Open Access
13 January 2024 Modeling and measurement of a miniature piezoelectric ceramic tube fiber scanner for two-photon endomicroscopy
Conghao Wang, Huilan Liu, Junjie Wang, Qiang Fu, Yanhui Hu, Yuqian Gao, Xinlei Luo, Aimin Wang, Lishuang Feng
Author Affiliations +
Abstract

To enable optical biopsy in clinical applications, it is essential to miniaturize fiber-optic two-photon endomicroscopy (TPEM). This study used theoretical modeling and experimental measurements on a 1-mm-outer-diameter piezoelectric ceramic tube (PZT) fiber scanner for TPEM. After determining resonant modes, the effects of the driving voltage, PZT length, PZT inner diameter, fiber cantilever length, and fiber eccentricity on the fiber’s first- and second-order resonant characteristics were investigated. A 2.7-mm endomicroscopic probe was also integrated, and its two-photon imaging capability was validated using ex-vivo mouse heart and brain tissues. This study’s findings contribute to the advancement of compact nonlinear endomicroscopy.

1.

Introduction

Endoscopic optical coherence tomography,1,2 confocal endomicroscopy,35 and two-photon endomicroscopy (TPEM)6,7 have made significant advances, providing optical biopsy techniques with potential applications in gastrointestinal diseases. Because of its high spatiotemporal resolution, increased penetration depth, and label-free imaging of tissue organs, TPEM has received much attention.810

TPEM probes are generally divided into two scanning schemes: proximal and distal scanning.11,12 The distal scanning scheme incorporates two- or three-dimensional scanning devices into a miniature probe, resulting in superior imaging quality and representing the prevalent implementation strategy. Microelectromechanical (MEMS) mirrors and piezoelectric ceramic tube (PZT) fiber scanners are common scanning actuators in this TPEM scheme.13,14 Unlike MEMS mirrors, which require additional reflecting mirrors or prisms within the probe configuration for light-path deflection, the PZT-driven double-cladding fiber (DCF) scheme reduces integration challenges in miniature endomicroscopic probes. This common-path configuration makes TPEM more practical for clinical endomicroscopic applications.15

PZT fiber scanners’ structure and electrical characteristics significantly impact the integration sizes and imaging parameters (e.g., frame rate and field of view) of miniature TPEM probes, making them critical components. Several research groups have reported the development of compact nonlinear probes with integrated outer diameters (ODs) of 2.8 mm (or even smaller).10,1621 However, some limitations were encountered, primarily due to the ODs of the PZT actuators and imaging objectives, which hampered further miniaturization of endomicroscopic probes.9,13,2224 Moreover, researchers attempting to develop miniature TPEM platforms compatible with clinical instrument channels face engineering and technical challenges as they strive to reduce probe size. A critical step in addressing this challenge is the development of PZT fiber scanners with smaller ODs (1 to 1.5 mm).

Li et al.25 modeled and analyzed a square PZT with four piezoelectric plates with an OD of 2  mm in 2011. They investigated its resonant modes and scanning capabilities, confirming their findings using a confocal imaging platform. Li et al.26 expanded on the potential for raster-scanning imaging in 2012 by combining two resonant modes on the same square PZT fiber scanner. In 2019, Tekpınar et al.27 performed modal analysis on a piezoelectric bimorph fiber scanner, confirming its suitability for Lissajous, raster, and spiral scanning. Wang et al.28 investigated the resonant modes and scanning range of a reverse-fixed 3.2-mm-OD PZT fiber scanner designed for TPEM in 2022. However, the effects of PZT parameters on the scanning characteristics of this scanner were not discussed. Compared to square PZT and piezoelectric bimorphs, the tubular four-quadrant PZT fiber scanner remains the preferred choice. However, theoretical analyses and experimental validations for miniature tubular PZT fiber scanners remain limited.

