Next Article in Journal
Effect of Surface Roughness on the Electrical Performances of CPW Transmission Lines Used in Future Ultra-High Frequency Applications
Previous Article in Journal
Study on In-Situ Tool Wear Detection during Micro End Milling Based on Machine Vision
Previous Article in Special Issue
General Strategy toward Laser Single-Step Generation of Multiscale Anti-Reflection Structures by Marangoni Effect
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Fabricating Air Pressure Sensors in Hollow-Core Fiber Using Femtosecond Laser Pulse

1
College of Physics and Electronic Science, Hubei Normal University, Huangshi 435002, China
2
State Key Laboratory of High Field Laser Physics and CAS Center for Excellence in Ultra-Intense Laser Science, Shanghai Institute of Optics and Fine Mechanics (SIOM), Chinese Academy of Sciences (CAS), Shanghai 201800, China
3
Shanghai Key Laboratory of Modern Optical System, Engineering Research Center of Optical Instrument and System, Ministry of Education, School of Optical-Electrical and Computer Engineering, University of Shanghai for Science and Technology, Shanghai 200093, China
*
Author to whom correspondence should be addressed.
Micromachines 2023, 14(1), 101; https://doi.org/10.3390/mi14010101
Submission received: 26 October 2022 / Revised: 13 December 2022 / Accepted: 25 December 2022 / Published: 30 December 2022
(This article belongs to the Special Issue Advanced Nanomachining)

Abstract

:
In this paper, a hollow core fiber was spliced with standard single-mode fibers to form a fiber optic gas pressure sensor, and its sensing characteristics with single hole or multi-holes punched on the hollow core fiber with femtosecond laser pulses were investigated. The experiments demonstrate that the air pressure sensitivity of the single hole sensor was −3.548 nm/MPa, with a linearity of 99.45%, while its response times for air pressure’s rise and fall were 4.25 s and 2.52 s, respectively. The air pressure sensitivity of the ten-hole sensor was up to −3.786 nm/MPa, with a linearity of 99.47%, while its response times for air pressure’s rise and fall were 2.17 s and 1.30 s, respectively. Theoretical analysis and experimental results indicate that the pressure sensitivity of the sensor with an anti-resonant reflecting guidance mechanism mainly comes from the refractive index change of the air inside the hollow core fiber. The proposed device with multi-holes drilled by a femtosecond laser has the advantages of fabrication simplicity, low cost, fast response time, good structural robustness, high repeatability, high sensitivity to air pressure, and insensitivity to temperature (only 10.3 pm/°C), which makes it attractive for high pressure sensing applications in harsh environments.

1. Introduction

Air pressure has been a key parameter in industrial production and meteorological monitoring. Recently, fiber-optic sensors for air pressure have been widely studied because of their advantages of small size, high sensitivity, and fast response time. Various solutions of fiber-optic air sensors have been proposed, including the gas pressure sensor based on hollow-core photonic bandgap fiber [1,2,3,4,5], inner air-cavity [6,7,8,9], a combination of hollow-core fiber and photonic crystal fiber [10,11], dual side-hole fiber [12], twin-core fiber [13], double off-axis twisted deformation of single mode fiber [14], quartz capillary tube with an open cavity [15], hollow optical fiber with two cores [16], laser-heated silicon pillar attached to a fiber tip [17], fiber-tip polyvinyl chloride diaphragm [18], and multimode fiber filled with ultraviolet adhesive [19]. However, these sensors usually need complex manufacturing processes and expensive specialty fibers. The hollow-core fiber (HCF) is not only cheap, simple to fabricate, but also of better compatibility with standard single-mode fibers (SMFs), making it suitable for fiber optic sensing applications [20,21,22,23,24,25,26,27,28,29].
In this paper, we propose and demonstrate a gas pressure sensor based on an anti-resonant reflecting guidance mechanism with HCF perforated by a femtosecond laser. The proposed sensor is fabricated by splicing a section of HCF between two sections of standard SMFs, and the single hole or multi-holes punched on the PCF enable the gas to enter or leave the cavity. The measurement of response time shows that the air pressure response of the fiber sensor with ten holes is twice as fast as that of the sensor with only one hole. Moreover, the proposed air pressure sensors also appear high sensitivity and a relatively low temperature cross-sensitivity of 2.9 kPa/°C and could be of a high potential in monitoring the gas pressure in high-temperature and high-pressure environments.

