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

Wideband low confinement loss anti-resonant hollow core fiber with nested U-shape tube

Open Access Open Access

Abstract

We propose a nested U-shape tube anti-resonant hollow core fiber (UARF) that can effectively reduce the confinement loss (CL) as well as the loss oscillations. The key parameters of UARF have been optimized via numerical analysis. Simulation results show that the CL of proposed UARF is lower than 0.01dB/km over a 550 nm operational bandwidth range from 1.3 µm to 1.85 µm. This CL is nearly one order of magnitude lower than the nested anti-resonant nodeless fiber (NANF). Moreover, the loss ratio between higher-order modes to the fundamental mode is verified to be more than 100,000 over a ultrawide bandwidth of 1000 nm, which indicates its excellent single mode performance. The tolerance towards the structure deformation of UARF has been evaluated for the purpose of practical fiber fabrication. Thus, the proposed UARF has potential application in large capacity data transmission, nonlinear optics, gas sensing and so on.

© 2021 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Hollow core fibers (HCFs) by guiding the light over the air core have been extensively investigated, due to the advantages of small transmission delay, less Rayleigh scattering, and small nonlinear coefficient [18]. Meanwhile, the applications of HCF have been comprehensively explored for high peak power pulse delivery [9,10], gas sensing and nonlinear optics [1113], high energy pulse compression [14], mid-IR radiation delivery [15], high harmonic generation [16]. Besides, the intrinsic features of HCF enables its application in large capacity data transmission with the wavelength ranging from short band to long wavelength band (such as L-band to 2 micro meter) [1720]. According to the light guidance mechanism, HCFs can be divided into two categories [21]. One category is photonic bandgap fiber (PBGF) based on the photonic bandgap effect [22], where the light at specific frequency can be guided over the hollow core. PBGF is promising for the HCF related applications, but faces great challenges in terms of operational bandwidth and fabrication technique [21]. The other category is anti-resonant fiber (ARF) relying on a combination of antiresonance and the inhibited coupling effect, to obtain a reduction of transmission attenuation and an enhancement of operational bandwidth [21,2326].

Generally, reducing the transmission loss while increasing the operational bandwidth at the same time is quite challenging. The total loss of a HCF include surface scattering loss (SSL), the leakage loss (namely, the confinement loss, CL) and the bending loss (BL). Among the three losses, the SSL and CL are the most key concerns in current theoretical researches [21]. As for the PBGF, the main loss comes from the SSL [27], which is difficult to overcome due to the frozen effect arising in thermodynamic fluctuations of surface capillary waves [28]. The lowest loss of PBGF was experimentally reported of 1.72 dB/km [29]. Alternatively, as for the ARF, the loss is mainly referred to the CL determined by the power leakage from the core [21]. Meanwhile, the SSL of ARF is typically lower than that of PBGF because of the anti-resonance feature. Therefore, ARF is a potential candidate to simultaneously achieve both ultra-low loss and wide operational bandwidth [6,23]. In recent years, some pioneering work on the ARF design has been carried out. In 2017, the loss of 7.7dB/km for the single-ring ARF was characterized [30]. To further reduce the CL, the most promising way is to change the ARF structure with additional anti-resonant layers. In 2018, Gao experimentally reported a conjoined-tube based ARF including two D-shaped air holes, resulting to a minimum loss of 2 dB/km at 1512 nm [31]. On the other hand, the nested anti-resonant nodeless fiber (NANF) attracted great attentions with its loss reduction from 1.3 dB/km to 0.28 dB/km, after a joint optimization of both the structural parameters and the fabrication process [7,3234]. According to the light [34]. Furthermore, there also exist lots of theoretical investigations about novel ARF structures, for the purpose of further CL reduction [3539]. According to the light [39]. For example, double negative curvature anti-resonant fibers (D-ARF) were proposed with the CL lower than 0.1 dB/km over a wavelength range of 320 nm (from 1.28 µm to 1.85 µm) [39] or with a loss ratio of 100,000 between higher-order modes and the fundamental mode [37]. However, the additional use of glass structures will inevitably create more glass connections and thus enhance the Fano resonance significantly, resulting in a non-flat loss spectrum [21,40,41]. Due to the influence of glass nodes, the operational bandwidth such as D-ARF is narrower in comparison with single-ring ARF or NANF. Moreover, the fluctuation of the loss spectrum of ARF induced by the Fano resonance is more significant in the long wavelength range [40], which causes the difficulty to potential applications such as the new wavelength band optical communications. Therefore, it is desired to design a kind of fiber structure to enhance the operational wavelength range with ultra-low loss and low loss oscillations.