In this study, we performed modeling and experimental measurements on a 1-mm-OD PZT fiber scanner for the TPEM platform. After extracting the scanner’s resonant modes, we simulated the effects of the driving voltage, PZT length, PZT inner diameter, fiber cantilever length, and fiber eccentricity on the first- and second-order resonant characteristics in the theoretical section. We built a miniature PZT fiber scanner in the experimental section and measured its first- and second-order resonance properties. A 2.7-mm-OD TPEM probe was obtained by integrating this scanner into a lensed fiber-optic TPEM platform. The imaging capability of this probe was validated by two-photon imaging of ex-vivo mouse heart and brain tissues.

2.

Method

2.1.

Modeling and Resonant MODE Analysis of the PZT Fiber Scanner

Figure 1 shows a schematic diagram of a forward-fixed PZT fiber scanner. This scanner comprised three main parts: a PZT, a spacer, and a DCF designed for femtosecond pulses and fluorescence in a common-path configuration. The spacer was physically connected to the PZT, allowing for precise positioning of the fiber and the PZT element. Table 1 presents the detailed geometric and material parameters of the PZT, including an OD (ODPZT) of 1 mm, an inner diameter (IDPZT) of 0.4 mm, a length (LPZT) of 8 mm, and PZT-5A as the selected material. Table 2 indicates the structural and material parameters of the built forward-fixed PZT fiber scanner, with a fiber cantilever length (Lfiber) of 14.3 mm.

Fig. 1

A schematic diagram of the PZT fiber scanner (PZT: piezoelectric ceramic tube; DCF: double-cladding fiber).

JOM_4_1_014001_f001.png

Table 1

Parameters of the PZT mechanical structure.

PZT parameterValue
Outer diameter (ODPZT)1 mm
Inner diameter (IDPZT)0.4 mm
Length (LPZT)8 mm
MaterialPZT-5A

Table 2

Parameters of the PZT fiber scanner mechanical structure.

Scanner parameterValueScanner parameterValue
Fiber outer cladding diameter0.134 mmSpacer outer diameter1 mm
Fiber coating diameter0.35 mmSpacer inner diameter0.35 mm
Fiber outer cladding length (Lfiber)14.3 mmSpacer thickness0.5 mm
Fiber coating length8.0 mmSpacer materialCopper

We built the forward-fixed PZT fiber scanner using commercial software based on the finite element analysis method. The PZT geometric model was established with plane A as the reference plane, and the spacer and DCF geometric models were established with plane B as the reference plane. This device functioned as a model that combined solid mechanics and electrostatic fields. Furthermore, the four quadrants of the PZT’s outer wall were defined by the driving voltage conditions, the inner wall by the ground condition, and the bottom of the PZT by fixed constraints.

The PZT fiber scanner functions by driving the resonant fiber cantilever through the inverse piezoelectric effect. The scanner achieved maximum deflection at specific frequencies, denoted as resonant modes of different orders. Figures 2(a) and 2(b), respectively, illustrate the scanner’s first- and second-order resonant modes, which had corresponding theoretical resonant frequencies of 799.1 and 2981.9 Hz.

Fig. 2

PZT fiber scanner resonant mode analysis. (a) First-order resonant mode. (b) Second-order resonant mode.

JOM_4_1_014001_f002.png

2.2.