2. Fabrication of Sensors

Figure 1 shows a splicing process between a HCF with an inner diameter of 30 μm and a standard SMF (YOFC, 9 μm/125 μm). First, the coating layer of SMF is removed with a fiber stripper to expose its cladding layer, and the fiber is cleaved and cleaned to obtain a flat end face. The HCF is burned with a flame to remove its coating layer, and then is cleaved by a fiber cutting knife to obtain flat end faces as well. Then, both ends of the HCF are fused to the SMFs, resulting in a SMF-HCF-SMF structure, as illustrated in Step 3. As shown in Step 4, the air hole structure was fabricated by the femtosecond laser drilling in the hollow-core fiber fixed on a three-dimensional machining platform.
Figure 2a shows a microscopy image of a junction between a HCF with an inner diameter of 30 μm and a standard SMF. Figure 2b presents the interference spectra of the optical fiber sensor corresponding to a HCF with different lengths. It can be seen that the resonance peak intensity of the transmission spectrum gradually increases as the HCF length increases.
As shown in Figure 2b, the resonance peaks associated with the HCFs with the lengths of 1.4 mm and 2.5 mm are less than 5 dB, which are too low for accurate data analysis. When the length of the HCF extends to 10 mm, such fiber sensors are too big and not suitable for compact packaging. Fortunately, the resonance peak intensity of the HCF between 5 and 7 mm is greater than 10 dB, which fully meets the experimental requirements. In the subsequent experiments, the HCF with the length of about 6 mm was selected for fusion splicing with the SMF. The spliced structure was then subjected to laser drilling by a femtosecond laser (Coherent, Santa Clara, CA, USA) with the wavelength of 800 nm, the pulse width of 35 fs, and the laser energy of 50 μJ. The femtosecond laser pulses were focused on the middle top of the HCF with a 50× microscope objective, and a hole was ablated in less than 5 s with a programmed three-dimensional stage moving upward in the vertical direction at a speed of 10 μm/s. In addition, a refractive index matching liquid was applied to the HCF during both the femtosecond laser processing and microscopic imaging, which eliminated the adverse effects of spherical aberration on the experiments. The dimension (diameter and distance) accuracy of the small holes could be as small as 1 μm by use of a self-built micromachining station.
Figure 3 presents microscopy images of a single micro-hole in an HCF drilled by a femtosecond laser, showing the micro-hole has a round shape and a diameter of about 20 μm. It can be seen from Figure 3b that the micro-hole has almost vertical inner walls and the interior of the HCF is well connected to the outside atmosphere. Figure 4 presents microscopy images of ten micro-holes in an HCF drilled by a femtosecond laser. The ten micro-holes are equally arranged along the axial direction of the HCF, with a spacing of 100 μm between the holes and a diameter of 20 μm. All these images were taken by an Olympus BX 51 microscope under transmitted illumination. In order to verify the connectivity of the ten micro-holes, the HCF was filled with water to obtain a clear profile of the micro-holes, as shown in Figure 4b. Note that the dark region under the left micro-hole in Figure 4b is an interface between the air and water in the hollow core.

