between the mask and the sample as close to the designed value as possible.

Fig. 7

Optical efficiencies at various defocus locations of the (a) 1-D and (b) 2-D phase masks. They are calculated at different exposure planes and take the average (black lines). Insets of (b): simulated light intensity distributions at four planes separated by 0.3 mm at 100μm defocus (left) and +100μm defocus (right).

JM3_14_2_023507_f007.png

5.

Conclusions

Managing light intensities in 3-D space using broadband diffractive optics allows for a new and efficient technique to pattern microstructures on oblique and multiplane surfaces. Clearly, this technique can be extended to conventional 3-D lithography. Compared to scanning two-photon lithographic techniques, the reported method is based on a single optical exposure and effectively avoids high-power pulsed lasers and slow scanning schemes.34,35 Our technique can be readily adapted for high-throughput manufacturing. The diffractive phase mask allows for a large number of degrees-of-freedom, which permits generation of complex geometries in 3-D space. The technique currently suffers from crosstalk between the patterns as is evident in Figs. 3(d)3(g). This effect can be reduced by the use of smaller fabrication pixels, which will provide many more pixels, and hence more degrees-of-freedom for the optimization algorithm. Shrinking pixel size also helps in improving patterning resolution. Furthermore, our previous work in broadband diffractive optics20 indicates that with a larger number of pixels, the sensitivity of the projected pattern to pixel errors is also minimized. One challenge in the reported method is that the resolution in the Z-axis is limited by the DOF of the diffractive-optical mask. Distances between the exposure planes that are several multiples of this DOF are necessary to effectively separate different patterns. The DOF can be decreased by using smaller pixels and longer wavelengths. In addition, the computer-generated micro-optic device can be faithfully replicated, and thus mass-produced via roll-to-roll nanoimprint.18 The next step is to explore its vast capabilities in 3-D micropatterning.

Acknowledgments

The authors would like to thank Brian Baker at the Utah Nanofabrication Facility for assistance with grayscale lithography process, and Brian van Devener for help with atomic force microscopy (AFM) imaging. We also thank Apratim Majumder for assistance with the exposure setup. Jose Dominguez-Caballero and Ganghun Kim are also acknowledged for their earlier work on the optimization algorithm.

References

1. 

J. D. Joannopoulos, P. R. Villeneuve and S. H. Fan, “Photonic crystals: putting a new twist on light,” Nature, 386 143 –149 (1997). http://dx.doi.org/10.1038/386143a0 NATUAS 0028-0836 Google Scholar

2. 

M. Campbell et al., “Fabrication of photonic crystals for the visible spectrum by holographic lithography,” Nature, 404 53 –56 (2000). NATUAS 0028-0836 Google Scholar

3. 

R. S. Patti, “Three-dimensional integrated circuits and the future of system-on-chip design,” Proc. IEEE, 94 1214 –1224 (2006). http://dx.doi.org/10.1109/JPROC.2006.873612 IEEPAD 0018-9219 Google Scholar

4. 

N. Miki et al., “Multi-stack silicon-direct wafer bonding for 3-D MEMS manufacturing,” Sens. Actuators A, 103 194 –201 (2003). http://dx.doi.org/10.1016/S0924-4247(02)00332-1 SAAPEB 0924-4247 Google Scholar

5. 

E. Verpoorte and N. F. De Rooij, “Microfluidics meets MEMS,” Proc. IEEE, 91 930 –953 (2003). http://dx.doi.org/10.1109/JPROC.2003.813570 IEEPAD 0018-9219 Google Scholar

6. 

S. Egashira and E. Nishiyama, “Stacked microstrip antenna with wide bandwidth and high gain,” IEEE Trans. Antennas Propag., 44 1533 –1534 (1996). http://dx.doi.org/10.1109/8.542079 IETPAK 0018-926X Google Scholar

7. 

R. Want, “An introduction to RFID technology,” IEEE Pervasive Comput., 5 25 –33 (2006). http://dx.doi.org/10.1109/MPRV.2006.2 IPCECF 1536-1268 Google Scholar

8. 

D. Schurig et al., “Metamaterial electromagnetic cloak at microwave frequencies,” Science, 314 977 –980 (2006). http://dx.doi.org/10.1126/science.1133628 SCIEAS 0036-8075 Google Scholar

9. 

H. Chen, C. Chan and P. Sheng, “Transformation optics and metamaterials,” Nat. Mater., 9 387 –396 (2010). http://dx.doi.org/10.1038/nmat2743 NMAACR 1476-1122 Google Scholar

10. 

M. J. Booth et al., “Adaptive aberration correction in a confocal microscope,” Proc. Natl. Acad. Sci. U. S. A., 99 5788 –5792 (2002). http://dx.doi.org/10.1073/pnas.082544799 PNASA6 0027-8424 Google Scholar

11. 

C. Kim et al., “Integrated wireless interface based on the Utah electrode array,” Biomed. Microdevices, 11 453 –466 (2009). http://dx.doi.org/10.1007/s10544-008-9251-y BMICFC 1387-2176 Google Scholar

12. 

J. John et al., “Microfabrication of 3-D neural probes with combined electrical and chemical interfaces,” J. Micromech. Microeng., 21 105011 (2011). http://dx.doi.org/10.1088/0960-1317/21/10/105011 JMMIEZ 0960-1317 Google Scholar

13. 

A. W. Martinez, S. T. Phillips and G. M. Whitesides, “Three-dimensional microfluidic devices fabricated in layered paper and tape,” Proc. Natl. Acad. Sci. U. S. A., 105 19606 –19611 (2008). http://dx.doi.org/10.1073/pnas.0810903105 PNASA6 0027-8424 Google Scholar

14. 

T. Ito and S. Okazaki, “Pushing the limits of lithography,” Nature, 406 1027 –1031 (2000). http://dx.doi.org/10.1038/35023233 NATUAS 0028-0836 Google Scholar

15. 

J. A. Dominguez-Caballero, S. Takahashi and G. Barbastathis, “Design and sensitivity analysis of Fresnel domain computer generated holograms,” Int. J. Nanomanuf., 6 207 (2010). http://dx.doi.org/10.1504/IJNM.2010.034784 IJNNIR 1746-9392 Google Scholar

16. 

M. Galus et al., “Replication of diffractive-optical arrays via photocurable nanoimprint lithography,” J. Vac. Sci. Technol. B, 24 2960 –2963 (2006). http://dx.doi.org/10.1116/1.2363401 JVTBD9 0734-211X Google Scholar

17. 

K. S. Chen, I. K. Lin and F. H. Ko, “Fabrication of 3-D polymer microstructures using electron beam lithography and nanoimprinting technologies,” J. Micromech. Microeng., 15 1894 –1903 (2005). http://dx.doi.org/10.1088/0960-1317/15/10/015 JMMIEZ 0960-1317 Google Scholar

18. 

L. J. Guo, “Nanoimprint lithography: methods and material requirements,” Adv. Mater., 19 495 –513 (2007). http://dx.doi.org/10.1002/(ISSN)1521-4095 ADVMEW 0935-9648 Google Scholar

19. 

J. G. Goodberlet and B. L. Dunn, “Deep-ultraviolet contact photolithography,” Microelectron. Eng., 53 95 –99 (2000). http://dx.doi.org/10.1016/S0167-9317(00)00272-0 MIENEF 0167-9317 Google Scholar

20. 

G. Kim, J. A. Dominguez-Caballero and R. Menon, “Design and analysis of multi-wavelength diffractive optics,” Opt. Express, 20 2814 –2823 (2012). http://dx.doi.org/10.1364/OE.20.002814 OPEXFF 1094-4087 Google Scholar

21. 

P. Wang et al., “A new class of multi-bandgap high efficiency photovoltaics enabled by broadband diffractive optics,” Prog. Photovoltaic: Res. Appl., (2014). http://dx.doi.org/10.1002/pip.2516 PPHOED 1062-7995 Google Scholar

22. 

P. Wang et al., “Hyper-spectral imaging in scanning-confocal-fluorescence microscopy using a novel broadband diffractive optic,” Opt. Commun., 324 73 –80 (2014). http://dx.doi.org/10.1016/j.optcom.2014.03.044 OPCOB8 0030-4018 Google Scholar

23. 

G. Kim and R. Menon, “An ultra-small 3-D computational microscope,” Appl. Phys. Lett., 105 061114 (2014). http://dx.doi.org/10.1063/1.4892881 APPLAB 0003-6951 Google Scholar

24. 

P. Wang and R. Menon, “Optimization of periodic nanostructures for enhanced light-trapping in ultra-thin photovoltaics,” Opt. Express, 21 6274 –6285 (2013). http://dx.doi.org/10.1364/OE.21.006274 OPEXFF 1094-4087 Google Scholar

25. 

P. Wang and R. Menon, “Optimization of generalized dielectric nanostructures for enhanced light trapping in thin-film photovoltaics via boosting the local density of optical states,” Opt. Express, 22 A99 –A110 (2014). http://dx.doi.org/10.1364/OE.22.000A99 OPEXFF 1094-4087 Google Scholar

26. 

P. Wang and R. Menon, “Computational spectroscopy based on a broadband diffractive optic,” Opt. Express, 22 14575 –14587 (2014). http://dx.doi.org/10.1364/OE.22.014575 OPEXFF 1094-4087 Google Scholar

27. 

M. A. Seldowitz, J. P. Allebach and D. W. Sweeney, “Synthesis of digital holograms by direct binary search,” Appl. Opt., 26 (14), 2788 –2798 (1987). http://dx.doi.org/10.1364/AO.26.002788 APOPAI 0003-6935 Google Scholar

28. 

T. R. M. Sales and D. H. Raguin, “Multiwavelength operation with thin diffractive elements,” Appl. Opt., 38 (14), 3012 –3018 (1999). http://dx.doi.org/10.1364/AO.38.003012 APOPAI 0003-6935 Google Scholar

29. 

B. Kress and P. Meyrueis, Digital Diffractive Optics: An Introduction to Planar Diffractive Optics and Related Technology, John Wiley, New York (2000). Google Scholar

30. 

K. Totsu et al., “Fabrication of three-dimensional microstructure using maskless gray-scale lithography,” Sens. Actuators A, 130 387 –392 (2006). SAAPEB 0924-4247 Google Scholar

31. 

S. Okazaki, “Resolution limits of optical lithography,” J. Vac. Sci. Technol. B, 9 2829 –2833 (1991). http://dx.doi.org/10.1116/1.585650 JVTBD9 0734-211X Google Scholar

32. 

H. Wang et al., “Spot size and depth of focus in optical data storage system,” Opt. Eng., 46 (6), 065201 (2007). http://dx.doi.org/10.1117/1.2744063 OPEGAR 0091-3286 Google Scholar

33. 

B. J. Lin, “The k3 coefficient in nonparaxial λ/NA scaling equations for resolution, depth of focus, and immersion lithography,” J. Micro/Nanolith. MEMS MOEMS, 1 (1), 7 –12 (2002). http://dx.doi.org/10.1117/1.1445798 JMMMGF 1932-5134 Google Scholar

34. 

S. Maruo, O. Nakamura and S. Kawata, “Three-dimensional microfabrication with two-photon-absorbed photopolymerization,” Opt. Lett., 22 132 –134 (1997). http://dx.doi.org/10.1364/OL.22.000132 OPLEDP 0146-9592 Google Scholar

35. 

W. Zhou et al., “An efficient two-photon-generated photoacid applied to positive-tone 3-D microfabrication,” Science, 296 1106 –1109 (2002). http://dx.doi.org/10.1126/science.296.5570.1106 SCIEAS 0036-8075 Google Scholar

Biography

Peng Wang is currently a PhD candidate in the Electrical and Computer Engineering Department at the University of Utah. He received his bachelor’s degree in optical engineering from Zhejiang University, China, in 2011. His research is focused on modeling, fabrication, and characterization of diffractive micro-optics and its applications in photovoltaics, spectroscopy, imaging, and lithography. He is also interested in nanophotonics and computational optics.

Rajesh Menon received his SM and PhD degrees from MITl. His research is focused on nanofabrication, computation, and photonics to innovate in super-resolution lithography, metamaterials, diffractive optics, integrated photonics, photovoltaics, and computational optics. His research has yielded over 70 publications, 30 patents, and 2 companies. Among his honors are the NASA Early Stage-Innovations Award, the NSF CAREER Award, and the International Commission for Optics Prize. Currently, he is the director of the Laboratory for Optical Nanotechnologies at the University of Utah.

CC BY: © The Authors. Published by SPIE under a Creative Commons Attribution 4.0 Unported License. Distribution or reproduction of this work in whole or in part requires full attribution of the original publication, including its DOI.
Peng Wang and Rajesh Menon "Optical microlithography on oblique and multiplane surfaces using diffractive phase masks," Journal of Micro/Nanolithography, MEMS, and MOEMS 14(2), 023507 (22 May 2015). https://doi.org/10.1117/1.JMM.14.2.023507
Published: 22 May 2015
Lens.org Logo
CITATIONS
Cited by 14 scholarly publications.
Advertisement
Advertisement
KEYWORDS
Photomasks

Optical lithography

Photomicroscopy

Computer simulations

Lithography

Geometrical optics

Photoresist materials

Back to Top