Frequency Domain Analysis of the PZT Fiber Scanner

The effect of the driving voltage on the PZT fiber scanner’s first- and second-order resonant characteristics was investigated using frequency domain analysis. Driving signals were applied to a single pair of PZT electrodes (e.g., +X and X) to ensure motion along a single axis for single-axis scanning. As shown in Fig. 3(a), the first-order frequency domain characteristics were obtained at 775 to 825 Hz frequency range (with 1 Hz increments). Meanwhile, as illustrated in Fig. 3(b), the second-order frequency domain characteristics were at 2850 to 3150 Hz frequency range (with 5 Hz increments). Among the six simulated data groups obtained by varying the driving voltage U from ±12.5 to ±75  V with ±12.5  V increments, the first-order mode exhibited maximum scan ranges (D1) of 0.126, 0.253, 0.379, 0.506, 0.633, and 0.76 mm. Correspondingly, the second-order mode exhibited maximum scan ranges (D2) of 0.103, 0.206, 0.311, 0.415, 0.519, and 0.618 mm. To evaluate the PZT fiber scanner’s driving capability, the driving coefficient, denoted by σi, was introduced. The slope of the fitted line, representing the scanner’s scan range as a function of the driving voltage, was used to calculate the driving coefficient. In this case, i signifies different resonant orders. The calculated driving coefficients σ1 and σ2 corresponding to the first- and second-order resonant modes were determined to be 5.07 and 4.15  μm/V, respectively. According to these findings, the first-order driving coefficient σ1 was approximately 1.22 times greater than the second-order driving coefficient σ2 for an identical driving voltage.25

Fig. 3

Frequency domain characteristics of the PZT fiber scanner under different driving voltages. (a) First-order. (b) Second-order.

JOM_4_1_014001_f003.png

Following that, a systematic investigation of the PZT structural parameters, specifically the PZT length and inner diameter, was carried out to evaluate their effects on the resonant characteristics of the PZT fiber scanner. The PZT length was varied between 6 and 10 mm (in 1 mm increments). The frequency domain characteristics corresponding to the first- and second-order resonant modes are depicted in Figs. 4(a) and 4(b), respectively, under a driving voltage of ±50  V. The maximum scan ranges (D1) of the first-order resonant mode in the context of the five datasets obtained from this parameterized analysis were 0.348, 0.424, 0.506, 0.598, and 0.696 mm. Correspondingly, the maximum scan ranges (D2) of the second-order resonant mode were 0.225, 0.303, 0.415, 0.600, and 0.981 mm. The results showed that as the PZT length increased, the scanner’s resonant frequency decreased, tending toward lower frequencies. The resonant frequencies of the scanner’s first-order resonant mode were 803, 801, 800, 798, and 796 Hz. On the other hand, the resonant frequencies of the scanner’s second-order resonant mode were 3010, 2995, 2985, 2950, and 2905 Hz.

Fig. 4

Frequency domain characteristics of the PZT fiber scanner with different PZT lengths. (a) First-order. (b) Second-order.

JOM_4_1_014001_f004.png

Figures 5(a) and 5(b) depict the effect of the PZT inner diameter (IDPZT) on the first- and second-order characteristics of the PZT fiber scanner. Five datasets were generated by increasing the PZT inner diameter from 0.1 mm to 0.3 mm (in 0.05 mm increments). Regarding the first-order resonant mode, the maximum scan ranges (D1) of the first-order resonant mode were 0.371, 0.430, 0.506, 0.610, and 0.756 mm as the IDPZT varied. Correspondingly, the maximum scan ranges (D2) of the second-order resonant mode were 0.308, 0.354, 0.415, 0.495, and 0.611 mm. Notably, increasing IDPZT decreased the thickness of the PZT tube wall, resulting in an expanded scan range for both the first- and second-order resonant modes.

Fig. 5

Frequency domain characteristics of the scanner with different PZT inner diameters. (a) First-order. (b) Second-order.

JOM_4_1_014001_f005.png

Figures 6(a) and 6(b) depict the effect of fiber cantilever length on the PZT fiber scanner’s first- and second-order resonant characteristics was also investigated. The initial length of the fiber cantilever, denoted as Lfiber, was set to 14.3 mm (shown in Table 2). A length coefficient, denoted as η, was introduced to aid in this investigation. The length of the fiber cantilever was expressed within this framework as η×Lfiber. The simulated data set included five datasets obtained by increasing the length coefficient η by 0.05 increments from 0.9 to 1.1. The first-order resonant mode’s maximum scan ranges (D1) were 0.467, 0.490, 0.506, 0.520, and 0.526 mm. Meanwhile, the second-order resonant mode’s maximum scan ranges (D2) were 0.479, 0.441, 0.415, 0.395, and 0.380 mm. In this case, as the length coefficient η varied from 0.9 to 1.1, the scanner’s resonant frequency decreased, approaching lower frequencies. The resonant frequencies of the scanner’s first-order resonant mode were 996, 891, 800, 723, and 654 Hz. On the other hand, the resonant frequencies of the scanner’s second-order resonant mode were 3690, 3315, 2985, 2695, and 2450 Hz.