3. Theoretical Analysis

In this work, the sandwiched HCF between two SMFs can be considered as a single-layer anti-resonant reflecting optical waveguide (ARROW) [20,23]. The high refractive index cladding of HCF is regarded as a Fabry-Pérot etalon. When the wavelength of incident light is close to the resonance wavelength, light passes out of the high refractive index cladding, resulting in the periodic lossy dips in the transmission spectrum of HCF, as shown in Figure 2b. The propagating light at the anti-resonant wavelength (far away from the resonant wavelength) is internally reflected and confined in the hollow core of HCF, corresponding to the location of the periodic transmission peaks in Figure 2b.
The resonant wavelength can be simplified as the following expression [23]:
λ m = 2 d m n c l 2 n a i r 2 ,
where m is the resonant order, d is thickness of the cladding (~47.5 µm), ncl (~1.45), and nair (~1) are refractive indexes of the cladding and hollow core of the HCF, respectively.
According to Equation (1), we can derive the free spectral range (FSR):
FSR = λ m λ m + 1 2 d n c l 2 n a i r 2 ,
The light transmission mechanism of the HCF can be explained by the anti-resonance principle. The transmission spectrum of the air layer confined in the middle of the HCF is extremely sensitive to changes in external factors, such as the change of in the air pressure inside the cavity that affect the refractive index (RI) of the air.
S P = d λ m d P = 2 d n c l m n c l 2 n a i r 2 n c l P 2 d n a i r m n c l 2 n a i r 2 n a i r P + 2 n c l 2 n a i r 2 m d P
where P is the applied air pressure, and the influence of air pressure change on the cavity length of the optical fiber sensor can be ignored [30]. This is because the HCF is filled with air, the effective Young’s modulus is very small, and the longitudinal strain is almost determined by the outer cladding of silica. The refractive index changes of the HCF and the deformation of the tube wall can also be neglected [20].
Therefore, the drift of the resonance peak is mainly dominated by the change in the RI of the air. The sensitivity SP can be simplified into the following equation:
S P = d λ m d P = 2 d n a i r m n c l 2 n a i r 2 n a i r P ,
As the pressure increases, the RI of air in the cavity also increases according to the following trend [4].
n a i r = 1 + 2.8793 × 10 9 1 + 0.003671 × T P
Therefore, if the temperature is kept at 25 °C, the pressure sensitivity can be calculated to be about 3.398 nm/MPa at 1500 nm in theory. This value is in agreement with the experimental results (−3.548 nm/MPa, −3.786 nm/MPa), which indicates that the RI of the air played a major role in the observed pressure sensitivity.

4. Experimental Results and Analyses

4.1. Experimental Setup

The experimental setup for temperature and air pressure sensing is shown in Figure 5. The light is output from a Broadband Light Source (BBS, FL-ASE-EB-D-2-FC/2-APC), a spontaneous radiation source with a spectral range of 1250~1650 nm. After the light passes through the sensor, it is finally connected to an Optical Spectrum Analyzer (OSA, AQ6370D), with an accuracy of ±0.02 nm. The interferometer is connected to the BBS and OSA at both ends, while the device is in the straightened state. All measurements were repeated three times.

4.2. Air Pressure Experiment of Single-Hole Sensor

Figure 6a shows an initial spectrum of the single-hole gas pressure sensor, in which the length and the inner diameter of the spliced HCF is 6 mm and 30 μm, respectively. It can be seen that the resonance peak intensity of the interference spectrum gradually rises with the increase in wavelength. As a result, the resonance peak near 1550 nm (Dip1), with an interference intensity of about 12 dB, is selected as the object of study.
As illustrated in Figure 5a, the single-hole fiber optic sensing head is placed inside a chamber with four ports. Its two ports are individually connected to BBS and OSA, while the remaining two ports are connected to an air pump and a barometer, respectively. In order to keep the chamber airtight, the four ports were sealed with epoxy AB adhesive. By pressing the air pressure pump, the air pressure inside the airtight chamber can be increased from 0.1 MPa to 0.8 MPa, while the Dip1 wavelengths are recorded every 0.1 MPa after the spectrum is stabilized. The spectral drift with air pressure is presented in the inset at the lower left corner of Figure 6b. When the air pressure inside the airtight chamber decreases from 0.8 MPa to 0.1 MPa, the Dip1 wavelengths are also recorded every 0.1 MPa. As shown in Figure 6b, the linear fits of the response curves almost overlap when the air pressure rises and falls, indicating the good repeatability. When the air pressure is rising, the sensitivity of the Dip1 wavelength is −3.548 nm/MPa with a linearity of 99.45%; and when the air pressure is falling, the sensitivity of the Dip1 wavelength is −3.429 nm/MPa with a linearity of 99.54%. It is worthy to indicate that the consistency of response curves has been verified by repeated measurements, in which the standard deviation of the multiple measurement results is too small to be distinguished by the OSA used in this work.
As shown in Figure 6a, four adjacent valleys in the wavelength range near 1500 nm to 1550 nm are selected for analysis. They are 1493.6 nm, 1515.4 nm, 1538 nm, and 1561 nm. Their adjacent intervals are 21.8 nm, 22.6 nm, and 23 nm. The FSRs are calculated from Equation (2) as 22.7 nm, 23.36 nm, and 24 nm, and it can be seen that the experimental and theoretical results are very close to each other.