In this paper, we propose a novel nested U-shape tube structure of ARF (UARF) to reduce the CL and decrease the loss fluctuation induced by the Fano resonance, as well as enlarge the low CL operational bandwidth. After parameters optimization, we put forward a theoretical model of 5-tube UARF and realize an operational bandwidth over 550 nm span from 1.3 µm to 1.85 µm, with the CL lower than 0.01 dB/km. Meanwhile, the loss fluctuation between maximum and minimum CLs is less than 0.03 dB/km over the wavelength range from 1.2 µm to 2.2 µm. Our proposed UARF guarantees a significant performance improvement, in terms of both the CL and operational bandwidth, in comparison with existing ARF. Additionally, the loss ratio between higher-order modes and the fundamental mode is more than 100,000 for the proposed UARF. Those results indicate that the proposed UARF is a promising candidate for the HCF involved applications.

2. Nested U-shaped HCF structure

The cross-section of the proposed UARF is shown in Fig. 1. The cladding includes a unit of circular tube and a U-shape glass structure, which can be considered as a circular tube nested with a U-shape structure. From a geometrical view, the nested U-shape structure can be regarded as a combination of one semicircular layer and two straight layers, which are parallel to each other and tangent to the semicircular layer. The geometrical parameters of the UARF are denoted as the core radius Rcore, the tube gap g, the glass thickness t, the tube radius R, the radial separation z between the tube and the U-shape structure, the semicircular radius r, and the distance d between parallel layers of U-shape structure. Please note that, d is equal to 2r, and R is not equal to z + r in our design.

 figure: Fig. 1.

Fig. 1. Cross-section of proposed nested U-shape ARF (UARF) with geometrical parameters.

Download Full Size | PDF

It has been verified that the CL of ARF is related to the power leakage from air core to the cladding [23]. Changing the cladding structure of ARF with additional anti-resonant layer is an effective solution to reduce the power leakage in tubes. In addition, we identify that, the surrounding power leakage can be effectively suppressed, when anti-resonant layers are close to each other even though they are along the radial direction of the fiber. As shown in Fig. 1, the parallel glass layers of U-shape structure form a gap while the connection nodes are sufficiently far away from the fiber core [21], which is the key point to reduce both the CL and the Fano resonance induced loss fluctuation simultaneously.

In order to verify the feasibility of reducing the CL with the U-shape structure, the variation of power density of cladding boundary (partially from −30° to 30°) of NANF and UARF has been, respectively, shown in Fig. 2. The power density can reveal the power leakage of different fiber structures. The core diameter, the glass thickness, and the tube gap of two fibers are the same, and we also utilize the same number of anti-resonant layers for those two fibers for the purpose of a fair comparison. The insets of (i) and (ii) show the cross-section of corresponding fiber structures (red-mark is the partially calculated boundary), respectively. We can observe that, the power density of proposed UARF is generally lower than that of NANF, and the maximum power density of UARF is found less than 10−8 W/m2. Therefore, we can conclude that the U-shape structure is effective to reduce the CL in comparison with the nested structure.

 figure: Fig. 2.

Fig. 2. Calculated power density at the cladding (red-mark of structure cross-section) of NANF and the proposed UARF. The insets show the fiber structures and profiles of fundamental mode. For those two fibers, the core radius, the glass thickness, the tube gap, and the tube radius are chosen the same as 20 µm, 0.5 µm, 3.3 µm, and 16.2 µm, respectively. For the NANF, the radial separation is optimized to 15.45 µm with the radius of the small nested tube of 8.1 µm. For our proposed fiber, the corresponding one is optimized to 7 µm with the radius of the small semicircular of 5.83 µm.

Download Full Size | PDF

Then numerical investigations have been carried out by using the COMSOL Multiphysics based on finite element method (FEM). In order to obtain accurate simulation results, we set the maximum element size of glass and air regions to λ/6 and λ/4, respectively [21]. The minimum element number of glass regions is 5. In addition, we set a small penetration of t/2 of all cladding tubes into the jacket layer and smaller nested tubes into the outer tubes based on the practical consideration [38,42]. A perfectly matched layer (PML) is placed outside the jacket layer to enclose the simulation area, to obtain the imaginary part of the mode eigenvalue to calculate the CL of fiber [35]. The thickness of the jacket layer and PML are set as 3 µm and 10 µm respectively. The CL can be calculated [37]