Fig. 6

Frequency domain characteristics of the scanner with different fiber cantilever lengths. (a) First-order. (b) Second-order.

JOM_4_1_014001_f006.png

2.3.

Fiber Eccentricity Impact Analysis of the PZT Fiber Scanner

The resonant characteristics of the scanner differed along the X and Y axes due to asymmetry in the fabricated PZT fiber scanner.29 Thus, we also simulated the effect of fiber outer cladding eccentricity on resonant characteristics at various orders. Setting fiber cantilever eccentricity (Decfiber) values at 0, 20, and 40  μm, the first-order resonant frequencies for the X- and Y-axes exhibited differences of 0.4, 5.96, and 38.14 Hz, respectively, while the second-order resonant frequencies for the X- and Y-axes exhibited corresponding differences of 1.4, 11.2, and 69.4 Hz, as depicted in Figs. 7(a) and 7(b). Apparently, non-uniformity in fiber fabrication could result in differences in the scanner’s characteristics along the X- and Y-axes, posing challenges for image reconstruction.

Fig. 7

Frequency domain characteristics of the scanner with different fiber eccentricity values. (a) First-order. (b) Second-order.

JOM_4_1_014001_f007.png

3.

Experiment Results

3.1.

Measurement Analysis of the PZT Fiber Scanner

Using a four-quadrant tubular PZT with an OD of 1 mm, a forward-fixed PZT fiber scanner with a cantilever length of 14.3  mm was fabricated. Figure 8 shows a schematic representation of the static measurement platform for this PZT fiber scanner. A voltage amplifier (TD250, PiezoDrive) amplified the driving voltage signals before being applied to the tubular PZT fiber scanner via a signal generator (AFG1022, Tektronix). A CMOS camera (Panda, 4.2 M PCO) was used to record the first- and second-order resonant characteristics of the PZT fiber scanner.

Fig. 8

A schematic diagram of the PZT fiber scanner measurement platform (CL: coupling lens, 354850 B, Lightpath. DCF: double-cladding fiber. Objective: uPlan Apo 10×, Olympus. TL: tube lens, LA1027-A, Thorlabs. CMOS camera: 4.2 M PCO, Panda).

JOM_4_1_014001_f008.png

Figure 9(a) illustrate the first-order resonant characteristics of a PZT fiber scanner driven by a ±50  V voltage over a frequency range of 785 to 835 Hz. Meanwhile, Fig. 9(b) depicts the voltage-scan range curves corresponding to the resonance frequency of 810 Hz. The PZT fiber scanner’s first-order driving coefficients σ1 along the X- and Y-axes were measured at 5.34  μm/V. Similarly, Fig. 9(c) depicts the scanner’s second-order frequency scanning characteristics in the 3000 to 3250 Hz frequency range. Consequently, Fig. 9(d) illustrates the voltage-scan range curve corresponding to the resonance frequency of 3130 Hz. The PZT fiber scanner’s second-order driving coefficients σ2 along the X- and Y-axes were measured to be 3.86 and 3.24  μm/V, respectively. These experiment results agreed with the theoretical simulation results, indicating that at the same driving voltage, the driving coefficients for the first-order resonant mode are greater than those for the second-order resonant mode.

Fig. 9

Resonant characteristic results of the PZT fiber scanner. (a) Scan range versus scanning frequency curve for the first-order resonant mode. (b) Scan range versus driving voltage curve for the first-order resonant mode. (c) Scan range versus scanning frequency curve for the second-order resonant mode. (d) Scan range versus driving voltage curve for the second-order resonant mode.