4.3. Air Pressure Experiment of Ten-Hole Sensor

The transmission spectra of the ten-hole gas pressure sensor before and after femtosecond laser punching are shown in Figure 7a. The length of the HCF fused to the air sensor is 6.2 mm, and its inner diameter is still 30 μm. The perforated device was annealed at 600 °C to relieve the stress induced by laser ablation. The intensity of the transmission spectrum does not change obviously before and after annealing, while the wavelength of the resonant peak drifted significantly.
The sensitivity measurement of the ten-hole gas pressure sensor is almost the same as that of the single hole sensor. At first, the air pressure in the airtight chamber slowly rises from 0 MPa to 0.6 MPa, while the Dip2 wavelengths (see Figure 7a) are recorded every 0.1 MPa. The spectral drift with air pressure is presented in the inset at the lower left corner of Figure 7b. Then, the air pressure slowly decreases from 0.6 MPa to 0 MPa. At the same time, the Dip2 wavelengths are also recorded every 0.1 MPa. Figure 7b shows linear fits of the response curves. The linearities of the rising and falling curves are 99.43% and 99.47%, and the sensitivities of the Dip2 wavelength are −3.643 nm/MPa and −3.786 nm/MPa, respectively.

4.4. Comparison Experiments of Response Times of Sensors

Response time is an essential parameter to judge the performance of the fiber optic air pressure sensor. Here, the response time of one-hole and ten-hole air pressure sensors are measured with the experimental setup in Figure 5a. Firstly, the sensing head was placed in the airtight chamber, and the Dip1 wavelength was recorded when the air pressure was set at 0 MPa. Then, the chamber pressure was rapidly increased to 0.3 MPa, and the time of Dip1 drift after the pressure reach 0.3 MPa was recorded. After holding for 30 s, the chamber pressure was rapidly dropped from 0.3 MPa to 0 MPa, and the time of Dip1 drift after the pressure reach 0 MPa was recorded again. The above measurements were repeated six times, and the ramp-up and the ramp-down average response times are illustrated in Figure 8. It can be seen that the response speed of the ten-hole sensor is twice as fast as that of the single-hole sensor. We expect that the faster response speed is caused by the smaller resistance of air entering into the HCF with more holes. It is worthy to note that the response time would be shorter by machining an open microcavity directly on the fiber crossing the core layer, but this microcavity will seriously reduce the robustness of fiber sensor. Multiple micro-holes in an HCF offer a pathway to raise the response speed of air pressure sensors but keep its robustness to high pressure environment.

4.5. Temperature Experiment

Figure 5b shows a diagram of the temperature experimental setup of the single-hole sensor, its sensing head was placed in a temperature-controllable tubular heating furnace, and the temperature was gradually increased from 30 °C to 130 °C by the program setting, while the spectral data were recorded every 10 °C, as displayed in the upper left inset of Figure 9. When the temperature rose, the wavelength of the resonance valley Dip1 of the single-hole sensor was red-shifted, and its intensity barely changed. The linear fit of the resonance valley Dip1 is 98.37%, its temperature sensitivity is 10.3 pm/°C, with the cross-sensitivity of temperature being 2.9 kPa/°C (0.0103/3.548 = 0.0029 MPa/°C), indicating that the device has a very small cross-sensitivity to temperature changes.

5. Conclusions

Single-hole and ten-hole HCF gas pressure sensors were demonstrated by femtosecond laser drilling in the HCFs sandwiched between two SMFs. The air pressure sensing experiments were performed respectively, showing that the air pressure sensitivity of the single-hole sensor reached the maximum at −3.548 nm/MPa with a linear fit of 99.45%, and the air pressure sensitivity of the ten-hole sensor achieved the maximum at −3.786 nm/MPa with a linear fit of 99.47%. These experimental results were both close to the theoretical values. The air sensitivities of the single-hole and ten-hole gas pressure sensors are similar, but the response speed of the ten-hole fiber-optic sensor is approximately twice as fast as that of the single-hole one. In addition, the temperature sensitivity of the single-hole air sensor is only 10.3 pm/°C, and the cross-sensitivity of temperature and air pressure is 2.9 kPa/°C, indicating that the effect of temperature fluctuations on the sensor is negligible compared to its air pressure sensitivity. Due to the advantages of low cost, easy fabrication, high air pressure sensitivity but insensitive to temperature, small size, fast response time, and good structural robustness, the device with multi-holes can offer a wide range of applications in high-pressure measurements in extreme environments.