$$CL = {{40\pi \times {n_{imag}}} / {[{\ln (10) \times \lambda } ]}},$$
where nimag is the imaginary part of effective refractive index, λ is the free space wavelength. The glass thickness t is determined by the anti-resonance condition [23,35]
$$t = {{\lambda \times (M - 0.5)} / {2 \times \sqrt {n_{glass}^2 - n_{air}^2} }},$$
where nglass and nair are refractive indexs of glass and air, respectively. M (=1,2,3…) represents the order of antiresonance, and M = 1 is chosen in our investigation for the purpose of the operational bandwidth enhancement. The SSL can be estimated after calibration [21,27]
$$SSL = \eta F{({{\lambda / {{\lambda_0}}}} )^{ - 3}},$$
where F is the power overlap between the fundamental mode and the silica boundary, and the calibration factor η is 300 when λ0 is 1.55 µm [21]. Besides, the random microbend loss (RML) caused by the coupling between fundamental mode and LP11 mode can be calculated [37,43]
$$RML = \beta _0^2C(\Delta {\beta _{01}})\left( {\left\langle {0|{{x^2}} |0} \right\rangle - {{\left|{\left\langle {0|x |1} \right\rangle } \right|}^2}} \right),$$
where β0 is the fundamental mode propagation constant and Δβ01 is the difference of fundamental mode propagation constant and LP11 mode group propagation constant. Cβ01) is the power spectral density at Δβ01, $\left\langle {0|x |0} \right\rangle $ and $\left\langle {0|x |1} \right\rangle $ are the spot radius of fundamental mode and LP11 mode respectively.

To have a better understanding of the loss contributions of the ARF, we have done the theoretical analysis as shown in Fig. 3. The basic parameters of proposed UARF and the referred NANF are kept the same, i.e., the core radius Rcore = 20 µm, the tube gap g = 3.3 µm, the glass thickness t = 0.5 µm, and the tube radius R = 24 µm. The other optimized parameters of NANF and UARF are, z = 10.08 µm, r = 10.32 µm and rN = 12 µm. The total loss performances of the UARF and NANF have been shown as the scatter points with red and yellow color respectively. And the SSL (green dot-dashed line) of UARF is almost equal to the one of NANF. As can be seen from the figure, the SSL of NANF is the dominant loss and slightly higher than CL in the shorter wavelength range (< 1.51 µm). For the longer wavelength range (> 1.51 µm), the CL becomes the dominant loss contribution. This increase of CL is originated from the power leakage that caused by the wavelength away from the anti-resonance condition. Therefore, the CL should be further reduced to meet the demands on low total loss large capacity optical transmissions toward to the long wavelength range (L-band to 2 µm). It is obviously that the proposed UARF has the reduced total loss (red scattered points) in the entire 1000 nm wavelength range with a comparison of that of NANF. Meanwhile, the optimized CL is nearly one order of magnitude lower than that of NANF, which resulted in the total loss of UARF being lower than 0.1 dB/km in a wide wavelength range (from 1.51 µm to 2.2 µm).

 figure: Fig. 3.

Fig. 3. Calculated the losses of UARF and NANF. These scattered points are total loss of corresponding fiber. The dot-dashed line is SSL of UARF, almost identical to that of NANF.

Download Full Size | PDF

3. Optimizations and discussions

3.1 Parameter optimization of nested U-shaped structure

According to the basic analysis above, and to investigate the CL performance for our proposed fiber structure, we first carry out numerical simulations to obtain the relationship between the semicircular radius r or the radial separation z and CL. We consider two cases for our proposed fiber with 5-tube and 6-tube, respectively, which are two typical arrangements for the ARF design. In our simulations, some parameters are fixed for all fibers, i.e., the core radius Rcore = 20 µm, the tube gap g = 3.3 µm, and the glass thickness t = 0.5 µm. Then the tube radius R = 24 µm for 5-tube and R = 16.2 µm for 6-tube can be derived from the geometric relationship.

The CL with respect to the variation of semicircular radius r and radial separation z are illustrated in Fig. 4(a) and (b), respectively, when the operation wavelength is 1.55 µm. Once we set the initial value of z = 0.42R in Fig. 4(a), we observe that the semicircular radius r has a great impact on the CL. Under the condition of 0.4 > r/R > 0.65, the CL of LP01 modes (solid lines) reduces dramatically as r decreases. When 0.25 < r/R < 0.4 is satisfied, the CL has an erratic fluctuation, which is inferred from the Fano resonance arising in the closed parallel layers [44]. It’s worth noting that, as for the 5-tube configuration, the CL of LP11 mode (red dashed line) reaches a peak value at r/R = 0.43, which is caused by the coupling of LP11 mode and the cladding tube mode. As for the effect of the radial separation z, as shown in Fig. 4(b), the CL of LP01 mode (solid lines) is independent on z/R, when z/R is at the range of 0.3 to 0.6. Additionally, we find that 5-tube case has lower CL of LP01 mode and higher CL of LP11 mode than that of 6-tube case. For the 5-tube UARF, we take two values of semicircular radius into account, relating to two specific fiber configurations in Fig. 4(a). Type I configuration has the highest CL of LP11 mode, under the condition of r = 0.43R (10.32 µm). Type II configuration has the lowest CL of LP01 mode when r = 0.33R (7.92 µm) is satisfied. Due to the independence of the radial separation on the CL of LP01 mode within the range, we choose z = 0.42R for Type I and Type II fiber.

 figure: Fig. 4.

Fig. 4. Calculated confinement loss variation with (a) semicircular radius r under the condition of z = 0.42R, and (b) radial separation z under the condition of r = 0.5R. The tube radius R = 24 µm for 5-tube (red lines) and R = 16.2 µm for 6-tube (blue lines), and wavelength at 1.55 µm.

Download Full Size | PDF

Figure 5 shows the loss spectra of proposed fibers (Type I and Type II), the NANF [21] and the D-ARF [37], for the ease of loss performance comparison. The insets of Fig. 5 are LP01 mode field distribution of (i) NANF, (ii) D-ARF, (iii) Type I of UARF and (iv) Type II of UARF. As shown in Fig. 5, Type I has a wide operational bandwidth with the CL lower than 0.01 dB/km from 1.3 µm to 1.85 µm and a minimum CL of 0.0056 dB/km at 1.55 µm, which is significantly better than that of NANF and D-ARF. Moreover, within the whole wavelength range from 1.2 µm to 2.2 µm, the difference of the loss spectra between maximum CL and minimum CL is lower than 0.03 dB/km, indicating a flat CL performance for our proposed UARF. For Type II UARF, the minimum CL of LP01 mode is even lower than 0.001 dB/km because of the small semicircular radius r. However, its loss spectrum is seriously fluctuated due to the Fano resonance. Therefore, we select Type I (R = 24 µm, r = 10.32 µm and z = 12 µm) as the proposed fiber design for subsequent discussions. Obviously, the UARF shows the lower CL with the same operational bandwidth comparable to NANF. It means that the proposed UARF has the lower transmission loss and wider operational bandwidth.

 figure: Fig. 5.

Fig. 5. Calculated loss spectra of the proposed fiber (Type I and Type II) and other fibers. The corresponding fiber structures and LP01 mode field distributions are shown at the right-side, the core radius, tube gap and glass thickness of all these fibers are same, while the other parameters of D-ARF are the same as [37].

Download Full Size | PDF

3.2 Higher-order mode extinction ratio

For the HCF-based data transmission, the single-mode operation is one of the important requirements. Such feature can be characterized by the loss ratio between the higher-order modes to the fundamental mode, which also called as higher order mode extinction ratio (HOMER) [37,38]. In order to obtain robust single-mode operation, we can suppress the higher-order mode for the ease of the HOMER improvement. Figure 6(a) shows the effect of semicircular radius r on effective refractive index of each mode. Since the core radius is fixed, the effective refractive index of LP01 mode and LP11 mode are almost constant. On the contrary, the effective refractive index of the cladding mode (CM) increases with r, due to the deformation of U-shaped air cavity. When r/R is closed to 0.43, the effective refractive index of the CM matches that of LP11 mode, corresponding to the CL peak of LP11 mode in Fig. 4(a). As shown in Fig. 6(b), we have also calculated the effective refractive index difference between the CM and two LP modes of proposed Type I fiber. The refractive index difference of LP11 mode and the CM is almost zero over a wide wavelength range, indicating that LP11 mode has been effectively suppressed. Based on the optimal parameters of U-shape structure, the CL and HOMER curves with respect to the operation wavelength are presented in Fig. 6(c). During the CL calculation, three modes are investigated. In Fig. 6(c), the loss performances of LP11 mode and LP21 mode are obtained to determine the lowest CL and the HOMER. We can observe that the proposed fiber has a HOMER more than 100,000 (blue dashed line) over the wavelength range from 1.2 µm to 2.2 µm. Compared with the D-ARF [37], our proposed UARF not only achieves a similar loss performance, but also ensures an excellent single-mode operation over much wider wavelength range. The insets of Fig. 6 are the field distributions of (i) LP21 mode, (ii) LP11 mode, and (iii) LP01 mode. Obviously, part of higher-order modes couples with CM.

 figure: Fig. 6.

Fig. 6. (a) The effective refractive index of LP01 mode (red), LP11 mode (green) and the CM (gray dashed) with a variation of semicircular radius r. (b)The difference of the CM and LP mode with the variation of wavelength. (c) Loss spectra of LP01 mode, LP11 mode and LP21 mode, as well as the HOMER (dot-dashed line in right axis), mode field distributions of LP01 mode, LP11 mode and LP21 mode are at the right-side. The parameters of proposed fiber (b) and (c) are R = 24 µm, r = 10.32 µm and z = 12 µm.

Download Full Size | PDF

Finally, the comparison of confinement loss, operational bandwidth and HOMER between our proposed fiber and other typical reports has been shown in Table 1. The proposed fiber has realized a confinement loss of <0.01 dB/km over 550 nm operational bandwidth and the HOMER of >100,000 over 1000 nm range, which shows a significant improvement than other designed fibers theoretically.

Tables Icon

Table 1. Comparison between our proposed fiber with other low loss single-mode fibers.

3.3 Fabrication tolerance

Although the proposed UARF has characteristics of low CL, wide operational bandwidth and single-mode operation, the tolerance of the fiber fabrication is also pivotal for practical applications. During the fabrication of practical ARFs, the geometrical deviations are unavoidable [45].

Initially, we consider that the semicircular of U-shape is deformed due to imperfect air pressure. Figure 7(a) shows the impact of fabrication deformation on the CL performance, when the semicircular of the proposed fiber is collapsed into the unexpected ellipse. We define the radial direction of the ellipse as the axis a, and half of the distance between parallel layer as the axis b. Assuming that the axis b of the ellipse and radial separation does not change, the extent of ellipse can be defined by a/b (ideally a/b = 1). It can be seen that the change of a/b has little impact on the CL. When a/b > 1.1, the CL even slightly decreases with the increment of a/b. Therefore, the high tolerance of a/b variation from 0.6 to 1.4 is anticipated.

 figure: Fig. 7.

Fig. 7. (a) Calculated confinement loss with the variation of semicircular deformed into an ellipse that defined by a/b, and the insets shows a/b = 0.6 and a/b = 1.4, (b) the loss spectra with a variation of parallel layer deflection θ.

Download Full Size | PDF

Next, we consider that the structure of parallel layer cannot be strictly maintained. Figure 7(b) shows the influence on the loss spectrum when the parallel layer of U-shape structure is varied. We set θ as the angle that layers deflect from the horizontal (ideally θ = 0°), and assume that all structures are deflected at the same angle for the ease of discussion. As shown in Fig. 7(b), the variation of θ has significant impact on the loss spectrum, because the gap of U-shape structure cannot be effectively maintained. When θ = 15° is set, although the CL level of proposed fiber is obviously higher than that of θ = 0°, it is still lower than that of NANF (gray dashed line). Meanwhile, based on the simulation results, the CL performance will not be degraded when θ is a negative value. Thus, the tolerance towards the deflection angle θ for our proposed URAF can be reached at least 15° under the desired CL target.

4. Conclusion

We have proposed a nested U-shape tube anti-resonant HCF to mitigate the influence of the Fano resonance on the confinement loss and then to realize low loss ultrawide operational wavelength. After parameters optimization, the confinement loss of proposed HCF is lower than 0.01dB/km over 550nm bandwidth ranging from 1.3 µm to 1.85 µm. Moreover, more than 100,000 of the higher-order mode extinction ratio is obtained over a wide wavelength range from 1.2 µm to 2.2 µm, which shows the excellent wideband single-mode performance. The fabrication tolerance of proposed fiber is also investigated by taking the deformation of U-shape into account. We identify that the confinement loss is insensitive to the semicircular of U-shape but more sensitive to the parallel layer deflection angle. From a viewpoint of practical fabrication, it is recommended to keep the deflection angle less than 15° to maintain the performance of low loss ultrawide operational bandwidth. The results show that the proposed UARF has the potential applications in data transmission, nonlinear optics, gas sensing etc.

Funding

National Key Research and Development Program of China (2018YFB1801001); National Natural Science Foundation of China (62022029); Guangdong Introducing Innovative and Entrepreneurial Teams of “The Pearl River Talent Recruitment Program” (2019ZT08X340); Special Project for Research and Development in Key areas of Guangdong Province (2018B010114002).

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. J. C. Knight, J. Broeng, T. A. Birks, and P. S. J. Russell, “Photonic Band Gap Guidance in Optical Fibers,” Science 282(5393), 1476–1478 (1998). [CrossRef]  

2. W. Belardi and J. C. Knight, “Hollow antiresonant fibers with reduced attenuation,” Opt. Lett. 39(7), 1853–1856 (2014). [CrossRef]  

3. P. Uebel, M. C. Gunendi, M. H. Frosz, G. Ahmed, N. N. Edavalath, J. M. Menard, and P. S. J. Russell, “Broadband robustly single-mode hollow-core PCF by resonant filtering of higher-order modes,” Opt. Lett. 41(9), 1961–1964 (2016). [CrossRef]  

4. D. K. Wu, F. Yu, Y. Y. Liu, and M. S. Liao, “Dependence of Waveguide Properties of Anti-Resonant Hollow-Core Fiber on Refractive Index of Cladding Material,” J. Lightwave Technol. 37(21), 5593–5699 (2019). [CrossRef]  

5. D. K. Wu, F. Yu, and M. S. Liao, “Understanding the material loss of anti-resonant hollow-core fibers,” Opt. Express 28(8), 11840–11851 (2020). [CrossRef]  

6. W. Ding, Y. Y. Wang, S. F. Gao, M. L. Wang, and P. Wang, “Recent progress in low-loss hollow-core anti-resonant fibers and their applications,” IEEE J. Sel. Top. Quantum Electron. 26(4), 1–12 (2020). [CrossRef]  

7. H. Sakr, Y. Chen, G. T. Jasion, T. D. Bradley, J. R. Hayes, H. C. H. Mulvad, I. A. Davidson, E. N. Fokoua, and F. Poletti, “Hollow core optical fibres with comparable attenuation to silica fibres between 600 and 1100 nm,” Nat. Commun. 11(1), 6030 (2020). [CrossRef]  

8. V. Michaud-Belleau, E. N. Fokoua, T. D. Bradley, J. R. Hayes, Y. Chen, F. Poletti, D. J. Richardson, J. Genest, and R. Slavík, “Backscattering in antiresonant hollow-core fibers: over 40 dB lower than in standard optical fibers,” Optica 8(2), 216–219 (2021). [CrossRef]  

9. G. Humbert, J. Knight, G. Bouwmans, P. Russell, D. Williams, P. Roberts, and B. Mangan, “Hollow core photonic crystal fibers for beam delivery,” Opt. Express 12(8), 1477–1484 (2004). [CrossRef]  

10. X. Y. Zhu, D. K. Wu, Y. Z. Wang, F. Yu, Q. R. Li, Y. F. Qi, J. Knight, S. F. Chen, and L. L. Hu, “Delivery of CW laser power up to 300 watts at 1080 nm by an uncooled low-loss anti-resonant hollow-core fiber,” Opt. Express 29(2), 1492–1501 (2021). [CrossRef]  

11. P. S. J. Russell, P. Hölzer, W. Chang, A. Abdolvand, and J. C. Travers, “Hollow-core photonic crystal fibres for gas-based nonlinear optics,” Nat. Photonics 8(4), 278–286 (2014). [CrossRef]  

12. C. Mao, B. Huang, Y. Wang, Y. J. Huang, L. F. Zhang, Y. Shao, and Y. P. Wang, “High-sensitivity gas pressure sensor based on hollow-core photonic bandgap fiber Mach-Zehnder interferometer,” Opt. Express 26(23), 30108–30115 (2018). [CrossRef]  

13. W. Liu, Y. Zheng, Z. Wang, Z. X. Wang, J. Yang, M. X. Chen, M. Qi, S. U. Rehman, P. P. Shum, L. Zhu, and L. Wei, “Ultrasensitive Exhaled Breath Sensors Based on Anti-Resonant Hollow Core Fiber with In Situ Grown ZnO-Bi2O3 Nanosheets,” Adv. Mater. Interfaces, early access (2021).

14. F. Gérôme, K. Cook, A. K. George, W. J. Wadsworth, and J. C. Knight, “Delivery of sub-100fs pulses through 8 m of hollow-core fiber using soliton compression,” Opt. Express 15(12), 7126–7131 (2007). [CrossRef]  

15. A. Urich, R. R. J. Maier, F. Yu, J. C. Knight, D. P. Hand, and J. D. Shephard, “Flexible delivery of Er:YAG radiation at 2.94 µm with negative curvature silica glass fibers: a new solution for minimally invasive surgical procedures,” Biomed. Opt. Express 4(2), 193–205 (2013). [CrossRef]  

16. D. E. Couch, H. Dan, S. J. Backus, D. G. Winters, and H. C. Kapteyn, “Ultrafast 1 MHz vacuum-ultraviolet source via highly cascaded harmonic generation in negative-curvature hollow-core fibers,” Optica 7(7), 832–837 (2020). [CrossRef]  

17. J. J. Ackert, D. J. Thomson, L. Shen, A. C. Peacock, P. E. Jessop, G. T. Reed, G. Z. Mashanovich, and A. P. Knights, “High-speed detection at two micrometres with monolithic silicon photodiodes,” Nat. Photonics 9(6), 393–396 (2015). [CrossRef]  

18. H. Sakr, Y. Hong, T. D. Bradley, G. T. Jasion, J. R. Hayes, H. Kim, I. A. Davidson, E. N. Fokoua, Y. Chen, K. R. H. Bottrill, N. Taengnoi, N. V. Wheeler, P. Petropoulos, D. J. Richardson, and F. Poletti, “Interband Short Reach Data Transmission in Ultrawide Bandwidth Hollow Core Fiber,” J. Lightwave Technol. 38(1), 159–165 (2020). [CrossRef]  

19. Y. Hong, H. Sakr, N. Taengnoi, K. R. H. Bottrill, T. D. Bradley, J. R. Hayes, G. T. Jasion, H. Kim, N. K. Thipparapu, Y. Wang, A. A. Umnikov, J. K. Sahu, F. Poletti, P. Petropoulos, and D. J. Richardson, “Multi-Band Direct-Detection Transmission Over an Ultrawide Bandwidth Hollow-Core NANF,” J. Lightwave Technol. 38(10), 2849–2857 (2020). [CrossRef]  

20. W. H. Shen, J. B. Du, L. Sun, C. Wang, Y. Z. Zhu, K. Xu, B. L. Chen, and Z. Y. He, “Low-Latency and High-Speed Hollow-Core Fiber Optical Interconnection at 2-Micron Waveband,” J. Lightwave Technol. 38(15), 3874–3882 (2020). [CrossRef]  