JOM_4_1_014001_f009.png

3.2.

Two-Photon Endomicroscopy Probe Integration

A common-path fiber-optic TPEM platform was developed, as shown in Fig. 10(a). The double-cladding antiresonant fiber (DC-ARF) was a delivery fiber for femtosecond pulse and fluorescence signals. The 920-nm femtosecond pulse was coupled into the core of the DC-ARF and guided to the probe’s distal end. A photomultiplier tube (PMT) collected and detected two-photon-excited fluorescence signals. Previous reports contained more detailed information on the TPEM hardware platform.30 Figure 10(b) displays the integrated TPEM probe, with an OD of 2.7 mm, a rigid length of 24 mm, and a weight of 0.33 g. Figure 10(c) depicts the schematic of the miniature TPEM probe. The lensed fiber technique aided in realizing objective-lens-free two-photon endomicroscopic imaging,23,31,32 and detailed information on lensed fiber fabrication could be found in one Ref. 30. The lensed fiber had a working distance of 240  μm. The PZT fiber scanner was activated by dual-channel amplitude-modulated sinusoidal and cosine voltage waveforms, allowing for a two-dimensional spiral pattern. The scanner’s increased and decreased amplitude periods each had 512 turns. The scanner’s increased amplitude period was used for imaging, i.e., single-side scanning, at a calculated frame rate of 0.79 frame per second (fps). The PZT fiber scanner and lensed fiber were integrated into a three-segmented probe housing in this design, with a glass window at the end for protection.

Fig. 10

Fiber-optic scanning TPEM. (a) A schematic diagram of TPEM (DM: dichroic mirror, DMLP650R, Thorlabs. CL: coupling lens, 354850 B, Lightpath. DC-ARF: double-cladding antiresonant fiber. Filter: FF01-530/43-25, Semrock. FL: focusing lens, LA1951-A, Thorlabs. PMT: photomultiplier tube, H10770PA-40-SEL, Hamamatsu). (b) Photograph of the integrated probe. (c) A schematic diagram of the integrated probe.

JOM_4_1_014001_f010.png

Figure 11(a) shows the two-photon imaging results of a 380 nm fluorescent microsphere (G400 Fluoro-max, Thermo Scientific). The lateral resolution measurement results in Fig. 11(b) were obtained by extracting the horizontal intensity distribution of the microsphere and performing Gaussian fitting. This probe’s measured lateral resolution was 2.47  μm.

Fig. 11

(a) Two-photon imaging results of the fluorescent microsphere. (b) Lateral resolution measurement results. Blue circle: raw data. Red line: Gaussian curve fitting.

JOM_4_1_014001_f011.png

Figures 12(a) and 12(b) illustrate the results of two-photon endomicroscopic imaging of green-fluorescent-protein-expressing mouse ex-vivo heart and brain tissue, revealing the structural features of myocardial fibers and neuronal somas, respectively. In this case, the peak-to-peak driving voltages for the PZT fiber scanner along the X- and Y-axes were ±63.75 and ±56.25  V, respectively, yielding a field of view of 440  μm. These findings supported the miniature PZT fiber scanner’s suitability for the TPEM platform.

Fig. 12

Ex-vivo two-photon endomicroscopic imaging results. (a) Heart tissue. (b) Brain tissue.

JOM_4_1_014001_f012.png

4.

Conclusion

Theoretical modeling, experimental testing, probe assembly integration, and two-photon imaging verification were performed on a 1-mm-OD PZT fiber scanner in this work. We extracted the PZT fiber scanner’s resonant modes in the theoretical section. We ran simulations to determine the impact of driving voltage, PZT length, PZT inner diameter, fiber cantilever length, and fiber eccentricity on its first- and second-order resonant characteristics. We built a miniature PZT fiber scanner and measured its first- and second-order resonant characteristics in the experimental section. Furthermore, a 2.7-mm-OD lensed fiber-optic TPEM probe with two-photon imaging capability was integrated and validated using ex-vivo mouse heart and brain tissues. Theoretical support for the development of compact fiber-optic endomicroscopes was provided by this work. Prospectively, developing a miniature and high-speed PZT fiber scanner can broaden the application scope of multi-photon endomicroscopy techniques, especially in the dynamic recording of brain neuron activity in freely behaving animals.

