Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Microbubble-based optical fiber Fabry-Perot sensor for simultaneous high-pressure and high-temperature sensing

Open Access Open Access

Abstract

An all-silica Fabry-Perot interferometer (FPI) based on a microbubble for high-pressure and high-temperature measurements is proposed and demonstrated. The microbubble-based air cavity is fabricated using a hollow silica tube and a single-mode optical fiber for pressure sensing. The suitable thickness between the two end faces of the microbubble enables the silica cavity to be used for temperature sensing. The wavelength shift of the reflection spectrum versus pressure is linear, and the sensitivity reaches −5.083 nm/MPa at room temperature (20 °C) within the range of 0 − 4 MPa. The temperature sensitivity reaches 12.715 pm/°C within the range of 20 − 700 °C. The very low temperature-pressure cross-sensitivity of the two cavities indicates that the proposed FPI sensor offers great potential for simultaneous high-pressure and high-temperature measurements in harsh environments.

© 2022 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Fiber optic Fabry-Perot interferometers (FPIs) have many advantages, such as immunity to electromagnetic interference, corrosion resistance, and electrical insulation. Many methods have been proposed for fabricating FPIs, including arc discharge, film bonding, chemical and laser corrosion, and 3D printing [16]. These sensors have been widely used for strain, humidity, temperature, and pressure sensing [711]. In particular, FPI sensors deliver stable performance in harsh environments and are therefore suitable for application in the oil, aerospace, and steel industries [1214]. FPI pressure sensors are mainly divided into two types based on the deformation of both the end face and sidewall of the F-P cavity [12,1525]. Regardless of the type, narrowing the thickness is an effective way to improve the pressure sensitivity. The traditional fabrication methods, film bonding or welding is disadvantageous because of the manufacturing and technical complexities associated with these methods [2][6]. In recent years, microbubble-based devices have been widely studied because they obviate the need for bonding or welding processes and a very thin sensing area can be used, which can improve the pressure sensitivity and simplify the manufacturing process [2629].

Ma et al. [26] used a fusion splicer to weld at the single-mode optical fiber (SMF) with a well-cut hollow silica tube (HST) at its end face, and then formed a microbubble with a thick wall by arc discharging at a specific position in the HST. However, the sensor has difficulties realizing high-sensitivity pressure sensing because of the thick wall of the microbubble. In a subsequent study [27], they produced a microbubble structure with a thickness of 4.3 µm by tapering the gas-pressured HST while arc discharging using the fusion splicer based on the steps in the previous study. Owing to the small size of the microbubble, the pressure sensitivity of the sensor was only 315 pm/MPa, and the low temperature sensitivity was approximately 1.55 pm/°C. Liao et al. [28] proposed a microbubble sensor by fusing two SMFs together by coating their end faces with oil with a similar refractive index. As the oil was vaporized and expanded to a microbubble owing to the high temperature, a thin-end microbubble could be formed at the end face of the fiber. Although the thickness of the end face of the microbubbles further decreased in Liao’s study, the thickness of the microbubble was uneven, and the thinnest area was limited to the top part of the microbubble. This resulted in a low-pressure sensitivity of 1.036 nm/MPa and a low-temperature sensitivity of 1 pm/°C. Li et al. [29] formed a microbubble structure by arc discharging on the end face of a pressurized HST, then inserted the SMF into the microbubble, and finally welded them together. The pressure sensitivity is approximately −6.382 nm/MPa whereas the temperature coefficient is lower at approximately 0.17 pm/°C. Li improved the uniformity of the microbubble thickness, but the mechanical strength of the sensor decreased as the sidewall of the microbubble was very thin (∼0.5 µm), which limited the measurement range. It should be noted that the temperature response of the FPI results from the thermal expansion and thermo-optic effect of the F-P cavity, and most of these sensors are based on a single air cavity; thus, the changes in the cavity length are mainly caused by the thermal expansion of the microbubbles. Therefore, the temperature response of an FPI based on a single air cavity is negligible [26].

In all of the above microbubble manufacturing processes, the F-P cavity was fabricated to have a thinner end face. Few studies have attempted to improve the pressure sensitivity of sensors by thinning the side of the F-P cavity when fabricating the microbubble. In this paper, a novel microbubble-based FPI sensor is proposed and demonstrated. In contrast to the above reports, the sensor mainly realizes pressure sensing by deformation of the thin sidewall of the microbubble (air cavity). Despite the thin sidewall, the cross-sensitivity of the air cavity is very low. Moreover, the sensor realizes a compound cavity with the function of temperature sensing by exploiting the thermal expansion and thermo-optic effects of the thick part at the end of the microbubble (silica cavity). Compared to traditional sensors based on sidewall deformation [14][25], the proposed sensor has a compact structure and offers improved pressure sensitivity owing the significantly reduced thickness of the sidewall. Apart from this, no additional structures and processes are required for temperature sensing [30], which greatly improves the manufacturing efficiency. In addition, the sensor can operate stably in high-temperature environments owing to its all-silica structure. The results show that the pressure sensitivity of the sensor reaches −5.083 nm/MPa and the temperature sensitivity reaches 12.715 pm/°C. This means that the proposed sensor has great application prospects in harsh environments for high-temperature and high-pressure sensing.