Author Contributions

C.L.: writing—original draft preparation; W.T.: software and data curation; C.C.: formal analysis; Y.L.: writing—review and editing, supervision and funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

This work is supported by the National Natural Science Foundation of China (12004105), Program for Innovative Teams of Outstanding Young and Middle-aged Researchers in the Higher Education Institutions of Hubei Province (Grant No. T2020014 and T2021010), Open Fund of Laboratory of Solid-State Microstructures (Grant No. M35054), and the foundation of Graduate Innovation Research of Hubei Normal University (Grand No. 20220501).

Data Availability Statement

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.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Tang, J.; Zhang, Z.; Yin, G.; Liu, S.; Bai, Z.; Li, Z.; Deng, M.; Wang, Y.; Liao, C.; He, J.; et al. Long Period Fiber Grating Inscribed in Hollow-Core Photonic Bandgap Fiber for Gas Pressure Sensing. IEEE Photonics J. 2017, 9, 7105307. [Google Scholar] [CrossRef]
  2. Mao, C.; Huang, B.; Wang, Y.; Huang, Y.; Zhang, L.; Shao, Y. High-sensitivity gas pressure sensor based on hollow-core photonic bandgap fiber Mach-Zehnder interferometer. Opt. Express 2018, 26, 30108–30115. [Google Scholar] [CrossRef] [PubMed]
  3. Huang, B.; Wang, Y.; Mao, C. Temperature-Independent Gas Pressure Sensor with High Birefringence Photonic Crystal Fiber-Based Reflective Lyot Filter. Sensors 2019, 19, 5312. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Tang, J.; Yin, G.; Liao, C.; Liu, S.; Li, Z.; Zhong, X.; Wang, Q.; Zhao, J.; Yang, K.; Wang, Y. High-Sensitivity Gas Pressure Sensor Based on Fabry–Pérot Interferometer With a Side-Opened Channel in Hollow-Core Photonic Bandgap Fiber. IEEE Photonics J. 2015, 7, 6803307. [Google Scholar] [CrossRef]
  5. Zhang, Z.; He, J.; Dong, Q.; Bai, Z.; Liao, C.; Wang, Y.; Liu, S.; Guo, K. Diaphragm-free gas-pressure sensor probe based on hollow-core photonic bandgap fiber. Opt. Lett. 2018, 43, 3017–3020. [Google Scholar] [CrossRef] [Green Version]
  6. Talataisong, W.; Wang, D.N.; Chitaree, R.; Liao, C.R.; Wang, C. Fiber in-line Mach–Zehnder interferometer based on an inner air-cavity for high-pressure sensing. Opt. Lett. 2015, 40, 1220–1222. [Google Scholar] [CrossRef]
  7. Liu, Y.; Zhang, T.; Wang, Y.; Yang, D.; Liu, X.; Fu, H.; Jia, Z. Simultaneous measurement of gas pressure and temperature with integrated optical fiber FPI sensor based on in-fiber micro-cavity and fiber-tip. Opt. Fiber Technol. 2018, 46, 77–82. [Google Scholar] [CrossRef]
  8. Lin, H.; Liu, F.; Dai, Y.; Zhou, A. Cascaded Fiber Mach–Zehnder Interferometers for Sensitivity-Enhanced Gas Pressure Measurement. IEEE Sens. J. 2018, 19, 2581–2586. [Google Scholar] [CrossRef]
  9. Xu, B.; Liu, Y.; Wang, D.; Jia, D.; Jiang, C. Optical Fiber Fabry–Pérot Interferometer Based on an Air Cavity for Gas Pressure Sensing. IEEE Photonics J. 2017, 9, 7102309. [Google Scholar] [CrossRef]
  10. Quan, M.; Tian, J.; Yao, Y. Ultra-high sensitivity Fabry–Perot interferometer gas refractive index fiber sensor based on photonic crystal fiber and Vernier effect. Opt. Lett. 2015, 40, 4891–4894. [Google Scholar] [CrossRef]
  11. Cui, Y.; Jiang, Y.; Liu, T.; Hu, J.; Jiang, L. A Dual-Cavity Fabry–Perot Interferometric Fiber-Optic Sensor for the Simultaneous Measurement of High-Temperature and High-Gas-Pressure. IEEE Access 2020, 8, 80582–80587. [Google Scholar] [CrossRef]