Disclosures

The authors declare no conflicts of interest.

Code and Data Availability

The data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

Acknowledgments

This research was supported by the National Key Research and Development Program of China (2020YFB1312802), National Natural Science Foundation of China (61973019, 31830036, and 61975002), and the Academic Excellence Foundation of BUAA for PhD Students. The authors thank the anonymous reviewers for their insightful and professional comments on this work.

References

1. 

G. J. Tearney et al., “In vivo endoscopic optical biopsy with optical coherence tomography,” Science, 276 (5321), 2037 –2039 https://doi.org/10.1126/science.276.5321.2037 SCIEAS 0036-8075 (1997). Google Scholar

2. 

M. J. Gora et al., “Endoscopic optical coherence tomography: technologies and clinical applications [Invited],” Biomed. Opt. Express, 8 (5), 2405 –2444 https://doi.org/10.1364/BOE.8.002405 BOEICL 2156-7085 (2017). Google Scholar

3. 

M. Minsky, “Microscopy apparatus,” (1961). Google Scholar

4. 

J. M. Jabbour et al., “Confocal endomicroscopy: instrumentation and medical applications,” Ann. Biomed. Eng., 40 (2), 378 –397 https://doi.org/10.1007/s10439-011-0426-y ABMECF 0090-6964 (2012). Google Scholar

5. 

H. Li et al., “500 μm field-of-view probe-based confocal microendoscope for large-area visualization in the gastrointestinal tract,” Photonics Res., 9 (9), 1829 –1841 https://doi.org/10.1364/PRJ.431767 (2021). Google Scholar

6. 

W. Denk, J. H. Strickler and W. W. Webb, “Two-photon laser scanning fluorescence microscopy,” Science, 248 (4951), 73 –76 https://doi.org/10.1126/science.2321027 SCIEAS 0036-8075 (1990). Google Scholar

7. 

V. Kučikas et al., “Two-photon endoscopy: state of the art and perspectives,” Mol. Imaging Biol., 25 3 –17 https://doi.org/10.1007/s11307-021-01665-2 (2021). Google Scholar

8. 

W. R. Zipfel, R. M. Williams and W. W. Webb, “Nonlinear magic: multiphoton microscopy in the biosciences,” Nat. Biotechnol., 21 (11), 1369 –1377 https://doi.org/10.1038/nbt899 NABIF9 1087-0156 (2003). Google Scholar

9. 

D. R. Rivera et al., “Compact and flexible raster scanning multiphoton endoscope capable of imaging unstained tissue,” Proc. Natl. Acad. Sci. U. S. A., 108 (43), 17598 –17603 https://doi.org/10.1073/pnas.1114746108 (2011). Google Scholar

10. 

W. Liang et al., “Nonlinear optical endomicroscopy for label-free functional histology in vivo,” Light Sci. Appl., 6 (11), e17082 https://doi.org/10.1038/lsa.2017.82 (2017). Google Scholar

11. 

W. Göbel et al., “Miniaturized two-photon microscope based on a flexible coherent fiber bundle and a gradient-index lens objective,” Opt. Lett., 29 (21), 2521 –2523 https://doi.org/10.1364/OL.29.002521 OPLEDP 0146-9592 (2004). Google Scholar

12. 

C. M. Lee et al., “Scanning fiber endoscopy with highly flexible, 1 mm catheterscopes for wide-field, full-color imaging,” J. Biophotonics, 3 (5–6), 385 –407 https://doi.org/10.1002/jbio.200900087 (2010). Google Scholar

13. 