2. Operating principle and fabrication

Figure 1(a) shows a structural diagram of the microbubble-based FPI sensor with an ellipsoidal shape, where d and L3 are the diameters of the short and long axes, respectively. L1 and L2 are the lengths of the air and the silica cavities, respectively. t represents the thickness of the sidewall of the microbubble. L is the distance between the fusion point and outer end face (r3) of the sensor. The air cavity is formed by the end face of the SMF (r1) and the inner end face of the microbubble (r2), whereas the silica cavity is formed by the inner (r2) and outer (r3) end faces of the microbubble. The fusion point between the SMF and HST creates space in the microbubble hermetic. Therefore, when the surrounding gas pressure changes, the sidewall of the microbubble is deformed owing to the difference between the internal and external pressure of the sensor, resulting in a change of the air cavity length of the sensor. The finite element analysis (COMSOL multiphysics) was used to qualitatively simulate the deformation of the microbubble structure (d = 470 µm, L3 = 550 µm, t = 10 µm, L2 = 180 µm, and L = 1 cm) under a pressure of 4 MPa. The deformation scale factor of the sensor is set as 500 in COMSOL software to show the deformation clear. As expected, the deformation of the sensor is mainly concentrated in the microbubble of the head of the sensor, and it’s mainly caused by the deformation of the sidewall of the microbubble, as shown in Fig. 1(b).

 figure: Fig. 1.

Fig. 1. Diagram of the (a) structure of the sensor and (b) deformation of the microbubble at 4 MPa simulated by COMSOL software.

Download Full Size | PDF

Figure 2 shows the reflected electric fields of the three reflective interfaces of the sensor (r1, r2, and r3). Only the first-order reflected fields of the three interfaces were considered because of the low reflectivity of the silica/air surface. The total reflected electric field (Er) can be regarded as the sum of the three reflected electric fields and can be expressed as

$$\begin{array}{l} {E_r} = {E_i}[\sqrt {{R_1}} + (1 - {A_1})(1 - {\alpha _1})(1 - {R_1})\sqrt {{R_2}} {e^{ - j2\beta {L_1} + j\pi }}\\ + (1 - {A_1})(1 - {A_2})(1 - {\alpha _1})(1 - {\alpha _2})(1 - {R_1})(1 - {R_2})\sqrt {{R_3}} {e^{ - j\beta (2{L_1} + 2{L_2})}}]. \end{array}$$
where Ei is the input electric field and R1, R2, and R3 are the reflection coefficients of the three reflective interfaces (r1, r2, and r3), respectively, with ${R_1} = {R_2} = {R_3} = {({n_s} - {n_a})^2}/{({n_s} + {n_a})^2}$. A1, A2, and A3 represent the transmission loss factors of the three reflective interfaces. ${\alpha _1}$ and ${\alpha _2}$ are the loss factors of the air and silica cavities, respectively. The transmission loss of the cavities over a distance can be ignored. L1 and L2 are the lengths of the air and silica cavities, respectively. (L1 + L2) is the combined length of the air-silica-cavity, and $\beta$ is the propagation constant of the optical fiber, which is $\beta = 2\pi n/\lambda$. Because the refractive index of air is less than that of silica, a π-phase shift occurs at the reflection surface r2. From Eq. (1), we can obtain the normalized reflection spectrum ${R_{FP}}(\lambda )$ as follows:
$$\begin{array}{l} {R_{FP}}(\lambda ) = {\left|{\frac{{{E_r}}}{{{E_i}}}} \right|^2}\\ = {R_1} + {(1 - {A_1})^2}{(1 - {\alpha _1})^2}{(1 - {R_1})^2}{R_2}\\ {\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} + {(1 - {A_1})^2}{(1 - {A_2})^2}{(1 - {\alpha _1})^2}{(1 - {\alpha _2})^2}{(1 - {R_1})^2}{(1 - {R_2})^2}{R_3}\\ {\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} - 2\sqrt {{R_1}{R_2}} (1 - {A_1})(1 - {\alpha _1})(1 - {R_1})\cos (4\pi {n_a}{L_1}/\lambda )\\ {\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} + 2\sqrt {{R_1}{R_3}} (1 - {A_1})(1 - {A_2})(1 - {\alpha _1})(1 - {\alpha _2})(1 - {R_1})(1 - {R_2})\cos [4\pi ({n_a}{L_1} + {n_s}{L_2})/\lambda ]\\ {\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} - 2\sqrt {{R_2}{R_3}} {(1 - {A_1})^2}{(1 - {\alpha _1})^2}{(1 - {R_1})^2}(1 - {A_2})(1 - {\alpha _2})(1 - {R_2})\cos [4\pi (2{n_a}{L_1} + {n_s}{L_2})/\lambda ]. \end{array}$$
where na and ns are the refractive indices of the air and silica cavities, respectively, and λ is the working wavelength. Equation (2) indicates that a direct current (DC) component exists in the reflection spectrum. The three frequency components correspond to the three cavities of the sensor. In addition, the relationship between the cavity length and the free spectral range (FSR) is written as
$$FSR = \frac{{{\lambda ^2}}}{{2nd}}.$$
where d represents the length of the F-P cavity. Therefore, we can calculate the corresponding F-P cavity length of the sensor based on the FSR. The pressure is determined by measuring the optical path difference $(\Delta OP{D_P})$ of the air cavity:
$$\Delta OP{D_P} = 2{n_a}({L_1} + \Delta L) - 2{n_a}{L_1} = 2{n_a}\Delta L.$$

 figure: Fig. 2.