21. F. Poletti, “Nested antiresonant nodeless hollow core fiber,” Opt. Express 22(20), 23807–23828 (2014). [CrossRef]  

22. S. D. Hart, G. R. Maskaly, B. Temelkuran, P. H. Prideaux, J. D. Joannopoulos, and Y. Fink, “External reflection from omnidirectional dielectric mirror fibers,” Science 296(5567), 510–513 (2002). [CrossRef]  

23. C. Wei, R. Joseph Weiblen, C. R. Menyuk, and J. Hu, “Negative curvature fibers,” Adv. Opt. Photonics 9(3), 504–561 (2017). [CrossRef]  

24. W. Belardi and J. C. Knight, “Effect of core boundary curvature on the confinement losses of hollow antiresonant fibers,” Opt. Express 21(19), 21912–21917 (2013). [CrossRef]  

25. G. K. Alagashev, A. D. Pryamikov, A. F. Kosolapov, A. N. Kolyadin, A. Y. Lukovkin, and A. S. Biriukov, “Impact of geometrical parameters on the optical properties of negative curvature hollow-core fibers,” Laser Phys. 25(5), 055101 (2015). [CrossRef]  

26. M. S. Habib, O. Bang, and M. Bache, “Low-loss single-mode hollow-core fiber with anisotropic anti-resonant elements,” Opt. Express 24(8), 8429–8436 (2016). [CrossRef]  

27. P. Roberts, F. Couny, H. Sabert, B. Mangan, D. Williams, L. Farr, M. Mason, A. Tomlinson, T. Birks, J. Knight, and P. S. J. Russell, “Ultimate low loss of hollow-core photonic crystal fibres,” Opt. Express 13(1), 236–244 (2005). [CrossRef]  

28. J. Jackle and K. Kawasaki, “Intrinsic Roughness of Glass Surfaces,” J. Phys.: Condens. Matter 7(23), 4351–4358 (1995). [CrossRef]  

29. B. J. Mangan, L Farr, A Langford, P. J. Roberts, D. P. Williams, F. Couny, M. Lawman, M. Mason, S. Coupland, and R. Flea, “Low loss (1.7 dB/km) hollow core photonic bandgap fiber,” in Proc. Optical Fiber Communication Conference (OFC)2014, Paper PDP24.

30. B. Debord, A. Amsanpally, M. Chafer, A. Baz, M. Maurel, J. M. Blondy, E. Hugonnot, F. Scol, L. Vincetti, F. Gérôme, and F. Benabid, “Ultralow transmission loss in inhibited-coupling guiding hollow fibers,” Optica 4(2), 209–217 (2017). [CrossRef]  

31. S. F. Gao, Y. Y. Wang, W. Ding, D. L. Jiang, S. Gu, X. Zhang, and P. Wang, “Hollow-core conjoined-tube negative-curvature fibre with ultralow loss,” Nat. Commun. 9(1), 2828 (2018). [CrossRef]  

32. T. D. Bradley, J. R. Hayes, Y. Chen, G. T. Jasion, S. R. Sandoghchi, R. Slavik, E. N. Fokoua, S. Bawn, H. Sakr, I. A. Davidson, A. Taranta, J. P. Thomas, M. N. Petrovich, D. J. Richardson, and F. Poletti, “Record Low-Loss 1.3 dB/km Data Transmitting Antiresonant Hollow Core Fibre,” in 2018 European Conference on Optical Communication (ECOC) (IEEE, 2018), pp. 1–3.

33. T. Bradley, G. Jasion, J. Hayes, Y. Chen, and F. Poletti, “Antiresonant hollow core fibre with 0.65 dB/km attenuation across the C and L telecommunication bands,” in Proc. European Conference on Optical Communication (ECOC)2019, Paper PD3.1.

34. G. T. Jasion, T. D. Bradley, K. Harrington, H. Sakr, Y. Chen, E. N. Fokoua, I. A. Davidson, A. Taranta, J. R. Hayes, and D. J. Richardson, “Hollow Core NANF with 0.28 dB/km Attenuation in the C and L Bands,” in Proc. Optical Fiber Communication Conference (OFC)2020.

35. Z. Wang, J. J. Tu, Z. Y. Liu, C. Y. Yu, and C. Lu, “Design of weakly coupled two-mode hollow-core antiresonant fiber with low loss,” J. Lightwave Technol. 38(4), 864–874 (2020). [CrossRef]  

36. G. T. Jasion, D. J. Richardson, and F. Poletti, “Novel Antiresonant Hollow Core Fiber Design with Ultralow Leakage Loss Using Transverse Power Flow Analysis,” in Optical Fiber Communication Conference (2019).

