Vous êtes sur la page 1sur 14

514

IEEE JOURNAL OF SELECTED TOPICS IN QUANTUM ELECTRONICS, VOL. 10, NO. 3, MAY/JUNE 2004

MEMS Optical Scanners for Microscopes


Hiroshi Miyajima, Kenzi Murakami, and Masahiro Katashiro
AbstractMicroelectromechanical systems (MEMS) optical scanners have been around for more than two decades. Various applications have been presented, but few of them have advanced to the commercial level to date due to the difficulties of combination of optics and MEMS devices. This paper presents our activities of investigating MEMS scanner applications related to microscopic imaging. First, we started with developing a millimeter-sized one-dimensional scanner for commercially available laser scanning microscope. This microscope with the MEMS scanner is now commercially available. In order to take advantage of the miniaturization capability of MEMS, the next step was to miniaturize the whole optics together with the scanners. Miniaturized confocal microscope with a two-dimensional (2-D) scanner has been developed, and its feasibility and key issues are clarified. Additionally, an alternative 2-D scanner capable of scanning wide angle has been prototyped and fundamental characterization showed a promising result. Throughout the study, feasibility of MEMS optical scanners for microscopes has been demonstrated. Index TermsElectromagnetic, electrostatic, laser scanning microscope (LSM), optical scanner.

I. INTRODUCTION

ICROELECTROMECHANICAL systems (MEMS) optical scanner was first presented by Petersen [1], even before the word MEMS was created. Since then, various scanners have been presented. Possible applications are, for example, bar code scanners [2], imaging devices [3], and displays [4]. Similar devices can also be used for optical switching as in dense wavelength-division-multiplexing (DWDM) optical networking [5], and well-known digital micromirror device (DMD) [6] is considered for a switching device, i.e., switching each pixel as required for constructing the appropriate image. MEMS optical scanners have an advantage over other MEMS actuators due to the fact that it is unnecessary have extra force and its transmission to outside world. However, it has to be compatible with optics. A lot of conventional optics has their beam diameter of millimeter size in order to minimize diffraction and, therefore, either the MEMS optical scanner has to be relatively large or the optics have to be modified to accommodate the small beam size. In this paper, our activity in this field is presented. We have chosen the confocal laser scanning microscope (LSM) as one of the potential applications, because its high-resolution three-dimensional (3-D) imaging capability has gained attention for both industrial and biological use, and optical scanners are its key components. At first, we addressed this issue by accommodating MEMS to optics. We have developed a one-dimensional
Manuscript received October 31, 2003; revised January 20, 2004. The authors are with the Olympus Corporation, Tokyo 192-8512, Japan (e-mail: h_miyajima@ot.olympus.co.jp). Digital Object Identifier 10.1109/JSTQE.2004.828487

(1-D) electromagnetic optical scanner for a commercially available LSM, Olympus OLS1100, in order to figure out the feasibility of MEMS scanner for microscopic imaging [7]. Since the scanner has to be compatible with the conventional one used in the product, the mirror size has to be millimeter order. Once the feasibility is verified, the next step was to miniaturize the whole microscope. In this way, the advantage of the MEMS scanner is more appropriately utilized. We have miniaturized the whole microscope optics so that it can be used for endoscopic imaging [8]. A submillimeter-sized two-dimensional (2-D) electrostatic gimbal scanner together with precise assembly technology has been developed [9] to realize the confocal microscope as thin as approximately 3 mm. High-resolution images are obtained using the prototype, demonstrating the feasibility of the miniature LSM. Another submillimeter-sized 2-D electromagnetic gimbal scanner has been prototyped to obtain much wider scan angle than the aforementioned one [10]. Polyimide hinge and electromagnetic actuation made it possible to have an optical p-p scan angle (scan angle is used as optical p-p scan angle in the rest of this paper unless otherwise notified) of more than 100 . As described in this paper, scanner specification varies depending on its application. Since we have developed several types of scanners, choosing optimal scanner configuration for each application, using, for example, different principle and material has become possible. II. MILLIMETER-SIZED SCANNER FOR COMMERCIAL LSM A. Background LSM has been industrially used for inspecting electronic components and various materials. Olympus OLS1100 (Fig. 1), nm) as a standard optical equipped with Ar laser ( source, is featured with high-resolution inspection and highly repeatable measurement due to a specially designed optics and rigid chassis. In a confocal optics [11] where the specimen is illuminated by a small optical spot (Fig. 2), it is necessary to scan it two-dimensionally to obtain an image from observation area. Two optical scanners are used in OLS1100; one for horizontal and the other is for vertical scanning. A resonant scanner is used for horizontal scanning, because it is necessary to realize both fast scanning with a frequency of 4 kHz and sufficient scan angle. Since OLS1100 is used not only for inspection but also for measurement, required specifications of the scanning speed, accuracy, and durability are very high, and thus, a commercially available resonant scanner with high performance had been used. We started developing a millimeter-sized electromagnetic MEMS optical scanner using polyimide hinges [12], [13] as a

1077-260X/04$20.00 2004 IEEE

MIYAJIMA et al.: MEMS OPTICAL SCANNERS FOR MICROSCOPES

515

TABLE I MAJOR SPECIFICATIONS OF LSM HORIZONTAL SCANNER

Fig. 1.

Olympus confocal laser scanning microscope, OLS1100.

Fig. 3.

Schematic of the MEMS scanner.

Fig. 2. Schematic of confocal optics in OLS1100.