Fig. 2. Transmission path of the electric field in the FPI cavity.

Download Full Size | PDF

The relationship between the wavelength shift $(\Delta {\lambda _P})$ of the air cavity of the FPI sensor and the variation in $\Delta OP{D_P}$ is written as

$$\Delta {\lambda _P} = \frac{{\Delta OP{D_P}}}{{OP{D_P}}}\lambda = \frac{{{n_a}\Delta L}}{{{n_s}{L_2} + {n_a}{L_1}}}\lambda .$$

According to Eq. (5), the wavelength shift $(\Delta {\lambda _P})$ of the FPI sensor is directly proportional to the change in $\Delta OP{D_P}$. The $\Delta \lambda _P$ is linearly related to the variation in air cavity length $(\Delta L)$, as the change in the air refractive index $({n_a})$ can be ignored owning to the weak deformation of the air cavity. Similarly, the relationship between the temperature and wavelength shift of the silica cavity is given by:

$$\Delta {\lambda _T} = \frac{{\Delta OP{D_T}}}{{OP{D_T}}}\lambda = (\frac{{\Delta L}}{L} + \frac{{\Delta n}}{n})\lambda = (\beta + \partial )\lambda \Delta T.$$

The wavelength shift $(\Delta {\lambda _T})$ at high-temperature is mainly affected by the thermal expansion $(\beta )$ and thermo-optic $(\partial )$ coefficients.

In contrast to the heating mode of the traditional arc discharge fusion splicer, laser welding has the advantages of accurate power input with a high energy density, deep penetration, good directivity, and an accurate heating position [3133]. Therefore, a CO2 laser processing system (LZM-100, Fujikura, Tokyo, Japan) was used to prepare the sensor. This laser enables different positions of the microbubble to be accurately heated, and the fiber to be rotated during welding, which makes it possible to obtain a symmetric microbubble with a unique structure. Figure 3 illustrates the fabrication process of the sensor. First, an end-cut HST with outer and inner diameters of 320 µm and 130 µm, respectively, was affixed to the one side of the fiber clamp (EV 400, Fujikura, Tokyo, Japan) of the CO2 laser processing system. Then, its position was adjusted to ensure that the end of the HST was located in the laser heating zone, as shown in Fig. 3(a). Second, the end of the HST was manually welded to form a hermetic structure. Because the fiber could be rotated at a uniform speed during the welding process, a symmetrical sealed structure was formed when the laser power was 400 bit, the welding time was 30 s, and the rotating speed was 0.075 °/ms, as shown in Fig. 3(b). Third, a syringe was used to pump air into the HST to form and expand the microbubble while it was being continuously heated under the same laser power and rotating speed with the welding time of 20 s, as shown in Fig. 3(c). Fourth, the entire structure was moved such that the sidewall of the microbubble was positioned in the center of the heating zone. Thus, the microbubble was not uniformly heated when the laser power was 320 bit, the welding time was 20 s, and the rotating speed was 0.075 °/ms. The temperature of the action area of the CO2 laser decreased gradually from the center to the outside of the bubble. At the point at which the energy absorbed by the end face of the microbubble is equal to the energy emitted, the temperature in this area ceases to rise. This prevents the end face of the microbubble from transforming into a molten state to become thinner. As a result, the sidewall closer to the area being heated expands faster than the end of the microbubble. Consequently, a microbubble structure with a thin sidewall and thick end face is obtained, as shown in Fig. 3(d). The microbubbles with different sizes and sidewall thicknesses can be obtained by adjusting the welding time and laser power. Finally, the cleaved SMF was manually controlled for insertion into the HST at a specific location, whereupon the HST and fiber were welded together at a fixed position away from the fiber tip with the laser power of 305 bit, and the welding time of 35 s, as shown in Fig. 3(e) and 3(f). This procedure enabled an FPI sensor with a specified air cavity length to be obtained.

 figure: Fig. 3.

Fig. 3. Fabrication procedure of the sensor: (a) end-cut HST in the CO2 laser; (b) fusing one end of the HST; (c) attaching the gas-filled capillary; (d) adjusting the position to fuse the sidewall of microbubble; (e) fusing the SMF with the HST; (f) complete structure of the sensor.

Download Full Size | PDF

Microscope (BX53M Olympus, Tokyo, Japan) images of the three fabricated devices with the sidewall thicknesses (t) of 9.8, 12.6, and 35.9 µm, respectively, are shown in Fig. 4(a). The diameter of the short axis (d) was 528.4, 546.3, and 642.6 µm, respectively. The diameter of the long axis (L3) was 604.1, 632.3, and 726.6 µm, respectively. Because the length of the sensor (L) is 2.0 cm, the microbubbles and fusion points cannot be observed at the same time in the field of view of the microscope as the maximum field of view is 0.26 cm×0.33 cm. Therefore, we observed the state of the microbubbles and fusion points under the same magnification, respectively, and then stitched them together. Figure. 4(b) shows the corresponding FPI reflection spectra. Because of the compound cavity structure, a large envelope exists in the interference spectrum. Fast Fourier transform (FFT) was used to obtain the corresponding frequency components of the two cavities.

 figure: Fig. 4.

Fig. 4. (a) Microscope images of the three fabricated devices and (b) corresponding reflection spectra of the three fabricated samples, respectively.