S. Tang et al., “Design and implementation of fiber-based multiphoton endoscopy with microelectromechanical systems scanning,” Biomed. Opt. Express, 14 (3), 034005 https://doi.org/10.1117/1.3127203 BOEICL 2156-7085 (2010). Google Scholar

14. 

X. Duan et al., “MEMS-based multiphoton endomicroscope for repetitive imaging of mouse colon,” Biomed. Opt. Express, 6 (8), 3074 –3083 https://doi.org/10.1364/BOE.6.003074 BOEICL 2156-7085 (2015). Google Scholar

15. 

D. H. Hong et al., “Clinical feasibility of miniaturized Lissajous scanning confocal laser endomicroscopy for indocyanine green-enhanced brain tumor diagnosis,” Front. Oncol., 12 994054 https://doi.org/10.3389/fonc.2022.994054 FRTOA7 0071-9676 (2023). Google Scholar

16. 

M. T. Myaing, D. J. Macdonald and X. Li, “Fiber-optic scanning two-photon fluorescence endoscope,” Opt. Lett., 31 (8), 1076 –1078 https://doi.org/10.1364/OL.31.001076 OPLEDP 0146-9592 (2006). Google Scholar

17. 

Y. Zhao, H. Nakamura and R. J. Gordon, “Development of a versatile two-photon endoscope for biological imaging,” Biomed. Opt. Express, 1 (4), 1159 –1172 https://doi.org/10.1364/BOE.1.001159 BOEICL 2156-7085 (2010). Google Scholar

18. 

G. Ducourthial et al., “Development of a real-time flexible multiphoton microendoscope for label-free imaging in a live animal,” Sci. Rep., 5 (1), 18303 https://doi.org/10.1038/srep18303 SRCEC3 2045-2322 (2015). Google Scholar

19. 

D. Y. Kim et al., “Lissajous scanning two-photon endomicroscope for in vivo tissue imaging,” Sci. Rep., 9 (1), 3560 https://doi.org/10.1038/s41598-019-38762-w SRCEC3 2045-2322 (2019). Google Scholar

20. 

E. Pshenay-Severin et al., “Multimodal nonlinear endomicroscopic imaging probe using a double-core double-clad fiber and focus-combining micro-optical concept,” Light Sci. Appl., 10 (1), 1 –11 https://doi.org/10.1038/s41377-021-00648-w (2021). Google Scholar

21. 

H. Guan et al., “Multicolor fiber-optic two-photon endomicroscopy for brain imaging,” Opt. Lett., 46 (5), 1093 –1096 https://doi.org/10.1364/OL.412760 OPLEDP 0146-9592 (2021). Google Scholar

22. 

Y. Wang et al., “Four-plate piezoelectric actuator driving a large-diameter special optical fiber for nonlinear optical microendoscopy,” Opt. Express, 24 (17), 19949 –19960 https://doi.org/10.1364/OE.24.019949 OPEXFF 1094-4087 (2016). Google Scholar

23. 

F. Akhoundi et al., “Compact fiber-based multi-photon endoscope working at 1700 nm,” Biomed. Opt. Express, 9 (5), 2326 –2335 https://doi.org/10.1364/BOE.9.002326 BOEICL 2156-7085 (2018). Google Scholar

24. 

A. Lombardini et al., “High-resolution multimodal flexible coherent Raman endoscope,” Light Sci. Appl., 7 (1), 10 https://doi.org/10.1038/s41377-018-0003-3 (2018). Google Scholar

25. 

Z. Li, Z. Yang and L. Fu, “Scanning properties of a resonant fiber-optic piezoelectric scanner,” Rev. Sci. Instrum., 82 (12), 123707 https://doi.org/10.1063/1.3671290 RSINAK 0034-6748 (2011). Google Scholar

26. 

Z. Li and L. Fu, “A resonant fiber-optic piezoelectric scanner achieves a raster pattern by combining two distinct resonances,” Rev. Sci. Instrum., 83 (8), 086102 https://doi.org/10.1063/1.4739770 RSINAK 0034-6748 (2012). Google Scholar

27. 

M. Tekpınar, R. Khayatzadeh and O. Ferhanoğlu, “Multiple-pattern generating piezoelectric fiber scanner toward endoscopic applications,” Opt. Eng., 58 (2), 023101 https://doi.org/10.1117/1.OE.58.2.023101 (2019). Google Scholar

28. 

C. Wang et al., “Modeling and analysis of reverse-fixed piezoelectric tube fiber scanner for two-photon endomicroscopy,” Proc. SPIE, 12560 1256008 https://doi.org/10.1117/12.2652027 PSISDG 0277-786X (2023). Google Scholar

29. 

H. C. Park et al., “High-speed fiber-optic scanning nonlinear endomicroscopy for imaging neuron dynamics in vivo,” Opt. Lett., 45 (13), https://doi.org/10.1364/OL.396023 OPLEDP 0146-9592 (2020). Google Scholar

30. 

C. Wang et al., “Lensed fiber-optic two-photon endomicroscopy for field-of-view enhancement,” Photonics, 10 (3), 342 https://doi.org/10.3390/photonics10030342 (2023). Google Scholar

31. 

H. Bao and M. Gu, “A 0.4-mm-diameter probe for nonlinear optical imaging,” Opt. Express, 17 (12), 10098 –10104 https://doi.org/10.1364/OE.17.010098 OPEXFF 1094-4087 (2009). Google Scholar

32. 

J. B. Kim et al, “Objective-lens-free confocal endomicroscope using Lissajous scanning lensed-fiber,” J. Opt. Microsyst., 1 (3), 034501 https://doi.org/10.1117/1.JOM.1.3.034501 (2021). Google Scholar

Biography

Conghao Wang is a PhD candidate at the School of Instrumentation and Optoelectronic Engineering, Beihang University. His research interests include multi-photon endomicroscopic techniques.

Huilan Liu is an associate professor at the School of Instrumentation and Optoelectronic Engineering, Beihang University. Her research interests include optical sensing and MOEMS techniques.

Junjie Wang is a postdoctoral fellow at the College of Future Technology, Peking University. His research interests include multi-photon microscopic techniques.

Qiang Fu is a senior engineer at Beijing Transcend Vivoscope Biotech Co., Ltd.

Yanhui Hu is a senior engineer at Beijing Transcend Vivoscope Biotech Co., Ltd.

Yuqian Gao is a senior engineer at Beijing Transcend Vivoscope Biotech Co., Ltd.

Xinlei Luo is a senior engineer at Beijing Transcend Vivoscope Biotech Co., Ltd.

Aimin Wang is an associate professor at the School of Electronics, Peking University. His research interests include ultrafast fiber lasers and multi-photon microscopic techniques.

Lishuang Feng is a professor at the School of Instrumentation and Optoelectronic Engineering, Beihang University. Her research interests include integrated optics and MOEMS techniques.

CC BY: © The Authors. Published by SPIE under a Creative Commons Attribution 4.0 International License. Distribution or reproduction of this work in whole or in part requires full attribution of the original publication, including its DOI.
Conghao Wang, Huilan Liu, Junjie Wang, Qiang Fu, Yanhui Hu, Yuqian Gao, Xinlei Luo, Aimin Wang, and Lishuang Feng "Modeling and measurement of a miniature piezoelectric ceramic tube fiber scanner for two-photon endomicroscopy," Journal of Optical Microsystems 4(1), 014001 (13 January 2024). https://doi.org/10.1117/1.JOM.4.1.014001
Received: 21 September 2023; Accepted: 21 December 2023; Published: 13 January 2024
Advertisement
Advertisement
KEYWORDS
Ferroelectric materials

Scanners

Ceramics

Modeling

Endomicroscopy

Two photon imaging

Optical microsystems

RELATED CONTENT


Back to Top