  12. Lin, H.; Liu, F.; Guo, H.; Zhou, A.; Dai, Y. Ultra-highly sensitive gas pressure sensor based on dual side-hole fiber interferometers with Vernier effect. Opt. Express 2018, 26, 28763–28772. [Google Scholar] [CrossRef] [PubMed]
  13. Li, Z.; Liao, C.; Wang, Y.; Xu, L.; Wang, D.; Dong, X.; Liu, S.; Wang, Q.; Yang, K.; Zhou, J. Highly-sensitive gas pressure sensor using twin-core fiber based in-line Mach-Zehnder interferometer. Opt. Express 2015, 23, 6673–6678. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Liu, Y.; Lin, H.; Dai, Y.; Zhou, A.; Yuan, L. Fiber In-Line Mach–Zehnder Interferometer for Gas Pressure Sensing. IEEE Sens. J. 2018, 18, 8012–8016. [Google Scholar] [CrossRef]
  15. Yang, Y.; Wang, D.; Xu, B.; Wang, Z. Optical fiber tip interferometer gas pressure sensor based on anti-resonant reflecting guidance mechanism. Opt. Fiber Technol. 2018, 42, 11–17. [Google Scholar] [CrossRef]
  16. Yang, X.; Zhao, Q.; Qi, X.; Long, Q.; Yu, W.; Yuan, L. In-fiber integrated gas pressure sensor based on a hollow optical fiber with two cores. Sens. Actuators A Phys. 2018, 272, 23–27. [Google Scholar] [CrossRef]
  17. Liu, G.; Han, M. Fiber-optic gas pressure sensing with a laser-heated silicon-based Fabry–Perot interferometer. Opt. Lett. 2015, 40, 2461–2464. [Google Scholar] [CrossRef]
  18. Zhang, Z.; Liao, C.; Tang, J.; Bai, Z.; Guo, K.; Hou, M.; He, J.; Wang, Y.; Liu, S.; Zhang, F. High-Sensitivity Gas-Pressure Sensor Based on Fiber-Tip PVC Diaphragm Fabry–Pérot Interferometer. J. Light. Technol. 2017, 35, 4067–4071. [Google Scholar] [CrossRef]
  19. Chen, W.P.; Wang, D.N.; Xu, B.; Zhao, C.L.; Chen, H.F. Multimode fiber tip Fabry-Perot cavity for highly sensitive pressure measurement. Sci. Rep. 2017, 7, 368. [Google Scholar] [CrossRef] [Green Version]
  20. Hou, M.; Zhu, F.; Wang, Y.; Liao, C.; Liu, S.; Lu, P. Antiresonant reflecting guidance mechanism in hollow-core fiber for gas pressure sensing. Opt. Express 2016, 24, 27890–27898. [Google Scholar] [CrossRef]
  21. Xu, B.; Liu, Y.M.; Wang, D.N.; Li, J.Q. Fiber Fabry–Pérot Interferometer for Measurement of Gas Pressure and Temperature. J. Light. Technol. 2016, 34, 4920–4925. [Google Scholar] [CrossRef]
  22. Gao, H.; Jiang, Y.; Zhang, L.; Cui, Y.; Jiang, Y.; Jia, J.; Jiang, L. Antiresonant mechanism based self-temperature-calibrated fiber optic Fabry–Perot gas pressure sensors. Opt. Express 2019, 27, 22181–22189. [Google Scholar] [CrossRef]
  23. Bao, L.; Dong, X.; Shum, P.P.; Shen, C. High sensitivity liquid level sensor based on a hollow core fiber structure. Opt. Commun. 2021, 499, 127279. [Google Scholar] [CrossRef]
  24. Fu, D.; Liu, X.; Shang, J.; Sun, W.; Liu, Y. A Simple, Highly Sensitive Fiber Sensor for Simultaneous Measurement of Pressure and Temperature. IEEE Photonics Technol. Lett. 2020, 32, 747–750. [Google Scholar] [CrossRef]
  25. Zhang, L.; Jiang, Y.; Gao, H.; Jia, J.; Cui, Y.; Ma, W.; Wang, S.; Hu, J. A diaphragm-free fiber Fabry-Perot gas pressure sensor. Rev. Sci. Instrum. 2019, 90, 025005. [Google Scholar] [CrossRef]
  26. Xu, B.; Wang, C.; Wang, D.N.; Liu, Y.; Li, Y. Fiber-tip gas pressure sensor based on dual capillaries. Opt. Express 2015, 23, 23484–23492. [Google Scholar] [CrossRef]
  27. Chen, M.-Q.; Wei, H.-M.; Zhao, Y.; Lei, X.-H.; Krishnaswamy, S. Temperature insensitive air-cavity Fabry-Perot gas pressure sensor based on core-offset fusion of hollow-core fibers. Sens. Actuators A Phys. 2019, 298, 111589. [Google Scholar] [CrossRef]