37. X. Chen, X. Hu, L. Yang, J. Peng, H. Li, N. Dai, and J. Li, “Double negative curvature anti-resonance hollow core fiber,” Opt. Express 27(14), 19548–19554 (2019). [CrossRef]  

38. M. S. Habib, J. Antonio-Lopez, C. Markos, A. Schülzgen, and R. Amezcua-Correa, “Single-mode, low loss hollow-core anti-resonant fiber designs,” Opt. Express 27(4), 3824–3836 (2019). [CrossRef]  

39. K. Shaha, A. Khaleque, and M. I. Hasan, “Low loss double cladding nested hollow core antiresonant fiber,” OSA Continuum 3(9), 2512–2524 (2020). [CrossRef]  

40. L. Vincetti and V. Setti, “Extra loss due to Fano resonances in inhibited coupling fibers based on a lattice of tubes,” Opt. Express 20(13), 14350–14361 (2012). [CrossRef]  

41. A. N. Kolyadin, A. F. Kosolapov, A. D. Pryamikov, A. S. Biriukov, V. G. Plotnichenko, and E. M. Dianov, “Light transmission in negative curvature hollow core fiber in extremely high material loss region,” Opt. Express 21(8), 9514–9519 (2013). [CrossRef]  

42. W. Belardi and J. C. Knight, “Hollow antiresonant fibers with low bending loss,” Opt. Express 22(8), 10091–10096 (2014). [CrossRef]  

43. E. N. Fokoua, Y. Chen, D. J. Richardson, and F. Poletti, “Microbending effects in hollow-core photonic bandgap fibers,” in ECOC 2016; 42nd European Conference on Optical Communication, (VDE, 2016), pp. 1–3.

44. M. F. Limonov, M. V. Rybin, A. N. Poddubny, and Y. S. Kivshar, “Fano resonances in photonics,” Nat. Photonics 11(9), 543–554 (2017). [CrossRef]  

45. A. D. Fitt, K. Furusawa, T. M. Monro, C. P. Please, and D. J. Richardson, “The mathematical modelling of capillary drawing for holey fibre manufacture,” J. Eng. Math. 43(2/4), 201–227 (2002). [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 (7)

Fig. 1.
Fig. 1. Cross-section of proposed nested U-shape ARF (UARF) with geometrical parameters.
Fig. 2.
Fig. 2. Calculated power density at the cladding (red-mark of structure cross-section) of NANF and the proposed UARF. The insets show the fiber structures and profiles of fundamental mode. For those two fibers, the core radius, the glass thickness, the tube gap, and the tube radius are chosen the same as 20 µm, 0.5 µm, 3.3 µm, and 16.2 µm, respectively. For the NANF, the radial separation is optimized to 15.45 µm with the radius of the small nested tube of 8.1 µm. For our proposed fiber, the corresponding one is optimized to 7 µm with the radius of the small semicircular of 5.83 µm.
Fig. 3.
Fig. 3. Calculated the losses of UARF and NANF. These scattered points are total loss of corresponding fiber. The dot-dashed line is SSL of UARF, almost identical to that of NANF.
Fig. 4.
Fig. 4. Calculated confinement loss variation with (a) semicircular radius r under the condition of z = 0.42R, and (b) radial separation z under the condition of r = 0.5R. The tube radius R = 24 µm for 5-tube (red lines) and R = 16.2 µm for 6-tube (blue lines), and wavelength at 1.55 µm.
Fig. 5.
Fig. 5. Calculated loss spectra of the proposed fiber (Type I and Type II) and other fibers. The corresponding fiber structures and LP01 mode field distributions are shown at the right-side, the core radius, tube gap and glass thickness of all these fibers are same, while the other parameters of D-ARF are the same as [37].
Fig. 6.
Fig. 6. (a) The effective refractive index of LP01 mode (red), LP11 mode (green) and the CM (gray dashed) with a variation of semicircular radius r. (b)The difference of the CM and LP mode with the variation of wavelength. (c) Loss spectra of LP01 mode, LP11 mode and LP21 mode, as well as the HOMER (dot-dashed line in right axis), mode field distributions of LP01 mode, LP11 mode and LP21 mode are at the right-side. The parameters of proposed fiber (b) and (c) are R = 24 µm, r = 10.32 µm and z = 12 µm.
Fig. 7.
Fig. 7. (a) Calculated confinement loss with the variation of semicircular deformed into an ellipse that defined by a/b, and the insets shows a/b = 0.6 and a/b = 1.4, (b) the loss spectra with a variation of parallel layer deflection θ.

Tables (1)

Tables Icon

Table 1. Comparison between our proposed fiber with other low loss single-mode fibers.

Equations (4)

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

C L = 40 π × n i m a g / [ ln ( 10 ) × λ ] ,
t = λ × ( M 0.5 ) / 2 × n g l a s s 2 n a i r 2 ,
S S L = η F ( λ / λ 0 ) 3 ,
R M L = β 0 2 C ( Δ β 01 ) ( 0 | x 2 | 0 | 0 | x | 1 | 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.