LSM horizontal scanner. Although it showed excellent shock resistance and durability, required resonant frequency and scan angle were not satisfied due to its compliance and low factor. On the other hand, a single crystal silicon hinge was used and successfully satisfied the specifications. The developed MEMS scanner has been commercialized as a part of Olympus OLS1200 (renamed from OLS1100 since the MEMS scanner installation). B. Major Specifications This particular MEMS scanner is designed so that it can be installed into our existing LSM product. Table I shows major specifications of the horizontal scanner. Mirror size is large (millimeter order) for MEMS because of the millimeter order optical

beam size, common in conventional microscope optics. High resonant frequency is needed for fast image acquisition. The scan angle should be changeable between 2.1 and 16 to cover 16 X optical zoom (featured in the LSM) with a margin. Mirror should be stiff enough to maintain its flatness against residual stress of the films deposited on the mirror plate and dynamic deformation while the scanner is in operation. Since LSM is used for precise measurement, scan angle should be controlled accurately and maintained consistently over long time. Considering this, there had been no suitable MEMS scanner for the LSM. Asada et al. has presented a millimeter-sized MEMS scanner [14], but its thin compliant mirror is unacceptable for the LSM specifications of mirror flatness. Most of other MEMS scanners are too small for the LSM and capability of closed-loop control is unclear. Furthermore, reliability of MEMS scanner has not been discussed until just recently [15][17]. C. Principle, Structure, and Design Fig. 3 shows the schematic of our new scanner and Fig. 4 shows the operation principle [18]. It is necessary to obtain high resonant frequency, i.e., high hinge stiffness and sufficient scan angle, inversely proportional to the hinge stiffness, at the same time. This tradeoff is solved by utilizing high factor of single crystal silicon hinge. Mirror flatness should be maintained and, thus, the whole wafer thickness (300 m) is used as a moving plate where a reflective Al film as a mirror is formed. Stiffening the moving plate leads to increased moment of inertia of the

516

IEEE JOURNAL OF SELECTED TOPICS IN QUANTUM ELECTRONICS, VOL. 10, NO. 3, MAY/JUNE 2004

Fig. 6. Cross sectional process flow at the dashdotted line A-A in Fig. 3. Fig. 4. Principle of the scanner operation.

Fig. 5. Cross sectional view of the magnets and the moving plate.

mirror plate. 7- m-thick Cu-electroplated driving coil and improved magnetic circuit to obtain magnetic flux density of approximately 0.6 T at the driving coil, almost twice as much as the previous one [13], contribute to obtain satisfactory torque. Output from the sensing coil, where magnetic flux density was increased to 0.5 T, was also increased by the improved magnetic circuit. Fig. 5 shows the cross-sectional view of the magnets and the moving plate after the magnetic circuit improvement. The coil is located lower than the previous one, between the magnets where the magnetic flux density much higher than that of the previous one in which the coils were located above the top surface of the magnets. At the same time, the mirror is located sufficiently close to the top surface of the magnets in order not to have the reflected beam been occluded by the magnets. One of the key issues for this scanner development was to guarantee its reliability. It was questioned whether the silicon hinge could survive the required operational load. As a result of a relatively large, thick mirror plate to be resonated at kilohertz-order frequency, the hinge has to be considerably stiff. Hinge dimensions were carefully determined to realize the resonant frequency, scan angle, and reliability. Stress generated in the hinge at the maximum scan angle has to be low enough to maintain durability. In order to reduce stress without changing the stiffness, hinge length should be increased, and either hinge width or thickness should also be increased. An silicon-on-insulator (SOI) wafer with a 100- m-thick device layer is used in this scanner and the hinge thickness equals to that of the device layer (see Fig. 6). The thickness is determined by the capability of the hinge etching process to have smooth etched

sidewall. Since the hinge thickness is determined, widening the hinge is necessary, but it can lead to nonlinear hinge characteristics [19]. In finalizing the hinge dimensions, starting with short hinge length and adjusting the hinge width to maintain the resonant frequency, both static and dynamic hinge characteristics are tested. When the hinge is short, it is broken before being deflected to required angle. Finally, 2.5-mm-long hinge is found to be sufficient for satisfying the specification. The hinge width is approximately 0.85 mm. Since the measurement accuracy of the LSM depends on the scan angle stability, it is necessary to maintain the scan angle by closed-loop control. Controller using the feedback from the sensing coil output for maintaining resonance and amplitude was also developed with this scanner. Resonance is maintained by applying driving signal being synchronized with the sensing coil output, and the driving signal amplitude is determined to keep the sensing coil output amplitude constant. This controller is also capable of varying the scan angle, inevitable for realizing a zoom function of the LSM. Actually, required magnetic flux density is determined to obtain sufficient sensing coil output when the scan angle is minimal (i.e., maximum zoom), and, thus, the size and the design of the magnetic circuit are also determined. This is the main reason why relatively large magnets are used in this scanner (see Fig. 8). If the scanner operation had not required variable scan angle, the magnet size could have been much smaller. D. Fabrication Fig. 6 shows a simplified process flow as a cross-sectional view at dash-dotted line A-A in Fig. 3. Using a double-side polished SOI wafer, first SiO is thermally grown and the backside (handle wafer) is patterned to form Si etching mask for making moving plate and fixed frame. On the frontside (device layer), Al is sputtered and patterned to form sensing coil and feedthroughs on the hinge to connect moving coils to stationary bonding pads. Plasma chemical vapor deposited (CVD) SiO is used for isolation and passivation between metal layers. It is also patterned, together with thermal SiO , to form Si etching mask for the moving plate, fixed frame and the hinges. Driving coil fabrication consists of seed layer deposition, thick photoresist mold patterning and subsequent Cu-electroplating, and removing mold and unnecessary seed layer. Fig. 7(a) shows a scanning electron microscopy (SEM) of the driving coil. The moving plate and the fixed frame are made both by bulk anisotropic etching of the handle wafer and Deep-reactive ion etching (DRIE) of the device layer, and

MIYAJIMA et al.: MEMS OPTICAL SCANNERS FOR MICROSCOPES

517

Fig. 7. (a) SEM of the Cu-electroplated coil. (b) SEM of the hinge and moving plate, showing the wet-etched and dry-etched structure. (c) Top view of the assembled scanner chip.

Fig. 9. Fig. 8. Perspective view of the scanner.

Resonant frequency distribution.

the torsion bar hinges are made only from the device layer by DRIE. Fig. 7(b) shows an SEM of the hinge and the edge of the moving plate. Finally, Al reflective mirror and a protective SiO layer are deposited on the backside. After the wafer processing, a separated die is assembled onto a metal base with adhesive. Electrical connections are made by wire bonding between electrode pads of the scanner chip and a printed circuit board (PCB) which is also glued onto the base. Fig. 7(c) shows a top view of the die at this step after encapsu-

lating the bonding wires with resin. Note that the die is attached to the base on its mirror side although the schematic drawing (Fig. 3) shows that the adhesion is made on its coil side. Electrical lead wires are soldered to the PCB and, finally, the base and the magnetic circuit consisting of a magnetic yoke and a pair of permanent magnets are assembled to complete the scanner. Fig. 8 shows a perspective view of the assembled scanner. E. Characterization and Discussion Fig. 9 shows the resonant frequency distribution of 37 scanner samples fabricated from two wafers. These are

518

IEEE JOURNAL OF SELECTED TOPICS IN QUANTUM ELECTRONICS, VOL. 10, NO. 3, MAY/JUNE 2004

TABLE II CHARACTERIZATION RESULTS (AVERAGE OF 37 SAMPLES)

Fig. 10. angle.

Driving current, sensing output, and Q factor dependence upon scan

Fig. 11.

Scan angle fluctuation from the start of operation.

Fig. 12. Resonant frequency change during the life test.

measured at a scan angle of 16 . This result includes the following: variation between wafers, process variation in one wafer and between wafers, and resonant frequency change during assembling process. Resulting resonant frequency is densely distributed, showing the excellent reproducibility of the fabrication. Fig. 10 shows the averaged scanner characteristics with two different (Max. and Min. in the specification) scan angles. The driving current and sensing output are shown at vertical axis as their normalized values using those at 2.1 . Sensing output is substantially proportional to the scan angle, showing that this sensing principle is appropriate. Also shown in Fig. 10 is that factor reduces when the scan angle increases. As a result, driving current is not proportional to the scan angle (extrapolated dash-dotted line does not cross the origin of the graph). Approximately twice as much driving current is necessary at the scan angle of 16 than that obtained by assuming their relationship proportional. Since this scanner has relatively high factor, a slight change of the resonant frequency would result in drastic reduction of the scan angle. Also due to the factor variation depending on various conditions, it is necessary to apply closed loop control for controlling the scan angle. Fig. 11 shows the scan angle fluctuation from the start of its operation at a closed-loop-controlled constant scan angle. After approximately 15 min from the start, its fluctuation is within 0.02% during the rest of the time. This 15 min, as the time needed for stabilization, is much less than that with the currently used resonant scanner. It is an important improvement for the

Fig. 13. LSM image (pitch = 10 (b) With MEMS scanner.

m).

(a) With conventional scanner.

LSM characteristics since the users can expect shorter warm-up time for precise measurement. Table II shows averaged characterization results. Combining all the results, required specifications are satisfied. Since it is of utmost importance for commercial products to be durable enough, it has been tested experimentally. Although single crystal silicon has been said to have no fatigue, there have been few data about this issue. We have made life test equipment which allows 24 scanners to be tested, and the data are automatically stored using personal computer. Resonant frequency change, automatically taken with the test apparatus, is mostly less than 0.03% during the 10 000 h test as shown in Fig. 12, corresponding to over 140 billions of cycles. The driving coil resistance change is approximately 1.1% increase with respect to the initial value over the 10 000 h period, showing that the possible

MIYAJIMA et al.: MEMS OPTICAL SCANNERS FOR MICROSCOPES

519

Fig. 14.

A 3-D LSM image of a Cu-electroplated structure.

change due to the stress on across-the-hinge conductors as well as environmental effect is negligible (change of less than 5% is acceptable) [7]. Twenty-four samples have been tested (six are shown in Fig. 12), and no sample has been failed yet except for four samples (one is included in the six in Fig. 12) broken by being mishandled during periodic measurement. Shock resistance is not very important for this application, because it is unlikely for the LSM to suffer from external shock in its normal operation. The requirement is determined mainly to sustain the shock during the transportation, and this scanner has passed the required test. Those results prove the sufficient durability for commercialization. F. Characterization Using the LSM Product Finally, the scanner is installed onto the LSM product and images of a line-and-space test pattern are obtained. Fig. 13 shows the LSM images obtained using the conventional scanner [Fig. 13(a)] and our new MEMS one [Fig. 13(b)]. The pitch of the test pattern is 10 m. No significant difference is seen between the two, demonstrating that our new scanner is at least equivalent to the conventional scanner. Also noteworthy is the significant reduction of audible noise. A preliminary measurement result shows that the new MEMS scanner had a noise level of 67 dB, whereas the current scanner had that of 95 dB. After being installed in the LSM, the sound generated by the new scanner is almost inaudible even when the zoom is 1 X (i.e., with maximum scan angle).

As a demonstration of the LSM possible application, Fig. 14 shows a 3-D image of a Cu-electroplated structure similar to the driving coil shown in Fig. 7(a). This image is obtained by capturing number of confocal images while changing the focus by moving the objective, and by reconstructing then into 3-D image. 100 X objective lens and 1 X optical zoom are inspection condition. Not only the lateral dimensions but also the vertical dimension can be obtained, demonstrating the capability of the LSM for MEMS structure characterization. As described earlier, the MEMS electromagnetic optical scanner for the LSM has been successfully developed. It has not only satisfied all the specifications but also improved some such as quick scanning stabilization and audible noise reduction. It has been commercialized as a part of our LSM product, Olympus OLS 1200. Key factor for commercialization has been reproducible fabrication, stable scan angle due to the closed-loop control, and experimentally verified reliability.

III. MINIATURE CONFOCAL LSM A. Background LSM has several advantages over conventional optical microscope as described previously, but it is relatively large and expensive mainly due to the complicated optics and scanning mechanism. If it is miniaturized sufficiently to be incorporated into endoscope for in vivo inspection in human body, it is expected to be useful for real time diagnosis of cancerous cells

520

IEEE JOURNAL OF SELECTED TOPICS IN QUANTUM ELECTRONICS, VOL. 10, NO. 3, MAY/JUNE 2004

TABLE III MAJOR SPECIFICATIONS OF CONFOCAL MICROSCOPE

Fig. 15. Photograph of forceps coming out from channel of endoscope. Miniature LSM is expected to be in the similar way.

Fig. 16.

Schematic drawing of the miniaturized confocal system.

without biopsy. Possibility of this LSM usage was proved by successful in vitro observation of the cells using a conventional LSM by Inoue et al. [20]. MEMS has been considered to be an enabling technology of miniature LSM, and some papers have been presented [21][24]. Except for the one with scanning lens [24], observation direction is perpendicular to the longitudinal direction of their scanning head, which in most cases is identical to the direction of the optical fiber axis transmitting optical beam from the optical source to the scanning head. However, in the case of endoscopic inspection, it is preferable for the scanning head to observe the same direction as the longitudinal direction of the scanning head. Additionally, the scanning head size has to be sufficiently compact as compared with the endoscope channel, where it should be inserted (concept is shown in Fig. 15). A compact packaging of the scanning head is inevitable. We have developed a miniature LSM capable of observing the longitudinal direction. Our newly developed optics and MEMS 2-D scanner realized the observation in the longitudinal direction. Precise assembly and packaging technique are the key for realizing the compact size to be fit into the endoscope channel for in vivo inspection. B. Configuration Fig. 16 shows a schematic drawing of the confocal system. The scanning head consists of a thin tube including an objective lens, an electrostatic 2-D MEMS scanner, and the tip of an optical fiber. The reasons why electrostatic actuation is used are as follows. First, the mirror size and the scan angle are relatively small, and thus the electrostatic gap can be small. It is expected that reasonable driving voltage can achieve required scan angle. Second, permanent magnet is necessary for electro-

magnetic actuation, and it is difficult to accommodate it into a thin scanning head to be fit into the endoscope channel. The through hole formed in the mirror center is bored to the backside. A glass molded double side aspherical lens is employed as the objective lens. The lens suppresses the aberration caused by off-axial beam. Its numerical aperture (NA) is 0.48 at the wavelength of 400 nm. A fixed mirror is fabricated on the center of the lens. The through hole of the MEMS mirror and the fixed mirror are concentric with the optical beam from the fiber. A laser diode (LD) and a photomultiplier are employed as the optical source and the detector, respectively. Optical sources having wavelength of 400 and 680 nm are prepared, because the performance of the LSM depends on it. The beam from the LD is led to the scanning head by way of a fiber coupler. The divergent ratio of the fiber coupler is 50:50. In the scanning head, the beam from the fiber is divergently directed on the fixed mirror by way of the through hole formed in the MEMS scanner. The beam reflected by the fixed mirror is scanned two-dimensionally by the MEMS gimbal scanner. The scanned beam is focused on the sample by the objective lens. The reflected light goes back along the same way into the fiber and is led to the photomultiplier by way of the fiber coupler. The imaging computer generates a driving signal of the MEMS scanner and displays the confocal image with the detected signal. In this optical design, the core of the fiber in the scanning head is used as a confocal pinhole. In case of this design, the mode field diameter should be determined by the confocal design instead to the wavelength. Consequently, a multimode fiber is used. Table III shows a specification of the LSM.

MIYAJIMA et al.: MEMS OPTICAL SCANNERS FOR MICROSCOPES

521

Fig. 17.

Schematic of MEMS gimbal mirror.

Fig. 19. (a) SEM of stuck prevention tips. (b) Perspective view of gimbal scanner.

Fig. 18. Simplified process flow.

The distance from the objective lens to the sample is defined as the focal length. For fixing the scanning head on the sample, it may be contacted directly to the sample. The field of view depends on the lens NA, the lens size, and the scan angle. In case of in vivo inspection, the frame rate is important, because the body always moves by a pulsation. At least 20/s of the frame rate is needed for in vivo imaging. C. MEMS Gimbal Scanner Fig. 17 shows the schematic of the gimbal scanner. Mirror electrodes are also used as the reflective mirror. A silicon wafer with a high resistivity of 3 k is used as the mirror substrate, and it is contacted with the driving electrodes by several via contacts to generate potential difference in the mirror substrate. Hinges, an isolation layer, and a through hole on the mirror are simultaneously fabricated from a silicon nitride (Si N ) layer. The mirror flatness is maintained by controlling the residual stress in Si N . Since Si N is an insulating material, the feedthroughs can be deposited directly on the hinges. The through hole made in Si N is thin enough to minimize the optical loss caused by mirror tilting. Fig. 18 shows a simplified cross sectional process flow. In Fig. 18(a), a cavity with several stuck prevention tips is formed before fusion bonding of the bottom silicon layer with the top silicon layer. The top silicon is originally a device layer of a
Fig. 20. Results of pull-in analysis and the measurement.

SOI wafer. Consequently, a handle wafer of the SOI wafer is removed after the bonding. In Fig. 18(b), the through hole at the bottom silicon layer is formed. After deposition of the Si N on the top silicon layer, torsion bar hinges, isolation layer, and several via holes are patterned. Al is sputtered for forming the driving electrodes (mirror), via contacts, and feedthroughs. In Fig. 18(c), the mirror and gimbal ring are released by chemical dry etching (CDE). Fig. 19(a) shows a photograph of the stuck prevention tip and, Fig. 19(b) shows a photograph of the MEMS gimbal scanner. Two pits formed on the fixed frame are etched to the bottom silicon layer and are contacted to the ground line. Fig. 20 shows a pull-in analysis result for the gimbal ring and a measured result in the same condition. In the results, the scan angle of 12 is obtained at a driving voltage of 73 V. The

522

IEEE JOURNAL OF SELECTED TOPICS IN QUANTUM ELECTRONICS, VOL. 10, NO. 3, MAY/JUNE 2004

Fig. 21.

Influence of the residual stress on

Q factor and the mirror flatness.

Fig. 23.

Assembly parts in the scanning head.

Fig. 22.

Relationship between reflectance and wavelength. Fig. 24. Cross-sectional drawing of the scanning head.

experimental pull-in voltage is 78 V, a little higher than that obtained by the analysis. The mirror flatness depends on the residual stress of Si N . Therefore, it should be controlled for flattening the mirror surface. Silicon-rich condition is usually used for reducing the residual stress [25]. Furthermore, the residual stress can be also finely tuned by controlling a deposition temperature. It changes from tensile to compressive with the increasing temperature. Fig. 21 shows the influence of the residual stress on the mirror flatness and factor. In order to improve of the mirror flatness, the residual stress should be controlled to zero. On the other hand, the sufficient factor is obtained at the tensile condition, because the hinges are also formed by Si N , and a compressive hinge lowers factor. Therefore, it is necessary to control the residual stress precisely. Typical specifications are the mirror flatness of 80 nm (peak-to-valley), factor of 20. As a result of appropriate stress control, mirror flatness is within the required specification. Scan angle of 12 is achieved at the mirror resonance with a driving voltage of approximately 50 V. D. Improvement of Mirror Reflectance It is preferable not to use fluorescent material for in vivo observation, because potentially hazardous material might be included. On the other hand, except for few successful cases without using fluorescence [20], its use with conventional LSM has been common. Therefore, our LSM needs high signal-tonoise ratio (SNR) as compared with the conventional LSM in order to detect generally low reflection from tissues.

In order to increase the signal, reflectance of the MEMS mirror and the fixed mirror is improved. A dielectric multilayered film consisting of TiO and MgF with their thickness of 82 and 55 nm, respectively, was deposited on the Al mirror. Fig. 22 shows the relationship between the reflectance and the wavelength. In the analysis, index of refraction of high-temperature-deposited film is used whereas TiO and MgF are deposited under low temperature for maintaining the Al reflectance. Therefore, the measured data are different from the analyzed ones. However, the measured reflectance is clearly and nm) improved at the needed wavelength (i.e., as compared with the Al-only mirror. When the reflectance is improved from 85% to 90%, the signal level increases by approximately 25% since the optical beam is reflected totally four times before going back into the fiber. After deposition of the dielectric multilayered film, the flatness of the MEMS mirror is degraded by approximately 10%, but stays within the specification. The driving characteristics of the MEMS scanner are not changed. E. Assembly of the Scanning Head Fig. 23 shows all the parts in the scanning head. The objective lens is fixed in a lens holder, and the fiber is inserted into the capillary of the ferule. The tip of ferule is sharply cut and inserted to the backside of the MEMS scanner (see Fig. 24). The MEMS scanner is fixed on a cylindrical glass base with a D-cut region. Five conductive columns are respectively formed on five bonding pads of the MEMS scanner. A flexible printed circuit (FPC) with five conductive lines is glued on the D-cut region, and the top of the FPC is electrically connected to the bonding

MIYAJIMA et al.: MEMS OPTICAL SCANNERS FOR MICROSCOPES

523

Fig. 25.

Principle of noise reduction using opening formed on fixed mirror.

Fig. 27.

Confocal image of Al line-and-space patterns on glass.

Fig. 26.

Photograph of the scanning head without metal bellows.

pads by bumping. The distance between the objective lens and the mirror is determined by a spacer. Fig. 24 shows a cross-sectional view of the scanning head. The cover frame with the glass plate is capped to the top of the thin tube. The working distance depends on the position of the cover frame, and, therefore, it has to be accurately controlled. The metal bellows is employed to protect the fiber and the electrical wires. The cover frame and the metal bellows are glued and totally sealed by a waterproof adhesive. In addition to enhance mirror reflectance, optical noise is reduced in order to improve the SNR, as shown in Fig. 25. A back reflection from the center of the fixed mirror going back directly to the fiber becomes unacceptable noise. Therefore, an opening on the fixed mirror is formed using a laser cutter in order to have the unnecessary beam pass forward. In the assembly, the fiber, the MEMS scanner, and the objective lens have to be assembled at a positioning error of less than few micrometers for achieving the targeted resolution. Therefore, the assembly is operated under a microscope capable of measuring the target position three-dimensionally. The fiber and the MEMS scanner are positioned by aligning the center of the through hole with the beam from the fiber. The objective lens and the fiber are positioned so that the noise generated in the scanning head is minimized. The noise is detected with the same system as that for detecting the confocal signal (see Fig. 16). During the assembly, the detected signal is considered to be due to the noise, because nothing exists at the focal plane. By monitoring the noise and minimizing it, high-precision assembly within an error of few micrometers has been achieved. F. Characterization and Discussion Fig. 26 shows a photograph of the scanning head without the metal bellows. The backside of the tube is filled with the

Fig. 28. Relationship between the resolutions and the wavelength in theory and measurement.

adhesive instead of the bellows. The length and the diameter of the tube including the cover fame are 8 and 3.3 mm, respectively. The driving frequency coinciding with the resonance is over 3000 Hz in horizontal direction. For achieving the required frame rate, the confocal image is displayed by interlacing, and the vertical scanning frequency is approximately 20 Hz, resulting in 150 150 pixels in a frame. The resonant frequency of the gimbal ring is approximately 840 Hz. Fig. 27 shows a confocal image of the Al patterns on the glass plate. A clear signal is obtained by the sample because Al has relatively high reflectance, and in addition, the difference of the reflectance in the line-and-space pattern is drastic. The pitch of pattern is 10 m. The field of view of 100 75 m and the frame rate of 21 are achieved in this imaging. Fig. 28 shows the lateral and the depth resolution in theory and a measurement with the wavelength of 400 and 680 nm. In the measurement, the lateral resolution is confirmed by using the variable line-and-space patterns formed on the glass and the depth resolution is confirmed by finding the defocus point. In case of the measurement with the wavelength of 400 nm, the lateral resolution of 0.5 m and the depth resolution of 2.9 m are accomplished. The measurement well agrees with the theory. This result shows successful high-precision assembly of the scanning head; otherwise, the resolution would have been much worse.

524

IEEE JOURNAL OF SELECTED TOPICS IN QUANTUM ELECTRONICS, VOL. 10, NO. 3, MAY/JUNE 2004

Fig. 29.

(a) Confocal image of tissues. (b) Photograph of the dyed tissues.

Fig. 29(a) shows the confocal image of pig tissues obtained using our miniature LSM, and Fig. 29(b) shows the photograph of dyed tissues obtained using the conventional microscope. The image of our miniature LSM is vague and has low contrast compared with the one of the conventional microscope, because the SNR is still unsatisfactory mainly due to the back reflection from the fixed mirror even after the noise reduction. However, the tissues are slightly imaged in Fig. 29(a), showing the feasibility of the miniature LSM for medical use if the optics is further improved in the future. In summary, both the feasibility and the key issue of the system have been clarified. IV. 2-D WIDE ANGLE OPTICAL SCANNER A. Background The MEMS scanners described earlier have their mechanical scan angle limited as small as few degrees. There is always a demand of wider scan angle to realize larger field of view. Although a wide-angle scanner cannot be easily combined with confocal optics, other applications including either image acquisition or image display can be considered. For these applications, charge-coupled device (CCD) or complementary metal-oxide-semiconductor (CMOS) image sensors and spatial light modulation devices like liquid crystal display (LCD) or DMD with many pixels are needed, but it is not easy to obtain high yield in such devices. If a scanner with a wide scan angle exists, it can be used as an alternative device for the similar purpose and can realize a compact scanning head. The scan angle is usually determined by the maximum stress or strain of the hinge material. There are also applications of the scanner with relatively slow scanning frequency. However, a drawback of Si or Si N hinges is that it is difficult to make them compliant due to their high stiffness. MEMS application becomes diverse and sophisticated if we can flexibly handle compliant materials other than silicon in fabrication processes. For example, a compliant material like polyimide is expected to bring a wider scan angle. Since current MEMS fabrication processes are mainly based on silicon, we need a new process development to introduce polyimide into MEMS device fabrication. In addition, achievement of higher specifications requires more complex structures to MEMS. We have developed a specific SOI wafer that has an embedded pattern of SiO . Based on our previous studies about polyimide-based hinges [12], [13], [19], the SOI wafer realizes an optical scanner with polyimide-based hinges and a multithickness structure of silicon, allowing the thin mirror to reduce the inertia. The optical scanner results in a higher resonant frequency and a wider scanning angle simultaneously. B. Specific SOI Wafer Fabrication An SOI wafer with a 10- m-thick device layer is thermally oxidized to 400 nm thick and patterned. The patterned side is fusion-bonded with another 300- m-thick conventional silicon wafer at 1000 C. Then, the whole handle wafer of the SOI is removed by wet etching. The newly created SOI wafer has 10- m-thick device layer, 300- m-thick handle wafer, and embedded patterns of SiO at the bonding boundary. In the fabrication process for the optical scanner mentioned later, the embedded pattern of SiO is used as etching mask of the device layer after making a cavity in the handle wafer of the SOI wafer (Fig. 30). To realize multithickness structure of silicon, we need this embedded mask. In particular, when patterns less than several micrometers are required on the device layer, it would be one and only solution [26]. Although the embedded pattern has a 400-nm step structure of SiO , there are few failures of bonding over the whole wafer [27]. In addition, no pattern deformation takes place at the bonded region. Fig. 31 shows the result of transmission electron microscopy (TEM) investigation on the bonding boundary. Islands of SiO are observed at silicon surface in the middle of two SiO steps and in the vicinity of the steps [Fig. 31(b), (c)]. There seem to be several amorphous regions within the periodic structure. We consider that these attribute to SiO . However,
Fig. 30. Schematic drawing of layer structure including polyimide film on active layer surface. (a) Layer structure and (b) cross section when wet etching from handle wafer surface are presented.

MIYAJIMA et al.: MEMS OPTICAL SCANNERS FOR MICROSCOPES

525

Fig. 31. Cross-sectional view of a bonded area. (a) Schematic drawing. (b)-(c) Si-Si bonded area. (d)-(e) Unbonded area. Polymer for TEM sample preparation fills unbonded space.

those SiO islands, as a result, do not obstruct wet etching from backside at the bonding boundary during fabrication processes of optical scanners. We speculate it is because the islands are very small and exist at just a low concentration. No amorphous region in the unbonded area is found [Fig. 31(d), (e)]. The brighter layer in the TEM photos is considered polymer for TEM sample preparation. When wet etching reaches the boundary, no decrease in etching speed is seen, indicating that no SiO residue exists around the surface of the unbonded silicon area. C. Optical Scanner
Fig. 32. SEM photograph of the 2-D optical scanner from mirror side.

By using this specific SOI wafer, we have prototyped an optical scanner. Fig. 32 shows a SEM viewed from its mirror side. Polyimide hinges support an inner moving plate and an outer moving plate. Each plate is 10- m thick, and the silicon frame to support the moving plates is much thicker. Coils for electromagnetic actuation are formed on the other side of the mirror.

The inner mirror is oscillated at its resonant frequency of 2.6 kHz, and the outer plate is driven at nonresonance at a frequency of 20 Hz. Scan angle of 100 is achieved and the relationships between driving current and scan angle are approximately linear (Fig. 33).

526

IEEE JOURNAL OF SELECTED TOPICS IN QUANTUM ELECTRONICS, VOL. 10, NO. 3, MAY/JUNE 2004

Fig. 33. Measured optical scan angle of the 2-D scanner as a function of driving current.

Polyimide-based hinges and thin-moving inertia have lead to these characteristics. The specific SOI we developed has greatly contributed to realize this structure. V. CONCLUSION Examples of MEMS scanner application to microscope are presented. Each has different specification and, therefore, different types of scanners are proposed. Optimal solution for each application has been provided by utilizing various materials and processing. Optimal combination of MEMS devices and optics is also found to be inevitable. This technology is applicable not only for microscopes but also for other optical equipments. ACKNOWLEDGMENT The author greatly thanks the colleagues of MEMS Technology Division, Olympus Corporation, for their contributions to the works described in this paper. REFERENCES
[1] K. E. Petersen, Silicon torsional scanning mirror, IBM J. Res. Develop., vol. 24, no. 5, pp. 631637, Sept. 1980. [2] M.-H. Kiang, O. Solgaard, R. S. Muller, and K. Y. Lau, Surface-micromachined electrostatic-comb driving scanning micromirrors for barcode scanners, in Proc. Microelectromechanical Systems, Feb. 1996, pp. 192197. [3] M. Ikeda, H. Goto, M. Sakata, S. Wakabayashi, K. Imanaka, M. Takeuchi, and T. Yada, Two dimensional silicon micromachined optical scanner integrated with photo detector and piezoresistor, in Tech. Dig. Transducers95 and Eurosensors IX, June 1995, pp. 293296. [4] M. O. Freeman, Miniature high-fidelity displays using a biaxial MEMS scanning mirror, in Proc. SPIE, vol. 4985, Jan. 2003, pp. 5662. [5] D. T. Neilson, V. A. Aksyuk, S. Arney, N. R. Basavanhally, K. S. Bhalla, D. J. Bishop, B. A. Boie, C. A. Bolle, J. V. Gates, A. M. Gottlieb, J. P. Hickly, N. A. Jackman, P. R. Kolodner, S. K. Korotky, B. Mikkelsen, F. Pardo, G. Raybon, R. Ruel, R. E. Scotti, T. W. Van Blarcum, L. Zhang, and C. R. Giles, Fully provisioned 112 112 micro-mechanical optical crossconnect with 35.8 Tb/s demonstrated capacity, in Tech. Dig. OFC2000, vol. PD-12, Mar. 2000, pp. 710. [6] L. J. Hornbeck, Digital light processing and MEMS: Reflecting the digital display needs of the networked society, in Proc. SPIE, vol. 2783, Aug. 1996, pp. 213.

[7] H. Miyajima, N. Asaoka, T. Isokawa, M. Ogata, Y. Aoki, M. Imai, O. Fujimori, M. Katashiro, and K. Matsumoto, A MEMS electromagnetic optical scanner for a commercial confocal laser scanning microscope, J. Microelectromech. Syst., vol. 12, no. 3, pp. 243251, 2003. [8] K. Murakami, A. Murata, T. Suga, H. Kitagawa, Y. Kamiya, M. Kubo, K. Matsumoto, H. Miyajima, and M. Katashiro, A miniature confocal microscope with MEMS gimbal scanner, in Proc. 11th Int. Conf. SolidState Sensors, Actuators, and Microsystems (Transducers03), Boston, MA, June 2003, pp. 587590. [9] K. Murakami, Y. Kamiya, K. Karatsu, H. Miyajima, and M. Katashiro, A MEMS gimbal scanner for a miniature confocal microscope, in Proc. IEEE/LEOS Optical MEMS-2002, Lugano, Switzerland, Aug. 2002, pp. 910. [10] M. Katashiro, M. Arima, K. Tokuda, T. Okamura, K. Matsumoto, and H. Miyajima, Specific SOI wafer with embedded pattern of silicon dioxide and its application for two-dimensional wide angle optical scanner (in Japanese), IEEJ Trans. Sensors Micromachines, vol. 123-E, no. 5, pp. 152157, 2003. [11] T. Wilson, Ed., Confocal Microscopy. New York: Academic Press, 1990. [12] H. Miyajima, N. Asaoka, M. Arima, Y. Minamoto, K. Murakami, K. Tokuda, and K. Matsumoto, An electromagnetic optical scanner with polyimide-based hinges, in Tech. Digest, Transducers99, Sendai, Japan, June 1999, pp. 372375. , A durable, shock-resistant electromagnetic optical scanner with [13] polyimide-based hinges, IEEE J. Microelectromech. Syst., vol. 10, pp. 418424, June 2001. [14] N. Asada, M. Takeuchi, V. Vaganov, N. Belov, S. Hout, and I. Sluchak, Silicon micro-optical scanner, in Tech. Dig. Transducers99, Sendai, Japan, June 1999, pp. 778781. [15] A. S. Dewa, J. W. Orcutt, M. Hudson, D. Krozier, A. Richards, and H. Laor, Development of a silicon two-axis micromirror for an optical cross-connect, in Tech. Dig. Solid-State Sensor and Actuator Workshop, Hilton Head Island, SC, June 2000, pp. 9396. [16] D. W. Wine, M. P. Helsel, L. Jenkins, H. Urey, and T. D. Osborn, Performance of a biaxial MEMS-based scanner for microdisplay applications, in Proc. SPIE, vol. 4178, Aug. 2000, pp. 186196. [17] A. Neukermans, Applications of MEMS switches in optical networks, in Tech. Dig. 4th Int. Topical Workshop on Contemporary Photonic Technologies, Tokyo, Japan, Jan. 1517, 2001, pp. 1922. [18] H. Miyajima, N. Asaoka, T. Isokawa, M. Ogata, Y. Aoki, M. Imai, O. Fujimori, M. Katashiro, and K. Matsumoto, Product development of a MEMS optical scanner for a laser scanning microscope, in Proc. Microelectromechanical Systems, Las Vegas, NV, Jan. 2024, 2002, pp. 552555. [19] H. Miyajima, T. Arikawa, T. Hidaka, K. Tokuda, and K. Matsumoto, A study on nonlinear torsional characteristics of polyimide hinges, in Tech. Dig., Transducers01, Munich, Germany, June 1014, 2001, pp. 14021405. [20] H. Inoue et al., A novel method of virtual histopathology using laserscanning confocal microscopy in-vivo with untreated fresh specimens from the gastrointestinal mucosa, in Proc. Endoscopy, vol. 32, 2000, pp. 439443. [21] D. L. Dickensheets and G. S. Kino, Silicon-micromachined scanning confocal optical microscope, IEEE J. Microelectromech. Syst., vol. 7, pp. 3847, Mar. 1998. [22] U. Hofmann, S. Muehlmann, M. Witt, K. Dorschel, R. Schutz, and B. Wagner, Electrostatically driven micromirrors for a miniaturized confocal laser scanning microscope, in Proc. SPIE, vol. 3878, Sept. 1999, pp. 2938. [23] W. Piyawattanametha, P. R. Patterson, G. D. J. Su, H. Toshiyoshi, and M. C. Wu, A MEMS noninterferometric differential confocal scanning optical microscope, in Tech. Dig. Transducers01, Munich, Germany, June 2001, pp. 590593. [24] S. Kwon and L. P. Lee, Stacked two dimensional micro-lens scanner for micro confocal imaging array, in Proc. Microelectromechanical Systems, Jan. 2002, pp. 483486. [25] L. Kiesewetter, J. Zhang, D. Houdeau, and A. Steckenborn, Determination of youngs moduli of micromechanical thin films using the resonance method, Sens. Actuators A, Phys., vol. 35, no. 2, pp. 153159, 1992. [26] H. Toshiyoshi and H. Fujita, Electrostatic micro torsion mirrors for an optical switch matrix, J. Microelectromech. Syst., vol. 5, no. 4, pp. 231237, 1996. [27] T. Martini, S. Hopfe, S. Mack, and U. Gosele, Wafer bonding across surface steps in the nanometer range, Sens. Actuators, vol. 75, pp. 1723, 1999.

MIYAJIMA et al.: MEMS OPTICAL SCANNERS FOR MICROSCOPES

527

Hiroshi Miyajima received the M.S. and Ph.D. degrees in precision machinery engineering from the University of Tokyo, Tokyo, Japan, in 1987 and 2004, respectively. Since joining Olympus Optical Company, Ltd., Tokyo, in 1987, he has been with the Optical Memory Department. From August 1992 to November 1994, he was a Visiting Research Scholar with Case Western Reserve University, Cleveland, OH. Since returning to Olympus in 1994, he had been engaged in research and development of optical microsensors and MEMS optical scanners. His current research interest is in various MEMS actuators for optical applications. Dr. Miyajima is a member of the Institute of Electrical Engineers of Japan.

Kenzi Murakami received the M.S. degree in mechanical engineering from the University of Electro-Communications, Tokyo, Japan, in 1990. Since Joining Olympus Optical Company, Ltd., Tokyo, in 1991, he has been engaged in research and development of bulk micromachining and optical MEMS devices.

Masahiro Katashiro received the B.S. degree in physics from Kyusyu University, Fukuoka, Japan, in 1983. Since joining Olympus Optical Company, Ltd., Tokyo, Japan, in 1983, he has been with the R&D center. From 1991 to 1998, he was involved in development project of solid-state imaging devices. After completion of the project, he has been engaged in MEMS application research.

Vous aimerez peut-être aussi