Download Full Size | PDF

Sample 2 was used as an example. First, the power unit of the interference spectra was converted to “mW,” as shown in Fig. 5(a). This spectrum was then converted to the spatial frequency spectrum by FFT. The latter spectrum contains three distinct peaks, which represent the air, silica, and the air-silica cavities, respectively, as shown in Fig. 5(b). The interference spectrum of the air and silica cavities can be derived using the narrow-band filtering method. Because the two reflecting interfaces of the silica cavity (r2 and r3) are curved, it is difficult to keep the two interfaces strictly parallel; hence, the interference spectrum of the silica cavity is relatively weak, as shown in Fig. 5(c) and (d). Based on the results in Fig. 5, and using Eq. (3), we can calculate the length of the two cavities of these sensors. The calculated air cavity lengths (L1) of samples 1, 2, and 3 are 60.1, 71.3, and 75.1 µm, respectively, with the corresponding silica cavity lengths (L2) demodulated as 227.8, 196.8, and 196.7 µm respectively.

 figure: Fig. 5.

Fig. 5. (a) Reflection spectrum of sample 2; (b) FFT of reflection spectrum of sample 2; frequency component of the (c) air cavity and (d) silica cavity.

Download Full Size | PDF

3. Experiments and results

The setup for the high-temperature pressure measurements is shown in Fig. 6. A nitrogen gas cylinder with a pressure regulator (QY-C1-10-Y, resolution ±0.02 MPa, Festo (China), Shanghai, China) and a digital pressure meter (700G08, Fluke, WA, USA) were used for pressure measurement. A tube furnace was used for temperature measurements, and an optical fiber sensing interrogator (SM125, Micron Optics Inc., Atlanta, GA, USA) with a wavelength range of 1510–1590 nm and a resolution of 5 pm was used for signal demodulation. The sensor head was placed in a sealed alumina ceramic tube located in the central area of the heating zone of the tube furnace heating, and the reflection of the sensor was monitored in real time.

 figure: Fig. 6.

Fig. 6. Schematic diagram of the system to measure the high-temperature-pressure performance of the FPI sensor.

Download Full Size | PDF

The results of the pressure test of the air cavity of the three samples, measured in the range from 0 − 4 MPa in increments of 0.5 MPa at 20 °C, are shown in Fig. 7. The maximum pressure sensitivity reaches −5.083 nm/MPa of the three samples. As the length of the air cavity decreases with an increase in the gas pressure, the wavelength becomes shorter. According to a previous report [29], the pressure sensitivity of a sensor is related to the diameter and thickness of the microbubble in that the pressure sensitivity of the sensor increases as the diameter of the microbubble increases. Compared with their study, in our study the pressure sensitivity of the sensor with the thinnest sidewall thickness and the shortest microbubble axis was the highest. Therefore, the pressure sensitivity of the sensor can be inferred to increase as the thickness of the sidewall of the microbubble decreases.

 figure: Fig. 7.

Fig. 7. Wavelength shifts of the three sensors with respect to the pressure at room temperature.

Download Full Size | PDF

A detailed analysis of the performance of Sample 2 was warranted. The test results of a one-round pressure rise and fall experiment of the two cavities of the sensor for the pressure range 0 − 4 MPa are shown in Fig. 8(a). The pressure sensitivity of the air cavity was −4.776 nm/MPa and shows a good linear relationship. As expected, the relative shift of wavelength for the silica cavity was hardly dependent on the pressure change, and the maximum wavelength shift was less than −0.615 nm, as shown in Fig. 8(b).

 figure: Fig. 8.

Fig. 8. Wavelength shifts for the (a) air cavity and (b) silica cavity of sample 2 in the range 0 − 4 MPa at 20 °C, respectively; wavelength shifts for the (c) air cavity and (d) silica cavity of sample 2 from 20 − 700 °C in atmospheric pressure, respectively.

Download Full Size | PDF

The high-temperature performance of the sensor is shown in Fig. 8(c) and 8(d). The sensor was tested in the temperature range of 20 − 700 °C in steps of 50 °C, and the temperature at each measurement point was maintained for 10 min to allow the temperature in the ceramic tube to stabilize. In terms of the air cavity, the change in the refractive index at high temperatures can be ignored, and thus thermal expansion plays a major role in the temperature sensitivity. For the silica cavity, the temperature sensitivity is mainly attributable to the change in the thermally induced refractive index at the thick silica end [34]. Because the thermal expansion coefficient of the SMF is slightly larger than that of the HST, the fiber extends beyond the HST when the temperature increases, resulting in a shorter air cavity length [29]. This explains the wavelength shift to shorter wavelengths. For the air cavity, the linearity of the wavelength shift of the reflection spectrum versus temperature is poor, and the process of increasing and decreasing the temperature is characterized by obvious hysteresis, as shown in Fig. 8(c). Considering the small difference in the thermal expansion coefficient between the SMF and HST, the wavelength shift corresponding to the air cavity should not be greater than 6 nm. Moreover, the temperature response of the air cavity in the range from 20 − 700 °C has poor linearity and obvious hysteresis. This phenomenon is similar to the high-temperature annealing behavior of long-period gratings in pure silica microstructure fibers which was previously reported [35]. We consider this to have been caused by the residual stress in the welding of the SMF and HST during the fast heating and cooling steps of the laser fusion process [36]. Figure 8(d) shows that the reflection spectrum of the silica cavity shifted to longer wavelengths owing to the thermo-optic effect of silica at high temperatures. The temperature sensitivity of the silica cavity was 12.715 pm/°C, and the wavelength shift had a good linear relationship with the temperature.

Figure 9 shows the temperature response of the air cavity when the sensor was heated to 900 °C and maintained at this temperature for 3 hours, after which the temperature was gradually returned to room temperature. The temperature response of the air cavity of the sensor was significantly improved. The wavelength exhibits a very slight shift of less than −0.550 nm, as shown in Fig. 9. Therefore, the annealing experiment at 900 °C is considered to eliminate the residual stress at the fusion point, which largely cancels the hysteresis and the excessive wavelength shift of the air cavity.

 figure: Fig. 9.

Fig. 9. Temperature response of air cavity after annealing at 900 °C.

Download Full Size | PDF

The stability of the sensor was assessed by carrying out three rounds of pressure tests by increasing the temperature from 20 − 700 °C in the steps of 100 °C. As shown in Fig. 10, three indicators were used to measure the working stability of the sensor: the repeatability error, hysteresis, and linear error [37]. The sensor operates stably in the range 20 − 600 °C with small repeatability and linear errors, and minimal hysteresis (≤ 4%). The sensor maintains good linearity at 700 °C, but the hysteresis and repeatability worsen (> 4%), as shown in Fig. 10(h). Moreover, the pressure sensitivity of the sensor decreases with increasing temperature, as shown in Fig. 10(i). This is due to the temperature dependence of the Young's modulus of the HST. The Young's modulus of silica increases as the temperature increases, such that the microbubble is difficult to deform at high temperatures, resulting in a decrease in pressure sensitivity [38][39]. The decrease in the pressure sensitivity caused by the increased temperature would lead to an increase in the testing error, which would limit the measurement range [25].

 figure: Fig. 10.

Fig. 10. Results of three rounds of pressure tests of sample 2 in the range (a)−(h) 20 − 700 °C in increments of 100 °C; (i) pressure sensitivities of sample 2 in the range 20 − 700 °C.

Download Full Size | PDF

The working performance of the sensor after exposure to high temperatures was verified by conducting three rounds of pressure tests on the three sensors at 20 °C. As shown in Fig. 11, the sensors operate stably and normally, and the pressure sensitivities of samples 1 and 2 are almost equal to those before high-temperature treatment. Therefore, the sensors can be inferred to continue to exhibit good working performance at high temperatures.

 figure: Fig. 11.

Fig. 11. Results of three rounds of pressure tests of the three samples in the range 0 − 4 MPa at room temperature after high-temperature treatment at 700 °C.

Download Full Size | PDF

The experimental results presented above indicate that the sensor developed in this study can be used for simultaneous pressure and temperature measurements. The temperature-pressure cross-sensitivity is an important indicator of the pressure characteristics of the sensor at high temperatures, and the cross-sensitivity of the two cavities of the sensor can be inferred from the following matrix [40]:

$$\left[ {\begin{array}{{c}} {\Delta P}\\ {\Delta T} \end{array}} \right] = \frac{1}{Q}\left[ {\begin{array}{{cc}} {{K_{2T}}}&{ - {K_{1T}}}\\ { - {K_{2P}}}&{{K_{1P}}} \end{array}} \right]\left[ {\begin{array}{{c}} {\Delta {\lambda_1}}\\ {\Delta {\lambda_2}} \end{array}} \right].$$
where K1T and K1P are the temperature and pressure sensitivities of the air cavity, respectively, K2T and K2P are the temperature and pressure sensitivities of the silica cavity, respectively, Δλ1 and Δλ2 are the wavelength shifts corresponding to the air and the silica-cavities, respectively, and ΔP and ΔT are the changes in the pressure and temperature, respectively, in the environment where the sensor is located. Q is the determinant of the matrix, and is expressed as $Q = {K_{2T}}{K_{1P}} - {K_{1T}}{K_{2P}}$.

Because the air and silica cavities are insensitive to changes in the temperature and pressure, the temperature sensitivity of the air cavity and the pressure sensitivity of the silica cavity, K1T and K2P, respectively, both approximated 0. Taking the experimental results of the sensor at 600 °C as an example, the cross-sensitivity matrix of the sensor can be expressed as:

$$\left[ {\begin{array}{{c}} {\Delta P}\\ {\Delta T} \end{array}} \right] = \frac{1}{{0.051}}\left[ {\begin{array}{{cc}} {0.013}&0\\ 0&{ - 4.045} \end{array}} \right]\left[ {\begin{array}{{c}} {\Delta {\lambda_1}}\\ {\Delta {\lambda_2}} \end{array}} \right].$$

According to Eq. (8), the temperature-pressure cross-sensitivities of the air and silica cavities are almost 0, which ensures the accuracy of the pressure and temperature measurements of the sensor at high temperatures and pressures. The accuracy of the sensor could be further improved by increasing the pressure sensitivity of the sensor by further thinning the sidewall of the microbubble or by increasing the distance between the fusion point and the microbubble to reduce the cross-influence of the temperature and pressure of the sensor.

4. Conclusion

We proposed and experimentally demonstrated a novel FPI sensor. The sensor has a microbubble structure with a thin sidewall and thick end face for high-pressure and high-temperature monitoring. The experimental results showed that the pressure sensitivity of the sensor can reach −5.083 nm/MPa, and the temperature sensitivity can attain 12.715 pm/°C with a very low temperature-pressure cross-sensitivity. Compared with previous microbubble-based FPIs, the temperature sensitivity of the proposed device was greatly improved, and the pressure sensitivity was relatively high. Additionally, the mechanical strength of the proposed FPI sensor is enhanced owing to its compact size and the relatively thick sidewall of the microbubble, which makes subsequent packaging convenient and suitable for higher-pressure environments.

Funding

National Natural Science Foundation of China (61735009, 61975108, 62005153).

Disclosures

The authors declare no conflicts of interest.

Data Availability

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.

References

1. M. Q. Chen, H. M. Wei, Y. Zhao, X. H. Lei, and S. Krishnaswamy, “Temperature insensitive air-cavity Fabry-Perot gas pressure sensor based on core-offset fusion of hollow-core fibers,” Sens. Actuator A 298, 111589 (2019). [CrossRef]  

2. J. Yi, “Sapphire Fabry–Perot pressure sensor at high temperature,” IEEE Sens. J. 21(2), 1596–1602 (2021). [CrossRef]  

3. Q. Zhang, J. C. Lei, Y. Z. Chen, Y. J. Wu, C. Chen, and H. Xiao, “3D printing of all-glass fiber-optic pressure sensor for high temperature applications,” IEEE Sens. J. 19(23), 11242–11246 (2019). [CrossRef]  

4. W. H. Wang, W. N. Wei, S. X. Wu, Y. Q. Li, C. Y. Huang, X. Y. Tian, X. X. Fei, and J. Huang, “Adhesive-free bonding homogenous fused-silica Fabry–Perot optical fiber low pressure sensor in harsh environments by CO2 laser welding,” Opt. Commun. 435, 97–101 (2019). [CrossRef]  

5. Y. G. Liu, T. Zhang, Y. X. Wang, D. Q. Yang, X. Liu, H. Y. Fu, and Z. A. Jia, “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. 46, 77–82 (2018). [CrossRef]  

6. S. Pevec and D. Donlagic, “Miniature all-fiber Fabry–Perot sensor for simultaneous measurement of pressure and temperature,” Appl. Opt. 51(19), 4536–4541 (2012). [CrossRef]  

7. H. H. Yu, Y. Wang, J. Ma, Z. Zheng, Z. Z. Luo, and Y. Zheng, “Fabry-Perot interferometric high-temperature sensing up to 1200 °C based on a silica glass photonic crystal fiber,” Sensors 18(1), 273 (2018). [CrossRef]  

8. C. L. Zhao, L. Y. Hou, J. Kang, B. N. Mao, C. Y. Shen, and S. Z. Jin, “High-sensitivity hydraulic pressure sensor based on Fabry-Perot interferometer filled with polydimethylsiloxane film,” Rev. Sci. Instrum. 90(9), 095002 (2019). [CrossRef]  

9. A. M. Shrivastav, D. S. Gunawardena, Z. Y. Liu, and H. Tam, “Micro structured optical fiber based Fabry–Pérot interferometer as a humidity sensor utilizing chitosan polymeric matrix for breath monitoring,” Sci. Rep. 10(1), 6002 (2020). [CrossRef]  

10. L. Cai, J. Wang, M. Q. Chen, and X. Ai, “A high-sensitivity strain sensor based on an unsymmetrical air-microbubble Fabry-Pérot interferometer with an ultrathin wall,” Measurement 181, 109651 (2021). [CrossRef]  

11. X. Wang, S. Wang, J. F. Jiang, K. Liu, X. Z. Zhang, M. N. Xiao, H. Xiao, and T. G. Liu, “Non-destructive residual pressure self-measurement method for the sensing chip of optical Fabry-Perot pressure sensor,” Opt. Express 25(25), 31937–31947 (2017). [CrossRef]  

12. S. H. Aref, H. Latifi, M. I. Zibaii, and M. Afshari, “Fiber optic Fabry–Perot pressure sensor with low sensitivity to temperature changes for downhole application,” Opt. Commun. 269(2), 322–330 (2007). [CrossRef]  

13. J. F. Jiang, S. Wang, K. Liu, X. Z. Zhang, J. D. Yin, F. Wu, and T. G. Liu, “Development of optical fiber temperature sensor for aviation industry,” in 15th International Conference on Optical Communications and Networks (ICOCN), 2016, pp. 1–3.

14. X. L. Zhou, Q. X. Yu, and W. Peng, “Fiber-optic Fabry–Perot pressure sensor for down-hole application,” Optics and Lasers in Engineering 121, 289–299 (2019). [CrossRef]  

15. S. Ghildiyal, R. Balasubramaniam, and J. John, “Diamond turned micro machined metal diaphragm based Fabry Perot pressure sensor,” Opt. Laser Technol. 128, 106243 (2020). [CrossRef]  

16. X. Guo, J. C. Zhou, C. Du, and X. W. Wang, “Highly sensitive miniature all-silica fiber tip Fabry–Pérot pressure sensor,” IEEE Photonics Technol. Lett. 31(9), 689–692 (2019). [CrossRef]  

17. J. Wang, L. Li, S. C. Liu, D. Y. Wu, W. Wang, M. Song, G. J. Wang, and M. X. Huang, “Investigation of composite structure with dual Fabry–Perot cavities for temperature and pressure sensing,” Photonics 8(5), 138 (2021). [CrossRef]  

18. J. Liu, P. G. Jia, H. X. Zhang, X. D. Tian, H. Liang, Y. P. Hong, T. Liang, W. Y. Liu, and J. J. Xiong, “Fiber-optic Fabry–Perot pressure sensor based on low-temperature co-fired ceramic technology for high-temperature applications,” Appl. Opt. 57(15), 4211–4215 (2018). [CrossRef]  

19. L. Zhao, Y. D. Zhang, Y. H. Chen, and J. F. Wang, “Composite cavity fiber tip Fabry-Perot interferometer for high temperature sensing,” Opt. Fiber Technol. 50, 31–35 (2019). [CrossRef]  

20. Y. Li, B. Dong, E. Q. Chen, X. L. Wang, and Y. D. Zhao, “Heart-rate monitoring with an ethyl alpha-cyanoacrylate based fiber Fabry-Perot sensor,” IEEE J. Sel. Top. Quantum Electron. 27(4), 1–6 (2021). [CrossRef]  

21. H. Li, Q. C. Zhao, S. D. Jiang, J. S. Ni, and C. Wang, “F-P cavity and FBG cascaded optical fiber temperature and pressure sensor,” Chin. Opt. Lett. 17(4), 040603 (2019). [CrossRef]  

22. G. J. Wang, X. L. Liu, Z. G. Gui, Y. Q. An, J. Y. Gu, M. Q. Zhang, L. Yan, G. Wang, and Z. B. Wang, “A high-sensitive pressure sensor using a single-mode fiber embedded microbubble with thin film characteristics,” Sensors 17(6), 1192 (2017). [CrossRef]  

23. W. F. Guo, J. X. Liu, J. R. Liu, G. Wang, G. J. Wang, and M. X. Huang, “A single-ended ultra-thin spherical microbubble based on the improved critical-state pressure-assisted arc discharge method,” Coatings 9(2), 144 (2019). [CrossRef]  

24. T. T. Gang, M. L. Hu, Q. Z. Rong, X. G. Qiao, L. Liang, N. Liu, R. X. Tong, X. B. Liu, and C. Bian, “High-frequency fiber-optic ultrasonic sensor using air micro-bubble for imaging of seismic physical models,” Sensors 16(12), 2125 (2016). [CrossRef]  

25. J. T. Chen, H. M. Wei, L. Zhang, Z. F. Wang, G. L. Zhang, Z. Y. Chen, T. Y. Wang, and F. F. Pang, “Influence of temperature on all-silica Fabry-Pérot pressure sensor,” IEEE Photonics J. 13(3), 1–9 (2021). [CrossRef]  

26. J. Ma, J. Ju, L. Jin, W. Jin, and D. N. Wang, “Fiber-tip micro-cavity for temperature and transverse load sensing,” Opt. Express 19(13), 12418–12426 (2011). [CrossRef]  

27. J. Ma, J. Ju, L. Jin, and W. Jin, “A compact fiber-tip micro-cavity sensor for high-pressure measurement,” IEEE Photonics Technol. Lett. 23(21), 1561–1563 (2011). [CrossRef]  

28. C. R. Liao, S. Liu, L. Xu, C. Wang, Y. P. Wang, Z. Y. Li, Q. Wang, and D. N. Wang, “Sub-micron silica diaphragm-based fiber-tip Fabry–Perot interferometer for pressure measurement,” Opt. Lett. 39(10), 2827–2830 (2014). [CrossRef]  

29. Z. Li, P. G. Jia, G. C. Fang, H. Liang, T. Liang, W. Y. Liu, and J. J. Xiong, “Microbubble-based fiber-optic Fabry–Perot pressure sensor for high-temperature application,” Appl. Opt. 57(8), 1738–1743 (2018). [CrossRef]  

30. Y. G. Liu, X. Liu, T. Zhang, and W. Zhang, “Integrated FPI-FBG composite all-fiber sensor for simultaneous measurement of liquid refractive index and temperature,” Optics and Lasers in Engineering 111, 167–171 (2018). [CrossRef]  

31. S. H. Zhang, Y. F. Shen, and H. J. Qiu, “The technology and welding joint properties of hybrid laser-tig welding on thick plate,” Opt. Laser Technol. 48, 381–388 (2013). [CrossRef]  

32. J. M. Sun, X. Z. Liu, Y. G. Tong, and D. Deng, “A comparative study on welding temperature fields, residual stress distributions and deformations induced by laser beam welding and CO2 gas arc welding,” Mater. Des. 63, 519–530 (2014). [CrossRef]  

33. T. Jokinen and V. Kujanpää, “High power Nd: YAG laser welding in manufacturing of vacuum vessel of fusion reactor,” Fusion Eng. Des. 69(1–4), 349–353 (2003). [CrossRef]  

34. H. Y. Choi, K. S. Park, S. J. Park, U. Paek, B. H. Lee, and E. S. Choi, “Miniature fiber-optic high temperature sensor based on a hybrid structured Fabry–Perot interferometer,” Opt. Lett. 33(21), 2455–2457 (2008). [CrossRef]  

35. Y. J. Rao, D. W. Duan, Y. E. Fan, T. Ke, and M. Xu, “High-temperature annealing behaviors of CO2 laser pulse-induced long-period fiber grating in a photonic crystal fiber,” J. Lightwave Technol. 28(10), 1530–1535 (2010). [CrossRef]  

36. D. W. Duan, Y. J. Rao, W. P. Wen, J. Yao, D. Wu, L. C. Xu, and T. Zhu, “In-line all-fiber Fabry-Pérot interferometer high temperature sensor formed by large lateral offset splicing,” Electron. Lett. 47(6), 401–403 (2011). [CrossRef]  

37. “Verification regulation of pressure transducer (static),” JJG, 860 (2015).

38. I. Štubňa, A. Trník, and L. Vozár, “Thermomechanical analysis of quartz porcelain in temperature cycles,” Ceram. Int. 33(7), 1287–1291 (2007). [CrossRef]  

39. A. Hunger, G. Carl, and C. Rüssel, “Formation of nano-crystalline quartz crystals from ZnO/MgO/Al2O3/TiO2/ZrO2/SiO2 glasses,” Solid State Sci. 12(9), 1570–1574 (2010). [CrossRef]  

40. N. N. Dong, S. M. Wang, L. Jiang, Y. Jiang, P. Wang, and L. C. Zhang, “Pressure and temperature sensor based on graphene diaphragm and fiber Bragg gratings,” IEEE Photonics Technol. Lett. 30(5), 431–434 (2018). [CrossRef]  

Data Availability

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.

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (11)

Fig. 1.
Fig. 1. Diagram of the (a) structure of the sensor and (b) deformation of the microbubble at 4 MPa simulated by COMSOL software.
Fig. 2.
Fig. 2. Transmission path of the electric field in the FPI cavity.
Fig. 3.
Fig. 3. Fabrication procedure of the sensor: (a) end-cut HST in the CO2 laser; (b) fusing one end of the HST; (c) attaching the gas-filled capillary; (d) adjusting the position to fuse the sidewall of microbubble; (e) fusing the SMF with the HST; (f) complete structure of the sensor.
Fig. 4.
Fig. 4. (a) Microscope images of the three fabricated devices and (b) corresponding reflection spectra of the three fabricated samples, respectively.
Fig. 5.
Fig. 5. (a) Reflection spectrum of sample 2; (b) FFT of reflection spectrum of sample 2; frequency component of the (c) air cavity and (d) silica cavity.
Fig. 6.
Fig. 6. Schematic diagram of the system to measure the high-temperature-pressure performance of the FPI sensor.
Fig. 7.
Fig. 7. Wavelength shifts of the three sensors with respect to the pressure at room temperature.
Fig. 8.
Fig. 8. Wavelength shifts for the (a) air cavity and (b) silica cavity of sample 2 in the range 0 − 4 MPa at 20 °C, respectively; wavelength shifts for the (c) air cavity and (d) silica cavity of sample 2 from 20 − 700 °C in atmospheric pressure, respectively.
Fig. 9.
Fig. 9. Temperature response of air cavity after annealing at 900 °C.
Fig. 10.
Fig. 10. Results of three rounds of pressure tests of sample 2 in the range (a)−(h) 20 − 700 °C in increments of 100 °C; (i) pressure sensitivities of sample 2 in the range 20 − 700 °C.
Fig. 11.
Fig. 11. Results of three rounds of pressure tests of the three samples in the range 0 − 4 MPa at room temperature after high-temperature treatment at 700 °C.

Equations (8)

Equations on this page are rendered with MathJax. Learn more.

E r = E i [ R 1 + ( 1 A 1 ) ( 1 α 1 ) ( 1 R 1 ) R 2 e j 2 β L 1 + j π + ( 1 A 1 ) ( 1 A 2 ) ( 1 α 1 ) ( 1 α 2 ) ( 1 R 1 ) ( 1 R 2 ) R 3 e j β ( 2 L 1 + 2 L 2 ) ] .
R F P ( λ ) = | E r E i | 2 = R 1 + ( 1 A 1 ) 2 ( 1 α 1 ) 2 ( 1 R 1 ) 2 R 2 + ( 1 A 1 ) 2 ( 1 A 2 ) 2 ( 1 α 1 ) 2 ( 1 α 2 ) 2 ( 1 R 1 ) 2 ( 1 R 2 ) 2 R 3 2 R 1 R 2 ( 1 A 1 ) ( 1 α 1 ) ( 1 R 1 ) cos ( 4 π n a L 1 / λ ) + 2 R 1 R 3 ( 1 A 1 ) ( 1 A 2 ) ( 1 α 1 ) ( 1 α 2 ) ( 1 R 1 ) ( 1 R 2 ) cos [ 4 π ( n a L 1 + n s L 2 ) / λ ] 2 R 2 R 3 ( 1 A 1 ) 2 ( 1 α 1 ) 2 ( 1 R 1 ) 2 ( 1 A 2 ) ( 1 α 2 ) ( 1 R 2 ) cos [ 4 π ( 2 n a L 1 + n s L 2 ) / λ ] .
F S R = λ 2 2 n d .
Δ O P D P = 2 n a ( L 1 + Δ L ) 2 n a L 1 = 2 n a Δ L .
Δ λ P = Δ O P D P O P D P λ = n a Δ L n s L 2 + n a L 1 λ .
Δ λ T = Δ O P D T O P D T λ = ( Δ L L + Δ n n ) λ = ( β + ) λ Δ T .
[ Δ P Δ T ] = 1 Q [ K 2 T K 1 T K 2 P K 1 P ] [ Δ λ 1 Δ λ 2 ] .
[ Δ P Δ T ] = 1 0.051 [ 0.013 0 0 4.045 ] [ Δ λ 1 Δ λ 2 ] .
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.