  28. Liu, S.; Wang, Y.; Liao, C.; Wang, Y.; He, J.; Fu, C.; Yang, K.; Bai, Z.; Zhang, F. Nano silica diaphragm in-fiber cavity for gas pressure measurement. Sci. Rep. 2017, 7, 787. [Google Scholar] [CrossRef] [Green Version]
  29. Cheng, X.; Dash, J.; Gunawardena, D.; Htein, L.; Tam, H.-Y. Silicone Rubber Based Highly Sensitive Fiber-Optic Fabry–Perot Interferometric Gas Pressure Sensor. Sensors 2020, 20, 4927. [Google Scholar] [CrossRef]
  30. Jin, L.; Guan, B.-O.; Wei, H. Sensitivity Characteristics of Fabry-Perot Pressure Sensors Based on Hollow-Core Microstructured Fibers. J. Light. Technol. 2013, 31, 2526–2532. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of a fusion splicing process between SMF and HCF.
Figure 1. Schematic diagram of a fusion splicing process between SMF and HCF.
Micromachines 14 00101 g001
Figure 2. (a) A microscope image of a junction between a SMF and an HCF, (b) Interference spectrum corresponding to different lengths of HCF.
Figure 2. (a) A microscope image of a junction between a SMF and an HCF, (b) Interference spectrum corresponding to different lengths of HCF.
Micromachines 14 00101 g002
Figure 3. Microscopic images of single hole drilled in a HCF. (a) Front view; (b) Side view.
Figure 3. Microscopic images of single hole drilled in a HCF. (a) Front view; (b) Side view.
Micromachines 14 00101 g003
Figure 4. Microscopic images of ten micro-holes drilled in an HCF. (a) Front view; (b) Side view.
Figure 4. Microscopic images of ten micro-holes drilled in an HCF. (a) Front view; (b) Side view.
Micromachines 14 00101 g004
Figure 5. Schematic diagram of the experimental setup. (a) Air pressure sensing unit; (b) Temperature sensing unit.
Figure 5. Schematic diagram of the experimental setup. (a) Air pressure sensing unit; (b) Temperature sensing unit.
Micromachines 14 00101 g005
Figure 6. Single-hole sensor for air pressure. (a) Initial spectrum; (b) Spectral drift and linear fit.
Figure 6. Single-hole sensor for air pressure. (a) Initial spectrum; (b) Spectral drift and linear fit.
Micromachines 14 00101 g006
Figure 7. Air pressure sensing experiment data of ten-hole sensor. (a) Initial spectrum; (b) Spectral drift and linear fit.
Figure 7. Air pressure sensing experiment data of ten-hole sensor. (a) Initial spectrum; (b) Spectral drift and linear fit.
Micromachines 14 00101 g007
Figure 8. Comparison graphs of response times: (a) Response time of single-hole sensor; (b) Response time of ten-hole sensor.
Figure 8. Comparison graphs of response times: (a) Response time of single-hole sensor; (b) Response time of ten-hole sensor.
Micromachines 14 00101 g008
Figure 9. Spectral drift and linear fit of temperature sensitivity.
Figure 9. Spectral drift and linear fit of temperature sensitivity.
Micromachines 14 00101 g009
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Liu, C.; Tao, W.; Chen, C.; Liao, Y. Fabricating Air Pressure Sensors in Hollow-Core Fiber Using Femtosecond Laser Pulse. Micromachines 2023, 14, 101. https://doi.org/10.3390/mi14010101

AMA Style

Liu C, Tao W, Chen C, Liao Y. Fabricating Air Pressure Sensors in Hollow-Core Fiber Using Femtosecond Laser Pulse. Micromachines. 2023; 14(1):101. https://doi.org/10.3390/mi14010101

Chicago/Turabian Style

Liu, Changning, Wuqiang Tao, Cong Chen, and Yang Liao. 2023. "Fabricating Air Pressure Sensors in Hollow-Core Fiber Using Femtosecond Laser Pulse" Micromachines 14, no. 1: 101. https://doi.org/10.3390/mi14010101

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop